Introduction To FEM
Introduction To FEM
Ursula Kowalsky
Introduction to
Finite Element Methods
Introduction to Finite Element Methods
Dieter Dinkler · Ursula Kowalsky
Introduction to
Finite Element Methods
Dieter Dinkler Ursula Kowalsky
Institut für Statik und Dynamik Institut für Statik und Dynamik
Technische Universität Braunschweig Technische Universität Braunschweig
Braunschweig, Germany Braunschweig, Germany
© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Fachmedien
Wiesbaden GmbH, part of Springer Nature 2024
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher,
whether the whole or part of the material is concerned, specifically the rights of translation,
reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any
other physical way, and transmission or information storage and retrieval, electronic adaptation,
computer software, or by similar or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are
exempt from the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in
this book are believed to be true and accurate at the date of publication. Neither the publisher
nor the authors or the editors give a warranty, expressed or implied, with respect to the material
contained herein or for any errors or omissions that may have been made. The publisher remains
neutral with regard to jurisdictional claims in published maps and institutional affiliations.
This Springer Vieweg imprint is published by the registered company Springer Fachmedien
Wiesbaden GmbH, part of Springer Nature.
The registered company address is: Abraham-Lincoln-Str. 46, 65189 Wiesbaden, Germany
The textbook presented here gives an introduction to the Finite Element Me-
thod from an engineering point of view and offers various insights into the
background and the details of the procedures to get finite elements of different
characteristics, types of discretization, and the related methodological approa-
ches. It is not the intention to present the most efficient types of elements,
because the efficiency of elements depends on the criteria of evaluation, the
kind of programming, and very often on their applications.
The manuscript of this textbook has gradually evolved from lecture notes pre-
sented by the authors at the Technical University Braunschweig and Stuttgart
University to students from the engineering faculties. Thus the textbook is ad-
dressed to students and engineers, who are dealing with structural mechanics
and are interested to develop elements for their own applications. The authors
like to remind of Hermann Ahrens and to gratefully acknowledge his longtime
commitment in lecturing the fundamentals of the method and his support to
develop this kind of presentation of the Finite Element Method.
Finally we gratefully acknowledge Springer Vieweg publishers for the possibility
to publish this monograph and for the great support until going to press.
Glossary
Denomination of Elements
FOUNDATIONS 1
1 Introduction 3
1.1 Governing Equations and Approximate Solution . . . . . . . . . . . . . . . . . . . 7
1.2 General Aspects Concerning the Finite Element Method . . . . . . . . . . . 8
1.3 A Comparison of Exact Solution and Approximate Solution . . . . . . . . 11
2 Discretization of the Work Equation 19
2.1 Principle of Virtual Displacements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2 Principle of Virtual Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.3 The General Procedure to set up the Element Matrices . . . . . . . . . . . 28
2.4 Shape Functions and Convergence Criteria . . . . . . . . . . . . . . . . . . . . . . . 38
3 Structure and Solution of the System of Equations 51
3.1 Real and Virtual Nodal Displacements . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.2 Equations of Condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.3 Structure of the System of Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.4 Assembly of the Stiffness Matrix of the Entire System . . . . . . . . . . . . 55
3.5 The Storage and Solution of the System of Equations . . . . . . . . . . . . . 58
3.6 The Optimization of the Bandwidth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.7 The Fulfillment of Dirichlet Boundary Conditions . . . . . . . . . . . . . . . . . 61
4 Heat Conduction 63
4.1 Heat Conduction at One–Dimensional Description . . . . . . . . . . . . . . . . 63
4.2 Heat Conduction Regarding Two Spatial Dimensions . . . . . . . . . . . . . . 69
4.3 Example of Use . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
5 Membrane Structures 81
5.1 Rectangular Elements Regarding Plane Stress Situation . . . . . . . . . . . 81
5.2 Rectangular Element Comprising Modified Shear Strains . . . . . . . . . . 96
5.3 Plane Strain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
6 Bending Structures 101
6.1 Element Matrices Regarding Euler–Bernoulli Beams . . . . . . . . . . . . . . . 101
6.2 Kirchhoff Plate Element with 16 DOF . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
6.3 Kirchhoff Plate Element with 12 DOF . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
6.4 The 12 DOF Element Employing a Weak Conformity . . . . . . . . . . . . . 129
XII Table of contents
REFERENCES 427
INDEX 437
FOUNDATIONS
1 Introduction
Basis of analytical and computational studies are models, which usually deal
with partial differential equations or integral equations. The model equations
have to be developed with respect to the level of accuracy, that is of interest.
Engineering models, for which analytical solutions exist, often suffice for sim-
plifying investigations. However, general processes that are to be investigated
may be so complex, that a proper description has to include multiple dimensi-
ons in space with partially curved boundaries as well as the dimension in time,
and should take into account different nonlinearities. Therefore, exact soluti-
ons are only possible for special cases. Hence, a general methodology must be
available to get approximate solutions in order to investigate and to design
engineering structures in a proper way.
The classical Finite Element Method has been developed over decades from 1956
on. The mathematical foundation goes back to Trefftz [98] and Courant [29].
With the development of the first computers the method became of interest
for engineers. First pioneers of the method in the field of engineering were
Clough [26, 99] at UC–Berkeley, Zienkiewicz [105, 106] at Swansea University
and Argyris [4, 5] at Stuttgart University. Further textbooks for students have
been presented by Bathe [12], Hughes [47], Onate [76, 77], and many others.
Applying the idea to the investigation of wood may lead to polygonal elements
with a thin–walled geometry for the solid part and an internal polygonal ele-
ment for the description of fluids, cf. Figure 1-3.
Fig. 1-3 Thin–walled polygonal elements to model wood on the meso–scale [51]
Originally the method is founded on the publications by Cundall [33, 34], who
developed the method to describe the behavior of gravel like solids. The exten-
sion of the method offers the possibility to investigate the micro–structure of
continua even down to the dynamics of molecules.
The discretization of beam like structures with particles of different size and
related sieve line as well as different physical and chemical properties offers the
possibility to investigate the deformation behavior of structures with disconti-
nuities down to the micro–structure, cf. Figure 1-4.
❄
❄
✻
The publications on the development of the Finite Element Method deal with
different approaches as basis of the discretization. The textbook at hand ap-
plies the Principle of virtual Work with respect to virtual displacements and
virtual forces. Due to the variety of elements and applications, which can be
hardly discussed sufficiently in a textbook, this book focuses on the basic me-
thodology of the displacement–based formulation as well as mixed and hybrid
formulations in order to explain the different approaches and the fulfillment
of the governing equations. The method will be presented with respect to civil
engineering structures, since it facilitates clear understanding of the derivation.
Therefore we will limit ourselves to load–carrying and deformation behavior of
one– and two–dimensional structures as well as to heat conduction in one and
two dimensions.
All results of examples, tests, and benchmarks have been computed by original
investigations. Nevertheless, the correctness of numbers cannot be guaranteed.
1.1 Governing Equations and Approximate Solution 7
Both approaches are equivalent and are convertible into each other. The exact
solution fulfills the governing equations in the domain as well as at the boun-
daries and at the intersections between different integration regions. Thus all
other equations, to get an integral description of the problem, are fulfilled by
implication.
The exact solution is also named the strong solution. Concerning problems with
more than one dimension in space and arbitrary boundaries, exact solutions
generally turn out to be impossible.
Thus for the analysis of complex structures, such as plates, membranes or
shells, methods are required to approximately solve the governing equations. If
the modeling equation is equivalent to the governing equations, the identified
solution is an approximation of the exact solution. Hereby the equations applied
to get the solution are still fulfilled exactly. However, the solution itself is only
an approximation of the exact solution. Therefore, it no longer exactly fulfills
the governing equations at the differential element. Hence the approximate
solution is also named the weak solution. To get an approximate solution an
approach for the primary variable is chosen, which depends on a normalized
function describing the course of the variable in space, and a multiplier, which
scales the solution.
In the case of a displacement field u(x) the approach could be
u(x) = φ(x) · û .
8 1 Introduction
The function φ(x) is prescribed and is named the shape function, whereat poly-
nomials are usually chosen. The multiplier û is still unknown and is evaluated
by means of the respective modeling equation. This idea, developed by the
physicist Walter Ritz [88] to fulfill the Principle of Minimum Potential Ener-
gy, is the basis of all modern methods that apply integral work equations or
principles of energy to compute an approximate solution.
As an approximation to the exact solution, one may choose shape functions that
are defined regarding the whole structure. However, a closed–form approach on-
ly makes sense if the stiffnesses are constantly distributed with respect to the
domain and if the geometry of the structure is regular. The solutions concer-
ning more general problems exhibiting arbitrary geometries as well as arbitrary
boundaries, stiffnesses and external actions are limited employing an closed–
form approach.
therefore also the number of DOF may not only result in an improved
approximate solution but also the geometry may be taken into account
in more detail, e.g. if curved boundaries are to be approximated by po-
lygonal representation.
When applying the FEM, the degrees of freedom are evaluated by means of work
equations or principles of energy. Regarding the book at hand, the Principle of
virtual Work is applied because in general, it is also valid for non–conservative
systems. The description of the virtual work can be performed in the following
different ways:
Concerning the first part of the book at hand, the PvD provides the basis for the
development of finite elements, which are commonly accepted in engineering
practice. In addition, it is closely related to the displacement method applied
in structural analysis. Concerning sizing, the displacement–based formulation
has the disadvantage that the stresses are solely evaluated from a subsequent
analysis, whereby the differentiation of the displacement approach yields a less
smooth distribution.
Applying the PvD, the sum of the external work due to external actions and
of the internal work due to internal force variables vanishes at regarding the
whole system. Dividing the complete structure into finite elements allows for
the description of the virtual work of the entire system as the sum of the virtual
work of the single elements.
Regarding problems in practice and due to the relatively high amount of data
which is generated, the Finite Element Method cannot be applied without using
FE–programs. FE–programs vary only slightly at employing different types of
elements, therefore they are transferable to diverse problems with just few
modifications.
Even to analyse arbitrary structures comprising different components, element
formulations must be provided employing different types of elements. Such an
element library could differentiate between:
• the physics of the problem as bending, membrane behavior, heat conduc-
tion, fluid flow etc.;
• the geometry of the element as quadrilaterals, rectangles, triangles, bars,
etc.; and
• the approximate approach to describe the displacements as linear, qua-
dratic or higher order functions.
Currently, in addition to the basic FE–programming systems, separate codes
are common concerning pre–processing, what deals with the data generation
and control, as well as post–processing, which is neccessary for the visual re-
presentation of the solution and for the interpretation of results.
1.3 A Comparison of Exact Solution and Approximate Solution 11
H
case 1 units :
EA
x, u, l [m]
p(x) p(x) [ N/m ]
case 2 H [N]
EA
x,u EA [N]
l
a) Kinematics
ε = u,x in domain and
(1.1)
u(0) = 0 at the boundary,
b) Equilibrium
N,x = 0 case 1 in domain,
N,x = −p (x) case 2 in domain and
(1.2)
N (ℓ) = H case 1 at the boundary,
N (ℓ) = 0 case 2 at the boundary.
N = EA · ε in domain, (1.3)
12 1 Introduction
p x2
u(x) = · (ℓ · x − ) ,
EA 2
p
ε(x) = · (ℓ − x) ,
EA
N (x) = p · (ℓ − x) .
u(x) = φ(x) · û .
Here, û is the unknown multiplier and φ(x) is the given shape function, for
which a linear run is feasible, see Figure 1-6.
x,u
Φ(x)
u(x) = φ(x) · û ,
φ(x) = x/l ,
H ·l
û = .
EA
In this particular case the approximate solution regarding u(x) corresponds
to the exact solution. ε(x) is evaluated from a subsequent analysis following
Equation (1.1), and ε(x) equals the exact solution, too:
H
ε = u,x = φ(x),x · û = 1/l · û = .
EA
The evaluation of the longitudinal force following Equation (1.3)
H
N = EA · ε = EA · =H
EA
also yields the exact solution. Introducing the longitudinal force N (x) into the
equilibrium condition (1.2) and controlling the boundary condition confirms
the PvD to provide the analytically exact solution in this particular case.
p . l2 exact solution
u
2 EA approximation
x
Fig. 1-7 Bar – approximation and exact solution regarding the displacement
1 p·ℓ
ε = u,x = φ(x),x · û = · û = .
ℓ 2EA
u(x) being an approximation results in ε(x) being an approximation, too. Only
the mean value of the strains equals the strong solution, see Figure 1-8.
p.l
ε EA p.l exact solution
2 EA approximation
x
Fig. 1-8 Bar – approximation and exact solution regarding the strains
The evaluation of the longitudinal force, following the material Equation (1.3),
also yields an approximate solution
p·ℓ
N = EA · ε = .
2
1.3 A Comparison of Exact Solution and Approximate Solution 15
Here, only the mean value of the approximation equals the exact solution, too,
see Figure 1-9.
Fig. 1-9 Bar – approximation and exact solution regarding the longitudinal force
Although, in general, the courses of the state variables are only approximati-
ons of the strong solution, parts of the governing equations are exactly fulfilled.
The kinematics following Equation (1.1) as well as the material behavior follo-
wing Equation (1.3) are implicitly enclosed in the work equation and therefore
are exactly fulfilled. The equilibrium conditions (1.2) are not exactly fulfilled,
because the PvD is only a weak formulation of the equilibrium. Since the longi-
tudinal forces are constant, inside the domain the equilibrium condition yields
N,x + p(x) 6= 0 .
|{z}
0
employing a linear approach and solely one element, the solution dealing with
two degrees of freedom better approximates the exact solution but needs further
discussion.
Φ1(x)
Φ2(x)
x,u
0 1 2
2
3 p.l
8 EA
p . l2 exact solution
u
2 EA approximation
Fig. 1-11 Bar – approximation and exact solution regarding the displacements
Performing the computation of the other state variables, inside the elements the
kinematic conditions (1.1) and the material equation (1.3) are exactly satisfied
again. This yields constant strains ε(x) as well as longitudinal forces N (x),
whereat the analytical and the approximate solutions turn out to be identical
at the midpoints of the elements, cf. Figure 1-12. Since the shape functions do
not be differentiable, strains and stresses offer a jump at node 1.
3p . l
p.l
N 4 exact solution
p.l
approximation
4
Fig. 1-12 Bar – approximation and exact solution of the longitudinal force
1.3 A Comparison of Exact Solution and Approximate Solution 17
N,x + p(x) 6= 0 ,
|{z}
0
p·ℓ
N (ℓ) = 6= 0 .
4
The solution also offends against the intersection condition concerning the lon-
gitudinal force at node 1, where the jump is an indication of the error as well.
Remarks
The comparison of the exact solution and the approximation leads to the fol-
lowing fundamental statements regarding the PvD.
The main idea of FEM is to subdivide the whole structure into finite elements.
Thereby it is possible to describe the virtual work of the entire structure as
the sum of the virtual works performed at the single finite elements. This is
required to develop the system of equations employing the degrees of freedom
of the approach as unknowns,
X
δW = δWel = 0 . (2.1)
This yields a two–part task to be worked on:
This section systematically describes the formal procedure at the element level.
At first, the Principle of virtual Displacements (PvD) will be proven to be an
equivalent formulation of the equilibrium conditions with regards to a bar, as
an example. It is the basis for further derivations. As a second approach, the
Principle of virtual Forces (PvF) will be applied to the bar in order to explain
the equivalence of the principle to the kinematic conditions.
The static case is described here, where deformations and thus also stresses and
strains slowly increase simultaneously with loads from zero to the final value.
The bar is used to clearly and concisely represent the different parts of the
work previously defined. Here, the work theorem according to Equation (2.2)
is applied to a bar loaded by an external action p(x) as well as by an external
boundary action H(ℓ) as shown in Figure 2-1. Due to the real loadings p(x)
and H(ℓ), the deformation state characterized by u(x) and ε(x) arises as do
the longitudinal forces N (x).
p(x)
H
x
l
δu(x) u
u(x)
δε(x) ε
ε(x)
Regarding this overall state being achieved in two steps, the work theorem also
has to be fulfilled:
Z Z
X 1 1 1
W =− N · ε dx + p · u dx + [ H · u ]x=ℓ
2Z 2 2
1 1
− δN · δε dx + [ δP · δu ]x=ℓ
2Z Z 2
− N · δε dx + p · δu dx + [ H · δu ]x=ℓ = 0 . (2.4)
2.1 Principle of Virtual Displacements 21
Referring to Equation (2.4), the 1/2 – coefficient marks the active work of the
basic state regarded at first, and the additional virtual state added afterwards.
Both parts of the work must be zero individually because of Equation (2.3).
Thus the remaining virtual passive work, which the real stress state performs
on the virtual deformation added subsequently, must be zero, too:
Z Z
δW = − N · δε dx + p · δu dx + [ H · δu ]l = 0 . (2.5)
Figure 2-2 illustrates the meaning of the aforementioned parts of the work
Equation (2.4), assuming a linear correlation between the forces and the re-
lated deformations. The area determined by the straight and the x–axis can
be interpreted as the active work. The horizontally–lined triangles mark the
active work of the basis state as well as the active work of the virtual state, the
crosshatched rectangle marks the passive work according to Equation (2.5).
N δN
δε
Referring to Equation (2.5) the virtual state must only fulfill kinematics. Thus,
the virtual external action and the related virtual stress state, primarily in-
troduced to enable a better understanding of the basic principle, both are not
required any longer. Therefore, since only the virtual deformation state is re-
quired, this can be prescribed independently of any virtual load.
The following may explain which assumptions ensure the Principle of virtual
Displacements according to Equation (2.5) being equivalent to the equilibrium
condition. Assuming that kinematics according to Equation (1.1) also apply to
the virtual state
δε = δu,x in domain,
(2.6)
δu(0) = 0 at boundary,
Now, all governing equations as well as all boundary conditions are taken into
account, and the work equation describes the real problem completely and with-
out any approximation. Since the work is solely described by real and virtual
displacements, it is named displacement–based formulation. For the purpose of
practical analyses it is convenient to multiply the equation by −1. Thus, the
PvD according to the work equation (2.9) is applied to further considerations:
Z Z
− δW = u,x · EA · δu,x dx − p · δu dx − [ H · δu ]l = 0 . (2.9)
2.1 Principle of Virtual Displacements 23
However, it must be pointed out that the PvD integrally fulfills the equilibrium
conditions concerning the entire structure. This is also known as a weak form
of equilibrium. Kinematics and the material equations are still exactly fulfilled
at every point of the structure, because these governing equations are taken
into account implicitly by the PvD. Even the boundary conditions are fulfilled
differently. While the Neumann boundary conditions of the real state, being
equilibrium conditions, are fulfilled integrally by the work equation and named
as natural boundary conditions, the Dirichlet boundary conditions of the real
and of the virtual state must, in contrast, be fulfilled exactly by the displace-
ment field. Therefore, they are named as substantial boundary conditions.
However, the virtual work regarding the entire structure according to Equation
(2.9) is represented by a scalar quantity, which can also be described as the sum
of the works performed at single integration domains, which are the elements:
X
δW = δWel = 0 . (2.10)
Figure 2-4 applies the method of sections to the elements considered, in order
to visualize the internal forces and displacements at the respective nodes.
N A 1 B N N B 2 C
A B B H
p p
A 1 B B 2 C
Following Equation (2.9), the work equation that applies the PvD to every
single element is comprised of both internal and external work inside the domain
as well as work performed at the element boundaries and yields
Z Z
−δWel = δu,x · EA · u,x dx − δu · p(x) dx − [ δWel−bound. ] (2.11)
with the work performed at the element intersections, cf. Figure 2-4,
When adding up the elements’ work to the system’s work, the work at the
elements’ intersections is also added up. If the shape functions regarding the
displacements are continuous at the element intersections, the following applies:
whereby the respective terms of internal work at the element intersections inside
the structure compensate exactly for each other and can be ignored at the
element level.
In general the polynomial order of the shape functions regarding the displace-
ments results in discontinuous longitudinal forces at the elements’ intersections
and therefore Equation (2.14) might not be fulfilled exactly. However, if the
work is disregarded at element intersections, Equation (2.14) is fulfilled impli-
citly in its weak form by applying the PvD to the entire system in the same
way as all the other equilibrium conditions.
Here again, the work theorem is applied to the bar shown in Figure 2-5. The bar
is loaded by a distributed external action p(x) and a prescribed displacement
ue at the boundary.
ue
p(x)
δP
P
x
l δN = δP
Fig. 2-5 A bar loaded by a distributed external action and virtual load
The deformations are generated by the forces itselves, therefore a 1/2 – coeffi-
cient in Equation (2.15) marks the active work, see also Figure 2-6.
Virtual Work
In addition to the virtual action, the system is loaded by the real external
action p(x), ue now. This yields a total deformation δu(x) + u(x) associated
with the strains δε(x) + ε(x) as well as with the normal forces δN (x) + N (x).
Regarding this overall state being achieved in two steps, the work theorem also
has to be fulfilled:
Z
X 1 1
W =− δN · δε dx + [ δP · δu ]x=ℓ
2Z 2Z
1 1 1
− N · ε dx + p · u dx + [ N · ue ]x=0
2
Z 2 2
− δN · ε dx + [ δP · u ]x=ℓ + [ δN · ue ]x=0 = 0 . (2.16)
Referring to Equation (2.16), the 1/2 – coefficient marks the active work of the
virtual state regarded at first, and the additional real state added afterwards.
Both parts of the work must be zero individually because of Equation (2.15).
Thus the remaining virtual passive work, which the virtual stress state performs
on the real deformation added subsequently, must be zero, too:
Z
δW = − δN · ε dx + [ δP · u ]x=ℓ + [ δN · ue ]x=0 = 0 . (2.17)
Figure 2-6 illustrates the meaning of the aforementioned parts of the work
Equation (2.16), assuming a linear correlation between the forces and the re-
lated deformations. The area determined by the straight and the x–axis can
be interpreted as the active work of the overall state. The horizontally–lined
triangles mark the active work of the virtual state as well as the active work of
the real state, the crosshatched rectangle marks the passive work according to
Equation (2.17).
δP P
δu
Referring to Equation (2.17) the virtual state must only fulfill the equlibrium
conditions. Thus, the virtual deformations and strains, primarily introduced
to enable a better understanding of the basic principle, both are not required
any longer. Therefore, since only the virtual stress state is required, this can
be prescribed independently of any virtual deformation.
Applying kinematics
u,x − ε = 0 in domain,
(2.18)
u − ue = 0 at boundary x = 0 ,
in a weak formulation Equations (2.18) turn to
Z
δW = δN · ( u,x − ε ) dx − [ δN · ( u − ue ) ]x=0 = 0 . (2.19)
Referring to Equation (2.19) kinematics related to the real state is given with-
in the brackets of the integral. The boundary condition concerning the real
displacement u − ue = 0 at position x = 0 is considered in a weak sense as well.
At x = 0, the boundary terms eliminate each other except for the prescribed
displacement. At x = ℓ, the remaining bondary term belongs to the virtual
load δN u|ℓ = δP u|ℓ . Furthermore, if the virtual state fulfills the equilibrium
condition δN,x = 0,
Z
δW = − δN · ε dx + [ δN · ue ]x=0 + [ δP · u]x=ℓ = 0 (2.20)
yields the final statement. Comparing Equation (2.17) and Equation (2.20) it
becomes clear, that the Principle of virtual Forces is a weak formulation of
kinematics and thus may be applied to develop finite elements analogously to
the Principle of virtual Displacements. Thus, the following sentence applies:
The Principle of virtual Forces according to Equation (2.17) is equi-
valent to kinematics (2.18), if the virtual state fulfills the equilibri-
um conditions (1.2).
28 2 Discretization of the Work Equation
The benefit of the FEM is that every single element’s work can be described
applying a pattern that is identical for all elements. In the following, all indi-
vidual procedural steps are represented in detail up to the discretization of the
work equation, as well as the evaluation of the internal force variables.
After the choice of shape functions and the integration of the work equation
concerning the element domain, one gets the discrete form of the work equation,
which depends on the still unknown nodal displacements. The discretization of
the work equation takes place in three steps:
1. set up the work equation regarding a single element applying a displace-
ment–based formulation,
2. choose shape functions regarding the displacements, and
3. introduce the shape functions into the work equation and perform inte-
gration.
Furthermore the FE–procedure comprises:
4. sum up all elements’ work to get a description of the entire system and
computation of the nodal displacements as presented in Section 3, and
5. subsequent stress analysis to evaluate the internal force variables element
by element.
The general procedure is shown by investigating a bar e.g. and is transferred
to other problems as heat conduction, membranes, beams and plates in the
following sections.
2.3 The General Procedure to set up the Element Matrices 29
has to be taken into account, whereby this notation equals the work equation
represented at system level. It must be pointed out that the element’s work
given in Equation (2.21) does not include the work performed at the element
intersections. Therefore, the work equation at element level cannot be directly
used to evaluate the displacements.
boundary conditions may be met more easily. The following will show how
to transfer the general polynomial approach given in Equation (2.22) into the
descriptive approach given in Equation (2.23), cf. Figure 2-7.
uB
u(x)
uA
ΦA uA ΦB uB
= +
1 1
A B x
x ΦA = 1− xl ΦB = l
l l l
Fig. 2-7 Scaling of the approach with respect to the nodal displacements
x = 0 : u(0) = a0 + a1 · 0 = uA ,
(2.24)
x = ℓ : u(ℓ) = a0 + a1 · ℓ = uB
a0 = u A ,
a1 = (uB − uA ) /ℓ .
u(x) = uA + (uB − uA ) · x /ℓ ,
Since the nodal values are independent of the integration, the element’s total
work can be represented in a matrix notation:
− δWel = − δWA − δWB
Z " # " #
φA,x EA · φA,x φA,x EA · φB,x uA
= [ δuA δuB ] · { dx ·
φB,x EA · φA,x φB,x EA · φB,x uB
Z " #
φA · p(x)
− dx } . (2.27)
φB · p(x)
The element matrix is symmetrical only if the shape functions regarding the
real and the virtual displacements are identical.
We may start with
p(x) =p = const. ,
φA (x),x = − 1/ℓ = const. ,
φB (x),x = + 1/ℓ = const. .
32 2 Discretization of the Work Equation
The integrated representation of the element matrix equals the stiffness matrix
of the bar, as it has been established by the displacement method developed
from structural analysis. Therefore the FEM’s element matrix may also be
named as element stiffness matrix.
The entries of the element matrix illustrate the nodal forces, which evolve from
the real nodal displacements and which perform work on their corresponding
virtual nodal displacements.
Every single sum of a row represents the work of the element referring to a rigid
body motion performed with respect to the corresponding virtual displacement
state, wherein the nodal displacements uA and uB are the same size. Due to
the fact that a rigid body motion must not produce any forces, the work and
therefore every sum of a row regarding the element stiffness matrix is equal to
zero. Just as the sum of a column describes the nodal forces’ work performed
with respect to a virtual rigid body displacement.
ε = u(x),x
= φA (x),x · uA + φB (x),x · uB
1 1
= − · uA + · uB ,
ℓ ℓ
2.3 The General Procedure to set up the Element Matrices 33
2.3.2 Example
Section 1.3 gives an approximate solution regarding the example referred to in
Figure 1-5 and Figure 2-8, which is also compared to the analytical solution.
H
case 1
EA
p(x)
case 2
EA
x,u
l
[ uB ] = [ Hℓ/EA ] ,
[ N ] = [ H ].
Since the nodal degrees of freedom v as well as the virtual nodal values δv are
independent of x, they can be excluded from the integral of work
Z Z
− δWel = δvT · { [ ΩT · DT ] · E · [ D · Ω ] dx · v − ΩT · p dx } . (2.34)
B= D·Ω , BT = ( D · Ω )T = ΩT · DT
can be introduced to get a short form describing the field of strains. This yields
Z Z
− δWel = δvT · { BT · E · B dx · v − ΩT · p dx }
| {z } | {z }
K f
The matrix notation of the virtual work offers the big advantage of being ap-
plicable to all physical problems. Only the contents of the respective matrices
and vectors as well as the integration region must be tentatively adjusted to
the particular problem being investigated.
− δWelement = δvT { K v − f } .
2.3 The General Procedure to set up the Element Matrices 37
After adding up the works regarding all the elements to get the system’s work,
and applying the same matrix symbols at the system level, the work equation
follows to
− δWsystem = δvT { K v − f } = 0 .
Here, K is the stiffness matrix of the system, f is the load vector of the system
and v is the displacement vector of the system. The procedure of adding up
the work with regard to all elements is explained in more detail in Section 3.
According to common practice with respect to the literature, no difference is
made between element matrices and system matrices or between the respective
vectors. In the following vectors will be characterized by small letters, matrices
by capital letters.
In addition to the elements’ works, the boundary conditions as well as the
external boundary actions must be taken into account, but may be considered
at the system’s level afterwards. This is advantegeous, because it does not
perturbe the element by element procedure to get the system’s stiffness matrix
and the load vector.
After assembling the system of equations to compute the system’s displacement
vector, v may be evaluated numerically by means of a standard equation solver,
cf. Section 3.
σ = S·v, (2.37)
If special values of stresses are of interest the coordinates of the related positions
have to be taken into account for the computation of the stress matrix S̃. Thus
σ̃ = S̃ · v (2.38)
no longer describes the polynomial approach of the stresses, but the specified
values.
1. General Polynomials
The most simplified representation of the displacement field consists of power
terms of x, which are named monomials. Thus the complete approach is given
by a general polynomial of specific order, e.g. a quadratic approximation of the
displacement u(x) is given in Figure 2-9.
A B
u(x) = a0 x0 + a1 x1 + a2 x2
l
x0 1 Φ 1(x)
Φ 2 (x)
x1 l
l2 Φ 3 (x)
2
x
2. Legendre Polynomials
Another type of approximation without any physical meaning of the scaling
factors deals with Legendre Polynomials. Legendre Polynomials are shape func-
tions employing the special feature
Z
1 2 1 für i = j
φi (x) · φj (x) dx = · δij , δij = { ,
ℓ 2i + 1 0 für i 6= j
which means, that they are orthogonal with respect to the integral. This leads
to numerical advantages, when the virtual work equation has to be integrated.
These polynomials, however, do not exhibit any physically meaningful degree of
freedom. In this case the transformation of the scaling factors to scaling factors
with a physical meaning can be done after the integration of the virtual work.
40 2 Discretization of the Work Equation
A B
u(x) = a0 · φ0 + a1 · φ1 + a2 · φ2 + . . . ,
x
l l
φ0 (x) = 1 Φ 1(x)
1
1
x Φ 2 (x)
φ1 (x) =
ℓ
Φ 3 (x)
1
1 x2
φ2 (x) = ( 3 2 − 1)
2 ℓ
1
2
Fig. 2-10 Legendre Polynomials
3. Lagrange Polynomials
If shape functions are chosen that employ nodal displacements at designated
coordinates as scaling factors, so–called Lagrange Polynomials are obtained. In
Figure 2-11 quadratic Lagrange Polynomials are depicted.
A C B
u(x) = uA φ1 + uB φ2 + uC φ3 l l
2 2
x x2
φ1 (x) = 1 − 3 +2 2 Φ 1 (x)
ℓ ℓ 1
x x2 1
φ2 (x) = − +2 2 Φ 2 (x)
ℓ ℓ
x x2 1 Φ 3 (x)
φ3 (x) = +4 −4 2
ℓ ℓ
The three scaling factors of the quadratic approach of the displacement field are
the displacements uA , uB , uC at the element nodes A, B, C. The shape functions
φ1 , φ2 , φ3 no longer keep single power terms, but functions, that describe the
influence of the related nodal displacement within the element. Thus, Lagrange
Polynomials are functions, which comprise the value of 1 in the node, that
keeps the unknown displacement, and a value of 0 in all other nodes.
4. Hermite Polynomials
Hermite Polynomials employ the nodal displacements as well as their higher
derivatives at the domain’s boundaries and further nodes inside the domain as
scaling factors. They are of advantage, if the intersection conditions between ele-
ments are described by means of higher derivatives of the displacement course.
Applying the cubic Hermite Polynomials according to Figure 2-12, this means:
u(x) = uA φ1 + u,x A φ2 + uB φ3 + u,x B φ4 .
A B
x
l
2 3
x x
φ1 (x) = 1 − 3 +2 3
ℓ2 ℓ 1 Φ1 (x)
x2 x3
φ2 (x) = x − 2 + 2 1 Φ2 (x)
ℓ ℓ
x2 x3
φ3 (x) = 3 −2 3 1 Φ3 (x)
ℓ 2 ℓ
x2 x3
φ4 (x) = − + 2 1 Φ4 (x)
ℓ ℓ
Criterion of Conformity
The criterion of conformity requires continuity and moreover continuous diffe-
rentiability of the shape functions within the element up to the strains, which
are required to have a continuous course inside the element’s domain. Otherwi-
se additional work must be taken into account when applying the PvD. At the
element’s intersections the following conditions must be fulfilled and the same
is valid for the system’s boundary conditions:
2.4 Shape Functions and Convergence Criteria 43
There might be some elements which cannot fulfill all element intersection con-
ditions a priori without increasing the polynomial order of the shape functions
considerably. Thus, when applying the work equation, it might be more mea-
ningful to explicitly consider the work performed by the internal force variables
at the element intersections.
Since such displacement states may occur at any area inside a structure, de-
pending on the situation, the shape functions must fulfil this criterion right
from the start and must be able to describe these displacement fields a priori.
displacement
w and rotation
The demand from the strains and curvatures not to occur in the case of a
rigid–body displacement is generally fulfilled regarding plane structures such
2.4 Shape Functions and Convergence Criteria 45
In contrast, for curved structures like circular arcs and shell structures the
rigid body criterion is hardly to fulfil, because of the special kinematic relations
applying to curved structures. Regarding circular arcs, the strain ε additively
follows from the first derivative of u with respect to the curved coordinate s and
from the zero order derivative of w. If the same shape functions, for example
linear courses, are employed to describe both displacements, the condition of
rigid–body displacement without occuring strains cannot be fulfilled at every
single point, see Figure 2-15. Instead a linear strain and thus a linear stress
field occurs.
exact solution linear shape functions for u and w
πs s s
w(s) = wB · sin , w(s) = (1 − ) · wA + · wB ,
2ℓ ℓ ℓ
πs s s
u(s) = uA · cos , u(s) = (1 − ) · uA + · uB ,
2ℓ ℓ ℓ
w 1 1 s
ε = u,s + = 0. ε = − · uA + · · wB 6= 0 .
R ℓ R ℓ
A uA A*
s wA = 0
w
uB = 0
u R
s=l wB wB = uA
B B*
For example:
• u(x, y) = a0 + a1 · x + a2 · y
is a complete linear approach and
• u(x, y) = a0 + a1 x + a2 y + a3 x2 + a4 xy + a5 y 2
is a complete quadratic approach.
1
x y
x2 xy y2
x3 x2y xy 2 y3
x4 x3y x2y 2 xy 3 y4
Triangular elements need complete approaches, because they should have in-
variant properties, since they may have different positions in the x–y–plane.
Furthermore, complete approaches match the number of possible nodes inside
triangles, because three unknowns are correlated to the three corners and six
unknowns are correlated to the corners and the midpoints of the edges.
They are invariant against rotation of the coordinate system by 90o and multi-
ple thereof. Regarding two–dimensional plane structures, the approaches may
be taken from the respective one–dimensional structure therefor.
Applying the procedure for each physically meaningful nodal degree of freedom,
the respective nodal coordinates are successively introduced into the general
relation (2.39). This yields a system of equations which is comprised of the
relations between the physically meaningful nodal degrees of freedom v and
the scaling factors a of the general polynomial by numbers:
uA ψ̃ A
uB = ψ̃ B · a
... ...
v = Ψ̃ · a . (2.40)
Thus, the matrix Ψ̃ transforms the scaling factors a to the physically meaning-
ful nodal degrees of freedom v. Reversing Ψ̃ according to Equation (2.40) in a
second step, one obtains
−1
a = Ψ̃ ·v = G·v. (2.41)
Thus the vector of scaling factors a regarding Equation (2.39) may be described
directly as employing the nodal degrees of freedom. It follows that
The matrix G is named as scaling matrix and scales the general polynomial
according to Equation (2.39) with respect to the physically meaningful nodal
degrees of freedom. The product ψ · G now describes the physically meaningful
shape functions Ω.
degrees of freedom
v T = [ uA uB uC ] aT = [ a0 a1 a2 ]
scaling matrix
v = Ψ̃ · a
uA ψ(x = 0) 1 0 0
uB = ψ(x = ℓ) · a = 1 ℓ ℓ2 · a
2
uC ψ(x = ℓ/2) 1 ℓ/2 ℓ /4
−1
a = Ψ̃ ·v = G·v
1 0 0
−1
G = Ψ̃ = −3/ℓ −1/ℓ 4/ℓ
2 2
2/ℓ 2/ℓ −4/ℓ2
shape functions
u=ψ·G·v
1 0 0
Ω = ψ · G = 1 x x2 · −3/ℓ −1/ℓ 4/ℓ
2/ℓ2 2/ℓ2 −4/ℓ2
h 2 2 2
i
= (1 − 3 x + 2 x2 ) (− x + 2 x2 ) (4 x − 4 x2 )
ℓ ℓ ℓ ℓ ℓ ℓ
50 2 Discretization of the Work Equation
It is an essential fact that the integration of the element stiffness matrix and
of the load vector can be simplified employing the scaling matrix and may
be generally more efficiently established. Therefore, integration of the element
stiffness matrix K and the load vector f may be performed by employing the
general polynomial. After the integration procedure, the stiffness matrix as well
as the load vector are multiplied by the scaling matrix, which now can be done
numerically
Z
δvT · K · v = δvT · GT · ψ T · DT · E · D · ψ dA · G · v , (2.43)
Z
δvT · f = δvT · GT · ψ T · p dA . (2.44)
Even if the scaling matrix appears to be a detour from computing the physically
meaningful shape functions at first, it is often more advantageous as well as
numerically more efficient to follow the procedure given in Equations (2.43)
and (2.44), with regard to the programming of the stiffness matrix.
3 Structure and Solution of the System of Equations
1 2 3 elements
1 2 3 4 nodes
u (x)
possible qualitative final solution
1 1 22 2 33 3 4
final solution inside single elements
Fig. 3-1 Shape functions for the course of displacements at a tensile bar
Since the Principle of virtual Work is applied for each element, the virtual dis-
placements are also described for each element separately. As per Section 2.1.3,
kinematics at the element intersections must also be fulfilled by the virtual
displacements. This is only possible, if the virtual nodal displacements of the
elements are the same at the common nodes of neighboring elements, and thus
become virtual nodal displacements of the system. Hence the virtual work at
the common nodes of neighboring elements can be summarised.
Each row of the system of equations comprises the virtual work of the real
stress state on the respective virtual displacement, whereat the virtual nodal
displacements occur as multipliers of the corresponding work equations. Thus
the work equation can also be taken as the condition of equilibrium of the nodal
forces acting in the direction of the virtual nodal displacements.
Accordingly, the conditions of equilibrium do not need to be fulfilled strongly
at all differential elements, but only approximately employing shape functions
and nodal displacements.
i i i
δuA kAA kAB fA
i i
=K = f
δuB i i i
kBA kBB fB
Since the virtual work of every single element is computed independently from
each other, the summation of the virtual work of the elements to the total virtu-
al work of the system has to satisfy the kinematic conditions of the Principle of
54 3 Structure and Solution of the System of Equations
u1 u2 u2 u3 u3 u4
1 1
δ u1 K f
δu 2
2 2
δu 2 K f
δu 3
3 3
δu 3 K f
δu 4
Fig. 3-3 The virtual work without considering the transition conditions
The transition conditions of deformation are fulfilled for the final state at the
element intersections, if the actual displacements ui at the common nodes of
the neighboring elements are the same and therefore considered as the nodal
displacements of the system:
u2 is same for elements 1 and 2,
u3 is same for elements 2 and 3.
