Sunteți pe pagina 1din 121

FINITE ELEMENT ANALYSIS

AS COMPUTATION
FINITE ELEMENT ANALYSIS AS COMPUTATION

What the textbooks don't teach you about finite element


analysis

Gangan Prathap
Preface
These course notes are specially prepared for instruction into the principles
underlying good finite element analysis practice. You might ask: Why another
book on FEA (Finite Element Analysis) and the FEM (Finite Element Method) when
there are hundreds of good textbooks in the market and perhaps thousands of
manuals, handbooks, course notes, documentation manuals and primers available on
the subject.

Some time earlier, I had written a very advanced monograph on the analytical and
philosophical bases for the design and development of the finite elements which
go into structural analysis software packages. Perhaps a hundred persons work on
element development in the world and out of these, maybe a dozen would probe
deep enough to want to explore the foundational bases for the method. This was
therefore addressed at a very small audience. In the world at large are several
hundred thousand engineers, technicians, teachers and students who routinely use
these packages in design, analysis, teaching or study environments. Their needs
are well served by the basic textbooks and the documentation manuals that
accompany installations of FEM software.

However, my experience with students and teachers and technicians and engineers
over two decades of interaction is that many if not most are oblivious to the
basic principles that drive the method. All of them can understand the
superficial basis of the FEM - what goes into the packages; what comes out; how
to interpret results and so on. But few could put a finger on why the method
does what it does; this deeper examination of the basis of the method is
available to a very few.

The FEM is now formally over thirty-five years old (the terminology "finite
element" having been coined in 1960). It is an approximate method of solving
problems that arise in science and engineering. It in fact originated and grew
as such, by intuition and inspired guess, by hard work and trial and error. Its
origins can be traced to aeronautical and civil engineering practice, mainly
from the point of view of structural engineering. Today, it can be used, with
clever variations, to solve a wide variety of problems in science and
engineering. There are billions of dollars worth of installed software and
hardware dedicated to finite element analysis all over the world and perhaps
billions of dollars are spent on analysis costs using this software every year.
It is therefore a remarkable commercial success of human imagination, skill and
craft.

The science of structural mechanics is well known. However, FEA, being an


approximate method has a uniqueness of its own and these curious features are
taken for closer examination in this course. The initial chapters are short ones
that examine how FEM is the art of numerical and computational analysis applied
to problems arising from structural engineering science. The remaining chapters
elucidate the first principles, which govern the discretisation process by which
"exact” theory becomes numerical approximations.

We shall close these notes with a telescopic look into the present scenario and
the shape of things to come in the future.

iii
Acknowledgements: I am extremely grateful to Dr T S Prahlad, Director, N.A.L.
for his encouragement and support. Many colleagues were associated at various
stages of writing this book. They include Dr B P Naganarayana, Dr S Rajendran
and Mr Vinayak Ugru, Dr Sudhakar Marur and Dr Somenath Mukherjee. I owe a great
deal to them. Mr Arjun provided secretarial help and also prepared the line
drawings and I am thankful to him. I am also thankful to Latha and Rahul for
their patience and understanding during the preparation of this manuscript.

Gangan Prathap
Bangalore
2001

iv
Contents

Preface iii

1. Introduction: From Science to Computation 1

2. Paradigms and some approximate solutions 11

3. Completeness and continuity: How to choose shape


functions? 28

4. Convergence and Errors 33

5. The locking phenomena 45

6. Shear locking 57

7. Membrane locking, parasitic shear and incompressible


locking 79

8. Stress consistency 97

9. Conclusion 111

v
Chapter 1

Introduction: From Science to Computation


1.1 From Art through Science to Computation

It will do very nicely to start with a description of what an engineer does,


before we proceed to define briefly how and where structural mechanics plays a
crucial role in the life of the engineer. The structural engineer's appointed
office in society is to enclose space for activity and living, and sometimes
does so giving time as well - the ship builder, the railway engineer and the
aerospace engineer enable travel in enclosed spaces that provide safety with
speed in travel. He did this first, by imitating the structural forms already
present in Nature.

From Art imitating Life, one is led to codification of the accumulated


wisdom as science - the laws of mechanics, elasticity, theories of the various
structural elements like beams, plates and shells etc. From Archimedes' use of
the Principle of Virtual Work to derive the law of the lever, through Galileo
and Hooke to Euler, Lagrange, Love, Kirchhoff, Rayleigh, etc. we see the
theoretical and mathematical foundations being laid. These where then copiously
used by engineers to fabricate structural forms for civil and military
functions. Solid and Structural Mechanics is therefore the scientific basis for
the design, testing, evaluation and certification of structural forms made from
material bodies to ensure proper function, safety, reliability and efficiency.

Today, analytical methods of solution, which are too restricted in


application, have been replaced by computational schemes ideal for
implementation on the digital computer. By far the most popular method in
Computational Structural Mechanics is that called the Finite Element Method.
People who use computational devices (hardware and software) with hardly any
reference to experiment or theory now perform design and Analysis. From
advertising claims made by major fem and cad software vendors (5000 sites or
installations of one package, 32,000 seats of another industry standard vendor,
etc.); it is possible to estimate the number of fem computationalists or
analysts as lying between a hundred thousand and two hundred thousand. It is to
them that this book is addressed.

1.2 Structural Mechanics

A structure is "any assemblage of materials which is intended to sustain loads".