The virtual work thus exhibits the scheme shown in Figure 3-4.
u1 u2 u3 u4
1 1
δ u1 K f
δu 2
2 2
δu 2 K f
δu 3
3 3
δu 3 K f
δu 4
Fig. 3-4 Conditions of deformation for the actual nodal displacements are fulfilled
3.4 Assembly of the Stiffness Matrix of the Entire System 55
u1 u2 u3 u4
1
δ u1 K f1
δu 2 K2 f2
δu 3 K3 f3
δu 4
Now the virtual work fulfills all kinematic conditions at the element intersecti-
ons. The columns of the stiffness matrix are to be multiplied by the unknown
real nodal displacements. Each row is to be multiplied by the respective virtual
displacement, whereupon the values of the variables δu1 to δu4 do not influence
the results, since they can be excluded from the related equation.
This information is summarized in the connectivity matrix. For the tensile bar
in Figure 3-1, this yields the connectivity matrix in Table 3.1:
node
element A B
1 1 2
2 2 3
3 3 4
The element stiffness matrix K of a single element may split up into the sub–
matrices klm , cf. Figure 3-6. The sub–matrices of the element matrix corre-
sponding to the nodal unknowns of the element are square matrices of the size
n×n, corresponding to the n real unknowns per node as well as to the n virtual
unknowns. The real and the virtual unknowns may comprise only one nodal
3.4 Assembly of the Stiffness Matrix of the Entire System 57
unknown as in the case of heat conduction, cf. Section 4, two nodal unknowns
as in the case of membranes, cf. Section 5, or four nodal unknowns as in the
case of plates, cf. Section 6.
Element matrix
vA vB vC vD
Fig. 3-6 Correlation of the node–numbers and the element stiffness matrix
1 2 3 4 5 6 7 8 9
1
2 2 2 2
2 k AA k AB k AD k AC
2 2 2 2
3 k BA k BB k BD k BC
4
2 2 2 2 2
K = 5 k DA k DB k DD k DC
2 2 2 2
6 k CA k CB k CD k CC
7
The procedure of building up the left hand side of the entire system of equa-
tions can be directly implemented into a finite element program, if the storage
considers all elements of the connectivity matrix. It can be transferred to the
system’s matrix and to the load vector as shown in Figure 3-8.
1 2 3 4 5 6 7 8 9
1
X X X X 1 X
2
X X X X 2 X
3 3
X X X X 4 X+
4 + + + +
X X X X X
K= 5 + + + + f= 5 +
6 6
7 + + + + 7 +
8 + + + + 8 +
9 9
X element 1 + element 3
element 2 element 4
Fig. 3-8 The assembly of the system matrix and the load vector
BW = 5 · ( nodal unknowns )
z }| {
⊗ ∗ ∗ ∗ ⊗ ∗ ∗ ∗
∗ ⊗ ∗ ∗ ∗ ∗ ⊗ ∗ ∗ ∗ ∗
∗ ⊗ ∗ ∗
−→
⊗ ∗ ∗
∗ ∗ ⊗ ∗ ∗ ∗ Utilisation of ⊗ ∗ ∗ ∗
∗ ∗ ∗ ∗ ⊗ ∗ ∗ ∗ ∗
symmetry and ⊗ ∗ ∗ ∗ ∗
∗ ∗ ∗ ⊗ ∗ bandwidth
∗ ⊗ ∗ ∗
∗ ∗ ⊗ ∗
−→
⊗ ∗
∗ ∗ ∗ ∗ ⊗ ∗ ⊗ ∗
∗ ∗ ∗ ⊗ ⊗
Since a symmetric system of equations does not only reduce the storage ca-
pacity needed but also considerably reduces the computing time, the order of
magnitude of the effort proportionally accounts for n · B W 2 , whereat n is the
number of unknowns of the entire system.
rated into the system matrix as related to their current system node numbers.
Hence, arbitrary allocation of the node numbers results in the system matrix
without any diagonal structure as shown in Figure 3-10–left. Since the compu-
ting time is related to n · BW 2 and the needed storage capacity with respect
to n · BW a renumbering of the nodes is necessary, which may significantly
3.7 The Fulfillment of Dirichlet Boundary Conditions 61
decrease the storage capacity required as well as the computing time needed.
The system matrix presented in Figure 3-10–right particular clearly shows the
effects of optimization of node numbering with regard to the bandwidth.
1. The values of the components of the respective rows and columns are set
to zero which is similar to cancelling out the related rows and columns.
Since this yields a singular matrix, the values on the related main diago-
nals are to be set to ’1’, whereby the related unknown will be formally
computed to zero.
2. A fixed constraint can be considered to be a spring with high stiffness.
The virtual work of the spring’s force can be introduced into the system
matrix by adding the spring stiffness to the value of the respective main
diagonal component.
1. Before the related row and column are replaced by zero values, the related
column is to be multiplied by the prescribed value of the unknown and
is to be subtracted from the right–hand side. Afterwards, the addressed
row and column are set to zero and the element at the main diagonal is
set to be ’1’. The value of the corresponding element in the load vector
is then replaced by the prescribed value.
2. The product of the prescribed displacement value and the spring stiffness
is to be stored in the load vector.
It should be noted that the boundary conditions does not affect the symmetry
and the storage of the system’s matrix either.
4 Heat Conduction
that can be designated as the first governing equation. The negative sign cap-
tures the heat flux from hot to cold. λ [W/m K] or as [J/s m K] is the material
parameter describing the thermal conductivity, cf. Figure 4-1.
z dx/2 dx/2
y T
dy
dT
x dz q (x) x
x
If the amount of heat is not totally transported through the domain, a part of
it may be stored inside the material. The stored energy ψ [ J/m3 ] depends on
the temperature as well as on the material density ρ [kg/m3 ], and on the heat
storage capacity or heat capacity c [W s/kg K] or c [J/kg K] respectively. This
relation leads to the second governing equation to describe the stored heat or
the stored energy
ψ(x, t) = ρ · c · T . (4.2)
At this, c describes the amount of heat that is required to warm up 1 kg of the
material by 1 K. The stored energy depends on the position x and the time t.
q-dq/2 q+dq/2
ψ , qv
0 x
dx
The third governing equation describes the balance of energy, cf. Figure 4-2. If
one part of the heat is transported and one part is stored, the sum of the parts
changed in space and in time needs to be balanced. Considering that the heat
is not only transported from the boundary into the domain but also develops
as qv (x, t) in the domain itself – for example due to chemical reactions as a
result of hydration as the concrete hardens – it applies
ψ̇ + q,x = qv . (4.3)
The balance of energy confirms nothing else but the fact that the sum of heat
storage and heat flux might equal the heat supplied to the system. Introducing
the first and the second governing equation into the balance of energy gives
(ρ · c · T )˙ − (λ · T,x ),x = qv .
4.1 Heat Conduction at One–Dimensional Description 65
In the following, only the stationary situation will be investigated, so the deri-
vatives with respect to the time disappear. Thus it applies
T − Tb = 0 , q − qb = 0 . (4.5)
Here, Tb and qb are the prescribed values at the boundary. In the case of tem-
perature and heat flux being combined at the boundary, the Cauchy boundary
condition needs to be fulfilled
q = −h · (T∞ − T ) (4.6)
with h [W/K m2 ] being the coefficient of heat transfer. This includes the influ-
ence of thermal convection at the boundary, when T∞ is the outside tempe-
rature. Here, T and q are not prescribed but the result of the complete heat
conduction process.
If radiation at the surface of a structure is taken into account, the boundary
condition deals with T 4 and needs algorithms to solve the nonlinearity. Thus,
this type of boundary condition is not discussed here.
− q,x + qv = 0 (4.7)
δT {−q,x + qv } = 0 . (4.8)
66 4 Heat Conduction
After integration over the entire domain, the weak form of the energy balance
equation follows to Z
δT {−q,x + qv } dV = 0 , (4.9)
which can be designated as weighted residual δR. Respective consideration of
the boundary conditions dealing with virtual heat fluxes and the surface S gives
Z Z
δR = δT {−q,x + qv } dV + {[ δq (T − Tb ) ] + [ δT (q − qb ) ]} dS = 0 . (4.10)
Taking into account Fourier’s model of heat conduction and presuming that
the temperature (T − Tb ) and the virtual temperature δT exactly fulfill the
boundary conditions, it follows
Z Z
−δR = {δT,xλ T,x − δT qv } dV + δT qb dS = 0 . (4.12)
This equation fully complies with the Principle of virtual Displacement regar-
ding the bar, thus the weak form may be interpreted as Principle of virtual
Temperature. If a Cauchy boundary condition is to be taken into account, the
weak form is to be extended accordingly:
Z Z
−δR = {δT,x λ T,x − δT qv } dV + δT [ −h(T∞ − T ) ] dS = 0 . (4.13)
The discretization, which is now enabled by applying the Finite Element Me-
thod, happens in complete analogy to the bar. However, here the unit of δR is
[W ] and converting by [W = N m/s] may result in [N m/s] only if the unit of
δT is 1.
T (x) = a0 + a1 · x
A B
x
l
1 TA ( 1 - x / l )
1 TB ( x / l )
Fig. 4-3 Shape functions for the description of the temperature field
ture field as
x x TA
[ T (x) ] = [ ( 1 − ) ( )] · ,
ℓ ℓ TB
u = Ω·v,
68 4 Heat Conduction
δu = Ω · δv .
Now, the integration of the weighted element residual can be performed ana-
logously to Section 2.3.2. Indroducing the shape functions into the weighted
residual gives
Z Z
− δRel = δvT · { [ ΩT · DT ] · E · [ D · Ω ] dx · v − [ ΩT ] · p dx } ,
B= D·Ω , BT = ( D · Ω )T = ΩT · DT
as
Z Z
− δRel = δvT · { BT · E · B dx ·v − ΩT · p dx } .
| {z } | {z }
K f
Employing the linear shape functions the residual can be written with element
matrix K, element heat production vector f, and element vector of nodal tem-
peratures v as
" # " # " #
λ/ℓ −λ/ℓ TA qV ℓ/2
− δRel = [ δTA δTB ] · { · − }.
−λ/ℓ λ/ℓ TB qV ℓ/2
If just one element 0 ≤ x ≤ ℓ is taken into account, the element matrix equals
the system matrix. Considering qV = 0 and Cauchy boundary conditions gives
" # " #
λ/ℓ −λ/ℓ T0
− δRsystem = [ δT0 δT1 ] · { ·
−λ/ℓ λ/ℓ T1
1 hℓ x hℓ x
T (x) = {[1 + (1 − )] T∞0 + [1 + · ] T∞1 } .
2 + h ℓ/λ λ ℓ λ ℓ
y x dq
qy - 2 y
dq
qx - 2 x
qv , ψ
dq
qx + 2 x
dq
qy + 2 y dz
dy
dx
c) weak form
Z Z Z
− δR = λ ( δT,x T,x + δT,y T,y ) dA − δT qV dA − δT [−h (T∞ − T )] ds .
Due to their clear arrangement, the element matrices are not derived from the
weak form, but, applying the matrix representation, are developed according
to the bar. Therefore, the matrix symbols need to be linked to the context
described by the governing equations.
4.2 Heat Conduction Regarding Two Spatial Dimensions 71
ly A lx
D B
x
y C
Fig. 4-5 Geometry and coordinate system regarding the rectangular element
1
A
1
B
1
D
1
C
If the primary nodal unknowns of the element are arranged in the vector v by
nodes
v T = TA TB TC TD , (4.19)
the shape functions are to be merged according to the matrix Ω by the sequence
of the primary nodal unknowns regarding the temperature field T
Ω= φA φB φC φD . (4.20)
Table 4.1 explicitly represents the shape functions and their derivatives with
respect to the coordinates.
4.2 Heat Conduction Regarding Two Spatial Dimensions 73
x y 1 y x
B lx
· (1 − ly
) lx
· (1 − ly
) lx
· (− l1 )
y
x y 1 y x 1
C lx
· ly lx
· ly lx
· ly
x y y x
D (1 − lx
)· l − l1 · ly
(1 − lx
) · l1
y x y
u = Ω·v, (4.21)
δu = Ω · δv . (4.22)
assigned to the real nodal temperatures and the rows to the virtual nodal tem-
peratures respectively. The symmetry of the integrand’s matrix entries follows
from its quadratic form and the choice of identical shape functions for the real
and virtual temperatures. Therefore the element matrix is symmetric, too.
* O B
E O *
BT BT E B
To get a clear view, the integrand is multiplied beforehand employing the fol-
lowing scheme and is integrated afterwards:
BT · E · B = λ · I . (4.26)
According to the Equations (4.25) and (4.26), the element matrix K follows
after the integration of I and is given by the representation below. As already
explained regarding bars, the sums of single columns as well as rows of the
4.2 Heat Conduction Regarding Two Spatial Dimensions 75
element matrix disappear, since a constant temperature field without heat flux
needs to be represented.
ℓ ℓ ℓ ℓy
2 ℓy + 2 ℓℓx −2 ℓy + ℓℓx − ℓy − ℓℓx ℓx
− 2 ℓx
ℓy
x y x y x y
ℓy ℓx ℓy ℓx ℓy ℓx ℓy
− ℓ − ℓℓx
−2 ℓ + ℓ 2 ℓ + 2 ℓ −2 ℓ
λ x y x y ℓ x y x y
K= ·
6 ℓy ℓy ℓy ℓy
− ℓ − ℓℓx ℓx
− 2 ℓx
ℓy
2 ℓx
+ 2 ℓx
ℓy
−2 ℓx
+ ℓx
ℓy
x y
ℓy ℓx ℓy ℓx ℓy ℓx ℓy ℓx
ℓ
− 2 ℓ
− ℓ
− ℓ
−2 ℓ
+ ℓ
2 ℓ
+ 2 ℓ
x y x y x y x y
qx + λ · T,x = 0 ,
qy + λ · T,y = 0
σ = E · ǫ,
σ = E · D· u, (4.28)
σ = E·D·Ω·v = E·B·v = S·v.
If the heat conduction matrix E is constant, the course of function of the heat
flux σ already follows with the course of function regarding B.
The special contents of the matrices required for the evaluation are known
from the derivation of the element matrices. Here the stress matrix S describes
the heat fluxes normalized to the nodal temperatures. Employing the shape
fuctions given in Section 4.2.4 yields
" #
φA,x φB,x φC,x φD,x
S = E · B = −λ · ,
φA,y φB,y φC,y φD,y
whereat the derivatives of the shape functions are included in Table 4.1. The
stress matrix S reveals the course of the heat fluxes inside the element, con-
stantly related to the respective direction and linearly perpendicular to it.
4.3 Example of Use 77
The nodal values σ̃ of the heat fluxes may be evaluated after the nodal coor-
dinates have been introduced into B. Thus the discrete stress matrix S̃ follows
with the discrete auxiliary matrix B̃:
σ̃ = E · B̃ · v ,
σ̃ = S̃ · v . (4.29)
In detail the heat fluxes according to the element nodes may be evaluated to
σ̃ = S̃ · v ,
σ(x = 0 , y = 0 ) S(x = 0 , y = 0 )
σ(x = ℓx , y = 0 ) S(x = ℓx , y = 0 )
= ·v,
σ(x = ℓx , y = ℓy ) S(x = ℓx , y = ℓy )
σ(x = 0 , y = ℓy ) S(x = 0 , y = ℓy )
0.50 1.25
0.125
o λ 2 = 0.12 B 0.125
22
0.50
o
6
A λ 1 = 0.08 0.50
0.25
y
22. 0
o
C
6. 0
6. 3
W/m2
0. 0
0. 0
W/m2
−15. 4
Membrane structures carry their loading along the main plane structure, so
the equilibrium conditions also may be addressed at this structural level. Thus,
instead of 3–dimensional coordinates and volume integrals, all governing equa-
tions as well as the work equations may be represented in the 2–dimensional
x–y–coordinate system, what may be equivalent to an integration with respect
to the third coordinate.
dx
x,u
px
y,v px σxx
dy
dy py
σxy
py σyx
dx
σyy
Fig. 5-1 Sign rule
fields u(x, y), v(x, y) [ m ], the stresses σxx (x, y), σyy (x, y), σxy (x, y), [ N/m2 ] as
well as Green’s strains εxx (x, y), εyy (x, y), εxy (x, y) [ 1 ].
a) Governing equations
K: εxx − u,x =0
εyy − v,y =0
εxy − 21 ( u,y + v,x ) =0
E: σxx ,x + σyx ,y + px =0
σyy ,y + σxy ,x + py =0
σxy − σyx =0
E
M: σxx − 1−ν 2
· (εxx + νεyy ) =0
E
σyy − 1−ν 2
· (εyy + νεxx ) =0
E
σxy − 1+ν
· εxy =0
b) Principle of virtual Displacements
Z
δW = (−δεxx σxx − δεyy σyy − 2 δεxy σxy + δu px + δv py ) dV = 0
c) Work equations: Instead of the integration with respect to the volume the
integration with respect to the z–coordinate usually is conducted a priori. As-
suming the thickness t to 1 m and employing the loadings px , py [ N/m2 ] this
leads to:
Z Z
E 1
−δWu = {δu,x (u, x + ν v,y ) + (1 − ν) δu,y (u, y + v,x )} dA − δu px dA ,
1 − ν2 2
Z Z
E 1
−δWv = {δv, y (v,y + ν u, x ) + (1 − ν) δv, x (u, y + v, x )} dA − δv py dA .
1 − ν2 2
σ =E·ε ,
σxx 1 ν εxx
E
σ
yy = ν 1 · εyy . (5.3)
1 − ν2
σxy 1 2εxy
2 (1 − ν)
Applying the governing equations from Section 5.1.1 kinematics may be repre-
sented by matrix notation, which is applied to virtual strains analogously:
ε = D·u ,
εxx ∂x " #
u
εyy = ∂y · . (5.4)
v
2εxy ∂y ∂x
The displacement vector u includes the displacements u(x, y) and v(x, y), which
can be described by means of shape functions and nodal degrees of freedom.
ly A lx
D B
x, u
y, v C
Fig. 5-2 Geometry and coordinate system regarding the rectangular element
84 5 Membrane Structures
At this point, a product approach, derived from the linear shape functions well–
known from the bar and being physically meaningful, achieves the goal faster.
Thus, the displacements u in x–direction and v in y–direction are approximated
by the unknown nodal displacements at nodes A to D and the bi–linear shape
functions φA (x, y), . . . which are correlated to the unknowns. Therefore, the
physically meaningful approach regarding the displacement fields u(x, y) and
v(x, y) yields:
u(x, y) = φA · uA + φB · uB + φC · uC + φD · uD ,
(5.6)
v(x, y) = φA · vA + φB · vB + φC · vC + φD · vD .
If the nodal degrees of freedom of the respective element are arranged in the
vector v by nodes,
vT = uA vA uB vB uC vC uD vD , (5.7)
the shape functions need to be arranged in the matrix Ω in the same order as
the nodal degrees of freedom in both displacement fields u
" #
φA φB φC φD
Ω= . (5.8)
φA φB φC φD
u = Ω·v. (5.9)
δu = Ω · δv . (5.10)
5.1 Rectangular Elements Regarding Plane Stress Situation 85
The bi–linear shape functions of the displacement fields are depicted in Figure
5-3. All shape functions φA , . . . are normalized to 1 at the nodes, linear in the
x– and y–direction and quadratic along the diagonal. Table 5.1 presents the
1
A
1
B
1
D
1
C
Fig. 5-3 Bi–linear shape functions regarding displacements at rectangular elements
x y 1 y x
B lx
· (1 − ly
) lx
· (1 − ly
) lx
· (− l1 )
y
x y 1 y x 1
C lx
· ly lx
· ly lx
· ly
x y y x
D (1 − lx
)· l − l1 · ly
(1 − lx
) · l1
y x y
B = D· Ω, (5.11)
86 5 Membrane Structures
that describes the strain field. Regarding the complete element, it follows
φA,x φB,x φC,x φD,x
B= φA,y φB,y φC,y φD,y . (5.12)
φA,y φA,x φB,y φB,x φC,y φC,x φD,y φD,x
The integrand employs a quadratic form that yields an 8 × 8-matrix. The mul-
tiplication scheme is represented in Figure 5-4. While the columns of the inte-
E B
* * 0
* * 0
0 0
*
BT
B TE B TE B
grand are linked to the real nodal degrees of freedom, the rows are linked to
the respective virtual nodal displacements. The symmetry of the integrand as
well as of the element stiffness matrix follows from the quadratic form of the
integrand and from the identical choice of shape functions regarding real and
virtual displacements.
5.1 Rectangular Elements Regarding Plane Stress Situation 87
For the sake of a clear view, the integrand is split into three parts according to
the terms occurring in the elasticity matrix
E νE E
BT · E · B = 2
· I1 + 2
· I2 + · I3 (5.14)
1−ν 1−ν 2(1 + ν)
with
φA,x φA,x φA,x φB,x φA,x φC,x φA,x φD,x
φA,y φA,y φA,y φB,y φA,y φC,y φA,y φD,y
φB,x φA,x φB,x φB,x φB,x φC,x φB,x φD,x
φB,y φA,y φB,y φB,y φB,y φC,y φB,y φD,y
I1 =
φC,x φA,x φC,x φB,x φC,x φC,x φC,x φD,x
φC,y φA,y φC,y φB,y φC,y φC,y φC,y φD,y
φD,x φA,x φD,x φB,x φD,x φC,x φD,x φD,x
φD,y φA,y φD,y φB,y φD,y φC,y φD,y φD,y
φA,y φA,y φA,y φA,x φA,y φB,y φA,y φB,x φA,y φC,y φA,y φC,x φA,y φD,y φA,y φD,x
φA,x φA,y φA,x φA,x φA,x φB,y φA,x φB,x φA,x φC,y φA,x φC,x φA,x φD,y φA,x φD,x
φB,y φA,y φB,y φA,x φB,y φB,y φB,y φB,x φB,y φC,y φB,y φC,x φB,y φD,y φB,y φD,x
φB,x φA,y φB,x φA,x φB,x φB,y φB,x φB,x φB,x φC,y φB,x φC,x φB,x φD,y φB,x φD,x
I3 =
φC,y φA,y φC,y φA,x φC,y φB,y φC,y φB,x φC,y φC,y φC,y φC,x φC,y φD,y φC,y φD,x
φC,x φA,y φC,x φA,x φC,x φB,y φC,x φB,x φC,x φC,y φC,x φC,x φC,x φD,y φC,x φD,x
φD,y φA,y φD,y φA,x φD,y φB,y φD,y φB,x φD,y φC,y φD,y φC,x φD,y φD,y φD,y φD,x
φD,x φA,y φD,x φA,x φD,x φB,y φD,x φB,x φD,x φC,y φD,x φC,x φD,x φD,y φD,x φD,x
The differentiated shape functions are given in Table 5.1. Because of the bi–
linear shape functions regarding the displacements, quadratic functions at ma-
ximum appear in the integrand. According to Equations (5.13) and (5.14), the
element stiffness matrix follows from the sum of the three sub–matrices after
integrating:
K = K1 + K2 + K3 .
In the following the matrices K1 , K2 , K3 are given by numbers.
88 5 Membrane Structures
ℓ ℓ ℓ ℓy
2 ℓy −2 ℓy − ℓy ℓx
x x x
ℓx ℓx
− ℓℓx −2 ℓℓx
2ℓ ℓy
y y y
ℓy ℓy ℓy ℓy
−2 ℓx 2ℓ
x ℓx
−ℓ
x
ℓx ℓx ℓx
− ℓℓx
2ℓ −2 ℓ
E ℓy y y y
K1 =
6(1 − ν 2 ) ℓy ℓy ℓy ℓy
− ℓx ℓx
2ℓ
x
−2 ℓ
x
− ℓℓx −2 ℓℓx 2 ℓℓx ℓx
y y y ℓy
ℓy ℓy ℓy ℓy
ℓ −ℓ −2 ℓ 2ℓ
x x x x
ℓx ℓx ℓx
−2 ℓ −ℓ ℓ
2 ℓℓx
y y y y
1 1 −1 −1
1 −1 −1 1
−1 −1 1 1
νE 1 −1 −1 1
K2 = 2
4(1 − ν )
−1 −1 1 1
−1 1 1 −1
1 1 −1 −1
−1 1 1 −1
4 ℓℓx 3 2 ℓℓx −3 −2 ℓℓx −3 −4 ℓℓx 3
y y y y
ℓ ℓ ℓℓ
3
4 ℓy 3 −4 ℓy −3 −2 ℓy
2 ℓy −3
x x x
x
ℓx ℓx ℓx ℓx
2 3 4ℓ −3 −4 ℓ −3 −2 ℓ 3
ℓy y y y
ℓy ℓy ℓy ℓy
−3 −4 ℓ −3 4 ℓ 3 2ℓ 3 −2 ℓ
E x x x x
K3 =
24(1 + ν)
−2 ℓx −3 −4 ℓℓx 3 4 ℓℓx 3 2 ℓℓx −3
ℓy y y y
ℓy ℓy ℓy ℓy
−3 −2 ℓ −3 2 ℓ 3 4ℓ 3 −4 ℓ
x x x x
−4 ℓx −3 −2 ℓ ℓx
3 2ℓℓx
3 4ℓ ℓx
−3
ℓy y y y
ℓ ℓ ℓ ℓ
3 2 ℓy 3 −2 ℓy −3 −4 ℓy −3 4 ℓy
x x x x
5.1 Rectangular Elements Regarding Plane Stress Situation 89
p = Ω · p̃ (5.16)
Thus it follows
Z Z Z
f = ΩT · p dA = ΩT · Ω · p̃ dA = ΩT · Ω dA · p̃ . (5.17)
Applying the matrix notation defined in Section 2.3.2, the computation of the
stresses σ regarding the respective element follows to
σ = E · ǫ,
σ = E · D· u, (5.19)
σ = E·D·Ω·v = E·B·v = S·v.
If the elasticity matrix E is constant, the courses of the stresses σ are already
determined with the course of the function of B.
The particular contents of the matrices required for the computation of stres-
ses are known from the derivation of the element matrices. By means of the
elasticity matrix E and the matrix of the differentiated shape functions B, the
stress matrix S can be determined. Employing the shape functions, cf. Section
5.1.4, the following applies:
φ νφA,y φB,x νφB,y φC,x νφC,y φD,x νφD,y
E A,x
S =E·B= νφA,x φA,y νφB,x φB,y νφC,x φC,y νφD,x φD,y .
1 − ν2
eφA,y eφA,x eφB,y eφB,x eφC,y eφC,x eφD,y eφD,x
As such, the abbreviation e = (1 − ν)/2 is used here. Table 5.1 comprises the
derivatives of the shape functions. It can be seen from the stress matrix S
that the course of the function inside the element is constant in the respective
direction of differentiation, and is linear perpendicular to it. Assuming that
Poisson’s ratio is not zero, additional linearly–constant terms have to be su-
perposed. Therefore, the course of the function is generally bi–linear, unlike as
at the bar.
92 5 Membrane Structures
The nodal values of the stresses σ̃ are obtained by inserting the coordinates of
the nodes into B. Applying the discrete matrix B̃, the stress matrix S̃ yields:
σ̃ = E · B̃ · v ,
σ̃ = S̃ · v ,
σ(x = 0 , y = 0 ) S(x = 0 , y = 0 )
σ(x = ℓx , y = 0 )
S(x = ℓx , y = 0 )
= ·v,
σ(x = ℓx , y = ℓy ) S(x = ℓx , y = ℓy )
σ(x = 0 , y = ℓy ) S(x = 0 , y = ℓy )
−ℓ1 − ℓν ν
1
x y ℓx ℓy
ν ν
−
ℓx − ℓ1 ℓx
1
ℓy
y
σxx e
− − ℓe e e
σyy ℓy
x ℓx ℓy
1
σxy A
−
ℓx
1
ℓx
− ℓν ν
ℓy
u
y
σxx ν ν v A
−
ℓx ℓx
− ℓ1 1
ℓy
σyy y
u
− ℓe − ℓe e e
σxy E x y ℓx ℓy v B
B
= 1 − ν2
· .
σxx − ℓν 1 ν
− ℓ1
u
y ℓx ℓy x
σyy
v C
ν 1
− ℓ1 ℓx ℓy
− ℓν
u
σxy C
y x
σxx
− ℓe e e
ℓy ℓx
− ℓe
v D
y x
σyy
− ℓν 1
− ℓ1 ν
y ℓx x ℓy
σxy ν
D
− ℓ1 ℓx
− ℓν 1
ℓy
y x
−ℓe e
ℓx
e
ℓy
− ℓe
y x
shear stress behavior are also present at the clamping cross section.
x = l/2
x, u PY
l B
3
y, v
l
l = 15 m, t = 1 m, E = 100 000 N/m2, ν = 0, PY = 2 N/m2
In the following, cf. Figure 5-6, the stresses σxx and σxy are depicted regarding a
mesh employing 24 × 8 elements without taking into account the antisymmetry.
The jumps at the element intersections are clearly visible, following from the
subsequent stress analysis executed element by element.
σxx
180. 0
−180. 0
σxy
24. 0
−7. 8
The approximation of the shear stresses that shows a change of sign at the
element intersections in the x–direction yet is particularly poor. An improve-
ment of these results employing the same number of elements as well as keeping
the order of shape functions unchanged may only be possible when applying a
special procedure e. g. according to Section 5.2.
94 5 Membrane Structures
The jumps in the distribution of stresses are depicted for the undisturbed area
in the center of the structure at x = 7. 50 m regarding meshes differently refi-
ned, see Figure 5-7, where only the upper half of the cantilever is represented.
The convergence of the normal stresses σxx is quite good, whereat the discreti-
zation error can be evaluated directly from the stress jumps. The shear stresses
converge very slowly and show particularly large jumps with changing signs in
some parts, so the results appear altogether questionable.
6 384 6 24 384 24 6
6
neutral axis
x = 7.5 ε
x = 7.5 + ε ε 0
σxx
σe x = 7,5 m
1,0
0,5
σxy
σe x = 7,5 m
1,0
0,5
v
ve
x = 15,0 m
1,0
0,5
Overall, the distribution of stresses regarding the cantilever may only be poorly
approached when employing a rectangular plane stress element with bi–linear
shape functions, even if a relatively fine discretization is applied. Reason for
this is the poor approximation of the shear stiffness inside the element, which
is clearly visible in this example. An improvement of the element behavior is
explained in Section 5.2.
96 5 Membrane Structures
εyx
A B
εxy
x uA = uB = −uC = −uD ,
εxy
y
vA = −vB = −vC = vD .
D C
εyx
multiplicative shape functions are sufficient for u(x, y), if they are constant in
the x–direction, and for v(x, y), if they are constant in the y–direction, in order
to correctly describe the shear strains inside the element.
5.2 Rectangular Element Comprising Modified Shear Strains 97
The second case, as given in Figure 5-10, describes a pure bending deformation
of the element whereby the strains in the x– as well as in the y–direction are
exactly represented, but additional shear deformations γAD and γBC occur,
showing changing signs and being non-physical. They arise because of the linear
shape functions that are employed, which cannot describe the element curvature
at pure bending. The drawback results in activation of shear stiffness in the case
of pure bending and therefore the respective bending deformation converges
more slowly against the exact solution.
A B A B
x x
y εxy y εxy
D C D C
∆u ∆u
The entries are constant except for the quotients of lengths and the sign.
ℓx ℓx
ℓy
1 ℓy
−1 − ℓℓx −1 − ℓℓx 1
y y
ℓy ℓy ℓy ℓy
1 1 − ℓ −1 − ℓ −1 ℓ
ℓx x x x
ℓx ℓ x ℓ x ℓ x
ℓy 1 ℓy
−1 − ℓy
−1 − ℓy
1
ℓy
−1 − ℓy −1 ℓy 1
ℓy
1 −
3·E ℓx ℓx ℓx ℓx
K3 = .
24(1 + ν)
− ℓx −1 − ℓx 1 ℓx
1 ℓx
−1
ℓy ℓy ℓy ℓy
ℓ y ℓ y ℓ y ℓ y
−1 − −1 ℓ 1 1 −ℓ
ℓx x ℓx x
ℓx ℓx ℓx ℓx
− −1 − 1 1 −1
ℓy ℓy ℓy ℓy
ℓy ℓy ℓy ℓy
1 ℓ
1 − ℓ
−1 − ℓ
−1 ℓ
x x x x
The stress matrix applied at subsequent analysis is given below regarding con-
stant shear stresses. This corresponds to an averaging of the shear stresses
according to Section 5.1.8. As such, e = (1 − ν)/2/2 is applied.
1 ν
−ℓ − ℓν 1
ℓx ℓy
x y
ν ν 1
− − 1
ℓx ℓy ℓx ℓy
σxx e
− − e −e e e e e
−ℓe
σyy ℓ y ℓ x ℓ y ℓ x ℓ y ℓ x ℓ y x
1 1 ν ν
σxy A − − u
ℓx ℓx ℓy ℓy
σxx ν ν 1 1 v A
− −
σ ℓx ℓx ℓy ℓy
yy e e e e e e e e u
− − − −
σxy E ℓy ℓx ℓy ℓx ℓy ℓx ℓy ℓx v B
B = · .
σ 1 − ν 2 ν 1 ν 1 u
xx
− ℓy ℓx ℓy
− ℓx
v
σyy
1 ν 1 ν C
σ
− ℓy ℓx ℓy
− ℓx
u
xy C e
− − e −e e e e e e
−ℓ v D
σxx ℓy ℓx ℓy ℓx ℓy ℓx ℓy x
σyy ν 1 1 ν
− ℓy ℓx
− ℓx ℓy
σxy D ν
− ℓ1 ℓ
− ℓν ℓ1
y x x y
−ℓe − ℓe −ℓe e
ℓx
e e
ℓy ℓx
e
ℓy
−ℓe
y x y x
5.2 Rectangular Element Comprising Modified Shear Strains 99
−180. 0
σxy
15. 0
2. 5
Remarks
Even if the selectively reduced integration allows for very good results here,
the results may be poor at other applications, cf. Section 24.2 and 24.4. Fur-
thermore, if the complete stiffness matrix would be integrated by a reduced
integration scheme, displacement fields, not comprising any strain energy, may
occur, which are named zero–enery modes. These displacement fields may be
recognized by an oscillation of the displacements in terms of an hourglass and
therefore are called hourglass modes.
100 5 Membrane Structures
σ = E·ε,
σxx 1−ν ν ν εxx
σyy
E
ν 1−ν ν
εyy
= · , (5.20)
σzz (1 + ν)(1 − 2ν) ν ν 1−ν εzz
σxy (1−2ν)/2 2εxy
ǫ = D· u,
εxx ∂x
εyy
∂y
u
= · . (5.21)
εzz v
2εxy ∂y ∂x
Except for the fact that the matrix B is comprised of an additional row employ-
ing zero entries, there are no significant differences in the further derivations,
therefore these are not shown here.
6 Bending Structures
Thin walled structures, which are loaded perpendicular to their main spatial
dimensions, carry their loading by bending. Modeling of bending structures
usually requires kinematic assumptions, which yield either the Euler–Bernoulli
beam theory at investigating 1–dimensional structures or the Kirchhoff theory
of plates for 2–dimensional structures. Investigating the bending behavior of
thin bending structures, shear deformations may be neglected, what requires
special finite elements, dealing with C1 –conformity at the element intersections.
Me
x e
V
ϕ p
z
z,w
M - dM M + dM
2 x 2
dQ dQ
Q- Q+
2 2
E: Q,x + pz =0
or : M,xx + pz = 0
M,x − Q =0
M : M − EI · (−κ) = 0
c) Work equation
Z Z
− δWd = δw,xx · EI · w,xx dx − δw · pz dx = 0 .
or in a short form
Z ℓ Z ℓ
T T
− δWd = δv { B · E · B dx · v − ΩT · p dx } . (6.2)
0 0
M = EI · (−κ) ,
σ = E · ǫ. (6.3)
The differentiation rule, necessary to compute the curvature κ from the de-
flection w, is given by the operator matrix D. Applying matrix notation, the
vectors of strains ε and of displacements u, which only comprise the deflection
w(x), yield
w(x) = a0 + a1 · x + a2 · x2 + a3 · x3 . (6.5)
To get a physically meaningful approach, it makes sense not only to use the
displacements wi , but also the rotations ϕ|i = w,x |i as degrees of freedom at
the element nodes, since, regarding convergence criteria, the requirement of
C1 –conformity may be easily fulfilled hereby
w(x) = φ1 · wA + φ2 · ϕA + φ3 · wB + φ4 · ϕB . (6.6)
The related shape functions φi , represented in Figure 6-2, are the Hermite–
Polynomials of the 3rd order. The courses of the functions are given as well
as the respective second derivatives. It is to be emphasized that the approach
is chosen to describe the deflection w(x), therefore the rotations are nodal
displacements variables only.
A B
i φi (x) φi (x),xx x
l
x2 x3
1 1−3 ℓ2
+2 ℓ3
− 6 ℓ12 + 12 ℓx3 1 Φ1 (x)
2 x3
2 x − 2 xℓ + ℓ2
− 4 1ℓ + 6 ℓx2 1 Φ2 (x)
2 3
3 3 xℓ2 − 2 xℓ3 6 ℓ12 − 12 ℓx3 1 Φ3 (x)
2 x3
4 − xℓ + ℓ2
− 2 1ℓ + 6 ℓx2 1 Φ4 (x)
vT = [ wA ϕA wB ϕB ]
104 6 Bending Structures
and the shape functions merged into the matrix of shape functions
Ω = [ φ1 φ2 φ3 φ4 ]
the approach to describe the deflection follows to
u=Ω·v (6.7)
and in analogy, regarding the virtual deflection, to
δu = Ω · δv . (6.8)
are already known. The matrix product, being part of the integrand, yields
φ1,xx φ1,xx φ1,xx φ2,xx φ1,xx φ3,xx φ1,xx φ4,xx
φ2,xx φ1,xx φ2,xx φ2,xx φ2,xx φ3,xx φ2,xx φ4,xx
BT · E · B = EI ·
.
φ3,xx φ1,xx φ3,xx φ2,xx φ3,xx φ3,xx φ3,xx φ4,xx
φ4,xx φ1,xx φ4,xx φ2,xx φ4,xx φ3,xx φ4,xx φ4,xx
The second derivatives of the shape functions are linear according to Figure
6-2. The products may be analytically integrated, what results in:
12 6
ℓ 3 ℓ 2 − 12
ℓ 3
6
ℓ 2
6 4 6 2
ℓ2 ℓ
− ℓ2 ℓ
K = EI · 12
. (6.11)
− 3 6 12 6
ℓ − ℓ 2 ℓ 3 − 2
ℓ
6 2 6 4
ℓ2 ℓ
− ℓ2 ℓ
6.1 Element Matrices Regarding Euler–Bernoulli Beams 105
If the nodal displacement variables are ordered by nodes, the element stiffness
matrix may be clearly arranged applying sub–matrices, which explicitly point
out the allocation of the stiffness entries to the element nodes
" #
kAA kAB
K= .
kBA kBB
The load vector corresponds to the supporting forces and moments of the beam
clamped at both ends and loaded by a constantly distributed external action.
A varying action p(x) may be approximated by a linear course of the function
inside the element. Applying the matrix Ωp
h x x i
Ωp = 1 − = [ φA φB ] ,
ℓ ℓ
employing linear shape functions according to Section 2.3.1 and the related
nodal values p̃ = [ pA pB ], it applies
p = Ωp · p̃ = φA · pA + φB · pB . (6.14)
106 6 Bending Structures
Thus the shear forces are constantly approximated inside an element, but allo-
cated both element nodes. The stress matrix regarding the shear forces as well
as the bending moments can be summarised and ordered by nodes
QA − 12
ℓ 3 − ℓ62 12
ℓ 3 − ℓ62
6 4 6 2
MA ℓ2 ℓ
− ℓ 2 ℓ
= EI · ·v. (6.16)
12 6 12 6
QB − ℓ3 − ℓ2 ℓ 3 − ℓ 2
MB − ℓ62 − 2ℓ ℓ
6
2 − 4
ℓ
In this formulation, the stress matrix is identical to the element stiffness matrix
according to Equation (6.11), except of the signs.
1 2 3 4
1 2 3 4 5
x
EI = 10,000 kNm2
z
l = 8.0 m
2.0 2.0 2.0 2.0
lc 1 p = 10 kN/m
lc 2 P = 16 kN
lc 3 w1 = 0.04 m
The discretization is chosen employing four elements and five nodes. The results
concerning the loading conditions two and three represented in Figure 6-4 are
identical to the exact analytical solution, since the approach for the deflection
is the solution of the homogeneous differential equation, and the particular
solution part disappears due to the external actions selected here. The results
108 6 Bending Structures
concerning loading condition one are characteristic for the FEM, being only an
approximate solution.
1 2 3 4 5
1 2 3 4
w1
20 mm
w2
w3
−76.67 = M5
lc 1: M −40.00 = Q 4
Q
−24.00 = M5
lc 2: M −11.00 = Q 4
Q
−18.75 = M5
lc 3: M −2.34 = Q 4
Q
Fig. 6-4 Single span beam – state variables regarding loading conditions 1 to 3
6.2 Kirchhoff Plate Element with 16 DOF 109
E: qx ,x + qy ,y + p =0
mxx ,x + mxy ,y − qx =0
myy ,y + mxy ,x − qy =0
or : mxx ,xx + 2mxy ,xy + myy ,yy + p = 0
M : mxx − B (−κxx − νκyy ) = 0
myy − B (−κyy − νκxx ) = 0
mxy − B (1 − ν) (−κxy ) = 0
E t3
with plate bending stiffness B=
12 (1 − ν 2 )
b) Principle of virtual Displacements
Z
− δWd = { δκxx mxx + 2δκxy mxy + δκyy myy + δw p } dA = 0
c) Work equation
Z
− δWd = B { δw,xx (w,xx + νw,yy ) + δw,yy (w,yy + νw,xx )
Z
+ δw,xy 2(1 − ν) w,xy } dA − δw p dA .
The matrix symbols are filled with physical content, if the governing equations
are interpreted in a corresponding manner.
σ = E ·ε,
mxx 1 ν −κxx
ν 1
myy = B · · −κyy . (6.19)
1
mxy 2
(1 − ν) −2κxy
The differentiation rule D to compute the strains ǫ from the deflection is defined
by the kinematic conditions
ε = D · u,
−κxx −∂xx
−κyy = −∂yy · [ w ] . (6.20)
−2κxy −2∂xy
The only independent state variable is the deflection surface w(x, y), which is
approximated by shape functions and nodal displacement variables
The shape functions assigned to the nodal displacement variables can be deve-
loped as a product approach employing Hermite–Polynomials that have been
already introduced regarding bending at Euler–Bernoulli beams. The combina-
tion of the respective four shape functions of the beam concerning the x– and
y–directions as a product approach yields 16 shape functions in total to describe
the deflection. If the nodal displacement variables vnode = [ w w,x w,y w,xy ]node
are chosen as degrees of freedom in the four corners of the rectangular element,
112 6 Bending Structures
the element intersection conditions can be fulfilled. Overall, the symmetric ap-
proach employs 16 degrees of freedom, ordered by the nodes here
vT = { [ w w,x w,y w,xy ]A
[ w w,x w,y w,xy ]B
[ w w,x w,y w,xy ]C
[ w w,x w,y w,xy ]D } . (6.22)
The shape functions assigned to the displacement variables of a single node are
depicted in Figure 6-6.
w w,x
w,y w,xy
The shape functions related to the displacement variables of the other nodes can
be obtained analogously, if those are combined with each other appropriately.
The order of the nodal displacement variables also defines the order of the
shape functions in the vector
Ω ={[ φ1 (x) · φ1 (y) φ2 (x) · φ1 (y) φ1 (x) · φ2 (y) φ2 (x) · φ2 (y) ]A
[ φ3 (x) · φ1 (y) φ4 (x) · φ1 (y) φ3 (x) · φ2 (y) φ4 (x) · φ2 (y) ]B
[ φ3 (x) · φ3 (y) φ4 (x) · φ3 (y) φ3 (x) · φ4 (y) φ4 (x) · φ4 (y) ]C
[ φ1 (x) · φ3 (y) φ2 (x) · φ3 (y) φ1 (x) · φ4 (y) φ2 (x) · φ4 (y) ]D }
or written in a short form related to the products of the shape functions
Ω = [ φA1 φA2 φA3 φA4
φB1 φB2 φB3 φB4
φC1 φC2 φC3 φC4
φD1 φD2 φD3 φD4 ]. (6.23)
6.2 Kirchhoff Plate Element with 16 DOF 113
If the more formal way is chosen, that is also correct and applies the scaling
matrix, at first the symmetric general polynomial approach is chosen:
w(x, y) = a1 + a2 x + a3 y
+ a4 x2 + a5 xy + a6 y 2
+ a7 x3 + a8 x2 y + a9 xy 2 + a10 y 3
+ a11 x3 y + a12 x2 y 2 + a13 xy 3
+ a14 x3 y 2 + a15 x2 y 3 + a16 x3 y 3 . (6.25)
The symmetry of this approach can be proved by Pascale’s triangle, cf. Figure
2-16. The transformation of the general polynomial into an approach employing
physically meaningful degrees of freedom is possible according to Section 2.4.3,
when the scaling is related to the nodal displacement variables according to
Equation (6.22). With the general polynomial
[ w(x, y) ] = ψ · a
v = Ψ̃ · a ,
Ω = ψ · G.
1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0
2 3 1
0 0 2 0 0 0 0 0 0 0 0 0 0 w
a1 − − lx lx
− lx
0 0 1 0 0 0 0 0 0 0 0 0 0 0
3
3
0 0 0 0 0 0 0 0 0
lx2
y y
ly2
w,x
0 − l2 0 − l1
0
1
a2 0
0 0 0 0 0 0 0 0 0 0 w
w,y
2
lx x x
0 − l23 l12
a3 3
0
3
a4 − ly2
0 0 0 0 0 0 0 0 0
2
2 w,xy A
x x lx x
0 − l32 − l2 − l1
a5 0
3 2 3
lx3
0 0 0 0 0 0 0 0 0
y y
ly2
a6
0 −l
0 0 w,x
1 2 1
w
w,y
0 0 0 0 0 0 0 0 0 0 0 0
2 3 2
y ly ly ly
a7 1
−
2 1
a8 0 − ly2 0 −l
0 0 0 0 0 0 0 0 0
= 2 w,xy B
3 2
lx lx x x
0 − l23 l12
a9 l 3
6 6 4 9 3 6 2 9 3 3 1 9 6 3
a10 w,x
0 0
x y x y x y x y x y x y x y x y x y x y x y
w
a11 w,y
2 1 2
a12 9
0 0 0 0 0 0 0 0 0 0 0
2
−
l2 l2 l l2 l2 l l l − l2 l2 l l2 − l2 l l l l2 l2 − l l2 − l2 l l l − l2 l2 − l l2 l2 l l l
a13 xy xy xy xy x y
a15 w,y
3 4 2 6 3 4 2 6 3 2 1 6 3 2 1
1
a16
w,xy D
a14 0
ly2
− − − − − −
w,x
6
− 3 2 − 2 2 − 3 − 2
x y x y x y x y x y x y x y x y x y x y
lx ly lx ly lx ly lx ly lx3 ly2 lx2 ly2 lx3 ly lx2 ly lx3 ly2 lx2 ly2 lx3 ly lx2 ly lx3 ly2 lx2 ly2 lx3 ly lx2 ly
4 2 2 1
6
l3 l3 l2 l3 l3 l2 l2 l2
− l2 l3 − l 4l3 − l23l2 − l 2l2 l26l3 − l 2l3 l23l2 − l 1l2 − l26l3 l 2l3 l23l2 − l 1l2 l26l3 l 4l3 − l23l2 − l 2l2
x y x y x y x y x y x y x y x y x y x y x y x y x y x y x y x y
− l34l3 l22l3 − l32l2 l21l2 l34l3 − l22l3 − l32l2 l21l2 − l34l3 − l22l3 l32l2 l21l2
xy xy xy xy xy x y
6 Bending Structures
6.2 Kirchhoff Plate Element with 16 DOF 115
B = D ·Ω. (6.26)
Ω
D
The submatrix BA comprising the shape functions φA1 , φA2 , φA3 , φA4 applies
for the node A, e.g., as follows
The differentiation of the shape functions with respect to the coordinates may
be seen more clearly applying product notation
−φ1 (x),xx φ1 (y) −φ2 (x),xx φ1 (y) −φ1 (x),xx φ2 (y) −φ2 (x),xx φ2 (y)
BA = −φ1 (x) φ1 (y),yy −φ2 (x) φ1 (y),yy −φ1 (x) φ2 (y),yy −φ2 (x) φ2 (y),yy .
−2φ1 (x),x φ1 (y),y −2φ2 (x),x φ1 (y),y −2φ1 (x),x φ2 (y),y −2φ2 (x),x φ2 (y),y A
116 6 Bending Structures
The integrand may be computed applying the scheme according to Figure 6-8
and is assigned to the corresponding element nodes.
A B C D
0
E ** ** 0 B
0 0 *
BT BT E B
All submatrices kAA , kAB , . . . are similarly arranged; only the shape functions
applied to build them are different.
6.2 Kirchhoff Plate Element with 16 DOF 117
whereat the multipliers of the integrals result from the elasticity matrix E. In
detail this yields
φA1 ,xx φA1 ,xx φA1 ,xx φA2 ,xx φA1 ,xx φA3 ,xx φA1 ,xx φA4 ,xx
+φA1 ,yy φA1 ,yy +φA1 ,yy φA2 ,yy +φA1 ,yy φA3 ,yy +φA1 ,yy φA4 ,yy
φ , φ , φA2 ,xx φA2 ,xx φA2 ,xx φA3 ,xx φA2 ,xx φA4 ,xx
A2 xx A1 xx
+φ , φ , +φ , φ , +φ , φ , +φ , φ ,
A2 yy A1 yy A2 yy A2 yy A2 yy A3 yy A2 yy A4 yy
IAA1 = ,
φA3 ,xx φA1 ,xx φA3 ,xx φA2 ,xx φA3 ,xx φA3 ,xx φA3 ,xx φA4 ,xx
+φA3 ,yy φA1 ,yy +φA3 ,yy φA2 ,yy +φA3 ,yy φA3 ,yy +φA3 ,yy φA4 ,yy
φ , φ , φA4 ,xx φA2 ,xx φA4 ,xx φA3 ,xx φA4 ,xx φA4 ,xx
A4 xx A1 xx
+φA4 ,yy φA1 ,yy +φA4 ,yy φA2 ,yy +φA4 ,yy φA3 ,yy +φA4 ,yy φA4 ,yy
φA1 ,xx φA1 ,yy φA1 ,xx φA2 ,yy φA1 ,xx φA3 ,yy φA1 ,xx φA4 ,yy
+φA1 ,yy φA1 ,xx +φA1 ,yy φA2 ,xx +φA1 ,yy φA3 ,xx +φA1 ,yy φA4 ,xx
φ , φ , φA2 ,xx φA2 ,yy φA2 ,xx φA3 ,yy φA2 ,xx φA4 ,yy
A2 xx A1 yy
+φ , φ , +φ , φ , +φA2 ,yy φA3 ,xx +φA2 ,yy φA4 ,xx
A2 yy A1 xx A2 yy A2 xx
IAA2 = ,
φA3 ,xx φA1 ,yy φA3 ,xx φA2 ,yy φA3 ,xx φA3 ,yy φA3 ,xx φA4 ,yy
+φA3 ,yy φA1 ,xx +φA3 ,yy φA2 ,xx +φA3 ,yy φA3 ,xx +φA3 ,yy φA4 ,xx
φ , φ , φA4 ,xx φA2 ,yy φA4 ,xx φA3 ,yy φA4 ,xx φA4 ,yy
A4 xx A1 yy
+φA4 ,yy φA1 ,xx +φA4 ,yy φA2 ,xx +φA4 ,yy φA3 ,xx +φA4 ,yy φA4 ,xx
φA1 ,xy φA1 ,xy φA1 ,xy φA2 ,xy φA1 ,xy φA3 ,xy φA1 ,xy φA4 ,xy
φA2 ,xy φA1 ,xy φA2 ,xy φA2 ,xy φA2 ,xy φA3 ,xy φA2 ,xy φA4 ,xy
IAA3 = .
φA3 ,xy φA1 ,xy φA3 ,xy φA2 ,xy φA3 ,xy φA3 ,xy φA3 ,xy φA4 ,xy
φA4 ,xy φA1 ,xy φA4 ,xy φA2 ,xy φA4 ,xy φA3 ,xy φA4 ,xy φA4 ,xy
Thus the stiffness matrix follows with three parts as well
1 ℓx ℓy ℓx ℓy
28080 14040 3960 1980 ·1
B·α
kAA1α = · 9360 1980 1320 ·ℓx ,
6300
720 360 ·ℓy
240 ·ℓx ℓy
1 ℓx ℓy ℓx ℓy
28080 3960 14040 1980 ·1
B·β
kAA1β = · 720 1980 360 ·ℓx ,
6300
9360 1320 ·ℓy
240 ·ℓx ℓy
1 ℓx ℓy ℓx ℓy
18144 9072 9072 1386 ·1
ν ·B·γ
kAA2 = · 2016 7686 1008 ·ℓx ,
6300
2016 1008 ·ℓy
224 ·ℓx ℓy
1 ℓx ℓy ℓx ℓy
9072 756 756 63 ·1
2 (1 − ν) · B · γ
kAA3 = · 1008 63 84 ·ℓx .
6300
1008 84 ·ℓy
112 ·ℓx ℓy
6.2 Kirchhoff Plate Element with 16 DOF 119
In this way, the stress matrix Sm = E · B and thus the courses of the plate’s
moments may be computed with
−φA1 ,xx − νφA1 ,yy "# "# "#
σ = E · B · v = B · −νφA1 ,xx − φA1 ,yy ·v.
B C D
−(1 − ν)φA1 ,xy A2 A3 A4
The stress analysis concerning the moments is performed at the corner nodes of
the element yielding the nodal moments, which can be evaluated by applying
the stress matrix S̃m explicitly given on the next page.
σ̃ = S̃m · v .
The shear forces may be computed analogously to the procedure presented for
the Euler–Bernoulli beam by means of the equilibrium conditions
" # " # " # mxx
qx mxx ,x + mxy ,y ∂x 0 ∂y
= = · myy = Dq · σ .
qy myy ,y + mxy ,x 0 ∂y ∂x
mxy
or in matrix notation
q = Dq · σ = Dq · ( E · B ) · v
= Sq · v .
The stress analysis concerning the shear forces is performed at the corner nodes
of the element as well. It applies
q̃ = S̃q · v ,
whereat S̃q is the stress matrix related to the shear forces at the element nodes.
It should be mentioned, that the shear forces may also be computed by means
of the equilibrium conditions and the first derivatives of the moments, if the
moments are described with a bi–linear approximation with respect to the cor-
ner values.
The moments, regarding the element nodes A to D, are evaluated applying the following stress matrix S̃m
including the abbreviation f = −(1 − ν)
6 4 4ν 2 2ν
ℓ2x
+ 6ν
ℓ2y ℓx ℓy
0 ℓx
0 0 0 0 0 0 ℓ2
0 ℓy
0
x y
− ℓ62 − 6ν
4 2ν 2
0 ℓ2 ℓx
0 0 0 0 0 0 0 ℓy
y x y
− 6ν − ℓ62
0 0 f 0 0 0 0 0 0 0 0 0 0 0
6ν 6 4ν
mxx
ℓ2 + ℓ2 ℓx ℓy 0
x
myy 6 4ν 2ν
0 + 6ν 0 0 0 0 0 0
ℓx x ℓy ℓ2 ℓy
0 0
6ν 2ν 6ν 4 2
mxx 0 0 0 0 0 0 0 0
−6 −2 0 0
+ ℓ62 ℓ
ℓx ℓx ℓ2x y x ℓy y ℓy
− 4ν − ℓ62
6.2 Kirchhoff Plate Element with 16 DOF
myy
A
0 0 0 0 0 0 f 0 0 0 0 0 0 0
mxy
− 2 − 0
6
mxx
0 0 0 0 + 6ν 0 0
0 0
ℓ2 y ℓ y ℓ2x ℓ2y x ℓ y x x
− 6ν 0 − 2ν − ℓ4 − 4ν − ℓ62 − ℓ2
myy
B ·v.
6ν
=B·
mxy 0 0 0 0 0
+ ℓ62 ℓ ℓ2 ℓ
y x y x
0 0
y ℓ2x y x
− ℓ62 − ℓ2 0 − 4ν − ℓ4 0 − 6ν − 2ν
mxx
0 0 0 0 0 0 0 0 0 0 f 0 0 0
0 0
myy
C
2 6 6µ 4
mxy D
0 0 0 0 0 0 0
ℓy ℓx
+ ℓ2 ℓx ℓ y
0 0
ℓ2y x ℓ2x y
0 − 2ν − ℓ62 − 4ν
6 2ν 6ν 4ν
0 0 0 0 0 0 0
− 6ν 0
+ ℓ62
ℓy y ℓ2x ℓx ℓ2x y ℓx y
0 − ℓ2 − 6ν − ℓ4
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 f
− 2 0
121
122
The shear forces, regarding the element nodes, are evaluated applying the related stress matrix S̃q
6 6 4 12 2
ℓ3 ℓ2 ℓ2
0 ℓy ℓ3x
0 0 0 0 0 0 0 0 ℓy
x y
− 12 − − ℓ62 − ℓ62
2 12
ℓx ℓy ℓx
0 0 ℓx
0 0 0 0 0
x ℓ3y y
− ℓ62 − ℓ62
qx
x y x
12 6 6 4 2
12
qy A 0 0 0 0 0 0 0 0
− 3 0 62− 62 4 0
qx
6 6 12
0 0 0 0 0 0
qy x ℓ3
− 12 0
12 6 6
qx
0 0
0 0 0 0 0 0 0
− ℓ62 − ℓ2 − − ℓ4 − 12 − ℓ62
B
qy
12 6 6
·v.
= B ·
0 0 0 0 0 0 0 0
0 0
ℓ3
qy 12 6 6
D
C
0 0 0 0 0 0 0
0 − ℓ2
0 − ℓ2 − ℓ62 − 12 −
x
12 6 2 12 6 6 4
0 0 0 0 0 0 ℓx
0 ℓx
0 − 62 − ℓ4
x clamped
G M ν = 0.2
hinged
l x = 10.0 m
l y = 10.0 m
lx
Representation of Results
The course of the deflection and the moments is depicted in Figure 6-10 regar-
ding the 16 DOF element at selected sections applying a discretization of 2 × 2
and 4 × 4 elements respectively. The deflection gives only little changes due to
mesh refinement, which indicates that the chosen approach may approximate
the deformation behavior very well.
The course of the bending moments is comprised of linear parts as well as of
cubic parts superposed because of lateral curvatures. The respective linear part
is represented only, which still shows significant changes in the case of mesh
refinement due to the second derivatives of the deflection surface. The quality
of the approximate solution of the respective discretization may be identified
from the jumps of the bending moments at the element intersections as well as
from the fulfillment of the boundary condition mxx = 0 at position G.
The twisting moment is directly computed employing the nodal displacement
variable w,xy , thus it is approximated as well as the deflection surface itself.
The course of the twisting moment is parabolic along the element boundaries
and given here as a polygonal line.
124 6 Bending Structures
section G − M: G M
0
w [mm]
G M
0
+ +
1.0
m xx [kNm/m]
section M − E: M E
0
5
w [mm]
M
0
+ E
3.0
m yy [kNm/m]
M
0
E
1.0 +
m xx [kNm/m]
section G − GE: G GE
0
− 1.0
m xy [kNm/m]
Investigation of Convergence
Table 6.1 represents the convergence behavior of the most important state va-
riables regarding the 16 DOF element, cf. Figure 6-11 and the comparison with
other elements is given in Section 10.4. As shown in Section 6.1.9, the deflection
exhibit only little errors, even when employing a small number of elements. Due
to the chosen shape functions the convergence behavior concerning the twisting
moment mxy is just as effective, because mxy may be directly computed from
the twisting w,xy , which is a nodal degree of freedom, but not from the deri-
vative of the deflection surface. The bending moments converge more slowly
against the exact solution, since the second derivative of the deflection surface
causes a roughening of the course of the corresponding state variable. The shear
forces converge even less effectively, since they are computed by applying the
third derivative of the deflection surface. Figure 6-11 shows the development
of the state variables z with respect to mesh refinement, whereat the reference
variables zexact are chosen from the 16 × 16 mesh.
z
z exact myy M
mxx M
w
1.0 mxy T
myy E
0.5
1 4 16 64 elements
16 36 100 324 unknowns
Fig. 6-11 Square plate – convergence behavior regarding the state variables
126 6 Bending Structures
0 0 1 0 0 0 0 0 0 0 0 0
3
a1 0 2 0 0 0 0 0 0 0 w
x x
x lx
− l32 − l2 − l1
a2 w,x
1 1 1 1
0 0 0 0
a3 x y y x lx ly lx x y lx ly ly w,y
− l 1l − l1 − l1 − l 1l
w
a4 3
0 0 0 0 0 0 0 0
y y ly2 y w,x
a5 − l32 − l2 − l1
A
2 1 1 w,y
a6
3 2 0 2 0 0 0 0 0 0 0
6.3 Kirchhoff Plate Element with 12 DOF
lx lx x lx w
− l23
a7
3 2 1 3
w,x
a8 0 0
2l 2l
=
y lx ly lx ly y x y x y
B
lx x y lx x y
0 − l23l − l 1l 0 − l23l − l 2l
·
a9 w,y
3 2 3 1
w
a10 0 0
lx ly2 lx ly x y x y lx ly2 x y x y lx ly
− l 3l2 0 − l 2l 0 − l 1l − l 3l2
a11 w,x
2 1 1
C
0 0 0 0 0 0 0 0
a13 w,y
2 2 1
0 0 0 0
3l 3l 2l
x y x y lx y x y x y x y lx y lx y
2 1 1 2
0 0
The shape functions regarding the 12 DOF element are depicted in Figure
6-12. The breach of the element intersection conditions is obvious, since the
shape functions related to the rotations become linear in one direction, and
thus exhibit a kink at the intersection to neighboring elements.
Example of Use
For comparison the following Table 6.2 shows the convergence behavior of the
12 DOF element. The deflection converges significantly less effectively in com-
parison to the 16 DOF element, since the clamping and the element intersecti-
on conditions with respect to the rotations cannot be exactly described, which
yields a more flexible structural behavior. In contrast, the moments as well as
the shear forces are relatively well approximated, because the omitted twisting
degrees of freedom w,xy does not influence the bending moments mxx , myy and
the shear forces qx , qy that much. It should be mentioned that the twisting
moment mxy now has to be computed at the element level with poorer conver-
gence, since the corresponding curvature w,xy is no nodal degree of freedom as
it is in the case of the 16 DOF element.
quadratic part
In order to satisfy the continuity with respect to the rotations w,n , the work
equation has to be extended by additional terms at the element boundaries,
which take into account the work perfomed by the bending moments mn on the
corresponding rotations. The bending moments are replaced by the material
equations, thus leading to a pure displacement–based formulation with
Z Z
− δWel.b. = δmn · (w,n quad − w,n lin )} ds + mn · (δw,n quad − δw,n lin )} ds
Z
= B · (δw,nn + ν δw,ss ) · (w,n quad − w,n lin )} ds
Z
+ B · (w,nn + ν w,ss ) · (δw,n quad − δw,n lin )} ds .
Applying the extended work equation, the continuity conditions for w,n and
δw,n can be fulfilled weakly at the system level, too.
G M ν = 0.2
hinged
l x = 10.0 m
l y = 10.0 m
lx
Fig. 6-14 Quadratic plate – system and loading, see Section 6.2.9
The deflection at the centre of the plate is given in Table 6.3 concerning the
element with 12 DOF, taking into account boundary integrals of work, and
applying it to different meshes. The results obtained by the modifications are
excellent, cf. Figure 6-15, where the results are compared to the different rectan-
gular elements with 16 DOF and 12 DOF respectively. The deflection converges
almost as well as for the element with 16 DOF.
6.4 The 12 DOF Element Employing a Weak Conformity 131
1.3
normalized centre deflection
1.0
1 4 16 64 256
number of elements
Because of the coordinate invariance, this must not influence the results. The
ξ–η–coordinates are normalised to the length of the element edges ℓAB and
ℓAC , so that the conditions
0≤ξ≤1 and 0≤η≤1
are fulfilled.
xc xa xb x
A
ya ( ξ = 0; η = 0 )
η ξ
yc C
( 0;1 )
yb B
( 1;0 )
y
With regard to the element domain, the shape functions and the integration
limits can be clearly formulated by applying local coordinates.
applying det = ac · bb − ab · bc .
det J = ac · bb − ab · bc .
Thus " #
ac −bc
J= (7.6)
−ab bb
holds as well as " #
−1 1 bb bc
J = .
det J ab ac
ξ = 0, η = 0
(xb − xa ) dξ ac dξ
ξ=0 η=0 dvξ = (yb − ya ) dξ = −bc dξ ,
A
0 0
(xc − xa ) dη −ab dη
η A dvη = (yc − ya ) dη = bb dη .
η=1 0 0
ξ=1 ξ
The vector product provides a third vector whereby its norm is related to the
value of the area
ex ey ez 0
dvξ × dvη = det ac dξ −bc dξ 0 = 0 .
−ab dη bb dη 0 [ ac · bb − ab · bc ] dξ dη
7.1 Local ξ–η–Coordinate System 139
dA = [ ac · bb − ab · bc ] dξ dη = det J dξ dη . (7.7)
At integration of the work related to the domain of the triangle, the integration
limits need to be adapted to the triangular geometry. Because of the crooked
boundaries, the upper integration limit is variable in the ξ–direction,
Z Z Z 1 Z 1−η
Integrand dx dy = Integrand · det J dξ dη .
y x 0 0
1 1
A B A B A B
1
C C C
over one edge of the element wherein its value is equal to one regarding the
opposite node
u = [ u(x, y) ] = Ω · v
uA
= [ φA φB φC ] uB ,
uC
140 7 Triangular Elements – Description of Geometry
v T = [ uA uB uC ]
φa = 1 − ξ − η, φb = ξ , φc = η .
Such, all shape functions are equivalent regarding the corner nodes A, B and
C, see Figure 7-3.
A (
λ a = 1, λ b = 0, λ c = 0 )
Ab Ac
yp P
C Aa
( 0, 0, 1 )
B
y ( 0, 1, 0 )
The triangle is subdivided into the three triangles Aa , Ab and Ac . The indices
correspond to the opposite corner points. The arbitrary position of P within
the triangle is set by means of the area coordinates
Aa Ab Ac
λa = , λb = , λc = , (7.9)
A△ A△ A△
7.2 Description Employing Area Coordinates 141
φA = λa , φB = λb , φC = λc .
The comparison with the shape functions depicted in Figure 7-3 shows that
they are identical.
144 7 Triangular Elements – Description of Geometry
λa , λb , λc .
A scaling to the corner and the mid points of the edges leads to, cf. Figure 7-5,
A 1
C
1
B
1
1/4
1 1/4 F
1 1/4
D
E
The basic idea when performing the element matrix as well as the load vector is
still the weak formulation of the energy balance equation applying the matrix
notation
Z Z
−δRel = δv · { Ω · D · E · D · Ω dxdy · v − ΩT · p dxdy} .
T T T
(8.1)
At first, the element matrices regarding elements with a linear temperature ap-
proach are derived, describing constant heat fluxes inside the element. Elements
employing a quadratic or even higher polynomial approach may be derived but
are not looked at here. Concerning linear shape functions, all convergence cri-
teria are fulfilled. The derivation of the element matrices does not essentially
change compared with rectangular elements, thus the presentation is focused
on the application of the local element coordinate system.
u = [T ] = Ω·v
v T = [ TA TB TC ] .
The shape functions chosen with Ω and their respective derivatives are:
φA = 1 − ξ − η φB = ξ φC = η
φA ,ξ = −1 φB ,ξ = 1 φC ,ξ = 0
φA ,η = −1 φB ,η = 0 φC ,η = 1
1 1
A B A B A B
1
C C C
whereby the negative sign is positioned in front of the matrix for convenience
here. Thus the matrix B, comprising the derivatives of the shape functions,
yields
B = D·Ω
"" # " # " # #
1 bb φA ,ξ + bc φA ,η
=− ·
det J ab φA ,ξ + ac φA ,η
A B C
8.1 A Linear Approach Related to the ξ–η–Coordinate System 147
1 −bb − bc bb bc
=− ·
det J −ab − ac A
ab B
ac C
" #
1 ba bb bc
=− · . (8.3)
det J aa ab ac
The matrix B turns out to be constant. Because the heat flux is also computed
by means of B, the heat flux is also constantly distributed inside the element.
Considering the chosen abbreviations for the coordinate differences, B may be
recognized now to be symmetric with respect to all element nodes.
It is remarkable that the order of the shape functions is reduced by one relating
to the whole element, not only in the respective direction like it does concerning
rectangular elements employing symmetric shape functions.
B is also constant here. When applying det J = 2A△ , the whole integrand
turns out to be constant. Thus, in this case, the element matrix is performed
by matrix multiplication BT · E · B analogously to the multiplication scheme
given in Section 4.2 and finally by multiplying the product by the factor
1 1
det J · = 2 A△ · = A△ .
2 2
Comparing to Section 4.2, the element matrix comprises three rows and three
columns, that are related to the virtual as well as to the real nodal temperatu-
res. This yields
b a · b a + aa · aa b a · b b + aa · ab b a · b c + aa · ac
λ
K= · b b · b a + ab · aa b b · b b + ab · ab b b · b c + ab · ac .
4A△
b c · b a + ac · aa b c · b b + ac · ab b c · b c + ac · ac
148 8 Triangular Elements to Describe Heat Conduction
p = [ qV ] = constant ,
this yields
Z Z
f= ΩT · p dA = ΩT dA · p .
A A
In the case of linear shape functions, the integral can be solved independently
of the chosen coordinate system. The external energy input is therefore equally
distributed to the three nodes by 1/3 each, because the shape functions are
equivalent
A△
3
A△
f =
· [ qV ] .
3
A△
3
A fluctuating external energy input may be described and taken into account
at performing the integration by means of the matrix of shape functions and
the nodal values of qV .
The stress matrix S follows, regarding the chosen linear shape functions to
" #
λ ba bb bc
S=E·B=− · .
2F△ aa ab ac
8.2 Example of Use 149
0.50 1.25
0.125
λ 2 = 0.12 B 0.125
22 o
0.50
6o
A λ 1 = 0.08 0.50
0.25
y
The results for linear shape functions, cf. Figures 8-3, 8-4 and 8-5, and for
quadratic shape functions, cf. Figures 8-6, 8-7 and 8-8, are plotted inside the
elements, whereat the heat fluxes are discontinous at the element intersections.
Figures 8-9, 8-10 and 8-11 represent the results after averaging them at the
nodes of neighboring elements.
The heat conduction process results in a stationary temperature field. The tem-
perature field is nearly linear in regions, where an almost linear heat conduction
without disturbance exists – see cross sections A and B. Thermal bridges de-
velop close to the outer corners, where the boundary conditions dominate the
temperature field inside the wall in two directions.
The heat fluxes qx [ W/m2 ] and qy [ W/m2 ] are correlated to the gradients of
the temperature field. Thus the heat flux is larger in the direction of the shorter
transport distance, cf. qy close to the cross sections A and B, and larger in the
region of the larger heat conduction coefficient – right part of the structure.
150 8 Triangular Elements to Describe Heat Conduction
22. 0
o
C
6. 0
5. 8
W/m2
0. 0
−0. 7
W/m2
−15. 4
22. 0
o
C
6. 0
8. 4
W/m2
−1. 2
0. 2
W/m2
−17. 2
22. 0
o
C
6. 0
8. 4
W/m2
−1. 2
Fig. 8-10 Brickwork, quadratic shape functions – averaged heat flux qx (x, y)
0. 2
W/m2
−17. 2
Fig. 8-11 Brickwork, quadratic shape functions – averaged heat flux qy (x, y)
9 Triangular Elements for Membrane Structures
In this section different approaches concerning the choice of the applied coor-
dinate system are represented when employing triangular elements for the de-
scription of plane stress linear elasticity. The procedure to derive triangular
elements for the description of linear elasticity is almost the same as for the
description of stationary heat conduction. Only the transformation of the par-
tial derivatives has to be extended to a vector notation regarding the primary
variables.
Again, basis of the development of the element stiffness matrix and the load
vector is the matrix notation of the work equation
Z Z
−δWel = δvT · { ΩT · DT · E · D · Ω dxdy · v − ΩT · p dxdy} . (9.1)
corner nodes A, B and C. Each linear function describes a plane tilted over one
edge of the element wherein its value is equal to one regarding the opposite
node. As represented in Section 8.1.1, the shape functions and their respective
derivatives are given with:
φA = 1 − ξ − η φB = ξ φC = η
φA ,ξ = −1 φB ,ξ = 1 φC ,ξ = 0
φA ,η = −1 φB ,η = 0 φC ,η = 1
1 1
A B A B A B
1
C C C
According to the plane stress situation, cf. Section 5.1, the displacements u(x, y)
and v(x, y) have to be approximated, which yields the nodal unknowns v of
linear shape functions
vT = [ uA vA uB vB uC vC ] .
Applying the differentiation rules to the shape functions yields the matrix B,
which is needed to compute the stiffness matrix,
B = D·Ω
bb φA ,ξ + bc φA ,η
1
= · ab φA ,ξ + ac φA ,η
det J
ab φA ,ξ + ac φA ,η bb φA ,ξ + bc φA ,η A B C
−bb − bc bb bc
1
= · −ab − ac ab ac
det J
−ab − ac −bb − bc A ab b b B ac b c C
ba bb bc
1
= · aa ab ac . (9.3)
det J
aa b a ab b b ac b c
The matrix B turns out to be constant. Because the strains and therefore also
the stresses are computed by means of B, they are constant inside the element,
too. Considering the chosen abbreviations for the coordinate differences, B may
be recognized now to be symmetric with respect to all element nodes.
1 1
det J · = 2 A△ · = A△ .
2 2
The element stiffness matrix may be summed up by three parts, which can be
assigned to the three different components of the elasticity matrix, as shown in
Section 5.1. With
K = K1 + K2 + K3
this yields
ba · ba ba · bb ba · bc
aa · aa aa · ab aa · ac
E bb · ba bb · bb bb · bc
K1 = · ,
4A△ (1 − ν 2 )
ab · aa ab · ab ab · ac
bc · ba bc · bb bc · bc
ac · aa ac · ab ac · ac
b a · aa b a · ab b a · ac
aa · b a aa · b b aa · b c
νE b b · aa b b · ab b b · ac
K2 = · ,
4A△ (1 − ν 2 ) ab · b a
ab · b b ab · b c
b c · aa b c · ab b c · ac
ac · b a ac · b b ac · b c
aa · aa aa · b a aa · ab aa · b b aa · ac aa · b c
b a · aa ba · ba b a · ab ba · bb b a · ac ba · bc
E ab · aa ab · b a ab · ab ab · b b ab · ac ab · b c
K3 = · .
8A△ (1 + ν) b b · aa bb · ba b b · ab bb · bb b b · ac bb · bc
ac · aa ac · b a ac · ab ac · b b ac · ac ac · b c
b c · aa bc · ba b c · ab bc · bb b c · ac bc · bc
For the same reasons as regarding rectangular elements, the sums over the row’s
components as well as over the column’s components of the element stiffness
matrix must turn out to be identical to zero.
9.1 Linear Shape Functions with Respect to ξ–η–Coordinates 157
1 1
A B A B A B
1
C C C
Analogously to Section 9.1.2 and taking into account Section 7.2.2, the operator
matrix D results in
P c
∂λi
∂λ
i=a ∂x i
∂
x
c
P ∂λi
D= ∂y =
∂
∂y λi
i=a
∂y ∂x P c c
P
∂λi ∂λi
∂
∂y λi
∂
∂x λi
i=a i=a
Pc
b ·∂
i=a i λi
c
1 P
= ai ·∂λi .
2A△
i=a
Pc c
P
ai ·∂λi bi ·∂λi
i=a i=a
9.3 Quadratic Approach Employing ξ–η–Coordinates 159
u(x,y) = a0 + a1 x + a2 y + a3 x2 + a4 xy + a5 y 2 ,
which can be checked by means of Pascale’s triangle, see Section 2.4.2. The
displacements of the corner nodes A to C and of the midpoints of the edges D
to F are chosen to represent the DOF of the physically meaningful approach,
cf. Figure 9-3.
x
A ( ξ = 0, η = 0 )
F ( 0, 0.5 )
C ( 0, 1 ) D ( 0.5, 0 )
η
E ( 0.5, 0.5 ) B ( 1, 0 )
y ξ
A 1
C
1
B
1
1/4
1 1/4 F
1 1/4
D
E
φD = 4ξ (1 − ξ − η) φE = 4ξη φF = 4η (1 − ξ − η)
φD ,ξ = 4 − 8ξ − 4η φE ,ξ = 4η φF ,ξ = −4η
φD ,η = −4ξ φE ,η = 4ξ φF ,η = 4 − 4ξ − 8η
The transformation, the partial derivatives and the notes related to the inte-
gration given in Section 9.1 are also valid here and are independent of the shape
functions. Thus the operator matrix D remains unchanged when compared to
Section 9.1.2. Because the number and particularly the courses of the shape
functions φi in Ω are changed, the differentiation rule D leads to different
derivatives regarding the matrix B:
bb φA ,ξ bb φB ,ξ ...
+ bc φA ,η + bc φB ,η ...
1
a φ
b A ξ, a φ ,
b B ξ . . .
B=D·Ω= · .
det J + ac φA ,η + ac φB ,η . . .
ab φA ,ξ bb φA ,ξ ab φB ,ξ bb φB ,ξ . . .
+ ac φA ,η + bc φA ,η + ac φB ,η + bc φB ,η ...
The matrix B is represented here only concerning the nodes A and B, for to the
nodes C to F the procedure has to be performed analogously. The derivatives
of the shape functions are represented explicitly in Figure 9-4. Since the matrix
B comprises linear functions, the subsequent stress analysis yields linear strains
and stresses inside the element domain.
... bb φD ,ξ + bc φD ,η ...
1
B= · ... 0 ... 7th column ,
2A△
... ab φD ,ξ + ac φD ,η ...
... bb φD ,ξ + bc φD ,η ...
E ...
E·B = · ν(bb φD ,ξ + bc φD ,η ) ... 7th column ,
2A△ (1 − ν 2 )
1−ν
... 2
(a b φD , ξ + a c φD , η ) . . .
.. .. ..
. . .
1
BT =
· 0 ab φE ,ξ + ac φE ,η bb φE ,ξ + bc φE ,η 10th row .
2A△
.. .. ..
. . .
The product of the 10th row and the 7th column gives
Z 1Z 1−η
1 E
K[10,7] = · ν(bb φD ,ξ + bc φD ,η )(ab φE ,ξ + ac φE ,η )
2A△ 2A△ (1 − ν 2 ) 0 0
1−ν
+ (ab φD ,ξ + ac φD ,η )(bb φE ,ξ + bc φE ,η ) · 2A△ dξdη .
2
When employing the derivatives of the shape functions given in Figure 9-4, it
follows
Z 1Z 1−η
E
K[10,7] = { ν [ bb (4 − 8ξ − 4η) + bc (−4ξ) ] [ ab 4η + ac 4ξ ]
2A△ (1−ν 2 ) 0 0
1−ν
+ [ ab (4 − 8ξ − 4η) + ac (−4ξ) ] [ bb 4η + bc 4ξ ] } dξdη .
2
Concerning integrals employing polynomials of an arbitrary order, the solution
can be analytically obtained by applying
Z 1Z 1−η
i! k!
ξ i · η k dξdη = .
0 0 (i + k + 2)!
Again, the element stiffness matrix K is divided according to the three different
parts of the elasticity matrix
K = K1 + K2 + K3 .
E
K1 = ·
4A△ (1 − ν 2 )
ba ba
0 aa aa
0 bb bb
1
ab ab . . . symmetric. . .