Every plant and animal, every artifact made by man or beast has to sustain
greater or less forces without breaking. The examination of how the laws of
nature operate in structural form, the theory of structural form, is what we
shall call the field of structural or solid mechanics. Thus, the body of
knowledge related to construction of structural form is collected, collated,
refined, tested and verified to emerge as a science. We can therefore think of
solid and structural mechanics as the scientific basis for the design, testing,
evaluation and certification of structural forms made from material bodies to
ensure proper function, safety, reliability and efficiency.

To understand the success of structural mechanics, one must understand its


place in the larger canvas of classical mechanics and mathematical analysis,

1
especially the triumph of the infinitesimal calculus of continuum behavior in
physics. The development and application of complex mathematical tools led to
the growth of the branch of mathematical physics. This therefore encouraged the
study of the properties of elastic materials and of elasticity - the description
of deformation and internal forces in an elastic body under the action of
external forces using the same mathematical equipment that was been used in
other classical branches of physics. Thus, from great mathematicians such as
Cauchy, Navier, Poisson, Lagrange, Euler, Sophie Germain, came the formulation
of basic governing differential equations which originated from the application
of the infinitesimal calculus to the behavior of structural bodies. The study is
thus complete only if the solutions to such equations could be found. For nearly
a century, these solutions were made by analytical techniques however, these
were possible only for a very few situations where by clever conspiracy, the
loads and geometries were so simplified that the problem became tractable.
However, by ingenuity, the engineer could use this limited library of simple
solutions to construct meaningful pictures of more complicated situations.

It is the role of the structural designer to ensure that the artifacts he


designs serve the desired functions with maximum efficiency, but putting only as
much flesh as is needed on the bare skeleton. Often, this is measured in terms
of economy of cost and/or weight. Thus for vehicles,low weight is of the
essence, as it relates directly to speed of movement and cost of operation. The
efficiency of such a structure will depend on how every bit of material that is
used in components and joints is put to maximum stress without failing. Strength
is therefore the primary driver of design. Other design drivers are stiffness
(parts of the structures must not have excessive deflections), buckling (members
should not deflect catastrophically under certain loading conditions), etc. It
is not possible to do this check for structural integrity without having
sophisticated tools for analysis. FEM packages are therefore of crucial
relevance here. In fact, the modern trend is to integrate fem analysis tools
with solid modeling and CAD/CAM software in a single seamless chain or cycle all
the way from concept and design to preparation of tooling instructions for
manufacture using numerically controlled machines.

1.3 From analytical solutions to computational procedures

1.3.1 Introduction

We have seen earlier in this chapter how a body of knowledge that governs the
behavior of materials, solids and structures came to exist. Gifted
mathematicians and engineers were able to formulate precise laws that governed
the behavior of such systems and could apply these laws through mathematical
models that described the behavior accurately. As these mathematical models had
to take into account the continuum nature of the structural bodies, the
description often resulted in what are called differential equations. Depending
on the sophistication of the description, these differential equations could be
very complex. In a few cases, one could simplify the behavior to gross
relationships - for a bar, or a spring, it was possible to replace the
differential equation description with a simple relation between forces and
displacements; we shall in fact see that this is a very simple and direct act
of the discretization process that is the very essence of the finite element
process. For a more general continuum structure, such discretizations
or simplifications were not obvious at first. Therefore, there was no recourse

2
Fig. 1.1 A simple bar problem under axial load

except to find by mathematical and analytical techniques, viable solutions to


the governing differential equations. This can become very formidable in cases
of complex geometry, loading and material behavior, and often were intractable;
however with ingenuity, solutions could be worked out for a useful variety of
structural forms. Thus a vast body of solutions to the elastic, plastic,
viscoelastic behavior of bars, beams, plates and shells has been built over the
last two centuries.

It was recognized quite early that where analytical techniques fail or were
difficult to implement, approximate techniques could be devised to compute the
answers. In fact, much before algebra and analysis arrived, in contexts more
general than solid or structural mechanics, simple approximation and
computational schemes were used to solve engineering and scientific problems. In
this chapter, we shall review what we mean by an analytical solution for a very
simple problem and then proceed to examine some computational solutions. This
will reveal to us how the discretization process is logically set up for such a
problem.

1.3.2 The simple bar problem

Perhaps the earliest application of the discretization technique as it appeared


in civil engineering practice was to the bar problem. The bar is a prismatic
one-dimensional structure which can resist only axial loads, in tension or in
compression, and cannot take any bending loads. Figure 1.1 shows its simplest
configuration - a cantilever bar. This is a statically determinate problem,
meaning that one can obtain a complete solution, displacements as well as
strains and stresses from considerations of equilibrium alone. We shall now
compute the analytical solution to the problem depicted in Fig.1.1 using very
elementary engineering mathematics.