1
0 − 3 ab aa 0
− bb ba
bc bc
3
3
1 1
1
0 − 3 ac aa 0 − 3 ac ab 0 ac ac
− bc ba 0 − 1 bc bb 0
8/3(bbbb +
3
4
0 b a b b 0 0 0
3 b b +b b )
b a a a
8/3(a a
b b +
4
4 4
0 0 0 0 0
bb ba
3 ab aa 3 aa ab a a +a a )
b a a a
3
9.3 Quadratic Approach Employing ξ–η–Coordinates
8/3(b c b c +
4 4 8
0 0 b b 0 b b 0 b b 0
3 c b 3 b c 3 a c bc bb +bb bb )
8/3(ac ac +
4 4 8
0 0 0 a c a b 0 a b a c 0 a a a c 0
3 3 3 a a +a a )
c b b b
8/3(b b
c c +
4 8 8
0 0 0 0
0 3 ba bc 0 3 bb bc 3 ba bb b c b a +b a b a )
8/3(a c a c +
4
4 4 8 8
0 a a 0 0 0 a a 0 a a 0 a a 0
bc ba
3 c a 3 a c 3 b c 3 a b ac aa +aa aa )
3
163
164
νE
K2 = ·
4A△ (1 − ν 2 )
0
0
1
0 − 3 b b aa 0
aa b a
ab b b 0 . . . symmetric. . .
−
1 1
1
0 − 3 b c aa 0 − 3 b c ab 0
3 ab b a 0
ac b c 0
4 4
0 b b a a 0 b a a b 0 0 0
3 3
1
4/3(a a ba +
− 3 ac ba 0 − 31 ac bb 0
4
0 0 0 0 0
3 aa b b a b +a
b b c bc )
4/3(aa bc
4
4
0 0 0 0 0
ab b a
3 b c ab 0 43 bb ac
+ac ba )
3
4/3(aabc 4/3(aa ba +
4
0 0 0 0
3 ac b b 0 43 ab bc 0
+a b )
c a ab bb +acbc )
4/3(ab bc 4/3(ab ba
4
0 0 0 0 0 0
3 b c aa 0 43 ba ac
+a b )
c b +aa bb )
4
0 0 0 0 0
3 ac b a 0 3 aa b c 0
+ac bb ) +ab ba ) ab bb )+ac bc )
4
9 Triangular Elements for Membrane Structures
E
K3 = ·
8A△ (1 + ν)
aa aa
ba ba
b a aa
bb bb . . . symmetric. . .
1
− 3 ab aa − 13 ab ba ab ab
1 1
ac ac
1
− 3 bb aa − 13 bb ba bb ab
b c ac bc bc
1
− 3 ac aa − 3 ac ba − 3 ac ab − 31 ac bb
8/3(ab ab +
1
0 0
− 3 bc aa − 13 bc ba − 31 bc ab − 31 bc bb
ab aa +aa aa )
4/3(aa ba + 8/3(bb bb +
4
0 0
3 ab aa 43 ab ba 43 aa ab 43 aa bb
ab ba +acbc ) bb ba +ba ba )
9.3 Quadratic Approach Employing ξ–η–Coordinates
4 4 4 4 8 4/3(ac ba 8/3(ac ac +
4
0 0
3 bb aa 43 ba ba 43 ba ab 43 ba bb
3 ac ab 3 ac b b 3 ab ac 3 ab b c 3 aa ac +aa bc ) ac ab +ab ab )
4 4 4 4 4/3(a b
a c 8 4/3(aa ba + 8/3(bcbc +
0 0
0 0
3 b c aa 3 bc ba 0 0 3 b a ac 3 b a b c 3 bb bc 3 ba bb
4
166 9 Triangular Elements for Membrane Structures
A ( λ a = 1, λ b = 0, λ c = 0 )
F ( 0.5, 0, 0.5 )
D ( 0.5, 0.5, 0 )
C ( 0, 0, 1)
y E ( 0, 0.5, 0.5 ) B ( 0, 1, 0 )
Employing area coordinates the quadratic shape functions to describe the dis-
placements given in Figure 9-4 and their respective derivatives follow to
The integration of the individual terms follows the integration rule related to
the product terms of the area coordinates as shown in Section 7.2.4
ZZ
i! j! k!
λia · λjb · λkc dλa dλb = .
(i + j + k + 2)!
Thus the evaluation of the entry K[10,7] regarding the stiffness matrix yields
ZZ
E
K[10,7] = [ ν (4ba λb + 4bb λa )(4ab λc + 4ac λb )
2A△ (1−ν 2 )
1−ν
+ (4aa λb + 4ab λa )(4bb λc + 4bc λb ) ] · dλa dλb
2
νE 1 2 1 1
= (16ba ab + 16ba ac + 16bb ab + 16bb ac )
2A△ (1−ν 2 ) 24 24 24 24
E 1 2 1 1
+ (16aa bb + 16aa bc + 16ab bb + 16ab bc )
4A△ (1+ν) 24 24 24 24
νE E
= 2
(aa bc + ac ba ) + (aa bc + ac ba ) .
3A△ (1−ν ) 6A△ (1+ν)
Of course, the stiffness matrix K equals the stiffness matrix employing the
ξ–η–coordinates derived in Section 9.3.1.
168 9 Triangular Elements for Membrane Structures
S̃ = E · B̃ .
The coordinates related to the nodes A, B and C need to be introduced into the
matrix B, as given in Section 9.4.1. Applying the abbreviation e = (1 − ν)/2,
the stress matrix explicitly follows to
σxxa 3ba 3νaa −bb −νab −bc −νac 4bb 4νab 0 0 4bc 4νac
σyya 3νb a 3aa 4νbb 4ab 0 0 4νbc
0
3eaa 4ebc
x = l/2
x, u PY
l B
3
y, v
l
l = 15 m, t = 1 m, E = 100 000 N/m2, ν = 0, PY = 2 N/m2
1 4 7 10 13 16 19 21 24 27 30 34 37
2 5 8 11 14 17 20 22 25 28 31 35 38
3 6 9 12 15 18 21 23 26 29 32 36 39
Table 9.1 Cantilever – displacement vB for different element types and meshes
σxx
σxy
The governing equations as well as the work equation concerning the Principle
of virtual Displacements are given in Section 6.2 regarding Kirchhoff ’s plate
theory. They are independent of the element formulation, so that they can
be adopted unchanged when considering triangular elements. The coordinate
system for the description of the triangular elements for plates can be taken to
be the same as the one introduced in Section 7.
Because of the requirements necessitated by C1 –conformity and the demand of
coordinate invariance, it becomes much more difficult to choose shape functions
and nodal unknowns concerning the triangular elements for Kirchhoff ’s plate
theory when compared to the membrane elements.
The C1 –conformity is achievable, but requires high effort and provides con-
forming elements with restricted applications. Thus, in the following, various
shape functions with ascending polynomial order are investigated, whereat li-
near shape functions are not able to describe constant curvatures, since these
are 2nd order derivatives of the bending surface w(x, y).
A complete quadratic approach (n = 2) employs 6 degrees of freedom and has
been applied first by Morley [73]. As shown in Figure 10-1, there are 3 DOF
per element edge, since [ w ] is chosen in the corner nodes and the slope [ w,n ]
is considered in the centre of the edges. Although 2 · 2 + 1 = 5 DOF are
formally needed at each edge, the solution converges towards the correct value
in the case of uniform meshes. Because of this restriction, the element cannot
be generally applied. The convergence is poor because the curvature is only
constantly approached inside the element domain.
n
w
s
A
w,n w,n
F D
C E B
w w
w,n
Fig. 10-1 Triangular plate element with non–conforming quadratic shape functions
M
C B
w w,x w,y w w,x w,y
Fig. 10-2 Triangular plate element with non–conforming cubic shape functions
10.1 Choice of Shape Functions and Nodal Unknowns 175
Complete shape functions of the 4th order employ 15 degrees of freedom. There-
fore, 2 · 4 + 1 = 9 DOF are required at each element edge. If [ w w,x w,y ] are
chosen in the corner nodes and [ w w,n ] in the centres of the edges, as shown in
Figure 10-3, a maximum of 8 degrees of freedom is obtained. Thus w could be
described as a polynomial of the 4th order along the element boundary and thus
w,s would subsequently be approached as a 3rd order polynomial. The nodal
degrees of freedom w,n relating to the corners and to the centres of the edges,
enable, at most, a description of the slope w,n with a quadratic polynomial.
n
w w, x w, y
s
A
w w, n w w, n
F D
C E B
w w, x w, y w w, x w, y
w w, n
Fig. 10-3 Triangular plate element with non–conforming 4th order shape functions
Complete shape functions of the 5th order employ 21 degrees of freedom. Thus
at each element edge 2 · 5 + 1 = 11 nodal unknowns are needed to satisfy the
compatibility conditions. If [ w w,x w,y w,xx w,xy w,yy ] are chosen as physically
meaningful unknowns in the corner nodes as well as [ w,n ] in the centres of the
edges, 13 DOF are obtained per element boundary and therefore all convergence
criteria can be satisfied.
n
A v={ [ w w,x w,y w,xx w,xy w,yy ]A
s
[ w w,x w,y w,xx w,xy w,yy ]B
F [ w w,x w,y w,xx w,xy w,yy ]C
D
Fig. 10-4 Triangular plate element with conforming 5th order shape functions
Shape functions of an even higher order can be chosen, but are not efficient
as the effort needed to compute the stiffness matrix increases sharply and,
moreover, the corresponding choice of unknowns creates further difficulties.
Because of the difficulties in the fulfillment of the C1 –conformity concerning
Kirchhoff ’s plate theory, triangular plate elements have been developed, which
go beyond the standard displacement–based formulation of the Principle of
virtual Displacements, cf. Sections 17, 18, 19.
w = ψ · a. (10.1)
The vector ψ contains the monomials of the chosen general polynomial in area
coordinates. The vector a contains the coefficients a1 −a21 of the polynomial. At
first the relationship between the coefficients ai and the physically meaningful
nodal unknowns v is to be formulated. This results in
v = Ψ̃ · a . (10.2)
a= G·v (10.3)
w (λa , λb , λc ) = ψ · a . (10.5)
For an approach of the 5th order, and following the definition of area coordi-
nates given in Section 7.2, this yields
ψ = [ λ5a λ4a λb λ4a λc λ3a λ2b λ3a λb λc λ3a λ2c λ2b λc λ2a
λ5b λ4b λc λ4b λa λ3b λ2c λ3b λa λc λ3b λ2a λ2c λa λ2b (10.6)
λ5c λ4c λa λ4c λb λ3c λ2a λ3c λa λb λ3c λ2b λ2a λb λ2c ].
To derive the scaling matrix, the derivatives of the deflection surface must be
evaluated with respect to the area coordinates.
178 10 Triangular Elements for Kirchhoff Plates
This results in
w,λa = [ 5λ4a 4λ3a λb 4λ3a λc 3λ2a λ2b 3λ2a λb λc 3λ2a λ2c 2λ2b λc λa
0 0 λ4b 0 λ3b λc 2λ3b λa λ2c λ2b
0 λ4c 0 3
2λc λa 3
λc λb 0 2λa λb λ2c ] · a ,
w,λa λa = [ 20λ3a 12λ2a λb 12λ2a λc 6λa λ2b 6λa λb λc 6λa λ2c 2λ2b λc
0 0 0 0 0 2λ3b 0
0 0 0 2λ3c 0 0 2λb λ2c ] · a,
w,λc λc = [ 0 0 0 0 0 2λ3a 0
3
0 0 0 2λb 0 0 2λa λ2b
20λ3c 12λ2c λa 12λ2c λb 6λc λ2a 6λc λa λb 6λc λ2b 2λ2a λb ] · a.
10.2 Complete Approach of the 5th Order 179
The first and the second row describe the transformation of the 1st order deri-
vatives as developed in Section 7.2.2. The rows three, four and five describe the
transformation of the 2nd order derivatives, whereat the 2nd order derivatives
of the area coordinates with respect to the x–y–coordinates occur, which de-
pend on the element geometry. In the case of a linear transformation as taken
into account at hand and given in Section 7.2.1 the 2nd order derivatives of the
area coordinates disappear and lead to a simplification of the scheme as follows.
After introducing the abbreviations for the derivatives of the area coordinates,
180 10 Triangular Elements for Kirchhoff Plates
w,x
w,y
= 1 ·
w,xx 2A∆
w,yy
w,xy
w,λa
w,
λb
ba bb bc w,λc
a a a w,
a b c λa λa
2 2 2
ba bb bc 2 aa ab 2 aa ac 2 ab ac w,λb λb . (10.7)
a2a a2b a2c 2 ba bb 2 ba bc 2 bb bc w,λc λc
ba aa bb ab bc ac ba ab +aa bb ba ac +aa bc bb ac +ab bc w,λa λb
w,
λa λc
w,λb λc
w|A = w (λa = 1, λb = 0, λc = 0)
= [ 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 ] · a.
1
w,x = · [ ba · w,λa + bb · w,λb + bc · w,λc ] ,
2A∆
and, introducing the derivatives of the approach, is evaluated for node A ap-
10.2 Complete Approach of the 5th Order 181
The scaling of the approach with respect to the slopes w,n in the centres of the
edges is more time–consuming, as the derivatives are defined with respect to
the global x–y–coordinates, but these must be transformed with respect to the
local element boundary coordinates. To identically define the unknowns w,n
in neighboring elements at assembling the system, the direction of the element
boundary coordinates n, s should be defined globally. The positive direction
of the normal with respect to the edge of the triangular element considered,
will be defined such that it encloses an angle of 0 ≤ β < π with the x–axis, as
shown in Figure 10-5. The angle between the normal and the x–axis is explicitly
evaluable when considering the differences in the coordinates of the respective
corner nodes.
x
A π yb − ya bc
tan( βd − )= =−
2 xb − xa ac
F D π yc − yb ba
βf tan( βe − )= =−
2 xc − xb aa
E βd
C π ya − yc bb
βe B tan( βf − )= =−
2 xa − xc ab
y
The transformation of the slopes w,x and w,y into the slopes w,s and w,n can
be performed with respect to the element boundary, as shown in Figure 10-6,
182 10 Triangular Elements for Kirchhoff Plates
by
π π
w,s = + cos( β − ) · w,x + sin( β − ) · w,y ,
2 2
π π
w,n = − sin( β − ) · w,x + cos( β − ) · w,y .
2 2
x
w,s
w, n
D
s w,y s
(β − π)
2
td td
y
nd w,x nd
Fig. 10-6 Definition of w,n and w,s according to w,x and w,y
By means of the derivatives w,x and w,y the last three rows of the matrix Ψ̃
can be computed as follows. Due to the transformation of coordinates the slope
w,n|D satisfies
w,n|D =
π 1
− sin( β − )· [ ba · w,λa + bb · w,λb + bc · w,λc ](λa = 21 , λb = 12 , λc =0)
2 2A∆
π 1
+ cos( β − ) · [ aa · w,λa + ab · w,λb + ac · w,λc ](λa = 21 , λb = 12 , λc =0)
2 2A∆
The entire Ψ̃–matrix employs 21 rows and columns and links the coefficients of
the general polynomial with the physically meaningful nodal degrees of freedom
The Ψ̃n –matrix is completely taken and has 3 rows and 21 columns. The in-
troduction of the abbreviations
π π 1
a = {−ba · sin(βd − ) + aa · cos(βd − )} · ,
2 2 32A∆
π π 1
b = {−bb · sin(βe − ) + ab · cos(βe − )} · ,
2 2 32A∆
π π 1
c = {−bc · sin(βf − ) + ac · cos(βf − )} · ,
2 2 32A∆
and consideration of the derivatives given in Section 10.2.2 yields the rows 19,
20 and 21:
5a 4a + b c 3a + 2b c 0 c
Ψ̃n = 0 0 0 0 0 0 0
5a b c + 4a 0 b 2c + 3a 0
5b c a + 4b 0 c 2a + 3b 0
5b 4b + c a 3b + 2c a 0 a
0 0 0 0 0 0 0
0 0 0 0 0 0 0
5c a b + 4c 0 a 2b + 3c 0 .
5c 4c + a b 3c + 2a b 0 b
can be computed numerically for each element in the FE–program when apply-
ing its respective coordinates. Here, the scaling matrix is also explicitly specified
with
G11 0 0 0
0 G22 0 0
0 0 G 33 0
G= .
← g7 →
← g →
14
← g21 →
Although the vectors gi are computed as rows 7, 14, 21 of G they are embed-
ded in the scaling matrix as rows 19, 20, 21 for simplicity. By employing the
10.2 Complete Approach of the 5th Order 185
0 0 0
−5(11c + 5b + 5a) aa (16c + 9b + 4a) − 5aab −ba (16c + 9b + 4a) + 5abb
−5(11c + 5a + 5b) −ab (16c + 9a + 4b) + 5baa bb (16c + 9a + 4b) − 5bba
0 0 0
aaa ab −a2a (1. 5c + b) ba aa (3c + 2b)−a(abba + aa bb ) aba bb − b2a (1. 5c + b)
baa ab −a2b (1. 5c + a) bb ab (3c + 2a)−b(abba + aa bb ) bbb ba − b2b (1. 5c + a)
1 0 0
0 1 0 .
0 0 1
186 10 Triangular Elements for Kirchhoff Plates
1 0 0 0 0 0
5 aa −ba 0 0 0
5 −ac bc 0 0 0
G22 = ,
10 4aa −4ba 0,5a2a −ba aa 0,5b2a
20 4(aa − ac ) −4(ba − bc ) −ac aa ac b a + aa b c −bc ba
10 −4ac 4bc 0,5a2c −bc ac 0,5b2c
1 0 0 0 0 0
5 ab −bb 0 0 0
5 −aa ba 0 0 0
G33 = .
10 4ab −4bb 0,5a2b −bb ab 0,5b2b
20 4(ab − aa ) −4(bb − ba ) −aa ab aa b b + ab b a −ba bb
10 −4aa 4ba 0,5a2a −ba aa 0,5b2a
Here, Kgen incorporates the coefficients of the general polynomial and will be
transformed into the element stiffness matrix K related to the physically mea-
ningful nodal unknowns by applying the scaling matrix G. Thus integration
is performed with regard to the general polynomial. This results in simpli-
fied programs and therefore in a substantial reduction in numerical cost. The
computation of the integrand is demonstrated for two different possibilities.
10.2 Complete Approach of the 5th Order 187
1st Approach
Here, the B–matrix is not evaluated with the shape functions Ω but with the
general polynomial given with respect to the area coordinates
B = D·ψ. (10.11)
E B
3 21
3
21 3 3 21 21
K = . dA .
21
21
21
21
21
GT BT G
The method of explicit integration to get the stiffness matrix is not performed
here, since the second approach is more efficient and clearly arranged.
2nd Approach
Here, to get the element stiffness matrix, the matrix operations in the integrand
are performed in advance with D and E, as per Section 6.2,
DT · E · D =
1−ν
B · { xx ∂∂xx + yy ∂∂yy + νxx ∂∂yy + νyy ∂∂xx + 4 xy ∂∂yx } . (10.12)
2
The multiplication yields a 1×1–matrix. For example, xx ∂ describes the partial
double differentiation with respect to x as related to the shape function in
front of it, and ∂xx describes the application of partial double differentiation
188 10 Triangular Elements for Kirchhoff Plates
with respect to x as related to the shape function behind it. When applying
this representation, it is possible to compute each term in the stiffness matrix
separately by applying the related differential operator. Such, the entry in the
8th row and the 10th column of the stiffness matrix Kgen can be performed, for
example, by applying the 8th shape function from Equation (10.6) in front of
the differential operator and by applying the 10th shape function from Equation
(10.6) behind the differential operator:
8th function : λ5b
10th function : λ4b λa
Z Z
Kgen [ 8, 10 ] = λ5b · DT · E · D · λ4b · λa · 2A∆ dλa dλb .
1 1 1 2 1 2 2
[ 6·7 5·6·7 5·6·7 4·5·6·7 4·5·6·7 4·5·6·7 3·4·5·6·7
]
1 1 1 2 1 2 2
[ 6·7 5·6·7 5·6·7 4·5·6·7 4·5·6·7 4·5·6·7 3·4·5·6·7
] }·p .
The computation of the element load vector concerning linearly varying actions
is not commonly accepted and hence is not shown here.
10.3 A Plate Element with 18 DOF 189
σ =E·D·Ω·v = E·B·v
σ (λi ) = E · D · ψ (λi ) · G · v ,
D · ψ (λi )
q = Dq · σ = Dq · (E · D · ψ (λi ) · G · v) = Sq · v ,
with Sq representing the stress matrix for shear forces. Explicitly introducing
the shape functions for the deflections yields
" # " #
qx ψ (λi ) ,xxx + ψ (λi ) ,yyx
= −B · ·G·v.
qy ψ (λ ) ,yyy + ψ (λ ) ,xxy
i i
The starting point of this derivation is the 21 DOF element previously described
with the unknowns
T
v21 ={ [ w w,x w,y w,xx w,xy w,yy ]A
[ w w,x w,y w,xx w,xy w,yy ]B
[ w w,x w,y w,xx w,xy w,yy ]C
[ [ w,n ]D [ w,n ]E [ w,n ]F ] }.
Concerning the 18 DOF element, the slopes in the normal directions w,n at
the centres of the edges are described as a linear combination of the nodal
unknowns at the corner points. Therefore, the nodal unknowns
T
v18 ={ [ w w,x w,y w,xx w,xy w,yy ]A
[ w w,x w,y w,xx w,xy w,yy ]B
[ w w,x w,y w,xx w,xy w,yy ]C }
suffice. The elimination is possible only if the remaining unknowns are able to
consistently fulfill the element intersection conditions along the element edges,
i.e. without a gap in w and in w,n .
The 5th order approach of the 21 DOF element fulfills all element intersection
conditions, i.e. for the slope w,n as a polynomial of the 4th order employing 5
DOF at each edge. Eliminating w,n at the centres of the edges, 4 DOF of the
quality [ w,n w,ns ] are left at the element corners, and a Hermite Polynomial of
the 3rd order can be defined at each edge relating to w,n . Therefor, the slopes
w,n at the centres of the edges are described by a linear combination of the
nodal unknowns [ w,x w,y w,xx w,yy w,xy ] at the corner points, which have
to be transformed to the edge coordinates n, s. Thus the element intersection
conditions for [ w w,s ], described as a polynomial of the 5th order, are still
fulfilled exactly. However, w,n is only described with a polynomial of the 3rd
order instead of the 4th order, what causes a small loss in accuracy but does
not provide any other disadvantage.
At first, the general polynomial gives
w,n = a0 + a1 · s + a2 · s2 + a3 · s3 ,
which is adapted to the nodal unknowns [ w,n w,ns ] at the corner points of the
edges. In this way, the slopes at the corner points w,s and w,n are described by
means of w,x and w,y , cf. Equation (10.8). Similarly w,ns may be computed.
In general, the relationship between the nodal unknowns of the elements with
18 DOF and 21 DOF is defined by
v21 = T · v18
10.3 A Plate Element with 18 DOF 191
The first 18 rows describe the identities of the nodal unknowns at the element
corners. The relationship between the slopes at the centres of the edges and
the nodal unknowns at the corners is established only by the last three rows.
Applying the transformation matrix T, the transformation rule to get the ele-
ment with 18 DOF from the element with 21 DOF yields
It is given explicitly here and is employed after the integration of the general
polynomial instead of the scaling matrix G21 with the main diagonal sub–
matrices G11 , G22 , G33 as given in Section 10.2.4. The vectors gi comprise 18
columns and are summarised into a matrix as follows. Here ℓa , ℓb , ℓc mean the
lengths of the respective element side and β holds for the angle between the
side and the x–axis, see Figure 10-5.
g7 · c
g14 · a =
g21 · b
192 10 Triangular Elements for Kirchhoff Plates
−5(11a + 5b + 5c) −ac (16a + 9b + 4c) bc (16a + 9b + 4c)
+5cab − sin β/2 −5cbb + cos β/2
0 0 0
−5(11a + 5c + 5b) ab (16a + 9c + 4b) −bb (16a + 9c + 4b)
−5bac − sin β/2 +5bbc + cos β/2
caa ac− a2c (1,5b + a) bc ac (3b + 2a) − caa bc cbc ba −b2c (1,5b + a)
− sin β cos β lc /8 −cac ba +(1−2 sin2 β)lc /8 + sin β cos β lc /8
aaa ac− a2a (1,5b + c) ba aa (3b + 2c) − aac ba aba bc −b2a (1,5b + c)
+ sin β cos β la /8 −aaa bc −(1−2 sin2 β)la /8 − sin β cos β la /8
0 0 0
0 0 0
−5(11c + 5b + 5a) aa (16c + 9b + 4a) −ba (16c + 9b + 4a)
−5aab − sin β/2 +5abb + cos β/2
−5(11c + 5a + 5b) −ab (16c + 9a + 4b) bb (16c + 9a + 4b)
+5baa − sin β/2 −5bba + cos β/2
0 0 0
2 2
aaa ab −aa (1,5c + b) ba aa (3c + 2b) − aab ba aba bb −ba (1,5c + b)
−aaa bb +(1−2 sin2 β)la /8
− sin β cos β la /8 + sin β cos β la /8 .
baa ab −a2b (1,5c + a) bb ab (3c + 2a) − bab ba bbb ba −b2b (1,5c + a)
+ sin β cos β lb /8 −baa bb −(1−2 sin2 β)lb /8 − sin β cos β lb /8
10.4 A Comparison of Standard Plate Elements 193
G M ν = 0.2
hinged
l x = 10.0 m
l y = 10.0 m
lx
The element meshes are depicted in Figure 10-9, whereat two triangles cover
the same area as one rectangular element.
Fig. 10-9 Square plate – meshes for rectangular and triangular elements
Since the solution employing the element comprised of 18 DOF only deviates
insignificantly from the solution employing the element with 21 DOF, just one
solution is presented here.
Concerning the general polynomial the difference between the rectangular ele-
ments comprised of 12 DOF and the rectangular element comprised of 16 DOF
is explained in Section 6.2.
194 10 Triangular Elements for Kirchhoff Plates
The following figures show the convergence behavior of state variables, which
indicate the quality of the elements. It is obvious, that due to the higher po-
lynomial degree the triangular element with 18 DOF yields superior results,
whereat the quality of the moments is of the same order as of the deflection.
Nonetheless, the rectangular elements converge to the right results as well but
very slowly regarding the moments.
As mentioned above the elements with 16 DOF respectively 18 DOF employ
curvatures as nodal DOF, what restricts their application to plates without
stiffeners and without discontinuities in stiffness distribution.
% error of wM
∆ 18
* 16
* 12 - weak conformity
* 12
5
0
20 100 500 dof
% error of mxxM
40
∆ 18
* 16
30 * 12 - weak conformity
* 12
20
10
0
20 100 500 dof
-10
-20
Fig. 10-11 Square plate – the convergence behavior of the moment mxxM
% error in m yyE
-5
-10
-15
∆ 18
* 16
-20
* 12 - weak conformity
* 12
-25
-30
Fig. 10-12 Square plate – the convergence behavior of the moment myyE
ISOPARAMETRIC ELEMENTS
11 Numerical Integration
When computing the element stiffness matrix and the load vector, an important
part is the solution of integrals like
Z Z Z
f (x) dx , f (x,y) dA and f (x,y,z) dV .
x A V
The integrand g(x) · h(x) or other products of functions are also included along
with the integrand f (x).
Concerning approaches of lower order, analytical integration can still be reaso-
nably managed, but for approaches of higher order, for example those shown
in Section 10, the effort increases enormously. For different problems as descri-
bed in Section 12 the analytical integration is not possible, and thus numerical
integration must be performed. Therefor different methods can be applied. Be-
sides the Newton–Cotes method, the Gauss–Legendre quadrature is also used
prevalently, since it offers the best possible solution for minimal effort. The
derivation of the Gauss–Legendre quadrature will not be shown in detail here
but its application in the FEM context. Basis for the numerical integration
scheme is the application of a normalized element coordinate system, as has
been described previously for the triangular elements in Section 7.
xA x0 xB
l/2 l/2
ξ = -1 ξ= 0 ξ = +1
A ξ B
coordinates to the local one, if x is exchanged by ξ. Assuming f (x) and g(ξ) are
normalized shape functions to approach the displacement field, the integration
can be changed to local coordinates with
x=x
Z B ξ=1
Z
ℓ
f (x) dx = g(ξ) dξ ,
2
x=xA ξ=−1
where the integration with respect to local coordinates can be separated from
the element geometry, if the shape functions are described in local coordinates,
too.
In the following, the rules for numerical integration with respect to cartesi-
an coordinates of rectangular elements and to area coordinates of triangular
elements are explained.
f (ξ i ) f ( ξ i+1)
ξ
−1 i 0 i +1 +1
n m = 2n − 1 ξi wi
1 1 0. 0 2. 0
√
2 3 ±1/ 3 1. 0
3 5 p 0. 0 8/9
± 3/5 5/9
4 7 ±0. 3399 8104 3585 0. 6521 4515 4863
±0. 8611 3631 1594 0. 3478 5484 5137
5 9 0. 0 0. 5688 8888 8889
±0. 5384 6931 0106 0. 4786 2867 0499
±0. 9061 7984 5939 0. 2369 2688 5056
6 11 ±0. 2386 1918 6083 0. 4679 1393 4573
±0. 6612 0938 6466 0. 3607 6157 3048
±0. 9324 6951 4203 0. 1713 2449 2379
This is true, since the product of two functions gives a new function, which has
to be integrated numerically by the same integration rule.
Example
Analytical Integration:
Z+1 Z+1
1 1 2
f (ξ) · g(ξ) dξ = (1 − ξ) · (ξ 2 ) dξ = [ ξ 3 − ξ 4 ]+1
−1 = .
3 4 3
−1 −1
202 11 Numerical Integration
√
Numerical integration with n = 2, ξi = ±1/ 3, wi = 1. 0:
Z+1
1 1 1 1 2
(1 − ξ) · (ξ 2 ) dξ = 1 · [ (− √ )2 − (− √ )3 ] + 1 · [ (+ √ )2 − (+ √ )3 ] = .
3 3 3 3 3
−1
Multi–dimensional integration
Similarly, integrals can be solved regarding coordinates in two directions. The
positions of the supporting points are shown in Figure 11-3 for n = 3 in ξ– and
η–directions. Here the summation is to be performed in both directions
Z1 Z1 n X
X n
f (ξ, η) dξ dη = f (ξi , ηj ) · wi · wj . (11.2)
−1−1 i=1 j=1
1 1
1 3/5
ξ
1 3/5
3/5 3/5
η
Fig. 11-3 Rectangular domains – supporting points regarding a square at n = 3
Triangles
Concerning the triangular elements, the integration is done using a single sum-
mation regarding the supporting points and the weighting factors, as shown in
Table 11.2
Z1 1−η
Z n
1 X
f (ξ, η) dξ dη = · f (ξi , ηi ) · wi . (11.4)
2 i=1
0 0
11.1 Numerical Integration Using Gauss–Legendre Quadrature 203
1 1 1
1 a a 3 3 3
1
C
B
A 1 1 1
a 2 2
0 3
c a 1 1 1
2 b 0 2 2 3
C c 1
0 1 1
b B 2 2 3
1 1 1
A
a 3 3 3
− 27
48
b 25
b 0. 6 0. 2 0. 2 48
3 a 25
c c 0. 2 0. 6 0. 2 48
C 25
d B d 0. 2 0. 2 0. 6 48
1 1 1
a 3 3 3
0. 225 000 000 0
b α1 β1 β1
A
e c β1 α1 β1 0. 132 394 152 7
c
5 a d d β1 β1 α1
f
C g e α2 β2 β2
b B
f β2 α2 β2 0. 125 939 180 5
g β2 β2 α2
xo x
2 lx
A (−1;−1) B(1;−1)
yo 2 ly ξ
D (−1;1) C ( ξ = 1; η = 1)
η
y
must also be given in local coordinates. In the following, symmetric linear and
quadratic shape functions and its derivatives are given with respect to local
coordinates.
A 1
2
1
4 1
D B
η ξ
C
vT = [ uA uB . . . uH uI | vA vB . . . vH vI ]
are subsequently given for the element employing quadratic shape functions.
The vector of the shape functions
φ = [ φA φB φC φD φE φF φG φH φI ]
1 1
φA = 4
ξ (1 − ξ) · η · (1 − η) φA,ξ = 4
(1 − 2ξ) · η · (1 − η)
φA,η = 14 ξ (1 − ξ)(1 − 2η)
φB = − 14 ξ (1 + ξ) · η · (1 − η) φB,ξ = − 14 (1 + 2ξ) · η · (1 − η)
φB,η = − 14 ξ (1 + ξ)(1 − 2η)
1
φC = 4
ξ (1 + ξ) · η · (1 + η) φC,ξ = 14 (1 + 2ξ) · η · (1 + η)
φC,η = 14 ξ (1 + ξ)(1 + 2η)
φD = − 14 ξ (1 − ξ) · η · (1 + η) φD,ξ = − 14 (1 − 2ξ) · η · (1 + η)
φD,η = − 14 ξ (1 − ξ)(1 + 2η)
φE = − 12 (1 − ξ 2 ) · η · (1 − η) φE,ξ = − 12 (−2ξ) · η · (1 − η)
φE,η = − 12 (1 − ξ 2 )(1 − 2η)
1
φF = 2
ξ (1 + ξ)(1 − η 2 ) φF,ξ = 12 (1 + 2ξ)(1 − η 2 )
φF,η = 12 ξ (1 + ξ)(−2η)
1
φG = 2
(1 − ξ 2 ) · η · (1 + η) φG,ξ = 12 (−2ξ) · η · (1 + η)
φG,η = 12 (1 − ξ 2 )(1 + 2η)
φH = − 12 ξ (1 − ξ)(1 − η 2 ) φH,ξ = − 12 (1 − 2ξ)(1 − η 2 )
φH,η = − 12 ξ (1 − ξ)(−2η)
φI = (1 − ξ 2 )(1 − η 2 ) φI,ξ = − 2 ξ (1 − η 2 )
φI,η = (1 − ξ 2 )(−2η) .
A
H E
1
I
D B
G F
η ξ
C
∂x ∂y ∂y ∂x
Here it holds det J = ∂ξ
· ∂η
− ∂ξ
· ∂η
= ℓx · ℓy .
and thus
φ,ξ 0
1 0 0 0 " #
J−1 0
φ,η 0
B=D·Ω= 0 0 0 1 · · (11.10)
0 J−1 0 φ,ξ
0 1 1 0
0 φ,η
or, briefly,
B = D · Ω = B1 · B2 · B3 . (11.11)
The elasticity matrix E, given in Section 5.1.3, and the matrix B, subdivi-
ded into three parts, yield the element stiffness matrix with respect to local
coordinates
Z Z 1Z 1
K= BT · E · B · dA = BT · E · B · det J dξdη . (11.12)
A −1 −1
B2 , B1 and E as well as det J = ℓx ·ℓy were discussed in Section 11.2.2. All the
terms are constant inside an element and hence they have the same numerical
value at each supporting point of the numerical integration. Thus the matrix
multiplication
can be performed in advance, what yields the element related constant matrix
ℓy
ℓ
0 0 ν
x
E 0 1−ν · ℓx 1−ν
0
2 ℓy 2
Ê = · . (11.14)
1 − ν2 0 1−ν 1−ν ℓy
· 0
2 2 ℓx
ℓx
ν 0 0 ℓ y
11.2 Numerical Integration Applied to Membrane Elements 209
The sequence of the nodal unknowns yields a clear partition of the matrix B3
and the stiffness matrix K into 4 sub–matrices. Table 11.3 illustrates the arran-
gement of the integrand of Equation (11.15) employing linear shape functions.
B3
φA,ξ φB,ξ φC,ξ φD,ξ 0 0 0 0
φA,η φB,η φC,η φD,η 0 0 0 0
0 0 0 0 φA,ξ φB,ξ φC,ξ φD,ξ
0 0 0 0 φA,η φB,η φC,η φD,η
ê11 0 0 ê14 ê11 φA,ξ ê11 φB,ξ . . . ê14 φA,η ê14 φB,η . . .
0 ê22 ê23 0 ê22 φA,η ê22 φB,η . . . ê23 φA,ξ ê23 φB,ξ . . .
0 ê32 ê33 0 ê32 φA,η ê32 φB,η . . . ê33 φA,ξ ê33 φB,ξ . . .
ê41 0 0 ê44 ê41 φA,ξ ê41 φB,ξ . . . ê44 φA,η ê44 φB,η . . .
Ê Ê · B3
φA,ξ φA,η 0 0
φB,ξ φB,η 0 0 (BT3 · Ê · B3 )δu u (BT3 · Ê · B3 )δu v
φC,ξ φC,η 0 0
φD,ξ φD,η 0 0
0 0 φA,ξ φA,η
0 0 φB,ξ φB,η (BT3 · Ê · B3 )δv u (BT3 · Ê · B3 )δv v
0 0 φC,ξ φC,η
0 0 φD,ξ φD,η
BT3 BT3 · Ê · B3
Table 11.3 Matrix scheme of the integrand to compute the stiffness matrix
11 Numerical Integration
11.2 Numerical Integration Applied to Membrane Elements 211
σ̃ = E · B̃ · v = S̃ · v
S̃ = E · B̃ = E · B1 · B2 · B̃3 = Ẽ · B̃3 .
E 1
Ê = · ·
1 − ν 2 det J
b2b + e a2b b b b c + e ac ab (e + ν) bb ab e bc ab + ν ac bb
b b b c + e ac ab
e a2c + b2c e bb ac + ν bc ab (e + ν) ac bc
· .
(e + ν) bb ab e bb ac + ν bc ab
e b2b + a2b e b b b c + ac ab
e bc ab + ν ac bb (e + ν) ac bc e bb bc + ac ab a2c + e b2c
11.3 Triangular Elements 213
When assembling B3 , three shape functions for the linear approach and six
shape functions for the quadratic approach are to be considered for u and v
respectively. Therefore, the integrand of the stiffness matrix
BT3 · Ê · B3
S̃ = E · B̃ = E · B1 · B2 · B̃3 = Ẽ · B̃3 ,
with
Ẽ = E · B1 · B2 .
Here, Ẽ is computed for the whole element to be evaluated
bb bc ν ab ν ac
E 1
Ẽ = · · ν bb ν bc ab ac . (11.16)
1 − ν 2 det J 1−ν 1−ν 1−ν 1−ν
2
ab 2
ac 2
bb 2
bc
The stiffness matrix regarding the general polynomial and the corresponding
load vector have to be transformed to the physically meaningful nodal unknows
by means of the scaling matrices G18 and G21 , which are not affected by the
numerical integration. The computation of the scaling matrix and the deter-
mination of the globally positive direction of the normal vector with respect
to each edge of the element is presented explicitly in Sections 10.2.4 and 10.3.
The matrix multiplications
K = GT
18 · Kgen · G18 as well as f = GT
18 · fgen
214 11 Numerical Integration
are performed after the numerical integration of the general polynomial has
been executed. With constantly distributed external actions, the integration
can also be performed explicitly, see Section 10.2.6.
rectangle quadrilateral
x x
(1;−1)
(−1;−1) (1;−1) linear (−1;−1)
functions
y ξ y ξ
x xa xe xb x
quadr.
y E B
functions a A
I F
yh H ξ
y ξ
yd D C
G
η
η
y
90o − ε ≤ ϕ ≤ 90o + ε,
x = φ(ξ, η) · x̃ ,
(12.1)
y = φ(ξ, η) · ỹ .
φ(ξ, η) = [ φA φB φC φD ... ]
and x̃ as well as ỹ represent the vectors of the coordinates of the element nodes
xa ya
xb yb
xc
ỹ = yc .
x̃ = ,
xd yd
.. ..
. .
12.1 Description of the Element Geometry 217
The matrix entries of the Jacobian matrix J may be computed applying Equa-
tions (12.1):
∂x ∂y
= φ,ξ · x̃ , = φ,ξ · ỹ ,
∂ξ ∂ξ
∂x ∂y
= φ,η · x̃ , = φ,η · ỹ .
∂η ∂η
In general the Jacobian matrix follows with
φA,ξ xa φA,ξ ya
+ φB,ξ xb + φB,ξ yb
" #
φ,ξ · x̃ φ,ξ · ỹ + ... + ...
J= = , (12.2)
φ,η · x̃ φ,η · ỹ φA,η xa φA,η ya
+ φB,η xb + φB,η yb
+ ... + ...
which yields the inverse of the Jacobian matrix
" # " #
i11 i12 −1 1 i22 −i12
J= , J = ·
i21 i22 det J −i21 i11
where k runs over all nodes and shape functions inside the element respectively.
In principle, the Jacobian matrix which is still dependent on ξ and η is not
constant. Hence, J−1 and det J cannot be evaluated for the entire element
but only for each separate coordinate pair ξi and ηj . Thus the integration
cannot be performed analytically any longer, and numerical integration needs
to be applied, e.g. the Gauss–Legendre quadrature. Normally, the number of
supporting points is chosen as in the case of a rectangular element.