1.3.2.1 Analytical solution

The problem is categorized as a boundary value problem. We presume here that for
this problem, the reader is able to understand that the governing differential
equation describing the situation is

EA u,xx = 0 (1.1)

where E is the Young’s modulus, A is the area of cross-section, u(x) is the


function describing the variation of displacement of a point, and ,x denotes
differentiation with respect to coordinate x. This is the equation of
equilibrium describing the rate at which axial force varies along the length of

3
the bar. In general, governing differential equations belong to a category
called partial differential equations but here we have a simpler form known as
an ordinary differential equation. Equation (1.1) is further classified as a
field equation or condition as one must find a solution to the variable u which
must satisfy the equation over the entire field or domain, in this case, for x
ranging from 0 to L. A solution is obviously

u = ax + b (1.2)

where a and b are as yet undetermined constants. To complete the solution, i.e.
to determine a and b in a unique sense for the boundary (support) and loading
conditions shown in Fig. 1.1, we must now introduce what are called the boundary
conditions. Two boundary conditions are needed here and we state them as

u = 0 at x = 0 (1.3a)
EA u,x = P at x = L (1.3b)

The first relates to the fact that the bar is fixed at the left end and the
second denotes that a force P is applied at the free end. Taking these two
conditions into account, we can show that the following description completely
takes stock of the situation:

u(x) = Px/EA (1.4a)


ε(x) = P/EA (1.4b)
σ(x)= Eε(x)= P/A (1.4c)
P(x) = Aσ(x) = P (1.4d)

where ε(x), σ(x) and P(x) are the strain, stress and axial force (stress
resultants) along the length of the bar. These are the principal quantities that
an engineer is interested in while performing stress analysis to check the
integrity of any proposed structural design.

1.3.2.2 Approximate solutions

It is not always possible to determine exact analytical solutions to most


engineering problems. We saw in the example above that an analytical solution
provides a unique mathematical expression describing in complete detail the
value of the field variable at every location in the body. From this expression,
other expressions can be derived which describe further quantities of practical
interest, e.g. strains, stresses, stress resultants, etc.

In most problems where analytical solutions appear infeasible, it is


necessary to resort to some approximate or numerical methods to obtain these
values of practical interest. One obvious method at this stage is to start with
the equations we have already derived, namely Equations (1.1) and (1.3). This
can be achieved using a technique known as the finite difference method. We
shall briefly describe this next.

4
Fig. 1.2 Grid for finite difference scheme

Other approximate solutions are based on what are called functional


descriptions or variational descriptions of the problem. These describe the
problem at a level which is more fundamental in the laws of physics than the
more specialized descriptions in terms of governing differential equations as
seen in Equations (1.1) and (1.3) above. We deal directly with energy or work
quantities and discretization or approximation is applied at that level. If we
can appreciate that the governing differential equations were derived
analytically from such energy or work principles using what is called a
variational or virtual work approach, then it would become easier to understand
that an approximation applied directly to the energy or work quantities and a
variation carried out subsequently on these terms will preserve both the physics
as well as the approximating or numerical solution in one single process. This
will be the basis for the finite element method as we shall see in subsequent
subsections.

1.3.2.3 Finite difference approximations

For a problem where the differential equations are already known, the technique
known as the finite difference method can be used to obtain an approximate or
numerical solution. In fact, before the finite element method was established,
it was this method which was used to obtain solutions for very complicated
problems in structural mechanics, fluid dynamics and heat transfer.

Figure 1.2 shows a very simple uniformly spaced mesh or grid of nodal
points that can be used to discretize the problem described by Equations (1.1)
and (1.3). The larger the number of nodal points used (i.e. the smaller the grid
spacing h), the greater will be the accuracy involved. The field variable is now
taken to be known (or very strictly, unknown, at this stage of the computation)
at these nodal points. Thus, u1,..., ui,..., un, are the unknown variables or
degrees of freedom. The next step is to rewrite the governing differential
equations and boundary conditions in finite difference form. We see that the
finite difference forms for u,x and u,xx are required. It is easy to show
(again, details are omitted, as these can be found in most books on numerical
analysis or fem) that at a grid point i.

u,x = (ui+1 – ui)/h (1.5a)


2
u,xx = (ui+1 – 2ui + ui-1)/h (1.5b)

5
We now attempt a solution in which four points are used in the finite difference
grid. The governing differential equations and boundary conditions are replaced
by the following discrete set of linear, algebraic equations:

u1 = 0 (1.6a)
u3 – 2u2 + u1 = 0 (1.6b)
u4 - 2u3 + u2 = 0 (1.6c)
u4 - u3 = Ph/EA (1.6d)

where h=L/3 and Equations (1.6a) and (1.6d) represent the boundary conditions at
x = 0 and L respectively. The reader can easily work out that the solution
obtained from this set of simultaneous algebraic equations is, u2 = PL/3EA,
u3 = 2PL/3EA and u4 = PL/EA. Comparison with Equation (1.4a) will confirm that
we have obtained the exact answers at the grid points. Other quantities of
interest like strain, stress, etc. can be computed from the grid point values
using the finite difference forms of these quantities.

The above illustrates the solution to a very simple one-dimensional


problem for which the finite difference procedure yielded the exact answer.
Generalizations to two and three dimensions for more complex continuous problems
can be made, especially if the meshes are of regular form and the boundary lines
coincide with lines in such regular grids. For more complex shapes, the finite
difference approach becomes difficult to use. It is in such applications that
the finite element method proves to be more versatile than the finite difference
scheme.

1.3.2.4 Variational formulation

We began this chapter with an analytical solution to the differential equations


governing the problem. We did not ask how these equations originated. There are
two main ways in which it can be done. The first, and the one that is most often
introduced at the earliest stage of the study of structural mechanics, is to
formulate the equations of equilibrium in terms of force or stress quantities.
Then, depending on considerations of static determinacy, make the problem
determinate by introducing stress-strain and strain-displacement relations until
the final, solvable, set of governing equations is obtained.