218 12 Isoparametric Elements
J, det J , J−1 , Ê
yields
Ê = BT2 · E1 · B2 · det J .
Similarly, at the computation of the stress matrix, only part of the matrix
multiplications to get Ẽ can be performed independently of the coordinates, at
which the stresses are to be evaluated. Again, the stresses may be determined
at both the supporting points or at the element nodes, cf. Section 11.2.3. Thus
the stress matrix, analogously taken from Section 11.2.3, yields
When employing a linear approach for the displacements, this can be achieved
if the terms, resulting from the entries (1,2), (1,3), (2,1), (2,2), (2,3), (2,4), (3,1),
(3,2), (3,3), (3,4), (4,2) and (4,3) of the Ê–matrix, are integrated numerically
applying a single supporting point. It means that the parts of the integrand
with respect to the terms u,y and v,x must be described to be constant inside
the element. The results given in Section 12.2.3 illustrate the improvements
concerning accuracy.
220 12 Isoparametric Elements
The reduced integration does not need to be considered at the subsequent stress
analysis, since correct nodal displacements are automatically followed by the
evaluation of correct stresses.
C
E = 1.0 N/mm 2
ν = 0.33
a
A a = 44 mm
b = 16 mm
c = 48 mm
Tables 12.2 and 12.3 present the convergence of the stresses σxx at the positions
A and B. The convergence behavior already noted regarding the displacements
is more clearly detected here. The best convergence can be observed for the
9–node element by applying selectively reduced integration, cf. Section 12.2.2.
9×9 17 × 17 33 × 33 65 × 65
El left right left right left right left right
1 0. 1907 0. 0947 0. 1676 0. 1149 0. 1499 0. 1228 0. 1397 0. 1261
2 0. 1497 0. 1619 0. 1411 0. 1468 0. 1353 0. 1382 0. 1321 0. 1335
3 0. 1349 0. 1438 0. 1298 0. 1333 0. 1289 0. 1300 0. 1287 0. 1291
4 0. 1546 0. 1412 0. 1354 0. 1328 0. 1302 0. 1299 0. 1291 0. 1290
5 0. 0987 0. 0960 0. 1126 0. 1154 0. 1204 0. 1230 0. 1245 0. 1261
6 0. 1207 0. 1435 0. 1263 0. 1331 0. 1281 0. 1299 0. 1285 0. 1290
222 12 Isoparametric Elements
9×9 17 × 17 33 × 33 65 × 65
El. left right left right left right left right
1 0. 1610 0. 1852 0. 1759 0. 1917 0. 1808 0. 1893 0. 1824 0. 1867
2 0. 1779 0. 1609 0. 1805 0. 1736 0. 1816 0. 1786 0. 1825 0. 1810
3 0. 1808 0. 1894 0. 1830 0. 1846 0. 1832 0. 1835 0. 1832 0. 1832
4 0. 1819 0. 1921 0. 1833 0. 1844 0. 1833 0. 1834 0. 1832 0. 1832
5 0. 1619 0. 1452 0. 1759 0. 1694 0. 1808 0. 1779 0. 1823 0. 1810
6 0. 1790 0. 1848 0. 1828 0. 1839 0. 1831 0. 1834 0. 1832 0. 1832
As known from rectangular elements the jump of the stresses at the intersection
of neighboring elements is an indication of the discretization error and the
convergence behavior. As expected, the convergence of stresses develops poorly
compared to the convergence of displacements. Because of the non–rectangular
element geometry, which normally yields poor shear stress approximations, the
distributions of the other stresses are also badly influenced. Furthermore, the
fulfillment of the Neumann boundary conditions needs a combination of all
stresses, what includes a transformation of the equilibrium conditions at the
boundary.
The course of the stresses σxx at the cross–section A–B is shown in Figure 12-3
for the mesh sized with 65 × 65 nodes. Due to the oblique boundaries at the
upper and the lower edge of the cantilever the course is not linear but slightly
curved.
B
48
x
B
16
y
44
sxx
- 0.1
- 0.2
0.2
0.1
0.0
Fig. 12-3 Cook’s cantilever – stress σxx at section A–B regarding 65 × 65 nodes
12.3 Quadrilateral Plate Elements 223
Figure 12-4 gives the distributions of the stresses σxx , σyy , σxy over the whole
structure.
Stress σxx Stress σyy Stress σxy
−0. 20 N/mm2 + 0. 20
A
y x (-1,-1)
z, w (1,-1)
B
(-1,1) η
D ξ
w
C
Fig. 12-5 Local coordinates of a quadrilateral plate element
224 12 Isoparametric Elements
At both the development of the stiffness matrix and the scaling matrix, and in
contrast to the membrane elements, the 1st as well as the 2nd order derivatives
of the deflection need to be transformed because of the curvature. The 1st and
the 2nd order derivatives with respect to the local coordinates ξ, η are trans-
formed into the respective derivatives with respect to the global coordinates x
and y altogether by matrix operation, cf. the following scheme
∂x ∂w
∂w ∂y
∂ξ ∂ξ ∂ξ ∂x
∂w ∂x ∂y ∂w
∂η ∂η ∂η ∂y
2 2 2 ∂2w
∂ w ∂ x ∂ y ∂x 2 ∂y 2 ∂x ∂y
∂ξ 2
=
∂ξ 2 ( ) ( ) 2( ) .
∂ξ 2 ∂x2
∂ξ ∂ξ ∂ξ ∂ξ
∂2w ∂2x 2 2
∂ y ∂x 2 ∂y 2 ∂x ∂y ∂ w
( ) ( ) 2( )
∂η 2 2 2 ∂y 2
∂η ∂η ∂η ∂η ∂η ∂η
2 2 2
∂ w ∂ x ∂ y ∂x ∂x ∂y ∂y ∂x ∂y ∂y ∂x ∂ 2 w
( ) ( ) ( + )
∂ξ∂η ∂ξ∂η ∂ξ∂η ∂ξ ∂η ∂ξ ∂η ∂ξ ∂η ∂ξ ∂η ∂x∂y
torily applicable, if regular element meshes are generated without any kinks at
the mesh lines.
w(x,y) = a1 + a2 ξ + a3 η
+ a4 ξ 2 + a5 ξη + a6 η 2
+ a7 ξ 3 + a8 ξ 2 η + a9 ξη 2 + a10 η 3
+ a11 ξ 3 η + a13 ξη 3 . (12.6)
In a first step the stiffness matrix and the load vector should be integrated
concerning the general polynomial. Four supporting points are needed in each
direction, because of the cubic order of the approach. The scaling of the co-
efficients of the general polynomial to the physically meaningful unknowns is
performed afterwards by means of the multiplication of the stiffness matrix as
well as of the load vector by the scaling matrix G12 .
The scaling matrix for the element with 12 DOF can be derived for the isopa-
rametric quadrilateral element with
[ w(x,y) ] = ψ 12 · a12
v12 = Ψ̃12 · a12
and
−1
a12 = Ψ̃12 · v12 = G12 · v12 .
Thus the approach for the element with 12 DOF is established by the scaling
matrix developed on the next pages. Since the continuity conditions of the rec-
tangular element with 12 DOF are not fulfilled as discussed in Section 6.3, it is
obvious, that the continuity conditions at the intersections of the isoparametric
element are not satisfied as well. Nonetheless, the element converges against the
correct solution.
The scaling matrix G12 with respect to the nodal degrees of freedom
vT = [ w w,x w,y ]A [ w w,x w,y ]B [ w w,x w,y ]C [ w w,x w,y ]D
may be developed by means of the Jacobian matrix and the following procedure.
Taking into account a quadrilateral element as represented in Figure 12-5 the
element geometry is described by the differences of the coordinates
a1 = 0. 25 · (−xA + xB + xC − xD )
a2 = 0. 25 · (+xA − xB + xC − xD )
a3 = 0. 25 · (−xA − xB + xC + xD )
b1 = 0. 25 · (−yA + yB + yC − yD )
b2 = 0. 25 · (+yA − yB + yC − yD )
b3 = 0. 25 · (−yA − yB + yC + yD ) .
For simplicity the entries of the inverse are named by ikl , cf. Section 12.1.
Since the Jacobian matrix is developed with respect to the transformation of
derivatives
∂ ∂
∂x i11 i12
= ∂ξ
∂ ∂
∂y
i21 i22 ∂η
it can be applied to describe the nodal degrees of freedom w,x and w,y by means
of the unknowns a12 of the general polynomial (12.6). The complete Ψ̃–matrix
is represented on the next page, whereat the rows have to be evaluated with
respect to the local coordinates of the respective node.
228
v = Ψ̃12 · a ,
1 −1 −1 1 1 1 −1 −1 −1 −1 1 1
0 i11 i12 3i11 2i11 + i12 i11 + 2i12 3i12
1 1 1 1
1 1 −1 −1 −1 −1 −1 −1
node A
−i11 + i12 −2i12 3i11 −2i11 + i12 i11 − 2i12 −3i11 + i12 −i11 + 3i12
−i21 + i22 −2i22 3i21 −2i21 + i22 i21 − 2i22 −3i21 + i22 −i21 + 3i22
Ψ̃12
1 1 1 1 1 1 1 1 1 1 1 1
0 i11 i12 2i11 i11 + i12 2i12 3i11 2i11 + i12 i11 + 2i12 +3i12 3i11 + i12 i11 + 3i12
node B
=
0 i21 i22 2i21 i21 + i22 2i22 3i21 2i21 + i22 i21 + 2i22 3i22 3i21 + i22 i21 + 3i22
1 1 1 1
1 −1 1 −1 −1 −1 −1 −1
−2i11 i11 − i12 3i11 −2i11 + i12 i11 − 2i12 3i11 − i12 i11 − 3i12
node D
−2i21 i21 − i22 3i21 −2i21 + i22 i21 − 2i22 3i21 − i22 i21 − 3i22
12 Isoparametric Elements
HYBRID QUADRILATERAL ELEMENTS
13 Hybrid Finite Elements
Applying the mixed formulation of the virtual work the displacements and the
stresses are employed both as primary variables. Thus shape functions are to
be chosen for the displacement variables as well as for the stress variables.
Thereby it is advantageous that the stress variables are directly computed and
no subsequent analysis is needed. Moreover, in particular regarding bending
problems, approaches of less polynomial order can be chosen concerning the
displacements likewise fulfilling the convergence criteria. A drawback might re-
sult from a higher number of degrees of freedom at every node.
x A interface
y z
a
B
d element domain b
D
c element boundary
C
Fig. 13-1 Element domain and interface
The classification into element domain and element interface results in multiple
possibilities at formulating and discretizing the work equations. Displacement
as well as stress fields are defined relating to the element domain and to the
element interfaces, therewith the conditions of equilibrium and of deformation
are to be fulfilled. Due to the different parts of the element, which are adressed
as domain and as interfaces, the formulation is named as hybrid element me-
thod. In total three different hybrid formulations are defined in literature, cf.
Figure 13-2:
1. HMM The hybrid–mixed model, whereby conditions of equilibrium as
well as of deformation are weakly fulfilled in the element domain. The
hybrid–mixed model comprises displacement as well as stress variables as
variables inside the element.
2. HSM The hybrid–stress model, whereby the conditions of equilibrium
are exactly fulfilled inside the element domain. This is equivalent to the
force method in structural analysis.
3. HDM The hybrid–displacement model, whereby the conditions of defor-
mation are exactly fulfilled inside the element domain. This is equivalent
to the displacement method in structural analysis.
The denotation goes by the degrees of freedom defined inside the element do-
main. Concerning the element interfaces normally displacement variables are
defined as primary variables of the system.
p(x)
He
c, EA
x, u
l
Fig. 13-3 Bar loaded by constantly distributed and concentrated external action
The governing equations are given already in Section 1.3.1, whereby hereinafter
the condition of equilibrium is extended by a tangential bedding force Fc . The
material equations describe the bedding force with the bedding modulus c and
the impressed strains εT from heating by T0 with the coefficient of thermal
expansion αT .
13.1 Mixed Formulation of Governing Equations 235
a) Kinematics
b) Equilibrium
c) Material Equations
fc − c · u = 0 , (13.3)
εel − N/EA = 0 , (13.4)
εT − αT · T0 = 0 . (13.5)
Condition of Deformation
Applying the formulation employing the displacements as primary variables,
the material equations as well as the governing equations related to kinematics
are introduced into the PvD and hence into the condition of equilibrium. Re-
garding the mixed formulation only the material equation related to bedding
is introduced into the condition of equilibrium.
The condition, which enforces the equality of strains due to heating and elasti-
city and due to kinematics, is named condition of deformation:
ε = εel + εT . (13.6)
Introducing the kinematics, Equation (13.1), and the material equations (13.4,
13.5) yield the condition of deformation in the domain, which is employed
to connect the displacements with the stress variables, and the displacement
boundary conditions:
1
u,x − · N − αT · T0 = 0 and u − ue = 0 . (13.7)
EA
Equilibrium (13.2) and the condition of deformation (13.7) are summarized
applying the matrix vector notation:
" # " # " # " #
−c ∂x u p 0
1 · + = . (13.8)
∂x − EA N −αT · T0 0
236 13 Hybrid Finite Elements
The internal work terms get a negative sign here, since the internal virtual force
variables and the strains act against each other, cf. Equation (13.9). Replacing
the real strains by the material equation, yields in a first step
Z ℓ
N
− δN · (u,x − − αT · T0 ) dx + [ δN · (u − ue ) ]bound. = 0 .
0 | EA {z }
−ε
Regarding the integral of the internal work the condition of deformation (13.7)
can be recognized now, including the displacement boundary condition in weak
form. The request to exactly fulfill the displacement boundary condition by the
approach itself yields the work equation of the PvF in the following represen-
tation, which is applied for the discretization:
Z ℓ
N
− δN · (u,x − − αT · T0 ) dx = 0 . (13.10)
0 EA
Concerning the work equations of the PvD and of the PvF, the displacements
and the forces are independent variables. Thus for both independent approaches
can be chosen in the context of finite element methods. Nonetheless, the boun-
dary conditions at the boundary of the structure and the interface conditions
between two elements are fulfilled in a different manner.
Choosing linear shape functions to describe the displacements and the forces,
already known from the formulation solely employing displacements as primary
238 13 Hybrid Finite Elements
variables, cf. Section 2.3.2 and Figure 13-4, u(x) and N (x) may be scaled to
nodal values as follows
u(x) = φA (x) · uA + φB (x) · uB ,
N (x) = φA (x) · NA + φB (x) · NB .
x x
φA =1-x/l φB =x/l
Concerning the virtual states δu, δN , the same shape functions are chosen
δu(x) = φA (x) · δuA + φB (x) · δuB ,
δN (x) = φA (x) · δNA + φB (x) · δNB .
Employing linear shape functions to describe N (x) according to the nodal de-
grees of freedom NA und NB , the conditions at the element interfaces Ni =
Ni+1 result in the fact, that concentrated external loadings must not act at
element interface nodes but are only reasonable at system boundaries.
stresses σ = Ωs · s σ = [ N (x) ]
Ωs = [ φA φB ]
sT = [ NA NB ]
δσ = Ωs · δs δsT = [ δNA δNB ]
kinematics ǫ =D·u D = [ ∂x ]
strains ǫ = ǫel + ǫT ǫ = [ǫ]
1
material equation ǫel = E−1 · σ E−1 = EA
strains due to heating ǫT = [ αT · T0 ]
bedding C = [c]
external action p = [ px ]
Bv = D · Ωv ,
BTv = (D · Ωv )T = ΩTv · DT ,
and may be applied in the same meaning. The work equations (13.11) and
(13.12) may be summarized applying the matrix notation, line by line extrac-
ting the virtual displacements and the virtual stresses:
Z " T # " # Z " T #
T T Ωv C Ωv ΩTv DT Ωs v Ωv p
− δW = δv δs { T T −1
dx · − T dx} .
Ωs D Ωv −Ωs E Ωs s Ω s ǫT
| {z } | {z }
element matrix load vector
Since the heat conduction is not included as part of the work equations, the
strains εT are considered as external action in the load vector.
240 13 Hybrid Finite Elements
Element Matrix
Concerning a generally applicable notation the representation arranging stres-
ses behind displacements, is convenient. However, applying the notation to
special element formulations, the arrangement with respect to nodal degrees of
freedom is more reasonable. Therewith the element matrix related to the bar
and employing linear shape functions follows to:
δuA : c·ℓ 1 c·ℓ 1 uA
3 − 2 6
− 2
δNA : 1 ℓ 1 ℓ N
− 2 − 3EA 2 − 6EA A
· . (13.13)
c·ℓ 1 c·ℓ 1
δuB : 6
2 3 2
u B
1 ℓ 1 ℓ
δNB : − 2 − 6EA 2 − 3EA NB
1 2 3 4 5
Fig. 13-5 Bar – geometry, loading, and discretization with four elements
13.2 Mixed Formulation of Work Equations 241
Employing linear shape functions for displacements and stresses the discreti-
zations with four, five, six, and seven elements yield the study of convergence
represented in Table 13.2, whereat the boundary condition of the force N |ℓ is
fulfilled in a weak sense.
number of nodes 1 2 3 4 5 6 7 8
N [ kN ] 5. 00 3. 75 2. 50 1. 25 0. 00
u [m] 0. 00 0. 95 1. 56 1. 99 2. 083
N [ kN ] 4. 90 4. 10 2. 90 2. 10 0. 90 0. 10
u [m] 0. 00 0. 77 1. 34 1. 76 2. 02 2. 083
N [ kN ] 5. 00 4. 16 3. 33 2. 50 1. 66 0. 83 0. 00
u [m] 0. 00 0. 65 1. 15 1. 58 1. 85 2. 04 2. 083
N [ kN ] 4. 95 4. 34 3. 52 2. 91 2. 09 1. 48 0. 66 0. 05
u [m] 0. 00 0. 56 1. 02 1. 41 1. 71 1. 92 2. 05 2. 083
The displacements at the nodes become partly the same values as at employing
the displacement-based element, whereby the displacements by chance partly
match the exact values.
The stresses approach the linear course of the exact solution quite well, depen-
ding on the number of elements. The stresses match the exact values in the
case of an equal number of elements but oscillate around the exact solution in
the case of an uneven number of elements, since the condition of deformation
Z
N
δN (u,x − ) dx = 0
EA
is fulfilled in a weak sense. This indicates some inconsistencies at the element
formulation. Nonetheless the element converges against the exact solution.
Oscillations may be avoided, if the approaches for the displacements and the
stresses are adapted to each other. This means, that a linear approach of the
displacements needs a constant approach for the stresses, and a quadratic ap-
proach of the displacements needs a linear approach of the stresses. In these
cases the results have the same quality as at employing the displacement–based
elements. Nonetheless, if the order of the shape functions for N is lower than the
242 13 Hybrid Finite Elements
order of the shape functions for the displacements, the number of element no-
des does not coincide with the number of stress variables. This means, that the
stresses can be dealt with as internal variables as present in the displacement–
based formulation.
The disadvantages concerning the bar element in the mixed formulation may
be summarized as follows:
A B
x
Fig. 13-6 Element of a bar loaded by distributed and nodal external actions
the system boundary. The interface condition related to the forces is fulfilled
strongly or weakly depending on the approach. Applying the work equation
related to the PvF (13.10) the conditions of deformation are weakly fulfilled
with respect to the domain but the boundary as well as the interface conditions
are strongly fulfilled concerning the displacements.
Applying hybrid elements both types of work equations are discretized in the
domain employing shape functions without fulfillment of the element interface
conditions. The element interface conditions related to kinematics and equili-
brium are separately looked at.
A Na a b Nb B
u E(x), NE(x)
Applying the PvD to the element domain including element boundaries yields
Z
δWd = {−δεE · NE − δuE · c · uE + δuE · p(x)} dx |domain
+ [−δua · Na + δub · Nb ]el.bound. = 0 . (13.15)
At the element intersection A between the elements j − 1, j and at the ele-
ment intersection B between the elements j, j + 1 the PvD describes the work
performed by the forces Na , Nb on the virtual displacements δuA , δuB of the
intersections, which are independent of δua , δub so far:
h i
δWd, A = δuA · −Nbj−1 + Naj + HA e
= 0, (13.16)
h i
δWd, B = δuB · −Nbj + Naj+1 + HB e
= 0. (13.17)
Equations (13.16) and (13.17) fulfill the equlibrium at the system level in a
weak sense, if the forces Na , Nb are replaced by the variables at the element
domain of the neighboring elements employing Equation (13.15). Due to the
elimination of the forces Na , Nb no system variables exist, which are related
to longitudinal forces at the interfaces. Thus, at the element level, introducing
kinematics into the PvD yields
Z
− δWd = {δu,x E · NE + δuE · c · uE − δuE · p(x)} dx 6= 0 , (13.18)
Applying this formulation of the virtual work, the element interface conditions
related to the longitudinal forces can only be fulfilled weakly, since the longi-
tudinal forces are defined solely at the element level and not at the interfaces,
cf. Equations (13.16) and (13.17). Thus no direct coupling to the neighboring
elements is present concerning N and δN , hence the PvF is to be arranged at
the element level only.
In this form the equation can be used to eliminate sE from the equation apply-
ing the PvD. Out of it follows a representation, only incorporating the degrees
of freedom at the system level v:
Z
T
− δWd = δv ΩTv C Ωv dx
Z Z Z
+ ΩTv DT Ωs dx · ΩTs E−1 Ωs dx }−1 · { [ ΩTs D Ωv · v − ΩTs εT ] dx
Z
− δvT ΩTv p dx . (13.24)
246 13 Hybrid Finite Elements
v T = [ uA uB ] at system level,
and sTE = [ Na Nb ] at element level,
13.3 Hybrid Discretization of Work Equations 247
which satisfy the continuity of the displacements but not of the stresses, gives
the element matrix according to Equation (13.21)
Hvv HTsv
cℓ cℓ
δuA : 3 6
− 12 − 12 uA
δuB :
cℓ cℓ 1 1
6 3 2 2
uB
· . (13.28)
−1 1 ℓ ℓ
δNa : 2 2
− 3EA − 6EA Na
ℓ ℓ
δNb : − 12 1
2
− 6EA − 3EA Nb
Hsv −F
The elimination of the longitudinal forces sE at the element level is performed
following Equations (13.25) and (13.26) with
" # " # " # " # " #
1 1
ℓ 3 6 Na − 12 12 uA 1 εT ℓ
− · 1 1 · + · = . (13.29)
EA
6 3
Nb − 12 12 uB 1 2
Solving the equations (13.29) with respect to the stresses gives in a first step
" # " # " # " # " #
Na EA 4 −2 − 12 12 uA 1 εT ℓ
= · ·{ · − }
Nb ℓ −2 4 − 21 12 uB 1 2
and after multiplication the final result
" # " # " # " #
Na EA −1 1 uA 1
= · · − EA εT . (13.30)
Nb ℓ −1 1 uB 1
Thus comparable to the element employing displacements only as primary va-
riables and linear shape functions, it follows Na = Nb . Hence, concerning N , a
constant approach is possible, leading to the same result here.
Introducing Equation (13.30) into the PvD, the stiffness matrix follows, accor-
ding to Equation (13.27), to
" # " # " #
1 1
T 3 6 EA 1 −1 uA
δv ·K·v = [ δuA δuB ]·{c·ℓ 1 1 + ℓ }· . (13.31)
6 3
−1 1 uB
In this case, the stiffness matrix of the hybrid–mixed method is identical by
chance to the matrix according to displacement–based elements, cf. Section
2.3.1, and thus leads to the same results applying the element to structural
analysis. Nonetheless, the hybrid–mixed method gives more freedom at the
choice of shape functions for displacements and stresses inside the element.
14 Hybrid–Mixed Plane Stress Elements
x A interface
y z
a
B
d element domain b
D
c element boundary
At the boundary ds the indices n and t indicate the normal and the tangential
direction respectively. Replacing the virtual strains by kinematics yields
Z
δWd = {−δu,x · σxx − δv,y · σyy − (δu,y + δv,x ) · σxy + δu · px + δv · py } dA
A Z
+ {δun σn + δut σt } ds = 0
s
or in matrix notation
Z Z Z
T T
− δWd = ( δu · D ) · σ dA − δu · p dA − δuT · σ e ds = 0 .
T
(14.7)
A A s
At each boundary, the stress vector σ e only comprises the tangential and the
normal component, which are related to the displacements un and ut . The
brackets in Equation (14.7) imply to which state variable the differential ope-
rator D is to be applied.
Thus the conditions of deformation are fulfilled weakly with the PvF in the
domain and are fulfilled strongly at the boundary. It follows in matrix notation
Z
− δWσ = δσ T · ( D · u − E−1 · σ ) dA = 0 . (14.8)
A
Regarding hybrid elements the element domain is distinguished from the ele-
ment boundary and the interface. In the mixed formulation the virtual work is
arranged according to the Principle of virtual Work given by the PvD and the
PvF, see Section 14.1 – without contour integrals. In addition the virtual work
at the element boundaries and at the interfaces is to be described, in a first
step without fulfillment of the conditions of continuity between the elements.
Figure 14-2 identifies the stress variables σn , σt , that perform work at the ele-
ment boundary b–c and at the interface B–C, as an example.
With the PvD and the PvF the displacements and the stressess are identified
to be the variables of description in the element domain and at the interfaces.
Hence, concerning the following discretization, different procedures are possible
at choosing adequate shape functions and at fulfilling the element continuity
conditions.
14.2 Work Equations of a Hybrid Plane Stress Element 253
x A
y z
a
B
d b
σn
D σt
c σt
σn
C
Fig. 14-2 Normal and tangential stresses at element boundary and interface
the contour integrals with respect to PvD are omitted, too, cf. Equation (14.10).
After having introduced kinematics into the work equations concerning real as
well as virtual strains, the work performed at element level follows to
Z
δWd = {−δu,x · σxx − δv,y · σyy − (δu,y + δv,x ) · σxy
+ δu · px + δv · py } dA |domain 6= 0 , (14.14)
Z
1 ν
δWσxx = − δσxx · {u,x − · σxx + · σyy } dA |domain = 0 , (14.15)
E E
Z
ν 1
δWσyy = − δσyy · {v,y + · σxx − · σyy } dA |domain = 0 , (14.16)
E E
Z
2 (1 + ν )
δWσxy = − δσxy · {(u,y + v,x ) − · σxy } dA |domain = 0 . (14.17)
E
Thereby it is essential that the displacements are defined with respect to the
element domain and to the interface and that the stresses are only defined with
respect to the element domain.
u = Ωv · v , or u = Ψu · au = Ψu · G · v ,
whereat G represents the scaling matrix. The approach for the sresses is chosen
as a general polynomial with respect to the element coordinates
σ = Ψs · as .
To start with, the work equation representing the PvD comprises the degrees
of freedom related to the system level v as well as the degrees of freedom as at
element level
Z Z
−δWd = δvT (ΩTv DT ) Ψs dA · as − ΩTv p dA . (14.18)
Applying the PvF, Equations (14.15) to (14.17), the following is valid just as
well
Z
−δWσ = δaTs { ΨTs (D Ωv ) · v − ΨTs E−1 Ψs · as } dA = 0 . (14.19)
Thus the degrees of freedom as may be computed from the nodal displacements
v solving
Z Z
as = { ΨTs E−1 Ψs dA }−1 · { ΨTs (D Ωv ) dA} · v . (14.20)
Alternative B
The calculation of the stresses, which is performed here by employing the shape
functions to describe the stresses and the PvF according to
σxx
σyy = σ = Ψs · as and as = F−1 · Hsv · v , (14.28)
σxy
yields more accurate results and is more efficient on top, if, after the discretiza-
tion of the work equation concerning the PvF, the product F−1 · Hsv is stored
element by element and is availabe for the subsequent stress analysis. After
the computation of as the stresses may be evaluated by means of the general
polynomials Ψs .
x A
y z local coordinates (ξ, η)
(-1,-1)
a
B
(-1,1)
d η ξ b (1,-1)
D
c
(1,1)
u = Ωv · v
δu = Ωv · δv .
260 14 Hybrid–Mixed Plane Stress Elements
The bi–linear approaches to describe the displacements u(ξ, η) and v(ξ, η) in-
corporate four nodal displacement variables each. Thereby, in two dimensions,
three rigid body motions – two translations and a rotation – may be described
as well as five displacement fields, which are available to describe the strains
εxx εyy εxy . Therefor an approach which balances stresses and strains should in-
corporate at least five degrees of freedom related to the stresses. At choosing the
approaches it should be claimed that the courses of the strain εxx = u(x, y),x
and thus also of the stress σxx are constant in x–direction. Concerning the
strain εyy = v(x, y),y and the shear strain 2εxy = u(x, y),y + v(x, y),x the
corresponding facts are valid. This is reached by employing the following non–
scaled approaches
σxx = a1 + a2 · η constant in ξ–direction,
σyy = a3 + a4 · ξ constant in η–direction,
σxy = a5 constant.
Therewith the approach to describe the stresses is given by
σ = Ψs · as
and aTs= [ a1 a2 a3 a4 a5 ]. By analogy the approach to describe the virtual
stresses might be chosen.
C
Fig. 14-4 Element nodes and coordinates – quadratic approaches
u(ξ, η) = a1 + a2 ξ + a3 η + a4 ξ 2 + a5 ξη + a6 η 2 + a7 ξ 2 η + a8 ξη 2
u(ξ, η) = Ψ8 · au .
vu T = [ uA uB uC uD uE uF uG uH ]
results in
1 −1 −1 1 1 1 −1 −1
1 1 −1 1 −1 1 −1 1
1 1 1 1 1 1 1 1
1 −1 1 1 −1 1 1 −1
Ψ̃8 =
1 0 −1 0 0 1 0 0
1 1 0 1 0 0 0 0
1 0 1 0 0 1 0 0
1 −1 0 1 0 0 0 0
−1
with the inverse representation G8 = Ψ̃8
−0,25 −0,25 −0,25 −0,25 0,50 0,50 0,50 0,50
0 0 0 0 0 0,50 0 −0,50
0 0 0 0 −0,50 0 0,50 0
−1 0,25 0,25 0,25 0,25 −0,50 0 −0,50 0
Ψ̃8 = .
0,25 −0,25 0,25 −0,25 0 0 0 0
0,25 0,25 0,25 0,25 0 −0,50 0 −0,50
−0,25 −0,25 0,25 0,25 0,50 0 −0,50 0
−0,25 0,25 0,25 −0,25 0 −0,50 0 0,50
Choosing an equivalent description for the displacement v(ξ, η) yields the ma-
trix notation concerning quadratic approaches for uT = [ u(ξ, η) v(ξ, η) ]
u = Ψ16 · au,v .
14.6 Quadratic Shape Functions 263
vT = [ uA vA uB vB uC vC uD vD uE vE uF vF uG vG uH vH ]
u = Ψ16 · G16 · v ,
= Ωv · v .
σ = Ψs · as
xo x
1 B
A
4 ϕξ
yo
2 ξ
ϕη
D
3
η C
y
a1 = 0. 25 · (−xA + xB + xC − xD )
a2 = 0. 25 · (+xA − xB + xC − xD )
a3 = 0. 25 · (−xA − xB + xC + xD )
b1 = 0. 25 · (−yA + yB + yC − yD )
b2 = 0. 25 · (+yA − yB + yC − yD )
b3 = 0. 25 · (−yA − yB + yC + yD ) .
σηη x x
σyy
σηξ σyx
1
5 σξξ
5
6
σηξ σξη
ϕξ σηη 6
σxy
σξξ ξ
σxx σξη
ϕξ
3 η ϕη
ξ
η ϕη
y y
The lengths of the edges, that are required for the transformation procedure,
are identified from the corner nodes of the sections. The coordinates of the
midpoint of the edges 1 and 3 are given with Figure 14-5. Thus the lengths
ℓ13 = ℓη ,
ℓ15 = ℓη sin ϕη = −2 · a3 ,
ℓ35 = ℓη cos ϕη = 2 · b3 ,
may be computed directly, since the coordinates of the corner node 5 are de-
termined by the midpoints of the edges 1 and 3. The coordinates of the node
6 may be determined with respect to corner node 5 by
∆y56 b1
tan ϕξ = = ,
∆x56 a1
∆x56 −a3
tan ϕη = = .
b3 − ∆y56 b3
Introducing
In a first step the conditions of equilibrium related to the x– and the y–direction
are assembled following Figure 14-6, and they are rearranged to transform the
stresses now. Thus matrix notation yields
ℓ36 ℓ56
σxx
ℓ
cos ϕξ ℓ
sin ϕη − ℓℓ36 sin ϕη − ℓℓ56 cos ϕξ σξξ
35 35 35 35
σ ℓ16 sin ϕξ ℓ56
cos ϕη ℓ16
cos ϕη ℓ56
sin ϕξ
σηη
yy ℓ15 ℓ15 ℓ15 ℓ15
= ℓ36 ℓ 56 ℓ 36 ℓ 56
.
σxy σξη
ℓ35 sin ϕξ − ℓ35 cos ϕη ℓ35
cos ϕη − ℓ sin ϕξ
35
σyx ℓ16
cos ϕξ − ℓ56 sin ϕη − ℓ16 sin ϕη ℓ56
cos ϕξ σηξ
ℓ15 ℓ15 ℓ15 ℓ15
Replacing the trigonometric functions as well as the length of the edges by the
differences of the coordinates and taking into account σxy = σyx results in
ℓ ℓ
a1 a1 ℓη a3 a3 ℓ ξ a1 a3 a1 a3 σξξ
σxx ξ η
1
ℓη ℓξ
σηη
σyy = · b1 b1 ℓ b3 b3 ℓ b1 b3 b1 b3 .
det J0 ξ η σξη
σxy ℓ ℓ
a1 b1 ℓη a3 b 3 ℓ ξ a1 b 3 b 1 a3 σηξ
ξ η
In a second step and following Section 14.5 the linear approach to describe the
local stresses is chosen to σξη = σηξ employing the 5 degrees of freedom βi
σξξ β1
1 0 0 η 0
σηη β2
0 1 0 0 ξ
= β3 (14.29)
σξη 0 0 1 0 0
β4
σηξ 0 0 1 0 0
β5
ℓ ℓ
β̃4 = β4 ℓη /det J0 , β̃5 = β5 ℓξ /det J0
ξ η
Example of Use 1
As a first example of use a benchmark is investigated at first published by Pian
and Sumihara, cf. Figure 14-7.
P2
2 2 1 1 4 P1
P2
A B C
y, v 2
x, u D P1
P2
1 1 2 3 3
Fig. 14-7 Cantilever published by Pian and Sumihara [81] – geometry and loading
14.8 Convergence Behavior of the Elements 269
Applying the finite element mesh according to Figure 14-7 the subsequent
Table 14.1 comprises the displacement results related to the loading case P1 =
1000 N . The three hybrid element formulations are compared to the analytical
solution, that is evaluated with respect to a linearly distributed loading in
y–direction p1 (y) = 2000 · ( y − 1 ) N/m.
vA vB vC vD
The stresses σxx related to P1 are represented in Figure 14-8. The results il-
lustrate the influence of the element geometry on the approximation of the
stresses. The quadratic approaches P-HMQ-8-13 yield the best results as ex-
pected. The stresses evaluated from the element P-HMQC-4-5 published by
Pian and Sumihara yield good results although an approach of low order is
employed. In contrast the bi–linear approach P-HMQ-4-5 is not adequate.
N/m2 σxx
-3000 P1 = 1000 N
analytical solution
P-HMQ-4-5
P-HMQ-8-13
P-HMQC-4-5
-2000
x
Fig. 14-8 Stress σxx at the upper boundary of the cantilever for P1
270 14 Hybrid–Mixed Plane Stress Elements
The second loading case is more difficult to be described by the given elements,
since the stresses σxx develop linearly along the x–axis, whereat the stresses
σxy are quadratically distributed along the y–axis. Due to the parabolic distri-
bution of the stresses σxy in y–direction the load is transferred to concentrated
nodal actions according to the PvD. The subsequent Table 14.2 comprises the
displacement results concerning the loading case P2 = 150 N . As in loading case
P1 the quadratic approach gives the best results, whereat the linear approach,
according to Section 14.5, is hardly acceptable.
vA vB vC vD
Applying the mesh according to Figure 14-7 the stresses σxx are represented
in Figure 14-9 related to the loading case P2 . As in loading case P1 the results
illustrate the strong influence of the element geometry on the approximation of
the stresses. The quadratic approach yields the best results and describes the
stresses σxx to be linearly distributed within the element without a larger gap at
the interface of neighboring elements. The stresses evaluated from the element
published by Pian and Sumihara yield good results, although an approach
of low order is employed. In contrast the bi–linear approach is not adequate,
since it leads to a stepwise approximation with larger gaps at the interfaces of
neighboring elements.
N/m2 σxx
-4000 P2 = 150 N
reference solution
P-HMQ-4-5
P-HMQ-8-13
P-HMQC-4-5
0
x
Fig. 14-9 Stress σxx at the upper boundary of the cantilever for P2
14.8 Convergence Behavior of the Elements 271
Example of Use 2
As a further example of use Cook’s can- c
tilever [28] is chosen, since comparable x
results are available. Without comple- B
y D 1N
b
tely representing the study of conver-
gence according to Section 12.2.3, the
results for different element formulati- C
ons are compared in Table 14.3 apply- E = 1.0 N/mm 2
ing meshes with 3 × 3 and 5 × 5 no- ν = 0.33
a
des. Concerning the hybrid elements the A a = 44 mm
respective subsequent stress analysis is b = 16 mm
performed each with alternative B. c = 48 mm
Me
e
x V
ϕ p
z
z,w dM dM
M- M+
2 x 2
dQ dQ
Q- Q+
2 2
a) Kinematics
in the domain w,xx − κ = 0 ,
at the boundary ϕ − ϕe = 0 , (15.1)
w − we = 0 .
b) Equilibrium
in the domain M,xx + pz = 0 ,
at the boundary Q−Ve = 0, (15.2)
M + Me = 0 .
Considering Eq. (15.6) the boundary conditions related to the force variables
are naturally fulfilled. Integrating by parts of the first expression and taking
into account δw,x = δϕ yields
Z
{−δw,x · M,x + δw · pz } dx + [δw · V e + δϕ · (M e + M )]bound. = 0 . (15.7)
is valid. Replacing the curvature by the material equation and thus by the
bending moment yields
Z
M
δM · (−w,xx − ) dx − [ δQ · (w − we ) + δM · (ϕ − ϕe ) ]bound. = 0 . (15.9)
EI
Concerning the domain the condition of deformation is weakly fulfilled by the
integral. The contour expressions take into account the displacement boun-
dary conditions. Integrating the first expression in the integral by parts and
introducing w,x = ϕ yields
Z
M
{δM,x · w,x − δM } dx − [ δQ · (w − we ) − δM · ϕe ]bound. = 0 . (15.10)
EI
A second integration by parts of the same term and replacing δM,x = δQ
results in
Z
M
{−δM,xx · w − δM } dx + [ δQ · we + δM · ϕe ) ]bound. = 0 . (15.11)
EI
In Eq. (15.9) the contour expressions comprise the boundary conditions related
to the displacement variables w and ϕ. After integration by parts the boundary
condition related to the rotation can only be fulfilled weakly applying Eq.
(15.10). In contrast the boundary condition related to the deflection may be
fulfilled strongly with w − we = 0. Repeated integration by parts yields weak
fulfillment of both boundary conditions applying Eq. (15.11).
Remarks
With both principles of work, applying the Equations (15.6), (15.7) and (15.8)
as well as (15.9), (15.10) and (15.11), different but equivalent formulations are
availabe, which may be chosen as a basis concerning the discretization.
276 15 Hybrid–Mixed Euler–Bernoulli Beam Elements
M,x l − M,x r = V e ,
Both terms, occuring at the left hand side, are implicitly incorporated in the
work equation. The external work δw · V e is taken into account by the load
vector.
Applying the PvF according to Eq. (15.13) the boundary conditions related to
δM and w are fulfilled strongly and the conditions related to δQ = δM,x as
well as to ϕ = w,x are fulfilled weakly. At a simply supported system boundary
the contour expression vanishes, since δM = 0 is valid. Regarding a clamped
system boundary a rotation of the support ϕe is to be considered with δM · ϕe
in the load vector. Concerning a fixed clamping it follows ϕe = 0.