The second method is one that originates from a more fundamental statement
of the principles involved. It recognizes that one of the most basic and elegant
principles known to man is the law of least action, or minimum potential energy,
or virtual work. It is a statement that Nature always chooses the path of
minimum economy. Thus, the law of least action or minimum total potential is as
axiomatic as the basic laws of equilibrium are. One can start with one axiom and
derive the other or vice-versa for all of mechanics or most of classical
physics. In dynamics, the equivalent principle is known as Hamilton's principle.

In structural mechanics, we start by measuring the total energy stored in


a structural system in the form of elastic strain energy and potential due to
external loads. We then derive the position of equilibrium as that state where
this energy is an extremum (usually a minimum if this is a stable state of
equilibrium). This statement in terms of minimum energy is strictly true only
for what are called conservative systems. For non-conservative systems, a more

6
general statement known as the principle of virtual work would apply. From these
brief general philosophical reflections we shall now proceed to the special case
at hand to see how the energy principle applies.

The total energy is stated in the form of a functional. In fact, most of


the problems dealt with in structural mechanics can be stated in a variational
form as the search for a minimal or stationary point of a functional. The
functional, Π, is an integral function of functions which are defined over the
domain of the structure. Thus, for our simple bar problem, we must define terms
such as the strain energy, U, and the potential due to the load, V, in terms of
the field variable u(x). This can be written as

L
U = ò 1/ 2 EA u 2 ,x dx
0 (1.7a)

V = P ux=L (1.7b)

Then,
Π ( υ(x )) =
L
ò0 1/ 2 EA u 2 ,x dx - P ux = L
(1.8)

Using standard variational procedures from the calculus of variations, we can


show that the extremum or stationary value of the functional is obtained as

δΠ = 0 (1.9)

If this variation is carried out, after integrating by parts and by regrouping


terms, we will obtain the same equation of equilibrium and boundary conditions
that appeared in Equations (1.1) and (1.3) above. Our interest is now not to
show what can be done with these differential equations but to demonstrate that
the approximation or discretization operation can be implemented at the stage
where the total potential or functional is defined.

1.3.2.5 Functional approximation

There are several ways in which an approximation can be applied at the level of
energy, virtual work or weighted residual statements, e.g. the Rayleigh-Ritz
(R-R) procedure, the Galerkin procedure, the least-squares procedure, etc. We
first choose a set of independent functions which satisfy the geometric or
essential (also called kinematic) boundary conditions of the problem. These
functions are called admissible functions. In this case, the condition specified
in Equation (1.3a) is the geometric boundary condition. The other condition,
Equation (1.3b) is called a nonessential or natural or force boundary condition.
The admissible functions need not satisfy the natural boundary conditions (in
fact, it can be shown that if they are made to do so, there can even be a
deterioration in performance). The approximate admissible configuration for the
field variable, say u (x ) , is obtained by a linear combination of these functions
using unknown coefficients or parameters, also known as degrees of freedom or
generalized coordinates. These unknown constants are then determined to satisfy
the extremum or stationary statement. Methods like the Rayleigh-Ritz procedure

7
(the one going to be used now), the Galerkin method, collocation methods, least
square methods, etc. all proceed with approximating fields chosen in this way.

For our problem, let us choose only one simple linear function, using a
polynomial form for the purpose. The advantages of using polynomial functions
will become clear later in the book. Thus, only one constant is required, i.e.

u (x ) = a1x
(1.10)

Note that this configuration is an admissible one satisfying the geometric


condition u (0 ) = 0 . By substituting in Equation (1.8) and carrying out the
variation prescribed in Equation (1.9), we are trying to determine the value of
a1 which provides the best functional approximation to the variational problem.
It is left to the reader to show that the approximate solution obtained is
u (x ) = Px/EA (1.11a)
ε (x ) = P/EA (1.11b)

σ (x ) = Eε(x) = P/A (1.11c)

P (x ) = Aσ(x) = P (1.11d)

Thus, for this problem, the approximate solution coincides with the exact
solution. Obviously, we cannot draw any conclusions here as to the errors
involved in approximation by the R-R procedure. In more complicated problems,
approximate solutions will be built up from several constants and admissible
functions and convergence to the exact solution will depend on the number of
terms used and the type of functions used.

The R-R procedure, when applied in a piecewise manner, over the element
domains that constitute the entire structure, becomes the finite element method.
Thus, an understanding of how the R-R method works will be crucial to our
understanding of the finite element method.

1.3.2.6 Finite element approximation

We shall now describe an approximate solution by the finite element method. We


shall use a form known as the displacement-type formulation (also called the
stiffness formulation). The simplest element known for our purpose is a two
noded line element such as shown in Fig. 1.3. This element is also known as the
bar, truss or rod element. It has two nodes at which nodal displacements u1 and
u2 are prescribed. These are also called the nodal degrees of freedom. Also
indicated in Fig. 1.3 are the forces P1 and P2 acting at these nodes.

Thus, any one-dimensional structure can be replaced by contiguous elements


placed from end to end. The only information communicated from element to
element is that stored as nodal degrees of freedom and the nodal forces. We can
now see how this piecewise Ritz approximation is performed.

8
Fig. 1.3 A two-node bar element

We first derive an expression for the approximate representation of the


field variable, u (x ) , within the region of the element by relating it to the
nodal degrees of freedom. It can be shown that,

u (x ) = u1N 1 + u2 N 2 (1.12)

where N1 =(1-ξ)/2 and N2 =(1+ξ)/2,where ξ=x/l. N1 and N2 are called the shape
functions. Note that this form is exactly the same as the linear function used
in our R-R approximation earlier, Equation (1.10), with generalized coordinates
ai being now replaced by nodal quantities ui. It is possible therefore to
compute the energy stored in a beam element as

é 1 - 1 ù ìïu1 üï
{u1 u2 } EA
ê ú í ý (1.13)
2l êë- 1 1úû ïîu2 ïþ

and the potential due to the applied loads as

{u1 u2 } ìíPP1 üý
î 2þ (1.14)

Thus, if a variation is taken over u1 and u2, we can write the equation of
equilibrium as
EA é 1 - 1ù ìu1 ü ìP1 ü
ê ú í ý = í ý
2l ë- 1 1û îu2 þ îP2 þ (1.15)

This matrix representation is typical of the way finite element equations are
assembled and manipulated automatically on the digital computer. We shall now
attempt to solve our bar problem using a finite element model comprising just
one element. Node 1 is therefore placed at the fixed end and node 2 coincides
with the free end where the load P2 = P is applied. The fixity condition at node
1 is simply u1 = 0 and P1 indicates the reaction at this end. It is very simple
to show from these that u2 = PL/EA. We can see that the exact solution has been
obtained.

What we have seen above is one of the simplest demonstrations of the


finite element method that is possible. Generalizations of this approach to two

9
and three dimensional problems permit complex plate, shell and three dimensional
elasticity problems to be routinely handled by ready made software packages.

1.4 Concluding remarks

In this chapter, we have briefly investigated how classical analytical


techniques, which are usually very limited in scope, can give way to
computational methods of obtaining acceptable solutions. We have seen that the
finite element method is one such approximate procedure. By using a very large
number of elements, it is possible to obtain solutions of greater accuracy. Over
the years, finite element theorists and analysts have produced a very large body
of work that shows how such solutions are managed for a great variety of
problems. This is in the first-order tradition of the method. The remaining part
of this book hopes to address questions like: How good is the approximate
solution? What kinds of situation affect these approximations in a trivial and
in not-so-trivial ways? Since solutions depend very much on the number of
elements used and the type of shape function patterns used within each element,
we must ask, what patterns are the best to assume? All these questions are
linked to one very fundamental question: In what manner does the finite element
method actually compute the solution? Does it make its first approximation of
the displacement field directly or on the stress field? The exercises and
exposition in this book are aimed at throwing more light on these questions.

10
Chapter 2

Paradigms and some approximate solutions


2.1 Introduction

We saw in the preceding chapter that a continuum problem in structural or solid


mechanics can either be described by a set of partial differential equations and
boundary conditions or as a functional Π based on the energy principle whose
extremum describes the equilibrium state of the problem. Now, a continuum has
infinitely many material points and therefore has infinitely many degrees of
freedom. Thus, a solution is complete only if analytical functions can be found
for the displacement and stress fields, which describe these states, exactly
everywhere in the domain of the problem. It is not difficult to imagine that
such solutions can be found only for a few problems.

We also saw earlier that the Rayleigh-Ritz (RR) and finite element (fem)
approaches offer ways in which approximate solutions can be achieved without the
need to solve the differential equations or boundary conditions. This is managed
by performing the discretization operation directly on the functional. Thus, a
real problem with an infinitely large number of degrees of freedom is replaced
with a computational model having a finite number of degrees of freedom. In the
RR procedure, the solution is approximated by using a finite number of
admissible functions ƒi and a finite number of unknown constants ai so that the
approximate displacement field is represented by a linear combination of these
functions using the unknown constants. In the fem process, this is done in a
piecewise manner - over each sub-region (element) of the structure, the
displacement field is approximated by using shape functions Ni within each sub-
region and nodal degrees of freedom ui at nodes strategically located so that
they connect the elements together without generating gaps or overlaps. The
functional now becomes a function of the degrees of freedom (ai or ui as the
case may be). The equilibrium configuration is obtained by applying the
criterion that Π must be stationary with respect to the degrees of freedom.

It is assumed that this solution process of seeking the stationary or


extremum point of the discretized functional will determine the unknown
constants such that these will combine together with the admissible or shape
functions to represent some aspect of the problem to some "best" advantage.
Which aspect this actually is has been a matter of some intellectual
speculation. Three competing paradigms present themselves.

It is possible to believe that by "best" we mean that the functions tend


to satisfy the differential equations of equilibrium and the stress boundary
conditions more and more closely as more terms are added to the RR series or
more elements are added to the structural mesh. The second school of thought
believes that it is the displacement field, which is approximated to greater
accuracy with improved idealization. The "aliasing" paradigm, which will be
critically discussed later, belongs to this school. It follows from this that
stresses, which are computed as derivatives of the approximate displacement
fields, will be less accurate. In this book however, we will seek to establish
the currency of a third paradigm - that the RR or fem process actually seeks to
find to best advantage, the state of stress or strain in the structure. In this

11
scenario, the displacement fields are computed from these "best-fit" stresses as
a consequence.

Before we enter into a detailed examination of the merits or faults of


each of these paradigms, we shall briefly introduce a short statement on what is
meant by the use of the term "paradigm" in the present context. We shall follow
this by examining a series of simple approximations to the cantilever bar
problem but with more and more complex loading schemes to see how the overall
picture emerges.