At the element interfaces the virtual bending moments are continuous. Here,
the condition δMl = δMr = δM is fulfilled strongly, in contrast, the continuity
condition related to the rotation
is fulfilled naturally. ∆ϕe corresponds to a sharp kink at the internal node here,
which is to be considered in the load vector by −δM · ∆ϕe , comparable to the
procedure at the system boundary.
Numerical results are represented in Section 21.2 and Section 21.3 with an
extension to shear deformations by means of simplified structures.
riables are defined at the interfaces, which are not available at the element
domain.
At Figure 15-2 a beam element is represented together with the interfaces A
and B, that are identical with the system nodes. Possible gaps ∆we between
element boundaries and interfaces are not considered.
Vei pz Vei+1
A B
i a b i+1
wi Ma Qa wE ME Mb Qb wi+1
ϕi ϕ i+1
Cutting cleanly of the element domain including the element boundaries a and
b from the interfaces A and B yields the bending moments Ma and Mb as well
as the shear forces Qa = (M,x )a and Qb = (M,x )b . The bending moments are
defined only related to the element domain and to the cut, but not related to
the interface as primary variables. As kinematic variable the deflection wE is
chosen. The nodal displacements and the nodal rotations are defined at the
system level to ensure the element continuity conditions with C1 –conformity.
Applying the PvD by analogy with Section 13.1 the equilibrium of forces is
formulated concerning the element. In addition the equilibrium of moments at
the interfaces – which means the system nodes i and i + 1 – is to be ensured
by employing the virtual rotations.
Z
δWd = {δκE · ME − δwE · c · wE + δwE · pz } dx
+ Ma · (−δϕA + δw,x |a ) − Mb · (−δϕB + δw,x |b )
+ Qa · (δwA − δwa ) − Qb · (δwB − δwb ) .
Applying the PvF at element level the conditions of deformation are fullfilled
yielding
Z
1
δWσ = {δME · κE + δME · · ME } dx
EI
+ δMa · (−ϕA + w,x |a ) − δMb · (−ϕB + w,x |b )
+ δQa · (wA − wa ) − δQb · (wB − wb ) = 0 .
is required with
Applying kinematics κE = wE ,,xx and accordingly δκE = δwE ,xx as well as in-
tegration by parts of the respective work expressions yields the work equations,
that are introduced into the discretization further on
Z
δw : − δWw = {δwE,x · ME,x + δwE · c · wE − δwE · pz } dx , (15.14)
As at the procedure concerning bars and plane stress structures the virtual
work related to the PvF disappears at element level, thus Equation (15.16) can
be taken to eliminate the bending moments ME .
[ wE ] = Ωv · w und wT = [ wa wb ] ,
T
[ M E ] = Ωs · s und s = [ Ma Mb ] .
280 15 Hybrid–Mixed Euler–Bernoulli Beam Elements
Along the lines of Section 13.2 and after the discretization of the virtual work
performed at the element domain the matrix notation of Equations (15.14),
(15.15) and (15.16) gives
" R T R T T # " # "R T #
δw : Ωv C Ωv dx Ωv D · D Ωs dx w Ωv p dx
R T T R · − .
δs : Ωs D · D Ωv dx − ΩTs E−1 Ωs dx s 0
Since the continuity conditions wA = wa , wB = wb , δwA = δwa , δwB = δwb
are fulfilled, the deflections as well as the rotations at the interfaces are defined
as nodal displacement variables
vT = [ wA wB ϕA ϕB ] .
Thereby the work performed by the nodal rotations can be introduced into the
work equation applying matrix notation. Employing linear shape functions to
describe wE and ME as well as the rotations ϕA , ϕB only being defined at the
interfaces and integrating the work performed at the element domain yields
Hvv HTsv
cℓ cℓ 1
δwA : 3 6 ℓ
− 1ℓ wA
cℓ cℓ
− 1ℓ 1
δwB : 6 3 ℓ
wB
δϕA :
+1
ϕA
. (15.17)
δϕB : −1
ϕB
1 ℓ ℓ
δMa :
ℓ − 1ℓ +1 − 3EI − 6EI =0
Ma
ℓ ℓ
δMb : − 1ℓ 1
ℓ
−1 − 6EI − 3EI Mb =0
Hsv −F
By analogy with the bar again, the calculation of the element stress resultants
s takes place applying the deformation condition
s = F−1 · Hsv · v. (15.18)
After introduction the stress resultants into the PvD the element stiffness ma-
trix and the load vector follow from eliminating the bending moments
−δWd = δvT [ Hvv + HTsv · F−1 · Hsv ] · v − fp . (15.19)
Replacing " #
6 −6 4ℓ 2ℓ EI
−1
F · Hsv = · (15.20)
−6 6 −2ℓ −4ℓ ℓ2
15.3 Element Stiffness Matrix 281
Numerical results are represented in Section 21.2 and Section 21.3 with an
extension to shear deformations by means of simplified structures.
16 Hybrid–Mixed Kirchhoff Plate Elements
x A
y z,w
ϕn(A,B)
ϕn(D,A)
a
p B
d b
qn
D qn
mt mt
mn mn
mt c mt
ϕn(B,C)
qn qn
ϕn(C,D)
C
To develop the mixed plate element Kirchhoff ’s plate theory is chosen, to illu-
strate the analogy with the mixed Euler–Bernoulli beam element.
p
x
y p
z,w dx
dy
dy
dx
m xy
m yy m xx
qx
m yx
qy
Again, the condition of equilibrium (16.2) may be represented with the matrix
symbols already known
−DT · σ + pz = 0 . (16.3)
Kinematics is described as a 2nd order system, too:
ǫ =D·u
−κxx −∂xx
−κyy = −∂yy · [ w ] . (16.4)
−2κxy −2∂xy
The material equation is applied in its inverse representation, thereby the fle-
xibility F = 1/B(1 − ν 2 ) = 12/Et3 is introduced as a material parameter. In
matrix notation this yields
ǫ = E−1 · σ,
−κxx F −νF mxx
−κyy = −νF F · myy . (16.5)
−2κxy 2(1 + ν)F mxy
16.1 Mixed Principles of Work Concerning Kirchhoff Plates 285
D · u − E−1 · σ = 0 ,
−∂xx F −νF mxx 0
−∂yy · [ w ] − −νF F · myy = 0 . (16.6)
−2∂xy 2(1 + ν)F mxy 0
The matrix symbols of the mixed formulation are determined from the gover-
ning equations, hence the work equations may be described correspondingly.
+ δw · p dA = 0 . (16.9)
A
The first term at the integral corresponds to the part of the work performed
by Kirchhoff ’s effective shear force, that is not captured by the shear force yet.
The last term may be integrated following the respective boundary. This yields
Z
δWbound. = −δw · mt ,s ds + δw · mt |corner .
s
The term comprising the work performed at the corner disappears, since ei-
ther the twisting moment at the free corner or the virtual displacement at
the supported corner is equal to zero. The remaining integrals are summarized
to Kirchhoff ’s effective shear force qKn employing the shear forces qx , qy and
qn implicitly incorporated in the work equation. Thus, the element continuity
conditions related to the effective shear forces following the contour coordinate
s of the element are also fulfilled weakly at the system level with
Z
δw[ (qn + mt ,s )l − (qn + mt ,s )r ] ds = 0 .
s
With Equation (16.9) the work equation related to the PvD follows to
Z
− δWd = {δw,x · mxx ,x + δw,y · myy ,y + δw,x · mxy ,y + δw,y · mxy ,x } dA
ZA Z
− δw,s · mt ds − δw · p dA = 0 (16.12)
s A
forming the basis of the discretization. Thus, besides the equilibrium in the
domain, the boundary conditions qKn = 0 at the free boundary as well as
mt = 0 at the free corner are fulfilled naturally. The virtual work performed
by line–shaped external actions in s–direction as well as work performed by
concentrated actions may be additively taken into account summing up the
external work at the interfaces in the load vector.
Regarding the PvD, the boundary conditions related to the bending moments
as well as to the virtual displacements are substantial ones and therefore are
to be fulfilled by the approaches.
Due to the first part being implictly considered already by the first contour
integral, the third contour integral of the work equation is to be evaluated
explicitly.
Thus the work equation related to the PvF yields
Z
δWσ = {δmxx ,x w,x − δmxx · F · ( mxx − ν myy )
+ δmyy ,y w,y − δmyy · F · ( ν mxx − myy )
+ (δmxy ,y w,x + δmxy ,x w,y ) − δmxy · 2(1 + ν) · F · mxy } dA
Z Z
− δmn ϕen ds − δmt w,s ds = 0 . (16.13)
At the positions indicated by ∗ and besides the domain integrals given here,
the contour integrals related to Eqns. (16.13) and (16.12) are to be evaluated.
u = Ωv · v with u = [ w(x, y) ] ,
Ωv = [ φA φB φC φD ] ,
vT = [ wA wB wC wD ] ,
mxx (x, y)
σ = Ωs · s with σ = myy (x, y) ,
mxy (x, y)
φ φ φ φ
A B C D
Ωs =
φA φB φC φD ,
φA φB φC φD
sT =
[ mxxA mxxB mxxC mxxD | myyA myyB myyC myyD | mxyA mxyB mxyC mxyD ] .
290 16 Hybrid–Mixed Kirchhoff Plate Elements
Applying the approach for the deflection and the moments to Equation (16.14)
directly leads to the element matrix. Therefore, corresponding to Equation
(16.14), the discretized representation of the virtual work is given summarizing
the terms to be integrated for the element matrix with φT = [ φA φB φC φD ]:
0 φT,x φ,x φT,y φ,y (φT,x φ,y + φT,y φ,x )∗
Z
φT,x φ,x −F φT φ νF φT φ 0
dA .
T T T
A φ,y φ,y νF φ φ −F φ φ 0
T T
(φ,x φ,y + φ,y φ,x )∗ 0 0 −2(1 + ν)F φ φ T
The domain integrals may be integrated analogously to the plane stress element
according to Section 14. Moreover, besides the domain integrals, the contour
integrals are to be evaluated, which are indicated by ( )∗ . Although the contour
integral is only to be evaluated at the free boundary, regarding the computa-
tion in practice, it may be more convenient to evaluate the integral in general
related to each element and to each element boundary. This results in an all–
purpose element matrix, that can be applied to all different possibilities of
support. Disregarding the free boundary the expressions cancel out each other
at assembling the elements to the entire structure.
Assembling the nodal variables in the vector z with
zT = [ wA mxxA myyA mxyA |
wB mxxB myyB mxyB |
wC mxxC myyC mxyC |
wD mxxD myyD mxyD ]
yields the element matrix A, which is given on the next page. Thus the Principle
of virtual Work is discretized on the element level as
−δW = δzT {A · z − f } .
Load Vector
Due to the linear approaches the computation of the load vector takes place
by analogy with the plane stress elements according to Section 14 and thus is
taken over. At the PvD distributed external actions are taken into account by
Z
δWd = δvT ΩT · p dA . (16.15)
A
0
A1
3
− F91
A2
− F92 − F91
0 0
− F93
A2 1
1 3
0 3
0+ 2
0
− A31
2 −1
2 A1
3 18 18
0 3
−F − F91
A2 F2 1 A2
0 . . . symmetric. . .
6 18 18 3
− A1 − F 1
− −F − F92 − F91
3
0 0 0 0
2 18
−F − 12 + 1 − F93
A2 A1 A2
0 3
0 6 3
0
0− 1
− A61 − 12 + 0 0 − 12
2 A1 A1
0 0
A=
6 36 36 6 3
1 A2
0 0
6 36 36 3
−F − A32 − F182 − F181 − F92 − F91
− A1 − F 1
3 1
0 0 0 0 0
16.1 Mixed Principles of Work Concerning Kirchhoff Plates
36
−F 0 + 12 − F183 2
−1 − F93
− A2 − F 2
A1 A2 1 A2 1 A2 1
0 0+ 0 +0 0 0+ 0
1
6 3 2 3 2 3 2
−2 + 0 0
− A61 − A31
A1 F1 2 A1 A1 2 A1
0 0 0
6 18 18 6 3 18 3
− −F − F361 − F362 − F181 − F − F91
F1 A2 F2 F1 A2
0 0 0
6 3
− 18 − A62 − F362 − F361 − 18 − 18 − F92 − F91
3 1 1 3
0 0 18 2
+0 0 0 2
0 0 18
0 0
0 − 12 −F − F363 0− −F − 12 + 1 − F93
− A32 − F182
291
292 16 Hybrid–Mixed Kirchhoff Plate Elements
l x = 10.0 m
l y = 10.0 m
lx
Although the element converges against the accurate solution, the mixed plate
element comprises some properties, which may limit the applications.
• Investigating oblique plates leads to problems with the fomulation of the
boundary conditions with respect to the moment mn at hinged or free
boundaries, since a transformation of the moment is needed there.
• The computation of stiffened plates is impossible, if the stiffeners have to
be discretized as well. In these cases the bending moment can not be a
nodal degree of freedom.
16.2 Hybrid–Mixed Rectangular Plate Element 293
x A
y
interface ’I’
z
a
B
D d element domain ’E’ b
1
1
c
C
1
1
Cutting cleanly the interfaces yields the shear forces and the moments at the
element boundaries E and at the interfaces I according to Figure 16-1. Incor-
porating the work performed at the transition between element boundary and
interface results in:
Z
δWd = {δw,xx · mxx + δw,yy · myy + 2δw,xy · mxy + δw · p} dA
AZ Z
− mn · (δw,n E − δϕn I ) ds − mt · (δw,s E − δw,s I ) ds
Z
+ qn · (δwE − δwI ) ds .
wT = [ wA wB wC wD ]
the matrix notation of the work equations follows – in a first step without the
parts related to external work –
R T T
δw : 0 0 Ωv D D Ωs dA + ( )∗ w
R T
δϕ : 0 0 Ωϕ Ωs ds · ϕ .
R R R
δs : ΩTs DT D Ωv dA + ( )∗ ΩTs Ωϕ ds − ΩTs E−1 Ωs dA s
296 16 Hybrid–Mixed Kirchhoff Plate Elements
Here, the expression ( )∗ alludes to the contour integrals, that are to be consi-
dered at the element boundary. The domain integrals
Z Z
(ΩTv DT ) (D Ωs ) dA and ΩTs E−1 Ωs dA
as well as the contour integrals ( )∗ may be directly taken from the mixed
formulation.
R
Integrating the additional work expressions at the interface ΩTϕ · Ωs ds and
employing linear approaches to describe ϕx , ϕy yields, regarding rectangular
elements, the matrix
A B C D
mxx myy mxy mxx myy mxy mxx myy mxy mxx myy mxy
ly ly
δϕxA : 3 6
lx lx
δϕyA : 3 6
−ly −ly
δϕxB : 3 6
lx lx
δϕyB : 6 3
.
−ly −ly
δϕxC : 6 3
−lx −lx
δϕyC : 3 6
ly ly
δϕxD : 6 3
−lx −lx
δϕyD : 6 3
To eliminate the moments related to the element domain the matrix notation
of the element matrix fits best:
" # " # " #
δv : 0 HTsv ∗ v f 6= 0 at element domain
· −
δs : Hsv ∗
−F s 0 = 0 at element domain
Thereby the contour integrals related to the element boundaries are taken into
account at the positions ∗. The elimination of the moments related to the
element domain yields the virtual work described with the PvD – including the
external work –
x clamped
G M ν = 0.2
hinged
l x = 10.0 m
l y = 10.0 m
lx
Table 16.2 summarizes the results concerning the state variables related to
the hybrid–mixed element employing 12 degrees of freedom. Comparing the
results to the results of the plate elements according to Sections 6.2, 6.3 as
well as Section 16.1.4 illustrates, that the convergence concerning the state
variables of the hybrid–mixed element is only marginally inferior despite of the
considerably lower polynomial order of the shape functions.
Employing the same number of degrees of freedom the results of the displace-
ment–based element comprising 12 degrees of freedom are slightly more accu-
rate, because of the higher order shape functions to describe w. This can be
observed explicitly at incorporating distributed external actions and bedding
respectively. Further differences occur regarding the expressions related to ben-
ding, too. But an advantage of the K-HMQ-4-16 element is the less effort at
integration of the element stiffness matrix due to the lower order approaches.
HYBRID TRIANGULAR PLATE ELEMENTS
17 Hybrid Triangular Plate Elements
From Section 11.1 it is known, that the triangular plate element employing a
cubic approach to describe w and therefor including 10 degrees of freedom –
as represented in Figure 17-1 – is not conform and converges to an incorrect
solution in particular cases.
w
x w,x
y z w,y w,s
A
w,n
D+w w
B w,x
C w,y
w
w,x
w,y
Concerning this plate element, at each boundary 6 degrees of freedom are availa-
ble. Due to the shape functions required to be cubic polynomials along the
boundary with respect to w employing four degress of freedom and to be qua-
dratic polynomials with respect to w,n with three degrees of freedom to fulfill
the continuity conditions by the approach itself, one degree of freedom is mis-
sing to ensure the C1 –conformity. Therefore, the slopes w,n are not continuous,
which causes a bad behavior of convergence. Subsequently different possibilities
are presented to ensure C1 –conformity, whereupon the cubic approach to des-
cribe the deflection provides a basis for the discretization in each case according
to Section 17.1.
u = [w] = ψ · a, (17.1)
v = Ψ̃ · a .
17.1 Cubic Approach for Triangular Plate Elements 303
Here, A represents the domain of the triangular element. The scaling matrix
G is defined as the reversed representation of Ψ̃
−1
G = Ψ̃ ,
1 0 0
3 ac −bc 0 0 0
3 −ab bb
1 0 0
0 3 aa −ba 0 0
G=
3 −ac bc
. (17.2)
1 0 0
0 0 3 ab −bb 0
3 −aa ba
−7 ab −ac bc −bb −7 ac −aa ba −bc −7 aa −ab bb −ba 27
[ kdd ] · wd + [ kdi ] · vi = [ fd ] ,
wd = [ kdd ]−1 · { [ fd ] − [ kdi ] · vi } .
Replacing wd in the first row, the stiffness matrix is lowered by one line and
one column to
are defined related to the scaled cubic approach. The element continuity con-
ditions concerning w and w,t are fulfilled by the approach, if [ w w,x w,y ]a, b, c
and [ w w,x wy ]A, B, C are equal, see Figure 17-2.
b
x A
y z a d +
a
D
x
F x c
x x d + b
x B
x
c E
C
Due to w being cubic, the first order derivative with respect to the boundary
coordinate s is described quadratically and is directly defined by w. Thus in
total four nodal displacements are allocated related to the tails of the element
boundary concerning the cubic course. Two degrees of freedom are left to li-
nearly describe the course of w,n along the boundary and at the interface. To
create an element continuity ensuring conformity a quadratic course also is nee-
ded with respect to w,n . The fulfillment of the element continuity conditions
with respect to w,n succeeds applying a hybrid formulation of the work related
to the element.
To ensure the element continuity conditions two principle methods of resolution
are possible:
is valid. The terms of virtual work that vanish at assembling the element for-
mulation, are not considered concerning Equation (17.3).
quadratic part
Applying the cubic approach according to Equation (17.1) the element conti-
nuity conditions related to the rotations w,n and related to the virtual rotations
δw,n cannot be fulfilled, since at the element boundary w,n is only described
linearly by the slopes w,n related to the corners. Therefore additional conditi-
ons and degrees of freedom are to be defined at the interface, that consider the
higher polynomial parts, cf. Figure 17-3. This goal can be reached following
different paths.
Fig. 17-4 The transformation of w,s and w,n with respect to w,x and w,y
After integration of the work employing the cubic approach to describe w ac-
cording to Section 17.1 matrix notation of the element stiffness matrix yields
" #
kww kwλ
K= .
kλw 0
Applying the complete element matrix and the respective load vector the bary-
centric–related deflection wd may be eliminated according to Section 17.1.2.
After allocation of the elements to the entire system of equations the element
continuity conditions related to the slopes arise as
Z
−δWλ|I = {δλI · (w,n |i − w,n |i+1 )} ds = 0 (17.6)
I
Test – Version A
As an example to test the formulation given above the square plate according
to Section 6.2.9 is investigated, cf. Figure 17-5.
x clamped
G M ν = 0.2
hinged
l x = 10.0 m
l y = 10.0 m
lx
In the subsequent Table 17.1 and Table 17.2 the behavior of convergence concer-
ning the element employing displacements as primary variables and additional
Lagrange multipliers is summarized with respect to different state variables.
Relating to the nodes, that are belonging to multiple elements and at differing
values, the results are given for the respective neighboring elements.
The comparison of the results applying the plate elements according to Sec-
tion 6.2.9 as well as to Sections 10.2 and 10.3 demonstrates, that the results
related to the hybrid element employing displacements as primary variables
and additional Lagrange multipliers exhibit worse convergence due to the lower
polynomial degree of the respective shape functions. In particular, the conver-
gence of the stress resultants is not sufficient. Nevertheless the results illustrate
that the method of Lagrange multipliers is appropriate to ensure the element
continuity conditions in principle requiring little effort.
310 17 Hybrid Triangular Plate Elements
Here, the work equation applying the PvD is extended to further terms of work
concerning the element boundaries and the interfaces. By analogy with the
hybrid–mixed plate element the rotations ϕn at the interfaces are chosen as
additional variables to ensure the equilibrium of moments mn at the element
interfaces. Thus the moments mn perform virtual work on the slopes w,n at
the element boundaries and on the rotations ϕn at the interfaces respectively.
Z
−δWd = B · {δw,xx (w,xx + νw,yy ) + δw,yy (w,yy + νw,xx )
A Z
+ 2 (1 − ν) δw,xy w,xy } dA − δw pz dA
Z Z A
The terms related to the virtual work, that vanishes at assembling the elements,
are not taken into account at Equation (17.7). This affects the work performed
by the twisting moment mt and by the shear forces qn or rather Kirchhoff ’s
effective shear force qKn = qn + mt ,s on the deflection w and on the rotation
w,s respectively.
At assembling the vitual work of the elements to the entire structure or rather
to the system of equations the condition of equilibrium related to the bending
moments at the element interface
Z
δϕn · ( mn (i) − mn (i+1) ) ds = 0
I
The continuity condition related to w,n between the element boundary and the
interface can be fulfilled applying the PvF at element level, when considering
the work performed by the virtual moment δmn on the rotation of the element
boundary as well as of the interface. Due to the linear part of w,n being fulfilled
already by the cubic approach to describe w, the rotation ϕn only has to fulfill
17.2 Hybrid–Displacement Elements Employing 10+3 DOF 311
the element continuity condition related to the quadratic part of w,n . This is
to be caught at formulating the element continuity condition with
Z Z
−δWσ = + δmn · ( w,n |quadr. − w,n |linear ) ds − δmn · ϕn ds . (17.8)
E I
Here, β is defined as the angle between the x–axis and the respective edge of
the triangle, cf. Figure 17-4, whereat sin β and cos β are computed by means
of the coodinate differences of the corners as given in Section 17.2.1:
p
∆x = xi+1 − xi , ∆y = yi+1 − yi and ℓi+1
i = ∆x2 + ∆y 2 ,
∆y ∆x
sin β = and cos β = .
ℓ ℓ
Introducing the material equations yields
mn = −B (w,xx + νw,yy ) sin2 β + (w,yy + νw,xx ) cos2 β
mn = −B ( w,nn + νw,ss ) ,
which results, at introducing the transformation rule for the derivatives, in the
same formulation related to the x–y–coordinates.
Furthermore the slope w,n is to be transformed at the element boundary. Ac-
cording to Section 17.2.1 it is obtained
Introducing mn and w,n into the work Equations (17.7) and (17.8), employing
the cubic approach to describe w, and a quadratic approach to describe ϕn
yields the stiffness matrix according to Section 17.1
" #
kww kwϕ
K=
kϕw 0
Thereby kww (size 10 × 10) represents the work performed at the element
domain and at the element boundary, kwϕ = kTϕw (size 3 × 10) represents the
work performed at the interface. To numerically integrate the work concerning
the element domain a 3–points Gauss–Legendre integration is sufficient, cf.
Section 12.1.
Test – Version B
As an example to test the formulation given above the square plate according
to Section 6.2.9 is investigated, see Figure 17-6.
x clamped
G M ν = 0.2
hinged
l x = 10.0 m
l y = 10.0 m
lx
In Table 17.3 and Table 17.4 the behavior of convergence concerning the hy-
brid element employing displacements as primary variables and 10+3 element
degrees of freedom is summarized with respect to different state variables. Re-
lating to the nodes M and T , that are belonging to multiple elements, the
results are given for the respective neighboring elements. The gap of the mo-
ments between different elements is an indication of the discretization error.
17.2 Hybrid–Displacement Elements Employing 10+3 DOF 313
At the element domain the work equation applying the PvD is extended to
the work at the element boundaries as well as at the interfaces. In contrast to
Section 17.2.2, at the element boundary the moment mn performs work on the
total rotation w,n |quadr. and at the interface the resultant moment mn performs
work on the rotations w,n |linear existing at the interface, without requiring an
additional variable. Thus the work equation applying the PvD yields
Z
−δWd = B · {δw,xx (w,xx + νw,yy ) + δw,yy (w,yy + νw,xx )
A Z
+ 2 (1 − ν) δw,xy w,xy } dA − δw pz dA
Z Z A
Test
As an example to test the formulation given above the square plate according to
Section 6.2.9 and Figure 17-7 is investigated. In Table 17.5 and Table 17.6 the
convergence behavior is summarized with respect to different state variables.
x clamped
l x = 10.0 m
l y = 10.0 m
lx
It is substantial, that the results converge against the reference solution, al-
though only the displacement variables [ w w,x w,y ] related to the corner nodes
are defined as degrees of freedom and the element continuity condition is only
fulfilled weakly concerning w,n . Comparison with the results obtained by the
plate element according to Section 17.2 demonstrates, that the results similarly
good converge against the reference solution. Only the clamping moment at
position E shows a slightly inferior convergence.
Although the element offers a very good performance regarding the presented
example, a general application is restricted to element geometries, where the
internal angles are limited to about α, β, γ < 100o.
18 Hybrid–Mixed Triangular Plate Elements
The governing equations related to Kirchhoff ’s plate theory are given in Sec-
tion 6. The weak formulation of the governing equations is developed applying
the hybrid–mixed principle of work according to Section 16, where the discreti-
zation employing rectangular elements is represented. Subsequently the work
equations are picked up again and are extended with respect to the demands
on triangular elements. By analogy with Section 17 the element is split into
element domain, element boundary and interface, cf. Figure 18-1.
b
x A
y z a d +
a
D
x
F x c
x x d + b
x B
x
c E
C
x
y z w ,s
w ,n
w ,s w ,n
s s
s
w ,n
w ,s
In the case the virtual displacements are continuous between the element boun-
dary and the interface, it follows
Z Z
qn · δw ds − qn · δw ds = 0
E I
and thus
Z
δWd = − { δw,x · mxx,x + δw,y · myy,y + δw,x · mxy,y + δw,y · mxy,x } dA
A
Z
− {mn · δw,n + mt · δw,s } ds . (18.1)
I
yields
Z
δWσ = {− δmxx [ κxx + F · ( mxx − ν myy ) ]
A
− δmyy [ κyy + F · ( ν mxx − myy ) ]
u = [ w ] = ψ u · G10 · v = Ωv · v
and the same may be arranged with respect to the virtual deflection δw.
σ = Ψσ · b ,
T
σ = [ mxx myy mxy ] ,
2 2 2
λa λb λc λa λb λa λc λb λc 0 0
λ2a λ2b λ2c λa λb λa λc λb λc
Ψσ = 0 0 ,
2 2 2
0 0 λa λb λc λa λb λa λc λb λc
bT = [ b1 b2 b3 b4 b5 b6 | b7 b8 b9 b10 b11 b12 | b13 b14 b15 b16 b17 b18 ] .
18.3 Element Stiffness Matrix and Load Vector 321
x
w,s
w, n
D
s w,y s
β
td td
y
nd w,x nd
The transformation of the moments [ mxx myy mxy ] into the moments [ mn mt ]
with respect to the contour coordinate yields
mn = mxx · sin2 β + myy · cos2 β − 2mxy · cos β · sin β ,
mt = (−mxx + myy ) · sin β · cos β + mxy · (cos2 β − sin2 β) .
Applying the transformation matrix Tv concerning the derivatives of the de-
flection surface and the transformation matrix Tσ concerning the moments,
matrix notation yields
Z " T T
# " #
T 0 (Ω v Tv ) (Tσ Ψσ ) v
−δW |I = δv δbT T T
ds .
G (Ψ T
σ σ )(T v Ω v ) 0 b
Elimination of Moment–DOF
Having integrated the work terms the degrees of freedom b may be numerically
eliminated. Introducing the abbreviations
Z Z
Hσv = (ΨTσ DTσ )(Dv Ωv ) dA + (ΨTσ TTσ )(Tv Ωv ) ds ,
ZA I
T −1
F= Ψσ E Ψσ dA ,
ZA
fp = ΩTv p dA ,
A
matrix notation gives
" # " # " #
δv : 0 HTσv v fp 6= 0 at element level
· −
δb : Hσv −F b 0 = 0 at element level
and the computation of b yields
b = F−1 Hσv · v . (18.5)
18.4 Fulfillment of the Continuity Conditions at the Interface 323
The expression inside the curly bracket of Equation (18.6) states the stiffness
matrix with respect to the element domain.
as well as
Z
δWσ = . . . − {δmn w,n + δmt w,s } ds .
I
By analogy with Section 17.2 the slopes w,n are described by a linear course
w,n |linear , given by the nodal displacement variables at the interface, as well as
by an additional quadratic part w,n |quadr. , which is assigned to the rotational
degree of freedom ϕn related to the midpoint of the interface. Thus the work
equations have to be transformed applying
Hence, at derivation of the Equations (18.5) and (18.6), the vector of the no-
dal displacement variables comprises the 10 degrees of freedom of the cubic
approach and the three rotational degrees of freedom related to the interface.
324 18 Hybrid–Mixed Triangular Plate Elements
as well as
Z
δWσ = . . . − {δmn w,n + δmt w,s } ds
I
are to be evaluated at the interface, whereat w,n again comprises linear and
quadratic parts. At evaluating the integrals related to the element domain and
to the interface the element continuity condition concerning w,n is fulfilled
weakly without employing further degrees of freedom. This effects the triangu-
lar element to comprise three nodal displacement variables [ w w,x w,y ] each
related to the corners of the element.
18.6 Test
To test the elements the square plate is investigated according to Section 6.2.9.
Taking into account the symmetry, a quarter of the plate is investigated ap-
plying different discretizations to illustrate the convergence of the results.
x clamped
G M ν = 0.2
hinged
l x = 10.0 m
l y = 10.0 m
lx
The elements presented in the sections above and related to the Kirchhoff ’s
plate theory apply the PvD in the representation
Z
−δWd = B · {δw,xx (w,xx + νw,yy ) + δw,yy (w,yy + νw,xx )
A Z
+ 2 (1 − ν) δw,xy w,xy } dA − δw pz dA = 0 . (19.1)
A
Here the internal work is described introducing the curvatures of the deflection
surface.
Nevertheless, the natural formulation of the work due to bending employs the
gradients of the rotations, since the curvatures are defined herewith. Thus,
without consideration of the Kirchhoff–hypothesis, it can be stated
Z
−δWd = B · {δϕx ,x (ϕx ,x + νϕy ,y ) + δϕy ,y (ϕy ,y + νϕx ,x )
A Z
1
+ (1 − ν) (δϕx ,y + δϕy ,x ) (ϕx ,y + ϕy ,x )} dA − δw pz dA . (19.2)
2 A
Applying this formulation the virtual work is completely described. The Kirch-
hoff–hypothesis is missing, without taking it into account, the constraints of
the rotations ϕx , ϕy and the deflection w are not considered, and the boundary
condition related to w cannot be fulfilled. To take into account the Kirchhoff–
hypothesis different paths of solution may be followed. These are in essence:
C B
w w,x w,y w w,x w,y
Fig. 19-1 Nodal degrees of freedom related to the deflection and the rotations
A ( λ a = 1, λ b = 0, λ c = 0 )
F ( 0.5, 0, 0.5 )
D ( 0.5, 0.5, 0 )
C ( 0, 0, 1)
y E ( 0, 0.5, 0.5 ) B ( 0, 1, 0 )
and a respective approach to describe the virtual rotations. Hereby, the inte-
gration of the virtual work can be evaluated with respect to the coefficients
a. The scaling of the general polynomial to the physically meaningful nodal
rotations [ ϕx ϕy ]A, B, C takes place by evaluating the approach with respect
to the coordinates of the element nodes. In matrix notation follows
vϕ = Ψ̃ϕ · aϕ
and in detail
" #
ϕ a
x 1 1
ϕ a
y 1 2
" #A
ϕ a3
x 1
ϕy 1 a4
" # B
a
ϕx 1 5
ϕy a6
1
" #C = · .
ϕx a7
1/4 1/4 1/4
ϕy 1/4 1/4 1/4 a8
" # D
a
ϕx 1/4 1/4 1/4 9
1/4 a10
ϕy
" # E 1/4 1/4
ϕx a11
1/4 1/4 1/4
ϕy 1/4 1/4 1/4 a12
F
Since, applying the PvD and the approaches to describe the rotations, the
Kirchhoff–hypothesis is not taken into account yet, it needs a second step to
fulfill the Kirchhoff–hypothesis. Therefor a cubic Hermite–polynomial is chosen
to describe w at the element boundaries, which is assigned to the deflections w
and to the rotations w,s according to the element nodes of the respective boun-
dary, see Figure 19-3. Thus it is feasible to describe the Kirchhoff–hypothesis
along the boundary employing the displacements w and the rotations ϕx , ϕy
and to fulfill the element continuity conditions related to w.
A
s
[ w w,s ] AB
[ w w,s ] CA
s
B
C [ w w,s ] BC s
whereat the shape functions φi are given in Figure 6-2. Employing this approach
the discrete fulfillment of the Kirchhoff–hypothesis takes place at the corners of
the element as well as at the midpoints of the edges, if it is possible to describe
the previous degrees of freedom related to the element rotations [ ϕxi ϕyi ] with
the new degrees of freedom [ w w,x w,y ] related to the displacement variables.
The evaluation of the Kirchhoff–hypothesis concerning the corner nodes yields
ϕt |0.5ℓ = − 1.5
ℓ
· w|i − 0. 25 · w,s |i + 1.5
ℓ
· w|i+1 − 0. 25 · w,s |i+1 (19.4)
the slopes w,n and w,s are connected to the slopes w,x and w,y of the nodal
degrees of freedom, see Figure 19-4.
x
w,s
w, n
D
s w,y s
β
td td
y
nd w,x nd
Fig. 19-4 The transformation of w,s and w,n with respect to w,x and w,y
332 19 Discrete Kirchhoff –Theory Element
The scaling of the nodal rotations with respect to the nodal degrees of freedom
of the deflection surface w is summarized in the scaling matrix Gw with
vϕ = Gw · vw .
The scaling employs the twelve nodal rotations vϕ and the nine degrees of
freedom vw at the three corner nodes A, B, C, which describe the deflection at
the element boundaries.
19.1 The Discrete Kirchhoff Triangular Element 333
The scaling matrix may be divided into submatrices with respect to the degrees
of freedom at the nodes A, B, C
" #
ϕx
ϕy
" #A
ϕx
ϕy
"
#B
ϕx
vwA
ϕy
" C
# = GwA GwB GwC · vwB .
ϕx
vwC
ϕy
"
#D
ϕx
ϕy
"
#E
ϕx
ϕy
F
The submatrices are given in detail with the abbreviations cIJ = cos β, and
sIJ = sin β, and the length ℓ of the edge IJ
0 1 0
0 0 1
0 0 0
0 0 0
0 0 0
w
0 0 0
GwA · vwA = · w,x
,
1.5
−
ℓ
cAB −0. 25 c2AB + 0. 5 s2AB −0. 75 cAB sAB
w,y A
1.5
−0. 25 s2AB + 0. 5 c2AB
− ℓ sAB −0. 75 cAB sAB
0 0 0
0 0 0
+ 1.5 c 2 2
AC −0. 25 cAC + 0. 5 sAC −0. 75 cAC sAC
ℓ
1.5
+ ℓ sAC −0. 75 cAC sAC −0. 25 s2AC + 0. 5 c2AC
334 19 Discrete Kirchhoff –Theory Element
0 0 0
0 0 0
0 1 0
0 0 1
0 0 0
w
0 0 0
GwB · vwB = · w,x ,
+ 1.5 cAB −0. 25 c2AB + 0. 5 s2AB −0. 75 cAB sAB
ℓ w,y B
+ 1.5 −0. 25 s2AB + 0. 5 c2AB
ℓ
sAB −0. 75 cAB sAB
− 1.5 cBC −0. 25 c2BC + 0. 5 s2BC
ℓ
−0. 75 cBC sBC
− 1.5
ℓ
sBC −0. 75 cBC sBC −0. 25 s2BC + 0. 5 c2BC
0 0 0
0 0 0
0 0 0
0 0 0
0 0 0
0 0 0
0 1 0
w
0 0 1
GwC · vwC = · w,x .
0 0 0
w,y C
0 0 0
1.5
+
ℓ
cBC −0. 25 c2BC + 0. 5 s2BC −0. 75 cBC sBC
1.5
−0. 25 s2BC + 0. 5 c2BC
+ ℓ sBC −0. 75 cBC sBC
− 1.5 cAC −0. 25 c2AC + 0. 5 s2AC −0. 75 cAC sAC
ℓ
− 1.5
ℓ
sAC −0. 75 cAC sAC −0. 25 s2AC + 0. 5 c2AC
a = Gϕ · Gw · vw = Gϕw · vw .
As in the case of Gw the scaling matrix Gϕw is divided with respect to the
19.1 The Discrete Kirchhoff Triangular Element 335
0 1 0
0 0 0
0 0 0
− 6ℓ cAB −1 − c2AB + 2 s2AB
−3 cAB sAB
+ 6ℓ cAC −1 − c2AC + 2 s2AC −3 cAC sAC
w
GϕwA · vwA
= 0 0 0
· w,x ,
0 0 1
w,y A
0 0 0
− 6ℓ sAB −1 − s2AB + 2 c2AB
−3 cAB sAB
0 0 0
+ 6ℓ sAC −1 − s2AC + 2 c2AC
−3 cAC sAC
0 0 0
0 0 0
0 1 0
0 0 0
6
+ ℓ cAB −1 − c2AB + 2 s2AB
−3 cAB sAB
0 0 0
6
w
2 2
= − ℓ cBC −1 − cBC + 2 sBC −3 cBC sBC
GϕwB · vwB · w,x ,
0 0 0
w,y B
0 0 1
0 0 0
6
+ sAB
ℓ −3 cAB sAB −1 − sAB + 2 c2AB
2
0 0 0
6
− ℓ sBC −3 cBC sBC −1 − sBC + 2 c2BC
2
336 19 Discrete Kirchhoff –Theory Element
0 0 0
0 0 0
0 1 0
0 0 0
6
− cAC −1 − c2 + 2 s2 −3 cAC sAC
ℓ AC AC w
6
+ cBC −1 − c2BC + 2 s2BC
−3 cBC sBC
GϕwC · vwC = ℓ · w,x .
0 0 0
w,y C
0 0 0
0 0 1
0 0 0
−6 s −3 c s −1 − s 2 2
ℓ AC AC AC AC + 2 cAC
6
+ ℓ sBC −3 cBC sBC −1 − sBC + 2 c2BC
2
The computation of the load vector with respect to distributed loadings p(x, y)
has to be performed by direct integration – without employing the shape func-
tions – and it has to be allocated to the deflections of the corner nodes. This is
necessary, because the deflection is not approached inside the element domain.
Thus the element load vector is given with
A
fT = [ 1 0 0 1 0 0 1 0 0 ] · ·p .
3
The subsequent analysis concerning the moments σ T = [ mxx myy mxy ]A, B, C
also takes place applying the scaling matrix Gϕw , since the stress matrix is
formulated employing the general approaches to describe the rotations
σ = E · D · Ψϕ · Gϕw · vw = Sϕ · Gϕw · vw .