2.2 What is a "paradigm"?

Before we proceed further it may be worthwhile to state what we mean by a


paradigm here. This is a word that is uncommon to the vocabulary of a trained
engineer. The dictionary meaning of paradigm is pattern or model or example.
This does not convey much in our context. Here, we use the word in the greatly
enlarged sense in which the philosopher T. S. Kuhn introduced it in his classic
study on scientific progress [2.1]. In this sense, a paradigm is a "framework of
suppositions as to what constitutes problems, theories and solutions. It can be
a collection of metaphysical assumptions, heuristic models, commitments, values,
hunches, which are all shared by a scientific community and which provide the
conceptual framework within which they can recognize problems and solve them
[2.2]. The aliasing and best-fit paradigms can be thought of as two competing
scenarios, which attempt to explain how the finite element method computes
strains and stresses. Our task will therefore be to establish which paradigm has
greater explanatory power and range of application.

Fig. 2.1 Bar under uniformly distributed axial load

12
2.3 Bar under uniformly distributed axial load

Consider a cantilever bar subjected to a uniformly distributed axial load of


intensity q 0 per unit length (Fig. 2.1). Starting with the differential
equation of equilibrium, it is easy to show that the analytical solution to the
problem is

(
u (x ) = (q 0 EA ) Lx − x 2 2 ) (2.1a)

σ (x ) = (q 0 A ) (L − x ) (2.1b)

Consider a one-term RR solution based on ur = a1x , where the subscript r denotes


the use of the RR approach. It is left to the reader to show that the
approximate solution obtained is

ur (x ) = (q 0 EA ) (Lx 2 ) (2.2a)

(
σ r (x ) = q 0 A ) (L 2 ) (2.2b)

We now compare this with an fem solution based on a two-noded linear element.
The solution obtained will be
uf (x ) = (q 0 EA ) (Lx 2 ) (2.3a)

(
σ f (x ) = q 0 A ) (L 2 ) (2.3b)

We see that the RR and fem solutions are identical. This is to be expected
because the fem solution is effectively an RR solution. We can note also the
curious coincidence where all three solutions predict the same displacement at
the tip. However, from Fig. 2.1 we can see that the ur and uf are approximate
solutions to u. It is also clear from Fig. 2.1 that σ r = σ f = σ at the mid-point
of the beam. It is possible to speculate that σ r and σ f bear some unique
relationship to the true variation σ.

Consider next what will happen if two equal length linear (i.e., two-
noded) bar elements are used to model the bar. It is left to the reader to work
out that the solution described in Fig. 2.2 will be obtained. First, we must
note that the distributed axial load is consistently lumped at the nodes. Thus
the physical load system that the fem equations are solving is not that
described in Fig. 2.1 or Fig. 2.2 as σ. Instead, we must think of a substitute
stairstep distribution σf, produced by the consistent load lumping process (see
Fig. 2.2) which is sensed by the fem stiffness matrix. Now, a solution of the
set of algebraic equations will result in σ f and uf as the fem solution.

We see once again that the nodal predictions are exact. This is only a
coincidence for this particular type of problem and nothing more can be read
into this fact. More striking is the observation that the stresses computed by

13
the finite element system now approximates the original true stress in a
stairstep fashion.

It also seems reasonable to conclude that within each element, the true
state of stress is captured by the finite element stress in a "best-fit" sense.
In other words, we can generalize from Figs. 2.1 and 2.2, that the finite
element stress magnitudes are being computed according to some precise rule.

Also, there is the promise that by carefully understanding what this rule is, it
will be possible to derive some unequivocal guidelines as to where the stresses
are accurate. In this example, where an element capable of yielding constant
stresses is used to model a problem where the true stresses vary linearly, the
centroid of the element yields exact predictions. As we take up further examples
later, this will become more firmly established.

A cursory comparison of Figs. 2.1 and 2.2 also indicates that in a general
sense, the approximate displacements are more accurate than the approximate
stresses. It seems compelling now to argue that this is so because the
approximate displacements emerge as "discretized" integrals of the stresses or
strains and for that reason, are more accurate than the stresses.

2.4 Bar under linearly varying distributed axial load

We now take up a slightly more difficult problem. The cantilever bar has a load
distributed in a linearly varying fashion (Fig. 2.3). It is easy to establish
that the exact stress distribution in this case will be quadratic in nature. We
can write it as

σ (x ) = æç q 0 L2 8A ö÷ æç 4 3 − 2ξ - æç 1 - 3ξ 2 ö÷ 3 ö÷ (2.4)
è ø è è ø ø
Some interesting features about this equation can be noted. A dimensionless
coordinate, ξ = 2x/L-1 has been chosen so that it will also serve as a natural
coordinate system taking on values -1 and 1 at the ends of the single three-node
bar element shown as modeling the entire bar in Fig.2.3. We have also very
curiously expanded the quadratic variation using the terms 1, ξ, (1-3ξ ). These
2

can be identified with the Legendre polynomials and its relevance to the
treatment here will become more obvious as we proceed further.

Fig. 2.2 Two element model of bar under uniformly distributed axial load.