19.3 Test 337
2ba 0 0 bb bc 0 ν · 2aa 0 0 ν · ab ν · ac 0
ν · 2ba 0 0 ν · bb ν · bc 0 2aa 0 0 ab ac 0
e · 2aa 0 0 e · ab e · ac 0 e · 2ba 0 0 e · bb e · bc 0
0 2bb 0 ba 0 bc 0 ν · 2ab 0 ν · aa 0 ν · ac
0 ν · 2bb 0 ν · ba 0 ν · bc 0 2ab 0 aa 0 ac ,
0 e · 2ab 0 e · aa 0 e · ac 0 e · 2bb 0 e · ba 0 e · bc
0 0 2bc 0 ba bb 0 0 ν · 2ac 0 ν · aa ν · ab
0 0 ν · 2bc 0 ν · ba ν · bb 0 0 2ac 0 aa ab
0 0 e · 2ac 0 e · aa e · ab 0 0 e · 2bc 0 e · ba e · bb
e = (1 − ν)/2 , B = E · t3 /12 (1 − ν 2 ) .
19.3 Test
As an example to test the formulation the square plate is taken according
to Section 6.2.9 and Figure 19-5. Exploiting the symmetry, only a quarter
of the plate is investigated applying different discretizations to illustrate the
convergence of the results.
x clamped
G M ν = 0.2
hinged
l x = 10.0 m
l y = 10.0 m
lx
The subsequent Tables 19.1 and 19.2 demonstrate the behavior of convergence
of different variables of state concerning the DKT element employing 9 degrees
of freedom related to the element. Regarding nodes, that belong to multiple
elements, the results are presented each with respect to the neighboring ele-
ments.
338 19 Discrete Kirchhoff –Theory Element
Comparing the results applying the plate elements according to Sections 17.2
and 18 illustrates, that the results obtained by the DKT element converge
somewhat inferior against the reference solution. In particular, to obtain suf-
ficiently accurate stress resultants, a more refined discretization is required.
Moreover the displacements converge against the reference values approaching
from above. The numerical efficiency can be seen as an advantage, since the
computation of the stiffness matrix may follow a 3–point–integration and an
analytical integration respectively.
20 Benchmark Concerning Triangular Plate Elements
Although all elements, that are described in Sections 17, 18, and 19, converge
against the reference solution, it is useful to compare the convergence behavior
and to discuss advantages and disadvantages. Therefor two benchmarks are
defined, which are applied throughout the textbook.
G M ν = 0.2
hinged
l x = 10.0 m
l y = 10.0 m
lx
The comparison of the numerical effort to get the results is partly possible by
taking into account the number of the degrees of freedom and the bandwith.
The effort to compute the element matrices is a second indicator, but it de-
pends strongly on kind as well as preparation of programming, if analytical
integration or numerical integration is applied. Thus, Table 20.1 compares the
element formulations with 9 and 9+3 degrees of freedom to the element for-
mulation given in Section 11 employing 18 degrees of freedom. The number of
elements, the number of unknowns, and the bandwidth are chosen for different
discretizations such, that the mesh lines are kept constant.
△9 △ 9 (+3) △ 18
mesh lines 2 3 5 9 2 3 5 9 2 3 5 9
elements 2 8 32 128 2 8 32 128 2 8 32 128
unknowns 12 27 75 243 17 43 131 451 24 54 150 486
bandwidth 12 15 21 33 17 23 35 59 24 30 42 66
% error of wM
K-DT-3-18
DKT-3-9
10
K-HMT-3-9
K-HMT-6-12
K-HDT-3-9
K-HDT-6-12
5 K-MT-6-12
0
20 100 500 dof
-5
-10
The 18 DOF element shows superior convergence with respect to all state
variables due to the employed shape functions of 5th order. Nonetheless, due to
the kinematic variables of second order, it is limited to benchmark applications.
The elements K-HDT-3-9 and K-HDT-6-12 converge very good as well, but the
last one employs 12 DOF and thus needs more effort to compute the stiffness
matrix and to solve the system of equations.
The convergence of the hybrid–mixed elements K-HMT-3-9 and K-HMT-6-12
is not as good with respect to the deflection, but convergence is better with
respect to the moments, since the formulations employ quadratic polynomials
to desrcibe the moments inside the element before the elimination process takes
place.
The DKT element and the K-MT-6-12 element converge a little inferior, but in
comparison to the other elements they need less effort to compute the element
matrix.
% error of mxxM
70
60 K-DT-3-18
DKT-3-9
50 K-HMT-3-9
K-HMT-6-12
40 K-HDT-3-9
K-HDT-6-12
K-MT-6-12
30
20
10
0
20 100 500 dof
-10
-20
% error in m yyE
-5
-10
-15
-20
-25
K-DT-3-18
DKT-3-9
-30 K-HMT-3-9
K-HMT-6-12
-35 K-HDT-3-9
K-HDT-6-12
-40 K-MT-6-12
Remarks
• All elements converge against the same solution.
• The element employing 18 DOF is not applicable to general problems.
• The convergence of the K-MT-6-12 published by Harvey–Kelsey [42] is
somewhat inferior in comparison to other elements.
• All elements with 9+3 DOF do not converge much better compared to
the correlated elements with 9 DOF and a weak formulation of the ele-
ment continuity conditions.
• Although the convergence is not equal with respect to all state variables,
the DKT element and the hybrid–mixed element K-HMT-3-9 show a good
overall performance with respect to convergence and effort.
20.2 Trapezoidal Plate Subjected to Distributed Loading 343
e
fre
ly
ν = 0.3
e
fre
lx = 12.0 in
ly = 8.48 in y S
x E clamped
lx
Remarks
• The elements incorporating midpoints at the edges significantly better
converge than the respective elements without midpoints at the edges.
• The deflection wM converges to values 0. 28 in ≤ wM ≤ 0. 29 in, never-
theless which element formulation is addressed. The deflection measured
from the experiment wM ≈ 0. 297 in, according to [27], is slightly larger
than the computed values, which may be ascribed to modeling inaccuracy.
• The clamping moments related to the elements K-DT-6-21, K-DT-3-18,
K-HMT-6-12, K-HMT-3-9 converge against values −7. 10 lbs. in/in ≤
myE ≤ −6. 90 lbs. in/in. The clamping moments related to the element
K-MT-6-12 slower converge, which also is regarded to some extent at the
tests represented in Section 17.
20.2 Trapezoidal Plate Subjected to Distributed Loading 345
Table 20.2 Rhombus plate – results for distributed loading p = 0. 26066 psi
units: w [ in ], m [ lbs.in/in ]
Table 20.3 Number of nodal displacement variables for the chosen discretizations
Figures 20-6 and 20-7 represent the convergence behavior by charts. The mo-
notonic convergence of the deflections is obvious, they tend to the same final
value. Due to the approach of fifth order the element K-DT-6-21 is the best,
nevertheless it employs kinematic degrees of freedom relating to second order
derivatives. The hybrid–mixed elements exhibit a very good convergence be-
havior that corresponds to the respective approach, in particular at regarding
the convergence of the clamping moments.
% error in w
-10.0
-20.0 K-DT-6-21
K-MT-6-12
-30.0
DKT-3-9
-40.0 K-HMT-6-12
K-HMT-3-9
-50.0
% error in m yyE
+10.0
20 100 500 1000 2000 dof
0.0
-10.0
-20.0 K-DT-6-21
K-MT-6-12
-30.0 DKT-3-9
K-HMT-6-12
K-HMT-3-9
x
w,x x
me
g
ments me [N m/m] may act along the beam axis, which can be interpreted as
couple of forces, see Figure 21-1–right hand side. Thus the governing equations
concerning beams including shear deformations γ are summarized as follows.
a) Kinematics
w,x − ϕ − γ = 0 ,
ϕ,x − κ = 0 .
b) Equilibrium
Q,x + p = 0 ,
M,x − Q − me = 0 .
c) Material Equations
GAS · γ − Q = 0 ,
EI · κ + M = 0 .
Shear deformations cause fundamental changes at developing a finite element
formulation, since the work equation applying the PvD is to be extended by
additional terms and the discretization of the work equation may effect non–
physical results to some extent, that are to be eliminated applying special
arrangements.
Introducing kinematics the primary variables concerning the PvD are determi-
ned with w and ϕ, which are independent from each other now and therefore
are to be described accordingly when choosing the shape functions.
Z
−δWd = {δϕ,x EI ϕ,x + (δw,x − δϕ) GAS (w,x − ϕ)} dx
Z
− δw · p dx = 0 . (21.1)
Due to the first derivatives of the variables the criteria of convergence offer
linear shape functions for both the deflection and the rotation employing the
nodal degrees of freedom [ wA ϕA wB ϕB ] inside the element, cf. Section 2.4.2
and Equation 21.2.
x x
w(x) = ( 1 − ) wA + wB ,
ℓ ℓ (21.2)
x x
ϕ(x) = ( 1 − ) ϕA + ϕB .
ℓ ℓ
21.1 Elements Employing Displacements as Primary Variables 351
x EI, GAS , l x
j
z,w
Due to the fact, that the rotation ϕ describes the rotation of the cross section
and not the inclination of the deflection line here, the realization of a clamping
is only possible with ϕ = 0, but is normally linked to a kink out of shear strain
because of w,x = γ + ϕ. Considering the boundary conditions wA = 0, δwA = 0
as well as ϕA = 0, δϕA = 0 yields the both equations
GAS GAS
δwB : + · wB − · ϕB =P, (21.4)
ℓ 2
GAS EI GAS · ℓ
δϕB : − · wB + + · ϕB = 0 , (21.5)
2 ℓ 3
with the solutions
!
Pℓ 1
ϕB = EI GAS ·ℓ
, (21.6)
2 ℓ + 12
!
1 ℓ2
wB = P ℓ + 2 . (21.7)
GAS 4EI + GA3S ·ℓ
352 21 Timoshenko Beam Elements
γ = w,x − ϕ (21.8)
this may lead to an unbalanced approach of the shear strains and the respective
virtual work. At choosing a linear approach to describe w and ϕ and due to the
shear strains being described employing a constant term w,x and a linear term
ϕ, this results in a serrated approximation of γ and thus of the shear forces.
One possibility to remedy this, is to choose an approach of one power more
to describe w than that to describe ϕ, whereby both parts of Equation (21.8)
comprise the same polynomial order.
Another possibility still is the choice of linear approaches to describe the varia-
bles w and ϕ, and though, instead of applying a linear course of ϕ to describe
the work performed on shear strains, only to take a constant average, see Figure
21-3.
w,x j
g
=
x
i i
At this case at hand the element matrix comprising selectively reduced integra-
tion differs from the element applying exact integration according to Equation
(21.3) only with regard to four components:
δwA : GAS GAS GAS GAS
ℓ 2
− ℓ 2
GAS EI GAS ℓ GA GAS ℓ
δϕA : − EI
2 ℓ
+ 4
− 2S ℓ
+ 4
. (21.9)
GA GAS GAS GAS
δwB :
− ℓ
S
− −
2 ℓ 2
GAS GAS ℓ GAS GAS ℓ
δϕB :
2
− EI
ℓ
+ 4
− 2
EI
ℓ
+ 4
After having introduced the boundary conditions, the computation of the dis-
placement variables of the cantilever according to Figure 21-2 yields the equa-
tions
GAS GAS
δwB : + · wB − · ϕB =P, (21.10)
ℓ 2
GAS EI GAS · ℓ
δϕB : − · wB + + · ϕB = 0 , (21.11)
2 ℓ 4
P ℓ2 1
ϕB = · , (21.12)
2 EI
1 ℓ2
wB = P ℓ · + . (21.13)
GAS 4EI
Concerning rigid shear behavior GAS → ∞, the part of the deflection related
to shear strains vanishes. The locking phenomenon does not exist. The element
stiffness matrices correctly describe the case of pure bending with respect to
the chosen approaches.
Example
Subsequently the results are represented referring to the example of a cantilever
beam loaded by a concentrated external action, cf. Firgure 21-4. It is essential
that, employing linear approaches, the course of the bending moment is con-
stant inside the element due to M = −EI · ϕ,x and also the course of the shear
force is constant because of the reduced integration of the work terms related
to shear behavior though Q = GAS · (w,x − ϕ) is valid.
P
EI = 4,000 kNm 2
x T
GA S = 200 kN
j
lx = 5m
z,w
P = 2 kN
M [kNm]
-8.75 -6.25 -3.75 -1.25
2.0 Q [kN]
the different orders of the terms being introduced into the condition of equili-
brium M,x − Q = 0. A constant approach to describe Q and accordingly δQ
inside the element eliminates the oscillations, moreover concentrated external
actions P may be put on all nodes of the element.
Example
At testing the mixed element gives excellent results, which are represented in
Figure 21-5. Subsequently the results are represented concerning a cantilever
beam loaded by a concentrated external action. Due to the linear approaches
356 21 Timoshenko Beam Elements
to describe the deflection w and the bending moment M the mixed element
yields the exact result applying the element matrix (21.18) and arbitrary shear
stiffnesses GAS > 0.
P
EI = 4,000 kNm 2
x T
GA S = 200 kN
j
lx = 5m
z,w
P = 2 kN
-10.0
M [kNm]
2.0 Q [kN]
By analogy with Section 15.3, the elimination of the bending moments yields
the deflection w and the rotation ϕ being system variables related to the in-
terface. Thus, instead of the boundary conditions of the mixed formulation
21.4 Convergence Behavior Concerning the Beam Elements 357
related to w and M , the boundary conditions related to w and ϕ are now the
substantial ones and are exactly fulfilled by the approaches at the interface.
p EI = 4,000 kNm 2
T GA S = 200 kN
x
j lx = 5m
z,w p = 2 kN/m
Figure 21-7 shows the moment as well as the shear force distribution concerning
the element employing displacements as primary variables and the mixed as well
as the hybrid–mixed element.
M d [kNm] M hm [kNm]
-19.531 -10.156 -3.906 -0.781 -25.00 -14.06 -6.25 -1.56 0.0
Q d [kN] Q hm [kN]
8.75 6.25 3.75 1.25 8.75 6.25 3.75 1.25
force is evaluated from the first order derivative of the bending moment in the
element, hence the courses are constant element by element.
Applying the hybrid–mixed element the bending moment is linearly approxi-
mated in the element domain and the shear force is computed from the first
order derivative of the bending moment.
Regarding the deflection the results are represented in Table 21.1. The ele-
ment employing displacements as primary variables yields, independent of the
number of elements and of the shear stiffness, the exact results concerning the
deflection wT . The mixed and the hybrid–mixed element yield identical results
but only an approximation concerning the deflection wT . Nevertheless, the de-
flection converge against the exact values at an increasing number of elements.
Introducing high values for the shear stiffness, all elements presented here con-
verge against the solution regarding the Euler–Bernoulli beam theory.
EI/GAS [ m2 ] 20 2 0,2
displacement–based elements 1 0. 1640625 0. 0515625 0. 0403125
2 0. 1640625 0. 0515625 0. 0403125
4 0. 1640625 0. 0515625 0. 0403125
8 0. 1640625 0. 0515625 0. 0403125
mixed elements 1 0. 1770833 0. 0645833 0. 0533333
2 0. 1673177 0. 0548177 0. 0435677
4 0. 1648763 0. 0523763 0. 0411263
8 0. 1642660 0. 0517660 0. 0405160
hybrid–mixed elements 1 0. 1770833 0. 0645833 0. 0533333
2 0. 1673177 0. 0548177 0. 0435677
4 0. 1648763 0. 0523763 0. 0411263
8 0. 1642660 0. 0517660 0. 0405160
22 Plate Elements Including Shear Deformations
t = 0.1 m
hi
p v = 0.2
GAS = 26,040 kN/m
e
fre
l x = 10.0 m
y hing
ed l y = 10.0 m
According to Kirchhoff ’s plate theory the twisting moments are unequal zero
at a simply–supported boundary. Regarding a simply–supported plate they
increase along the boundary up to the corner, where they effect a lift–off force.
Applying the Reissner–Mindlin plate theory the twisting moment at the simply–
supported boundary is equal zero, if the conjugated rotation can be reached
unrestrictedly. This is denoted as a soft bearing. At increasing the shear stiffness
the gradient of the twisting moment continuously rises in the direct neighbor-
hood of the simply–supported boundary, till the twisting moment approaches
Kirchhoff ’s solution at GAS → ∞. This vividly justified limit becomes visible,
when the twisting moment is equal to zero at the boundary and rises to a finite
value at a small distance to the boundary. This implies at the same time, that
a corresponding fine discretization is necessary in the boundary region.
In Figure 22-2 the bending and the twisting moments according to Kirchhoff ’s
plate theory are opposed to the results according to the Reissner–Mindlin plate
theory. Symmetries and antisymmetries of the stress resultants can be observed
for both plate theories. However, at the Reissner–Mindlin theory, the bending
moments are mxx 6= 0, myy 6= 0 along the boundaries and, due to the chosen
soft bearing, the twisting moments are mxy = 0.
mxx [ kN m/m ]
myy [ kN m/m ]
mxy [ kN m/m ]
Also the course of the shear forces mainly differs at the boundary region of the
plate, cf. Figure 22-3. Instead of the corner force according to Kirchhoff ’s plate
theory the shear forces qx , qy according to the Reissner–Mindlin theoriy reach
corresponding high values. Applying the Reissner–Mindlin theory, the boun-
dary conditions related to the shear forces qn = 0 and related to the moments
mn = 0, mt = 0 are ensured at the free boundary. Applying Kirchhoff ’s plate
theory, only the conditions related to mn = 0 and related to the Kirchhoff ’s
effective shear force qKn = 0 are fulfilled.
qx [ kN/m ]
qy [ kN/m ]
w [m]
Fig. 22-3 Reinforced concrete plate – bending surface and shear forces,
a.) left according to Kirchhoff’s plate theory,
b.) right according to the Reissner–Mindlin plate theory
362 22 Plate Elements Including Shear Deformations
y x
z
mxy
ϕy
myx
ϕx
a) Kinematics
κxx − ϕx,x = 0
κyy − ϕy,y = 0
2κxy − ϕx,y − ϕy,x = 0
γxz − w,x + ϕx = 0
γyz − w,y + ϕy = 0
Introducing the additional equations related to the shear strains, the rotations
ϕn , ϕt are independent of w and, accordingly, are to be regarded as independent
kinematic boundary conditions. Furthermore, since the shear strains do not
vanish, they need the shear forces as conjugated variables.
22.3 Weak Formulation of the Governing Equations 363
b) Equilibrium
qx,x + qy,y + p = 0
mxx,x + mxy,y − qx + mex = 0
myy,y + mxy,x − qy + mey = 0
Nevertheless the eqilibrium conditions are identical to the ones related to Kirch-
hoff ’s plate theory but analogously to the Timoshenko beam extended by ex-
ternal actions mex , mey . Three independent boundary conditions according to
mn , mt , qn may be forced at each boundary now, whereat external actions at
the boundaries could be applied as well.
c) Material equations
mxx + B(κxx + ν · κyy ) = 0
myy + B(κyy + ν · κxx ) = 0
mxy + B(1 − ν)κxy = 0
qx − GAS γxz = 0
qy − GAS γyz = 0
Hereby, the plate bending stiffness B and the shear stiffness GAS , that, accor-
ding to the Reissner–Mindlin theory, is modified by the shear correction factor
κQ compared to the complete cross section area, are determined as follows:
E · t3 E ·t
B= , GAS = κQ · .
12(1 − ν 2 ) 2(1 + ν)
The shear correction factor κQ is proposed by Reissner to be 5/6, but is chosen
differently, depending on the derivation and on the material.
The PvD can be transformed into the work equation, that bases the finite
element formulation, by incorporating the material equations and kinematics.
First of all, introducing the material equations yields
Z
δW = − B{δκxx (κxx + νκyy ) + δκyy (κyy + νκxx ) + 2δκxy (1 − ν)κxy )} dA
Z
− {GAS (δγxz γxz + δγyz γyz )} dA
Z
+ {δwp + δϕx mex + δϕy mey } dA = 0 .
Since the deflection w and the rotations ϕx , ϕy as well as the respective virtual
displacement variables are independent from each other, the work equation
must be fulfilled separately for all virtual displacement variables.
Z
δWw = − {GAS δw,x (w,x − ϕx ) + GAS δw,y (w,y − ϕy )} dA
Z
+ δwp dA = 0 , (22.2)
Z
δWϕx = − B{δϕx ,x (ϕx ,x + νϕy ,y ) + 2 (1 − ν) δϕx ,y (ϕx ,y + ϕy ,x )} dA
Z Z
+ GAS δϕx (w,x − ϕx ) dA + δϕx mex dA = 0 , (22.3)
Z
δWϕy = − B{δϕy ,y (ϕy ,y + νϕx ,x ) + 2 (1 − ν) δϕy ,x (ϕx ,y + ϕy ,x )} dA
Z Z
− GAS δϕy (w,y − ϕy ) dA + δϕy mey dA = 0 . (22.4)
In the following, the work equations (22.2), (22.3) and (22.4) provide the basis
for the discretization with the finite element method. Therefor the respecti-
ve approaches are to be chosen for the independent displacement variables
[ w ϕx ϕy ].
22.4 Quadrilateral Element Employing a Bi–linear Approach 365
u = Ω·v. (22.5)
δu = Ω · δv . (22.6)
Z
1−ν T 1−ν T
+ B BTx Bx + 2 By By νBTx By + 2 By Bx dA
1−ν T 1−ν T
νBTy Bx + 2 Bx By BTy By + 2 Bx Bx
However, employing the linear approaches chosen here, by analogy with the
Timoshenko beam element, the numerically caused locking–phenomenon arises,
that is effected by the transverse shear strains. With
γxz = w,x − ϕx bzw. γyz = w,y − ϕy
and employing linear approaches to describe w and ϕx a constant course con-
cerning w,x and a linear course concerning ϕx are added up, which causes
non–physically oscillations at integration and, as a consequence, an unwanted
stiffening of the element. This phenomenon may by avoided by selectively re-
duced integration of the respective terms of work.
Hence, the numerical integration of the work at element level is performed ap-
plying a 2 × 2 integration concerning the bending parts and a 1 × 1 reduced in-
tegration concerning all shear parts, what leads to an element, which is referred
to as RM-DQ-4-SRI. However, due to the selectively reduced integration, zero–
energy modes may exist. This means, that deformations may arise, that require
22.4 Quadrilateral Element Employing a Bi–linear Approach 367
The basic idea to stabilize the stiffness matrix, is denoted as hour–glass con-
trol referring to Kavangh and Key [54], Kosloff et al. [57], see also Belytschko
et al. [36, 18, 17]. Nonetheless, the choice of the stabilization matrix is diffi-
cult, depends on the model equations, and governs the convergence against the
reference solution.
One possibility consists in weighting the stiffness matrix K1×1 related to the
reduced integration of all parts by the matrix K2×2 related to the complete
integration of all parts. In the following this element is referred to as RM-DQ-
4-HC. With
K = K1×1 + ε · (K2×2 − K1×1 ) (22.7)
the results may be influenced substantially. The multiplier ε has to be adjusted
to the problem to be investigated. Employing ε < 0. 01 already leads to a signi-
ficantly slower convergence of the displacements against the reference solution.
Nevertheless, it is advantageous, that unwanted oscillations of moments and
shear forces are highly damped.
Belytschko et al. [17] develop another stabilization matrix applying only a
weighting of the shear stiffness and propose to connect the weighting to the
element geometry, they implement ε = r · t2 /A, whereat 0. 01 < ε < 0. 05 may
yield feasible results,
K = KB,2×2 + {KS,1×1 + ε · (KS,2×2 − KS,1×1 )} . (22.8)
In the following this formulation is named RM-DQ-BHC.
A far more effective method is the application of modified approaches for the
deflection surface w(x, y) as well as for the rotations ϕx (x, y), ϕy (x, y), which
are aligned to kinematics. In Section 22.4.1 bi–linear approaches are chosen
to describe the deflections and the rotations, that result in the phenomena
discussed before. Taking into account that, by analogy with the beam, both
terms describing the shear strains γxz = w,x − ϕx and γyz = w,y − ϕy should
comprise the same polynomial order, then, the terms w,x and w,y should be
described employing bi–linear approaches.
This can be managed, if a hybrid–displacement element is applied, that is
referred to as RM-HDQ-4-5. In addition to the bi–linear approaches to describe
[ w ϕx ϕy ] with respect to the element nodes a−d , which match the approaches
onto the interfaces A − D , additional quadratic approaches of w are chosen
within the element domain, that are assigned to the element nodes e − i subject
to Figure 22-5. This approach is referred to as method of incompatible modes
and discussed by Wilson et al. [104], Taylor et al. [96], Ibrahimbegovic and
Wilson [50], and Simo and Rifai [92].
x A
y
a
h e
i B
d b
D
g f
c
η ξ
C
Fig. 22-5 Element nodes related to the element boundaries and the interfaces
Subsequently the approaches to describe w(x, y) are given, that fulfill this con-
dition. As before, the bi–linear approaches are assigned to the degrees of free-
dom related to the element corner nodes:
1
φa = · ( 1. 0 − ξ ) · ( 1. 0 − η ) ,
4
1
φb = · ( 1. 0 + ξ ) · ( 1. 0 − η ) ,
4
1
φc = · ( 1. 0 + ξ ) · ( 1. 0 + η ) ,
4
1
φd = · ( 1. 0 − ξ ) · ( 1. 0 + η ) .
4
370 22 Plate Elements Including Shear Deformations
The quadratic approaches are hierarchically developed and are assigned to the
degrees of freedom related to the midpoints of the edges and to the center of the
element. It is essential, that the approaches φe , φi are quadratic in ξ–direction
and are linear in η–direction. In contrast, the approaches φf , φh are quadratic
in η–direction and linear in ξ–direction. The approach φi concerning the degree
of freedom related to the center node of the element is bi–quadratic. This yields
1
φe = · ( 1. 0 − ξ 2 ) · ( 1. 0 − η ) ,
2
1
φf = · ( 1. 0 + ξ ) · ( 1. 0 − η 2 ) ,
2
1
φg = · ( 1. 0 − ξ 2 ) · ( 1. 0 + η ) ,
2
1
φh = · ( 1. 0 − ξ ) · ( 1. 0 − η 2 ) ,
2
φi = ( 1. 0 − ξ 2 ) · ( 1. 0 − η 2 ) .
Hence the element comprises 9 degrees of freedom concerning w(x, y) and 4 de-
grees of freedom each to employ bi–linear approaches concerning the rotations
ϕx (x, y) and ϕy (x, y) inside the element domain, but 4 degrees of freedom on
the interfaces for w(s), ϕx (s), ϕy (s) each.
A
y x
z
p B
d b
D
σ σ
ϕx
σ c σ
ϕx ϕy
w
w ϕy
C
Fig. 22-6 Displacement variables and stress resultants at the hybrid element
22.5 Hybrid–Displacement Quadrilateral Element 371
if the shape functions at the interface ui, lin. as well as at the element domain
uE, lin. employ identical courses and variables. Because of the orthogonality
condition also
Z
δuTE, quadr. · σ ds = 0
E
is valid and has to be taken into account with respect to δWd, E = 0. This
yields
Z Z
δuTE, lin. · σ ds = δuTE, lin. · DT · E · D · (uE, lin. + uE, quadr. ) dA ,
E E
Z
0= δuTE, quadr. · DT · E · D · (uE, lin. + uE, quadr. ) dA ,
E
and the second equation can be applied to eliminate the variables of the qua-
dratic shape functions at the element level.
372 22 Plate Elements Including Shear Deformations
In a second step the degrees of freedom related to the quadratic approaches are
eliminated, so that the remaining element stiffness matrix K is solely assigned
to the degrees of freedom related to the corner nodes. With
" # " # " #
K12, 12 K12, 5 v12 f12 6= 0 at element level
· −
K5, 12 K5, 5 v5 05 = 0 at element level
Due to the final representation of the work equations the elimination of the
degrees of freedom v5 is possible without taking into account the respective
contour integrals. Thus the element continuity conditions may be ensured only
concerning the linear approaches.
σ = E · D· u,
u =Ω·v
The bending moments and the shear forces are computed inside the element
domain by means of the bi–linear shape functions only, without considering the
quadratic parts. The evaluation of the shape functions takes place concerning
the bending moments at the corner nodes and concerning the shear forces at
the center of the element in order to avoid unphysical oscillations.
22.6 Comparison of the Elements 373
22.6.1 Test 1
At first a cantilever plate loaded by a twisting moment is investigated accor-
ding to Figure 22-7. The main deformation behavior is related to twisting, but
in the vicinity of the load application a small part of the loading is carried via
bending. The shear deformations are negligible. As a reference concerning the
displacement variables and the twisting moment the values of the correspon-
ding torsion bar may be chosen:
5m
P
P
1m
x EI = 16,666 kNm2
y GA S = 10 7 kN
z, w
P = 1 kN
22.6.2 Test 2
As a second example the cantilever plate loaded by pure bending is investigated
according to Figure 22-8. The results may be calibrated to the reference values
of a corresponding beam, whereby the shear deformations are negligible here.
Concerning the beam and employing E = 2. 0·108 kN/m2 , G = 1. 0·108 kN/m2 ,
22.6 Comparison of the Elements 375
x EI = 16,666 kNm2
y GA S = 10 7 kN
z, w
P = 1 kN
22.6.3 Remarks
The chosen structure is statically determined, so that, applying the PvD, the
equilibrium is ensured and the resultant stress variables converge against the
correct values in almost all cases. The displacements naturally depend on the
stiffnesses, that strongly differ from each other in parts according to the element
formulations shown here, hence the following comments may be given.
22.6.4 Benchmark
As a benchmark to test the element, again, by analogy with the triangular
Kirchhoff plate elements according to Section 20, the unilaterally clamped
rhombus plate in Figure 22-9 is investigated referring to [103, 27]. All calculati-
ons are performed with the element employing modified approaches according
to Section 22.5.
ly
ν = 0.3
e
fre
lx = 12.0 in
ly = 8.48 in y S
x E clamped
GA S variable
lx
E · t3
B= = 1,878 lbf. in2/in ,
12 · (1 − ν 2 )
E·t
GAS = κQ · = κQ · 505,000 lbf ,
2 · (1 + ν)
whereby κQ acts as a parameter to scale the shear stiffness. Due to the slender-
ness of the plate, exhibiting a ratio ℓ / t = 12 / 0. 125 = 96, the shear deformati-
ons are negligible at realistic values of QAS . Thus, subsequently the parameters
378 22 Plate Elements Including Shear Deformations
100 wM [inch]
0.70
0.60
0.40
0.30
K-D-6-21 Kirchhoff
0.20 GA s = 50500 lbf
GA s = 505000 lbf
GA s = 5050000 lbf
The subsequent Tables 22.3 and 22.4 comprise the results concerning the de-
flections wM and the clamping moment myyE . In accordance with this study
the following statements can be made:
As stated in Section 22.4, the work equations (22.10), (22.11) and (22.12) are
basis for the discretization employing triangular elements. Therefor the respec-
tive approaches are to be chosen to describe the independent displacement
fields w, ϕx , ϕy .
w ϕx ϕy
A
w w
F D
C E B
w ϕx ϕy w w ϕx ϕy
Dx = [ ∂x ] , Dy = [ ∂y ] ,
Bxw = Dx · φ2A
φ2B φ2C φ2D φ2E φ2F , Byw = Dy · φ2A φ2B φ2C φ2D φ2E φ2F ,
2 2 2 2 2 2
φw = φA φB φC φD φE φF .
The derivatives of the shape functions related to the rotations may also be
summarized in matrix notation by
Bxϕ = Dx · φ1A φ1B φ1C , Byϕ = Dy · φ1A φ1B φ1C ,
φϕ = φ1A φ1B φ1C .
K=
BTxw Bxw + BTyw Byw −BTxw φϕ −BTyw φϕ
Z
GAS
−φTϕ Bxw φTϕ φϕ dA
−φTϕ Byw φTϕ φϕ
Z
1−ν T 1−ν T
+ B BTxϕ Bxϕ + 2 Byϕ Byϕ νBTxϕ Byϕ + 2 Byϕ Bxϕ dA .
1−ν T 1−ν T
νBTyϕ Bxϕ + 2 Bxϕ Byϕ BTyϕ Byϕ + 2 Bxϕ Bxϕ
σ m = Em · D · Ωϕ · vϕ = Sm · vϕ .
The subsequent stress analysis concerning the shear forces takes place applying
kinematics γxz = w,x − ϕx and γyz = w,y − ϕy . Matrix–vector notation yields
σ q = E q · D · Ω · v = Sq · v .
Due to the chosen approaches the courses of the shear forces are linear inside
the element, and in detail given by
wA
wB
wC
qxA 3ba −bb −bc 4bb 0 4bc −1 0 0 0 0 0
wD
qyA 3aa −ab −ac 4ab 0 4ac 0 0 0 −1 0 0 wE
qxB GA −b 3b −b 4b 4b 0 0 −1 0 0 0 0 wF
S a b c a c
= · · .
qyB 2A∆ −aa 3ab −ac 4aa 4ac 0 0 0 0 0 −1 0 ϕxA
ϕ
qxC −b −b 3b 0 4b 4b 0 0 −1 0 0 0 xB
a b c b a ϕ
xC
qyC −aa −ab 3ac 0 4ab 4aa 0 0 0 0 0 −1
ϕy A
ϕy B
ϕy C
384 22 Plate Elements Including Shear Deformations
ly
ν = 0.3
e
fre
lx = 12.0 in
ly = 8.48 in y S
x E clamped
GA S = 5.05 .105 lbf
lx ly
The subsequently given Table 22.5 illustrates the behavior of convergence re-
lated to the deflection wM and the clamping moment myyE employing w–
quadratic / ϕ–linear approaches and w–cubic / ϕ–quadratic approaches in
comparison. The clamping moment is evaluated separately concerning the three
elements meeting at the respective corner at positon E.
22.7 Displacement–Based Triangular Element 385
Remarks
Comparing the quadrilateral elements according to Section 22.4.2 to the trian-
gular element employing a quadratic / cubic approach it may be stated:
• Both types of elements converge against the reference solution.
• The triangular element discretizes the deflection employing a complete cu-
bic approach and the rotations employing a complete quadratic approach.
• The quadrilateral element discretizes the deflection as well as the rota-
tions employing bi–linear approaches whereby the deflection comprises
additional quadratic approaches being adapted to the rotations.
• At applying the same element mesh the triangular element exhibits round
about twice the number of degrees of freedom.
• Concerning the triangular element the displacement and the stress resul-
tant variables faster converge against the reference solution at taking into
account the same number of degrees of freedom.
• Concerning the triangular element the stress resultant variables converge
bottom–up against the reference solution. Concerning the quadrilateral
element the stress resultant variables converge top–down against the re-
ference solution.
22.8 Mixed Quadrilateral Element 387
Conditions of Equilibrium
The work equation applying the PvD describes the internal work of bending
and twisting moments as well as shear forces on the conjugated virtual strains
and curvatures. The external work takes into account the pressure p, and the
bending moments mex , mey at the boundary of the plate. In contrast to Kirch-
hoff ’s plate theory the twisting moment mexy can act on the conjugated virtual
rotations δϕx , δϕy at the boundary. Thus the Principle of virtual Displacements
includes the follwing terms
Z
δWd = {(δκxx mxx + δκyy myy + 2δκxy mxy ) − (δγxz qx + δγyz qy ) + δwp}dA
Z Z
+ {δϕx mex + δϕy mexy } dy + {δϕy mey + δϕx mexy } dx = 0 . (22.15)
The PvD may be extended by further external actions like concentrated loa-
ding P and line shaped loading qxe , qye if necessary. Furthermore, areal moments
mex , mey could be considered as well, but are neglected in the presentation for
simplicity. The PvD may be transformed by employing
Z Z
{δκxx mxx − δγxz qx }dA + δϕx mex dy
Z Z
= {δϕx ,x mxx − (δw,x − δϕx )(mxx ,x + mxy ,y )}dA + δϕx mex dy
Z Z
= {−δw,x (mxx ,x + mxy ,y ) + (δϕx mxx ),x + δϕx mxy ,y }dA + δϕx mex dy
Z Z
= {−δw,x (mxx ,x + mxy ,y ) + δϕx mxy ,y }dA + δϕx (mx + mex ) dy
and accordingly the work performed by the bending moments myy and by the
shear forces qy might be transformed. Thus only the moments [ mxx myy mxy ]
388 22 Plate Elements Including Shear Deformations
remain as primary variables of the actual stress state, whereas the virtual de-
flection δw and the virtual rotations δϕx , δϕy describe the virtual deformations:
Z
δWd = {− δw,x (mxx ,x + mxy ,y ) − δw,y (myy ,y + mxy ,x )
The work equation may be further simplified by integrating the work performed
on the rotations by the twisting moments
Z
{(δϕx mxy ,y + δϕy mxy ,x ) + (δϕx ,y + δϕy ,x ) mxy } dA
Z Z
= δϕx mxy dx + δϕy mxy dy .
This means, that all moments have to be treated as essential boundary conditi-
ons. It is remarkable, that the PvD only employs the moments [ mxx myy mxy ]
as primary variables and the virtual deflection to describe the weak formulation
of the equilibrium as presented by Kirchhoff ’s plate theory. This is valid, since
the shear deformations only modify the kinematic conditions and the material
equations.
22.8 Mixed Quadrilateral Element 389
Conditions of Deformation
The conditions of deformation related to the Reissner–Mindlin plate theory are
weakly formulated applying the Principle of virtual Forces. Incorporating the
bending flexibility FB and the shear flexibility FS and without the boundary
conditions concerning the deflection w, that are strongly fulfilled, it follows
Z
δWσ = + {δmxx (κxx + FBxxxx · mxx + FBxxyy · myy )} dA
Z
+ {δmyy (κyy + FByyxx · mxx + FByyyy · myy )} dA
Z
+ {2 δmxy (κxy + FBxyxy · mxy )} dA
Z Z
− {δqx (γxz − FSxx · qx )} dA − {δqy (γyz − FSyy · qy )} dA
Z
− {δmy (ϕy − ϕey ) + δmxy (ϕx − ϕex )} dx
Z
− {δmx (ϕx − ϕex ) + δmxy (ϕy − ϕey )} dy = 0 .
In contrast to Kirchhoff ’s plate theory, at the boundary the virtual twisting mo-
ments δmxy act on the conjugated rotations ϕx , ϕy , which can be constrained
by external rotations ϕex , ϕey .
Incorporating the governing equations concerning kinematics κxx = ϕx ,x and
κyy = ϕy ,y as well as κxy = (ϕx ,y + ϕy ,x )/2 and after integration by parts
the work performed on the rotations ϕx , ϕy , different contour integrals are
eliminated. It remains the work equation
Z
δWσ = − {δmxx ,x ϕx − δmxx (FBxxxx · mxx + FBxxyy · myy )} dA
Z
− {δmyy ,y ϕy − δmyy (FByyxx · mxx + FByyyy · myy )} dA
Z
− {(δmxy ,y ϕx + δmxy ,x ϕy ) − 2δmxy · FBxyxy · mxy } dA
Z Z
− {δqx (γxz − FSxx · qx )} dA − {δqy (γyz − FSyy · qy )} dA
Z Z
+ {δmy ϕey + δmxy ϕex } dx + {δmx ϕex + δmxy ϕey } dy = 0 .
Applying the equilibrium qx = mxx ,x + mxy ,y and qy = myy ,y + mxy ,x for the
virtual and the actual shear forces as well as kinematics γxz = w,x − ϕx and
390 22 Plate Elements Including Shear Deformations
γyz = w,y − ϕy , the primary variables [ w mxx myy mxy ] persist, which are
used to formulate the PvF in equivalence to the PvD
Z
δWσ = − {δmxx,x w,x − δmxx (FBxxxx · mxx + FBxxyy · myy )} dA
Z
− {δmyy ,y w,y − δmyy (FByyxx · mxx + FByyyy · myy )} dA
Z
− {δmxy ,x w,y + δmxy ,y w,x − 2δmxy · FBxyxy · mxy } dA
Z
+ {(δmxx ,x + δmxy ,y ) · FSxx · (mxx ,x + mxy ,y )
Boundary Conditions
The work equations (22.18) applying the PvD and (22.19) applying the PvF are
formulated employing the primary variables [ w mxx myy mxy ], whereby the
respective boundary conditions are to be fulfilled strongly by the approach. In
contrast to Kirchhoff ’s plate theory, this means, that the twisting moment along
the boundary is to be set to mxy = −mexy , whereas the boundary conditions
concerning the conjugated rotations ϕx , ϕy are fulfilled weakly, if the twisting
moment can freely evolve.