14
Fig. 2.3 Bar under linearly varying axial load

We shall postpone the first obvious approximation, that of using a one-


term series ur = a1x till later. For now, we shall consider a two-term series
ur = a1x + a2 x 2 . This is chosen so that the essential boundary conditions at x=0
is satisfied. No attempt is made to satisfy the force boundary condition at x=L.
The reader can verify by carrying out the necessary algebra associated with the
RR process that the solution obtained will yield an approximate stress pattern
given by
( )
σ r (x ) = q 0 L2 8A (4 3 − 2ξ ) (2.5)

This is plotted on Fig. 2.3 as the dashed line. A comparison of Equations (2.4)
and (2.5) reveals an interesting fact - only the first two Legendre polynomial
terms are retained. Taking into account the fact that the Legendre polynomials
are orthogonal, what this means is that in this problem, we have obtained σ r in
a manner that seems to satisfy the following integral condition:
ò-1 δσ r (σ r − σ ) dξ = 0
1
(2.6)
That is, the RR procedure has determined a σ r that is a "best-fit" of the true
state of stress σ in the sense described by the orthogonality condition in
Equation (2.6). This is a result anticipated from our emerging results to the
various exercises we have conducted so far. We have not been able to derive it
from any general principle that this must be so for stronger reasons than shown
here up till now.

15
( )
Let us now proceed to an fem solution. It is logical to start here with
the three-node element that uses the quadratic shape functions, N = ξ ξ − 1 2 ,
1

(
N2 = 1 − ξ
2
) and ( )
N 3 = ξ ξ + 1 2. We first compute the consistent loads to be placed

at the nodes due to the distributed load using Pi = ò N iq dx . This results in the
following scheme of loads at the three nodes identified in Fig. 2.3: P1 = q 0 L 6 ,
2

P2 = q 0 L2 3 and P3 = 0 . The resulting load configuration can be represented in the


form of a stress system shown as σf, represented by the dashed-dotted lines in
Fig. 2.3. Thus, any fem discretization automatically replaces a smoothly varying
stress system by a step-wise system as shown by σ f in Figs. 2.2 and 2.3. It is
this step-wise system that the finite element solution σ f responds to. If the
finite element computations are actually performed using the stiffness matrix
for the three-node element and the consistent load vector, it turns out, as the
reader can assure himself, that the computed fem stress, pattern will be

( )
(4 3 − 2ξ )
σ f x = æç q 0 L 8A ö÷
2 (2.7)
è ø
This turns out to be exactly the same as the σ r computed by the RR
process in Equation (2.5). At first sight, this does not seem to be entirely
unexpected. Both the RR process and the fem process here have started out with
quadratic admissible functions for the displacement fields. This implies that
both have the capability to represent linear stress fields exactly or more
complicated stress fields by an approximate linear field that is in some sense a
best approximation. On second thought however, there is some more subtlety to be
taken care of. In the RR process, the computed σ r was responding to a
quadratically varying system σ (see Fig. 2.3). We could easily establish
through Equation (2.6) that σ r responded to σ in a "best-fit" manner. However,
in the fem process, the load system that is being used is the σf system, which
varies in the stairstep fashion. The question confronting us now is, in what
manner did σ f respond to σf - is it also consistent with the "best-fit"
paradigm?

Let us now assume an unknown field σ = c0 + c1 ξ which is a "best-fit" of


the stairstep field given by σ f = 2q 0 L2 3 in 0 <x <L/2 and σf = 0 in L/2 <x <L.
We shall determine the constants c0 and c1 so that the "best-fit" variation
shown below is satisfied:

δ σ æç σ − 2q 0 L2 3 ö÷ dξ + æ ö
0 1
ò−1 è ø ò0 δ σ çè σ − 0 ÷ø dξ = 0 (2.8)

The reader can work out that this leads to

(
σ (x ) = q 0 L2 8A ) (4 3 − 2ξ ) , (2.9)

16
which is identical to the result obtained in Equation (2.7) by carrying out the
finite element process. In other words, the fem process follows exactly the
"best-fit" description of computing stress fields. Another important lesson to
be learnt from this exercise is that the consistent lumping process preserves
the "best-fit" nature of the stress representation and subsequent prediction.
Thus, σ f is a "best-fit" of both σ and σf! Later, we shall see how the best-fit
nature gets disturbed if the loads are not consistently derived.

It again seems reasonable to argue that the nodal displacements computed


directly from the stiffness equations from which the stress field σ f has been
processed can actually be thought of as being "integrated" from the "best-fit"
stress approximation. It seems possible now to be optimistic about our ability
to ask and answer questions like: What kind of approximating fields are best to
start with? and, How good are the computed results?

Before we pass on, it is useful for future reference to note that the
approximate solutions σ r or σ f intersect the exact solution σ at two points. A
comparison of Equation (2.4) with Equations (2.5) and (2.7) indicates that these
are the points where the quadratic Legendre polynomial, (1 − 3ξ )
2
vanishes, i.e.,
at ξ = ± 1 3 . Such points are well known in the literature of the finite element
method as optimal stress points or Barlow points, named after the person who
first detected such points. Our presentation shows clearly why such points
exist, and why in this problem, where a quadratic stress state is sought to be
approximated by a linear stress state, these points are at ξ = ± 1 3 . Curiously,
these are also the points used in what is called the two-point rule for Gauss-
Legendre numerical integration. Very often, the two issues are confused. Our
interpretation so far shows that the fact that such points are sampling points
of the Gauss-Legendre rule has nothing to do with the fact that similar points
arise as points of optimal stress recovery. The link between them is that in
both, the Legendre polynomials play a decisive role - the zeros of the Legendre
polynomials define the optimal points for numerical integration and these points
also help determine where the "best-fit" approximation coincides with a
polynomial field which is exactly one order higher.