In the case, where the boundaries do not coincide with the x– and y–coordinates,
the moments have to be transformed to the n– and t–directions of the boun-
dary, if mn or mt are prescribed. This fact is a major drawback, since the
transformation has to be conducted after assembling all elements by means of
the system’s equations.
Benchmark
As an example to test the formulation the square plate is investigated again
according to Section 6.2.9. Utilizing the symmetry, the quarter plate is investi-
gated in turn for different FE–discretizations, cf. Figure 22-13.
x clamped
l x = 10.0 m
l y = 10.0 m
lx
By reason of the Reissner–Mindlin theory the twisting moments and the con-
jugated rotations related to the boundaries GE–E as well as GE–G are to be
provided case–dependent. Here the rotation in direction of the boundaries is
set to zero, so that the twisting moment takes finite values at the boundaries,
what is named a strong support, cf. Table 22.8. Therefore, the boundary and
accordingly the symmetry conditions are provided as follows:
Equilibrium
By analogy with the mixed formulation the work equation applying the PvD
at element level follows to
Z
δWd = + {(δκxx mxx + δκyy myy + 2δκxy mxy ) − (δγxz qx + δγyz qy ) + δwp} dA
E
Z Z
− {δϕx mxx + δϕy mxy } dy − {δϕy myy + δϕx mxy } dx
E E
Z Z Z Z
+ δwqx dy + δwqy dx − δw qx dy − δw qy dx
E E I I
Z
e e
+ {δϕx (mxxE + mxxI ) + δϕy (mxyE + mxyI )} dy
I
Z
+ {δϕy (myyE + meyyI ) + δϕx (mxyE + mexyI )} dx .
I
The index E indicates the element domain and the index I denotes the inter-
face. In comparison to the mixed formulation, here the cutted cleanly shear
forces and moments perform work on the displacements related to the element
boundary E and to the interface I. Furthermore the external moments me
perform virtual work on the interface I.
By analogy with Section 22.8 the domain as well as the contour integrals may
be transformed introducing
Z Z
{δκxxmxx − δγxz qx }dA − δϕx mxx dy
Z Z
= {−δw,x (mxx ,x + mxy ,y ) + δϕx mxy ,y }dA + δϕx (mxx − mxx ) dy ,
so that only the deflection and the moments [ w mxx myy mxy ] as well as the
virtual deflection δw and the virtual rotations δϕx and δϕy are variables at
394 22 Plate Elements Including Shear Deformations
the element domain. If, moreover, the virtual deflections δw are continuous
at the transition between element boundary and interface, the Principle of
virtual Displacements comprises the following work terms related to the element
domain E and to the interface I:
Z
δWd = {−δw,x (mxx ,x + mxy ,y ) − δw,y (myy ,y + mxy ,x ) + δwp}dA
E
Z
+ {δϕx (mxxE + mexxI ) + δϕy (mxyE + mexyI )} dy
I
Z
+ {δϕy (myyE + meyyI ) + δϕx (mxyE + mexyI )} dx . (22.20)
I
It should be noted that the integrals at the boundary, which deal with the
rotations δϕx , δϕy and the twisting moments mxy , can be transferred with
qx
ϕx = w,x − γxz = w,x − = w,x − ( mxx ,x + mxy ,y )
GAs
and analogously with ϕy . Hereby the variables ϕ perpendicular to the bounda-
ry can be avoided.
Conditions of Deformation
By analogy with the mixed formulation the conditions of deformation are for-
mulated applying the PvF. Here, it has to be taken notice of the fact, that
the transition conditions for the displacements between element boundary and
interface are formulated with the virtual stress resultants. By analogy with
Section 22.8 and without impressed kink angles, that are not possible here, the
following is valid
Z
δWσ = + {δmxx (κxx + FBxxxx · mxx + FBxxyy · myy )} dA
E
Z
+ {δmyy (κyy + FByyxx · mxx + FByyyy · myy )} dA
E
Z
+ {2 δmxy (κxy + FBxyxy · mxy )} dA
E
Z Z
− {δqx (γxz − FSxx · qx )} dA − {δqy (γyz − FSyy · qy )} dA
E E
Z Z
− {δmyy ϕy + δmxy ϕx − δqy w} dx − {δmxx ϕx + δmxy ϕy − δqx w} dy
E E
Z Z
+ {δmyy ϕy + δmxy ϕx − δqy w} dx + {δmxx ϕx + δmxy ϕy − δqx w} dy
I I
= 0.
22.9 Hybrid–Mixed Quadrilateral Element 395
Boundary Conditions
The work equations (22.20) applying the PvD and (22.21) applying the PvF
employ the variables [ w mxx myy mxy ]E as well as the rotations [ ϕx ϕy ]I .
After discretization has happened related to the element domain, the moments
are eliminated numerically, so that a formulation is derived employing the dis-
placement variables [ w ϕx ϕy ] of the respective element nodes. This means,
that the boundary conditions may be set correspondingly to a formulation em-
ploying only displacements as variables. Nevertheless, it has to be taken into
account, that, due to the Reissner–Mindlin theory, the rotations in direction
of the boundary are independent variables.
Test
As an example to test the formulation the square plate according to Section
6.2.9 and Section 22.8 is investigated again. Utilizing the symmetry the quarter
plate is investigated in turn for varying FE–discretizations.
x clamped
G M ν = 0.2
hinged
l x = 10.0 m
l y = 10.0 m
lx
By reason of the Reissner–Mindlin theory the twisting moments and the con-
jugated rotations related to the boundaries GE–E as well as GE–G are to be
provided case–dependent.
In the case of strong support the boundary conditions are given in Table 22.12,
where the rotations at the hinged boundary GE–G and the clamped boundary
GE–E are set to be zero, so that the twisting moments take finite values at
the hinged boundary. Therefore, the boundary and accordingly the symmetry
conditions are provided as follows:
Table 22.12 Boundary Conditions for strong support – Tables 22.14, 22.15, 22.16
boundary w ϕx ϕy comment
GE–E 0 0 0 clamped
GE–G 0 ∗ 0 hinged
G–M ∗ ∗ 0 symmetry
E–M ∗ 0 ∗ symmetry
In comparison to the strong support, cf. Table 22.12, weak boundary conditions
are investigated, too. Table 22.13 gives the boundary conditions for a weak
support at the hinged boundary.
Table 22.13 Boundary Conditions for weak support – Tables 22.17, 22.18, 22.19
boundary w ϕx ϕy comment
GE–E 0 0 0 clamped
GE–G 0 ∗ ∗ hinged
G–M ∗ ∗ 0 symmetry
E–M ∗ 0 ∗ symmetry
A much less expensive method is to estimate the error after a one–time analysis
and to report its distribution. In mathematics, many possibilities are known
to estimate or indicate discretization errors. Concerning the FE–analysis two
types may be distinguished, these are the a priori and the a posteriori error
estimations [2, 40, 8, 100]. The benefit of the a priori error estimation is more
theoretical, since it investigates the convergence of various norms with respect
to the element sizes and the shape functions, in order to evaluate how fast an
error will decrease. At applying the FEM the a posteriori error estimation is
more useful, since it gives an evaluation of a specific FE–solution, what could
be the basis for a mesh refinement. In constrast to the error estimation an error
indication does not state the size of the error, but only indicates where an error
occurs what is sufficient in many applications.
In general, global and local errors must be distinguished. Normally, global error
indicators are worthless, since locating the error is not possible thereby. Thus,
no information is provided concerning the area where the mesh needs to be
refined. As an example, this is the case if the error concerning the deformation
energy is determined using two different meshes. If the total potential must
fulfill a minimum condition, the strain energy converges towards an absolute
minimum when performing mesh refinement.
Local error indicators work at an element level, so that information about the
parts of the domain where mesh adaptation is required can be immediately
received. Nevertheless, no absolute statements can be predicated upon local
error indicators, as the existing local FE–solution usually also depends on the
solutions at other parts of the domain.
The quality check of the approximate solution performed by a local indicator
yields poor information, if, for example at re–entrant angles of membranes, va-
lues of the solution towards infinity are to be approximated. Here, in general,
a significant local error develops, even though the solution might be approxi-
mated very well in the overall structure.
The literature offers many approaches to develop and to apply an error esti-
mation, what can not be discussed in detail here [2, 3, 6, 12, 40, 8, 100]. Thus
in order to give an introduction of the problem the following options may be
discussed:
3. The least squares method applies the available FE–solution for example
to identify the error in equilibrium.
4. By applying the Principle of virtual Displacements the available FE–
solution can be taken into account to improve stresses at element level,
which can be compared to the stresses from applying the material equa-
tions.
The development of the error indicator can be discussed by means of the sim-
plifying structure of Section 1.3, where a bar is loaded by a distributed external
action p and has to fulfill the Neumann boundary condition with respect to H.
Choosing a formulation of the error geared to least squares, this yields,
Z
E=ℓ [N,x + p ]2 dx + [ N − H ]2boundary .
ℓ
This corresponds to the global error of the FE–solution, which, however, has low
significance, as shown above. Alternatively, the integral related to the domain
can be defined as the addition of the errors related to the element domains and
element intersections, respectively.
∆ℓ−ε
Z Z+ε
E = Σel ∆ℓ [N,x + p ] dx + Σintersect. ∆ℓ [N,x + p ]2 dx + [ N − H ]2boundary .
2
ε −ε
Regarding ε → 0 one can simplify the second integral, since the axial external
action is not defined at the jump discontinuity. Apart from that, the derivatives
of the jumps ∆N at the element intersections can be described mathematically
exactly employing the Dirac–Delta function δ(x), see Figure 23-1.
N
x
Δl Δl Δl Δl
N,x
Δ N i · δ (x) <0
This yields
∆ℓ−ε
Z Z+ε
E = Σel ∆ℓ [N,x + p ] dx + Σintersect. δ(x)[ ∆N ]2 dx + [ N − H ]2boundary .
2
ε −ε
406 23 Error Estimation
δ(x) Z+ε
δ(x) dx = 1
−ε
-ε x 0 +ε x
Applying the definition of the Dirac–Delta function δ(x), see Figure 23-2,
further gives
∆ℓ−ε
Z
E = Σel ∆ℓ [N,x + p ]2 dx + Σintersect. ∆N 2 + [ N − H ]2boundary .
ε
Regarding the single element domains, or intersections, one may obtain the
amount of the local error. Thus the local error can be directly recognised from
the stress jumps at both the respective element intersections as well as from the
integral of the error of the differential equation regarding the element domain.
Simplification gives
The procedure may be extended to two or three dimensions, therefore the local
error can also be indicated for membrane elements. Regarding the element mesh
of a membrane given in Figure 23-3, jumps occur wih respect to σyy and σyx
related to the T–T line and with respect to σxx and σxy related to the V–V
line, which must disappear in the case of the exact solution.
σxy
M
T T
σyy σyx
σxx
However one should take into account, that jumps of normal and tangenti-
al stresses can emerge at every element boundary, which must be considered
separately. If the least squares method is separately applied to the equilibri-
um in the x–and in the y–direction, the points may be recognised, where the
stress jumps indicate an error. Regarding the x–direction, without considering
a system boundary, this yields
∆A−ε
Z
Ex = Σel ∆ℓx ∆ℓy [σxx ,x + σyx ,y + px ]2 dxdy
ε
Z Z
2 2
+ Σintersection ∆ℓx ∆σxx dy + Σintersection ∆ℓy ∆σyx dx ,
y x
entire system
pX
1 2 3 4 5
x, u
NFEM
1 2 3 4 5
single element
pX
N A B N N N
A B A FEM
= B FEM
= NFEM
i i+1
δuA 1
1 δ uB
By applying the PvD, the internal force variables NA and NB may be directly
computed at the element intersections. In accordance with Equations (2.11)
and (2.12), it holds that
Z
{N δu,x − p δu} dx + NA δuA − NB δuB = 0 .
stress analysis. Thus the internal force variables can be computed directly to
NA = NF EM + pℓ/2 ,
NB = NF EM − pℓ/2 .
Hence, the local error is given as the difference between the internal force
variables NA and NB and the longitudinal forces NF EM from stress analysis,
FA = NA − NF EM = + pℓ/2 ,
FB = NB − NF EM = − pℓ/2 .
It should be noted that the error evaluated by this procedure is not an absolute
error but a relative one, as the FE–approximation applied at the element–level
may be totally inaccurate.
Transfering Equation (23.1) to the membrane elements, the nodal forces with
respect to the x–and y–directions are directly obtained, and can be compared
to the stresses from applying the material equations. Analogously to the PvD,
it makes sense to integrate the stresses at the element intersections, taken from
stress analysis, to get the nodal forces.
Comparing the nodal forces from the stress analysis and from applying the
PvD, the following should be taken care of. If the equilibrium is separately
fulfilled for every node, see Figure 23-3, the error obtained from the stresses
σxx and σyy as well as σxy is more or less the same but with opposite signs, so
they cancel out each other with regard to the entire system. Therefore it seems
to be reasonable to divide the nodal forces into parts related to σxx , σyy and
σxy , and then to separately introduce the respective parts into the comparison.
This procedure may lead to an even better estimation of the local error.
410 23 Error Estimation
23.3 Mesh–Adaptation
Applying the error estimation to a FE-solution offers the opportunity to impro-
ve the solution by means of an adaptation of the FE-mesh. In general, a new
discretization can be performed introducing locally distinctive and solution–
matching mesh refinement respectively. This may considerably reduce the com-
putational cost. Modern computer–program systems allow for the indication of
errors and the making of suggestions for a re–discretization related to the error
distribution inside the structure. This procedure, named as mesh–adaptation,
may be performed by applying different strategies, changing the element sizes
by means of the h–adaptation or raising the polynomial order of the approach
by means of the p–adaptation in specified parts of the mesh without breaching
the conformity criterion with respect to the shape functions.
Regarding different cases, the strong solutions of the governing equations exhi-
bit singularities and high gradients with respect to the distribution of stresses,
which effect large stress discontinuities between neighboring elements and, as
a result, local error concentrations.
Singularities are, for example, jumps in the stress distribution, which may be
easily described by applying the PvD, if they occur at the element intersecti-
ons. Singularities also occur as non–real points of infinity inside the distribu-
tion of stresses, for example at nooks or due to concentrated external actions
regarding membranes, plates, and shell structures. In such cases, a proper des-
cription of the stress distribution around the points of infinity should be taken
care of. However, one should not try to directly approximate points of infinity,
as this neither makes sense regarding physics, nor is required when following
St.Venant’s principle.
Nonetheless, since singularities with high stress gradients are part of the model-
ling of the structural behavior, a proper finite element approach in the vicinity
of the singularities is necessary in order to minimize the discretization error. In
general this is possible with a refinement of the finite element mesh or with an
extension of the order of the shape functions. Thus it depends on the elements,
which are applied to solve the problem.
When applying the h–method the adaptation of the mesh deals with the ele-
ment size as explained above. Hereby, the mesh is refined by means of very
small elements at the regions where high stress gradients are observed. Gene-
rally these regions need a high number of elements to aproximate the solution
sufficiently, since shape functions of low order are employed.
Both methods to adapt the mesh require special modifications of the elements
at the intersections. At employing Lagrange Polynomials of different order,
the additional DOF of the higher order approach must be attached to the
lower order approach at the respective intersections, as they are no longer
independent DOF. Thus additional nodes are required regarding the element
employing the lower order approach. They are denoted as hanging nodes, see
Figure 23-6 in the case of rectangular elements. Hanging nodes are also needed
in the case of mesh refinement, concerning elements with the same polynomial
order but with discontinuing mesh lines, see Figure 23-6. Here the displacements
related to the additional nodes of the smaller element must be attached to the
displacements related to the respective edge of the larger element.
( K − λi · I ) φi = 0
5 6 7 8
0.001.l y
lx lx
Table 24.1 represents the eigenvalues of the stiffness matrix of the square ele-
ment in Figure 24-2. Some of the eigenvalues occur as pairs, since the related
deformation modes activate same stiffnesses with respect to the x– and the
y–coordinate. Regarding the elements P-DQ-4, P-DQ-4-SRI, P-HMQC-4-5
dealing with bi–linear shape functions the eigenvalues 6, 7, 8 describe the lon-
gitudinal and the shear stiffnesses. The eigenvalues 4, 5 are related to the
trapezoidal deformation modes, whereat the element P-DQ-4 is stiffer due to
the unphysical shear stiffnesses as described in Section 5.2.
24.1 The Eigenvalue Analysis 415
Due to the comparison of the eigenvalues, it has to be noted, that the bi–
linear deformation fields of the elements P-DQ-4, P-DQ-4-SRI, P-HMQC-4-5
are not represented by single modes of the elements P-DQ-9, P-DQ-9-SRI,
P-HMQ-8-13, but can be represented by a combination of modes. Thus the
corresponding eigenvalues do not coincide in all cases. For example the eigen-
value λ8 = 1. 0 of the bi–linear elements is correlated to the shear deformation
φ8 of Figure 24-1. The correspondig eigenvalue of the element P-HMQ-8-13
dealing with quadratic shape functions is λ8 = 0. 47948. This eigenvalue is cor-
related to φ8 superposed by a quadratic deformation field, which reduces the
416 24 Quality of Elements
shear–constraints at the corners of the deformation field. On the other side the
eigenvalue λ14 = 1. 8538 is correlated to φ8 as well, but is superposed by ano-
ther quadratic deformation field, which increases the constraints at the corners.
Table 24.2 represents the eigenvalues of the rectangular element in Figure 24-2,
whereby the lengths of the edges have a relation of ℓx /ℓy = 1000. Such elements
are applied, if beams are investigated without taking into account the beam
theory, or in the three dimensional case, if thin plates have to be investigated
as a three dimensional continuum.
In this case the eigenvalues λi = 0. 001 and λi = 1000. 0 of the bi–linear appro-
ximation describe the stiffnesses in the directions of the x– and y–coordinates,
which are correlated to ℓy /ℓx and ℓx /ℓy respectively. The eigenvalue λi = 500. 0
describes the shear stiffness, and λi = 333. 3 as well as λi = 0. 00033 characte-
rize the trapezoidal modes in both directions.
24.2 The Locking Phenomena 417
γ = w,x − ϕ
418 24 Quality of Elements
stiffnesses vary inside the elements and thus strongly influence their properties
with respect to the deformation modes. This may be discussed by means of
the eigenvalues of the case ℓx /ℓy = 1 given in Table 24.3 and those of the case
ℓx /ℓy = 1000 given in Table 24.4.
e e e = lx / 4
l x = 1.0 m, l y = 1.0 m, E = 1.0 N/m 2 , ν = 0.0
ly e e
0.001.l y
e lx e e lx e
Table 24.3 indicates, that the trapezoidal shape only effects a moderate varia-
tion of the eigenvalues in comparison to a square. However, the deviations from
the accurate values may be caused by additional stiffnesses, which are reason
of undesired stresses with respect to the deformation modes.
But, if the element is very thin in one direction, the eigenvalues change much
more and indicate the thickness locking for the elements P-DQ-4, P-DQ-4-SRI
with respect to the eigenvalues 4 and 5, for the element P-DQ-9 with respect
to the eigenvalues 8, 9 ,11, and for the element P-DQ-9-SRI with respect to
the eigenvalues 6, 7 and 8, cf. Table 24.4. Thus these elements are not able to
accurately describe the deformation and may lock the deformation modes, if
they are effected by external loads.
The only elements, which give reasonable results, are the P-HMQC-4-5 element
and the P-HMQ-8-13 element, whereat the linear approach is hardly able to
describe bending.
Table 24.4 Eigenvalues of membrane elements ℓx /ℓy = 1000 and ℓx1 /ℓx2 = 3
that in the case of Reissner–Mindlin plates the conditions γxz = ϕx − w,x and
γyz = ϕy − w,y have to be satisfied without numerical oscillations. Instead of a
selectively reduced integration this demand can be achieved employing shape
functions of different order for the description of the state variables w, ϕx , ϕy .
This idea is applied with a similar intention at Section 22.7 and 22.9, but
without a rigorous mathematical development as presented in [24].
As an example the shear strain γxz may be described in a quadrilateral element
by means of bi–linear shape functions for w(x, y) and ϕx (x, y) with
γxz (x, y) = w(x, y),x − ϕx (x, y)
X X
= φi (x, y),x · wi − φi (x, y) · ϕx i .
Thus the shear strain comprises a constant part and a linear part in x–direction,
what effects the shear locking phenomenon. Assuming a constant shear strain
with respect to the x–coordinate, it follows
X
γxz (y) = φ̄i (y) · γxz i ,
whereat φ̄i (y) are scaled to the nodal strains γxz i . Figure 24-4 illustrates the
course of the shear strain γxz (y), that is constant in x–direction but linear
in y–direction, and normalized with respect to the shear strains γ(xz)E and
γ(xz)G . The nodes E, G are defined as those nodes, where the shear strains
related to the bi–linear shape functions of the primary variables would become
the best value, which usually are the Gaussian supporting points. In the case
of bi–linear shape functions these are the midpoints of the related edges. Thus
the shear strains γxz i can be described with the nodal degrees of freedom of
the primary variables w, ϕx , ϕy of the corresponding element edge.
x γxzE
y z A A
E E
D B D B
G
γxzG G
C C
Fig. 24-4 Course φ̄E (y), φ̄G (y) of the assumed shear strain γxz
are able to eliminate the defiencies of the original approach and lead to similar
results as the selectively reduced intergration, but can be applied to other
locking phenomena, too.
γ = w,x − ϕ .
424 24 Quality of Elements
Employing linear shape functions φi for w and ϕ results into the problems
discussed in Section 21, due to the different orders of w,x and ϕ. Thus seve-
ral approaches may be applied to this problem, e. g. the selectively reduced
integration technique, the ANS–method, or the EAS–method.
The DSG–method deals with the integration of w(x),x concerning kinematics
Z
w(x) = (γ + ϕ) dx = wγ + wϕ .
Thereby wγ describes the difference between the total deflection w(x) and the
deflection wϕ due to bending and thus it is named shear strain gap.
In order to compute the element stiffness matrix an assumption has to be made
with respect to the course of the shear strains within the element. However,
the accumulated shear strain gap between the element nodes A and B can be
computed to
Z B
B
wγ A = (w,x − ϕ) dx .
A
Since the integral only depends on the values at the boundaries, employing
linear shape functions it follows
wB − wA
wγ B
A = { − (ϕB − ϕA )} · ℓ ,
ℓ
whereat both parts of wγ B
A have got the same polynomial order. Thus the course
of the shear strain can be described by means of a constant shape function
φγ = 1/ℓ with
γ = φγ · wγ B
A
7 8 x/l x y/l y
1 0.000 0.000
➃
2 1.000 0.000
5 6
3 0.166 0.166
➁ ➄ 4 0.750 0.250
➂ 5 0.333 0.666
4 6 0.666 0.666
3
7 0.000 1.000
1 ➀ 2
8 1.000 1.000
x
strain in y shear in x
node u[m] v[m] u[m] v[m]
1 0. 0 0. 0 0. 0 0. 0
2 — 0. 0 — 0. 0
7 — 1. 0 1. 0 0. 0
8 — 1. 0 1. 0 0. 0
426 24 Quality of Elements
strain in y shear in x
ele. node σxx σyy σxy σxx σyy σxy
1 1 0. 0089 0. 3373 0. 0749 −0. 0110 −0. 0112 0. 1234
2 0. 0083 0. 3347 −0. 0445 −0. 0980 −0. 0065 0. 1244
4 0. 0175 0. 2959 −0. 0376 0. 0455 0. 0076 0. 1651
3 0. 0240 0. 3963 0. 0636 0. 0608 0. 0054 0. 0706
2 1 −0. 0113 0. 3430 −0. 0968 0. 1629 0. 0415 0. 1367
3 −0. 0186 0. 5013 −0. 0913 0. 2126 0. 0509 0. 0993
5 0. 0117 0. 2621 0. 0327 −0. 0561 −0. 0147 0. 1657
7 0. 0168 0. 3500 0. 0366 −0. 0948 −0. 0229 0. 1367
3 3 0. 0204 0. 3471 0. 0112 0. 0475 0. 0097 0. 1227
4 0. 0234 0. 3392 −0. 0163 0. 0425 0. 0076 0. 1480
6 0. 0270 0. 3497 −0. 0065 −0. 0574 −0. 0175 0. 1329
5 0. 0224 0. 3312 0. 0194 −0. 0571 −0. 0164 0. 1488
4 5 0. 0259 0. 3472 −0. 0484 −0. 0378 −0. 0021 0. 1024
6 0. 0204 0. 3252 0. 0490 −0. 0379 −0. 0026 0. 1620
8 0. 0150 0. 3238 0. 0531 0. 0018 0. 0074 0. 1322
7 0. 0205 0. 3459 −0. 0525 0. 0020 0. 0078 0. 1322
5 2 0. 0126 0. 3375 −0. 0752 0. 1987 0. 0488 0. 1351
8 0. 0025 0. 3350 0. 0546 −0. 1569 −0. 0401 0. 1357
6 0. 0090 0. 3899 0. 0498 −0. 0977 −0. 0242 0. 0913
4 0. 0048 0. 2532 −0. 0811 0. 1995 0. 0500 0. 1357
Table 24.7 Results of the P-HMQC-4-5 element due to Pian - Sumihara [81]
strain in y shear in x
ele. node σxx σyy σxy σxx σyy σxy
all all 0. 0 0. 3333 0. 0 0. 0 0. 0 0. 1333
References
[1] ADINI, A., CLOUGH, R. W.: Analysis of PLate Bending by the Finite Ele-
ment Method. National Science Foundation, Grant G7337 (1960)
[2] AINSWORTH, M., ODEN, J.T.: A posterior error estimation in finite element
analysis. New York, John Wiley & Sons (2000)
[3] AINSWORTH, M., ODEN, J.T.: A unified approach to a posteriori error esti-
mation using element residual methods. Numerische Mathematik 65, pp 23–
50(1993)
[5] ARGYRIS, J. H., FRIED, I. and SCHARPF, D.: The TUBA family of plate
elements for the matrix displacement method. Aeronautical Journal 72 (1968)
[6] BABUŜKA, I.: Error bounds for finite element method. Numerische Mathe-
matik 16, Springer Verlag (1971)
[7] BABUŜKA, I.: The Finite Element Method with Lagrangian Multipliers. Nu-
merische Mathematik 20, Springer Verlag (1973)
[8] BABUŜKA, I., STROUBOULIS, T.: The finite element method and its relia-
bility. Oxford, Oxford Science Publications (2001)
[9] BABUŜKA, I., SURI, M.: Locking effects in the finite element approximation
of elasticity problems. Numerische Mathematik 62, pp. 439–463 (1992)
[10] BABUŜKA, I.; SZABÓ, B. A.: On the Rates of Convergence of the Finite
Element Method. International Journal of Numerical Methods in Engneering
18, pp. 323-341 (1982)
[11] BABUŜKA, I.: The p- and hp-Versions of the Finite Element Method: The
State of the Art. in: Finite Elements: Theory and Applications, edtrs. D. L.
Dwoyer, M. Y. Hussaini and R. G. Voigt. Proceedings of the ICASE Finite
Element Theory and Application Workshop 1986, Springer-Verlag, New York
(1988)
[12] BATHE, K. J.: Finite Elemente Methoden. Springer Verlag, Berlin (2002)
[13] BATHE, K. J., DVORKIN, E. N.: A Four-Node Plate Bending Element Based
on Mindlin–Reissner Theory and a Mixed Interpolation. International Journal
for Numerical Methods in Engineering 21, pp. 367-383 (1985)
© The Editor(s) (if applicable) and The Author(s), under exclusive license to
Springer Fachmedien Wiesbaden GmbH, part of Springer Nature 2024
D. Dinkler und U. Kowalsky, Introduction to Finite Element Methods,
https://doi.org/10.1007/978-3-658-42742-9
428 References
[16] BELL, K.: A refined triangular plate bending finite element. International
Journal for Numerical Methods in Engineering 1 (1969)
[17] BELYTSCHKO, T., TSAY, C.S., LIU, W.K. : A Stabilization Matrix for the
Bilinear Mindlin Plate Element. Computer Methods in Applied Mechanics and
Engineering, 29, pp. 313-327 (1981)
[19] BLETZINGER, K. U., BISCHOFF, M., RAMM, E.: A unified approach for
shear–locking–free triangular and rectangular shell finite elements. Computers
& Structures 75, pp. 321-334 (2000)
[20] BOGNER, F. K., FOX, R. L. and SCHMIT, L. A.: The generation of interele-
ment, compatibie stiffness and mass matrices by use of interpolation formulae.
Proc. 1st Conference on Matrix Methods in Structural Mechanics. Wright Pat-
terson Air Force Base, Ohio, AFFDL–TR–66–80, (1965)
[21] BOSSHARD, W.: Ein neues, vollverträgliches endliches Element für Platten-
biegung. International Association for Bridge and Structural Engineering, Bul-
letin 28, 27-40 (1968)
[23] BUCALEM, M. L., BATHE, K. J.: Higher-order MITC general shell elements.
International Journal for Numerical Methods in Engineering 36, pp. 3729-3754
(1993)
[24] BUCALEM, M. L., BATHE, K. J.: Finite Element Analysis of Shell Structures.
Archive of Computational Methods in Engineering 4 pp. 3–61 (1997)
[25] BUTLIN, G. and FORD, R.: A compatible triangular plate bending finite ele-
ment. University of Leicester, Engineering Department, Report 68–15 (1968)
References 429
[26] CLOUGH, R. W.: The finite element method in plane stress analysis. In:
Proceedings, 2nd Conference on Electronic Computation, A.S.C.E. Structural
Division. Pittsburgh, Pennsylvania (1960)
[27] CLOUGH, R. W., TOCHER, J. L.: Finite element etiffness matrices for ana-
lysis of plate bending. Proc. 1st Conference on Matrix Methods in Structural
Mechanics, Wright Patterson Air Force Base, Ohio, AFFD-TR-66-80, pp. 515–
545 (1965)
[29] COURANT, R.: Variational Methods for the Solution of Problems of Equi-
librium and Vibrations. Bulletin of the American Mathematical Society 49
(1943)
[30] COWPER, G. R., KOSKO, E., LINDBERG, G. M. and Olson, M. D.: Sta-
tic and Dynamic Applications of a High–Precision Triangular Plate Bending
Element. AIAA Journal 7, 1957–1965 (1969)
[32] CUTHILL, E., McKEE, J.: Reducing the bandwidth of sparse symmetric ma-
trices. ACM P-69, Proceedings of ACM National Conference, New York, pp.
157172 (1969)
[33] CUNDALL, P. A.: A computer model for simulating progressive large scale
movements in blocky rock systems. In: Proceedings of the Symposium of the
International Society of Rock Mechanics. Vol 1, II–8, France (1971)
[37] GALLAGHER, R. H.: Finite Element Analysis. Springer Verlag, Berlin (1976)
[39] GEORGE, P.: The advancing–front mesh generation method revisited. Inter-
national Journal for Numerical Methods in Engineering 58, pp. 1061–1089
(1994)
[41] GREENE, B. E., JONES, R. E., McLAY, R. W., STROME, D. R.: Generalized
variational principles in the finite-element method. AIAA Journal 6, pp. 1254–
1260 (1968)
[42] HARVEY, J. W., KELSEY, S.: Triangular Plate Bending Elements with En-
forced Compatibility. AIAA Journal 9, 1023–1026 (1971)
[43] HELLINGER, E.: Die allgemeinen Ansätze der Mechanik der Kontinua. In:
Die Encyklopädie der mathematischen Wissenschaften, Vol. IV/2, ed. F. Klein
und C. H. Müller, B. G. Teubner Verlag, Leipzig (1913)
[44] HU, H.C. Some Variational Principles in Elasticity and Plasticity. Acta Phy-
sica Sinica 10, pp. 259-290 (1954)
[46] HUGHES, T. J. R., COHEN, M., HAROUN, M.: Reduced and selective inte-
gration techniques in the Finite Element Analysis of plates. Nuclear Enginee-
ring and Design 46, 203–222 (1978)
[47] HUGHES, T. J. R.,: The Finite Element Method: Linear Static and Dynamic
Finite Element Analysis. Prentice Hall (1987), Dover (2000, ... 2012)
[48] HUGHES, T. J. R., TEZDUYAR, T. E.: Finite elements based upon mindlin
plate theory with particular reference to four-node bilinear isoparametric ele-
ment. In: T. J. R. HUGHES, D. GARTLING and R. L. SPILKER Edtrs: New
concepts in finite element analysis 44, The American Society of Mechanical
Engineers, pp. 81-106 (1981)
[53] IRONS, B. M., RAZZAQUE, A.: Experience with the patch test for con-
vergence of finite elements. In: The Mathematical Foundations of the Finite
Element Method with Applications to Partial Differential Equations, ed. A.
K. Aziz, Academic Press, New York, pp. 557–587 (1972)
[54] KAVANAGH, K. T., KEY, S. W.: A note on selective and reduced integration
techniques in the Finite Element Method. International Journal of Numerical
Methods in Engineering 4, pp. 148 (1972)
[55] KIRCHHOFF, G.: Über das Gleichgewicht und die Bewegung einer elastischen
Scheibe. Journal fuür die reine und angewandte Mathematik 4, de Gruyter,
Berlin (1850)
[58] KRÖPLIN, B. H., DINKLER, D.: A creep type strategy used for tracing the
load path in elastoplastic post buckling analysis. Computer Methods in Ap-
plied Mechanics and Engineering 32 (1982)
[60] KRÖPLIN, B. H., DINKLER D.: Dynamic versus static buckling analysis
of thin walled shell structures. Finite Element Methods for Plate and Shell
Structures - Vol 2: Formulation and Algorithm, eds. T.J.R. Hughes and E.
Hinton, Pineridge Press (1986)
[63] LO, S. H.: A New Mesh Generation Scheme for Arbitrary Planar Domains
International Journal of Numerical Methods in Engineering 21, pp. 1403–1426
(1985)
[64] LO, S. H.: Finite Element Mesh Generation. CRC Press, Taylor & francis
Group (2015)
[65] LÖHNER, R., PARIKH, P.: Three-dimensional grid generation by the advan-
cing front method. International Journal of Numerical Methods in Engineering
26, pp. 11351149 (1988)
[66] LÖHNER, R.: A 2nd Generation Parallel Advancing Front Grid Generator. in:
Proceedings of the International Meshing Roundtable Conference, Computer
Science (2013)
[67] McLAY, R. W.: A special variational principle for the finite element method.
AIAA Journal 8, pp. 533–534 (1969)
[69] MACNEAL, R.: A simple quadrilateral shell element. Computers & Structu-
res, pp. 175-183 (1978)
[71] MELOSH, R. J.: A stiffness matrix for the analysis in plate bending. Journal
of the Aerospace Sciences 28, pp. 34–42 (1961)
[74] NOOR, A. K., PETERS, J.: Nonlinear analysis via global-local mixed finite
element approach. International Journal for Numerical Methods in Enginee-
ring 15, pp. 1363-1380 (1980)
References 433
[75] NOOR, A. K., PETERS, J.: Mixed models and reduced/selective integrati-
on displacement models for nonlinear analysis of curved beams. International
Journal for Numerical Methods in Engineering 17, pp. 615-631 (1981)
[76] ONATE, E.: Structural Analysis with the Finite Element Method. Linear Sta-
tics: Volume 1: Basis and Solids. Lecture Notes on Numerical Methods in
Engineering and Sciences, Springer Verlag (2009)
[77] ONATE, E.: Structural Analysis with the Finite Element Method. Linear Sta-
tics: Volume 2: Beams, Plates and Shells, Lecture Notes on Numerical Methods
in Engineering and Sciences, Springer Verlag (2013)
[78] PARK, K. C., STANLEY, G. M.: A curved C0 shell element based on assumed
natural–coordinate strains. Journal of Applied Mechanics 53, Transactions of
ASME, pp. 278-290 (1986)
[79] PARK, K. C., PRAMONO, E., STANLEY, G. M., CABINESS, H. A.: The
ANS shell elements: earlier developments and recent improvements ((assu-
med natural–coordinate strain)). In: Analytical and Computational Models of
Shells, A. K. NOOR (Edtr.) ASME Special Publication CED–3, pp. 217-239
(1989)
[80] PIAN,T. H. H., TONG, P.: Basis of finite element methods for solid continua.
International Journal for Numerical Methods in Engineering 1, pp. 3–28 (1969)
[81] PIAN,T. H. H., SUMIHARA, K.: Rational approach for assumed stress finite
elements. International Journal for Numerical Methods in Engineering 20, pp.
1685–1695 (1984)
[84] REDDY, J. N.: On Complementary Variational Principles for the Linear Theo-
ry of Plates. Journal of Structural Mechanics 4, pp. 417-436 (1976)
[85] REISSNER, E.: A new Derivation of the Equations for the Deformation of
Elastic Shells. American Journal of Mathematics 63 (1941)
[86] REISSNER, E.: On the Theory of Bending of Elastic Plate. Journal of Ma-
thematics and Physics 23, pp. 184–191 (1944)
[87] REISSNER, E.: On a variational theorem for finite elastic deformations. Jour-
nal of Mathematical Physics 32, pp. 129–135 (1953)
434 References
[88] RITZ, W.: Über eine neue Methode zur Lösung gewisser Variationsprobleme
der mathematischen Physik. Journal für die reine und angewandte Mathematik
35 (1908)
[92] SIMO, J. C., RIFAI, S.: A Class of Mixed Assumed Strain Methods and the
Method of Incompatible Modes. International Journal for Numerical Methods
in Engineering 29, pp. 1595-1638 (1990)
[93] STRICKLIN, J. A., HAISLER, W., TISDALE, P., GUNDERSON, R.: A ra-
pidly converging triangular plate element. AIAA Journal 7, pp. 180–181 (1969)
[94] SUKUMAR, N., TABARRAEI, A.: Conforming polygonal finite elements. In-
ternational Journal for Numerical Methods in Engineering 61, pp. 2045–2066
(2004)
[98] TREFFTZ, E.: Ein Gegenst ück zum Ritzschen Verfahren. Proceedings of the
2nd International Congress for Applied Mechanics, Zrich, pp. 131-137 (1926)
[99] TURNER, M. J., CLOUGH, R. W., MARTIN, H. C., TOPP, L. J.: Stiffness
and Deflection Analysis of Complex Structures. Journal of the Aeronautical
Sciences 23 (1956)
[100] VERFÜHRT, R.: A review of a posterior error estimation and adaptive mesh
refinement techniques. Chichester, John Wiley & Sons (1996)
[101] WACHSPRESS, E. L.: A Rational Finite Element Basis. Academic Press, New
York (1975)
References 435
[102] WASHIZU, K.: On the variational principles of elasticity and plasticity. Aero-
elastic and Structures Research Laboratory, Massachuetts Institute of Tech-
nology, Technical Report 25-18 (1955)
[103] WILLIAMS, M. L.: Theoretical and experimental effect of sweep upon stress
and deflection in aircraft wings of high solidity – Part 5. California Institute
of Technology (1950)
[106] ZIENKIEWICZ, O. C., TAYLOR, R. L.: The Finite Element Method. Vol-
ume I: Basis and Fundamentals. 6th Edition, Butterworth–Heinemann, Oxford
(2005)
© The Editor(s) (if applicable) and The Author(s), under exclusive license to
Springer Fachmedien Wiesbaden GmbH, part of Springer Nature 2024
D. Dinkler und U. Kowalsky, Introduction to Finite Element Methods,
https://doi.org/10.1007/978-3-658-42742-9
438 Index
matrix notation, 47
stiffness matrix, 50
shape function, 8, 38
bar, 29
complete approach, 46
convergence, 42
convergence criteria, 38
general polynomials, 39
Hermite Polynomials, 41
Lagrange Polynomials, 40
Legendre Polynomials, 39
matrix notation, 35
Pascal’s triangle, 46
scaling matrix, 47
symmetric approach, 46
singularities, 410
solution
approximate solution, 7, 12
bandwidth, 58
exact solution, 12
storage of equations, 58
strong solution, 7
weak solution, 7
systems’ equations, 55, 58
virtual work, 28
matrix notation, 34