We shall now return to the linear Ritz admissible function, ur = a1x , to


see if it operates in the best-fit sense. This would be identical to using a
single two-node bar element to perform the same function. Such a field is
capable of representing only a constant stress. This must now approximate the
quadratically varying stress field σ(x) given by Equation (2.4). This gives us
an opportunity to observe what happens to the optimal stress point, whether one
exists, and whether it can be easily identified to coincide with a Gauss
integration point.

Again, the algebra is very simple and is omitted here. One can show that
the one-term approximate solution would lead to the following computed stress:

(
σ r (x ) = q 0 L2 8A ) (4 3 ) (2.10)

17
What becomes obvious by comparing this with the true stress σ(x) in Equation
(2.4) and the computed stress from the two-term solution, σ r (x ) in Equation
(2.5) is that the one-term solution corresponds to only the constant part of the
Legendre polynomial expansion! Thus, given the orthogonal nature of the Legendre
polynomials, we can conclude that we have obtained the "best-fit" state of
stress even here. Also, it is clear that the optimal stress point is not easy to
identify to coincide with any of the points corresponding to the various Gauss
integration rules. The optimal point here is given by ξ = 1 − 4 3 .

2.5 The aliasing paradigm

We have made out a very strong case for the best-fit paradigm. Let us now
examine the merits, if any, of the competing paradigms. The argument that fem
procedures look to satisfy the differential equations and boundary conditions
does not seem compelling enough to warrant further discussion. However, the
belief that finite elements seek to determine nodal displacements accurately was
the basis for Barlow's original derivation of optimal points [2.3] and is also
the basis for what is called the "aliasing" paradigm [2.4].

The term aliasing is borrowed from sample data theory where it is used to
describe the misinterpretation of a time signal by a sampling device. An
original sine wave is represented in the output of a sampling device by an
altered sine wave of lower frequency - this is called the alias of the true
signal. This concept can be extended to finite element discretization - the
sample data points are now the values of the displacements at the nodes and the
alias is the function, which interpolates the displacements within the element
from the nodal displacements. We can recognize at once that Barlow [2.3]
developed his theory of optimal points using an identical idea - the term
substitute function is used instead of alias.

Let us now use the aliasing concept to derive the location of the optimal
points, as Barlow did earlier [2.3], or as MacNeal did more recently [2.4]. We
assume here that the finite element method seeks discretized displacement
fields, which are substitutes or aliases of the true displacement fields by
sensing the nodal displacements directly. We can compare this with the
"best-fit" interpretation where the fem is seen to seek discretized
strain/stress fields, which are the substitutes/aliases of the true
strain/stress fields in a "best fit" or "best approximation" sense. It is
instructive now to see how the alternative paradigm, the "displacement aliasing"
approach leads to subtle differences in interpreting the relationship between
the Barlow points and the Gauss points.

Table 2.1 Barlow and Gauss points for one-dimensional case

p Nodes Gauss Barlow points


at u u ε ε points ‘best-fit’ aliasing
1 ± 1 ξ2
ξ ξ 1 0 0 0
2 0, ± 1 ξ3 ξ2 ξ2 ξ ± 1/√3 ± 1/√3 ± 1/√3

3 ±1/3, ± 1 ξ4 ξ
3
ξ3 ξ2 0, ± 3 5 0, ± 3 5 0, ± √5/3
1, ξ..,ξ
4
indicate polynomial orders

18
2.5.1 A one dimensional problem

We again take up the simplest problem, a bar under axial loading. We shall
assume that the bar is replaced by a single element of varying polynomial order
for its basis function (i.e. having varying no. of equally spaced nodes). Thus,
from Table 2.1, we see that p=1,2,3 correspond to basis functions of linear,
quadratic and cubic order, implying that the corresponding elements have 2,3,4
nodes respectively. These elements are therefore capable of representing a
constant, linear and quadratic state of strain/stress, where strain is taken to
be the first derivative of the displacement field. We shall adopt the following
notation: The true displacement, strain and stress field will be designated by
u, ε and σ . The discretized displacement, strain and stress field will be
designated by u , ε and σ . The aliased displacement, strain and stress field
will be designated by u , εa and σa. Nodal displacements will be represented by
a

ui.

We shall examine three scenarios. In the simplest, Scenario a, the true


displacement field u is exactly one polynomial order higher than what the finite
element is capable of representing - we shall see that the Barlow points can be
determined exactly in terms of the Gauss points only for this case. In Scenario
b, we consider the case where the true field u is two orders higher than the
discretized field u . The definition of an identifiable optimal point becomes
difficult. In both Scenarios a and b, we assume that the rigidity of the bar is
a constant, i.e., σ = D ε. In Scenario c, we take up a case where the rigidity
can vary, i.e., σ = D(ξ) ε. We shall see that once again, it becomes difficult
to identify the optimal points by any simple rule.

We shall now take for granted that the best-fit rule operates according to
the orthogonality condition expressed in Equation (2.6) and that it can be used
interchangeably for stresses and strains. We shall designate the optimal points
determined by the aliasing algorithm as ξa, the Barlow points (aliasing), and
the points determined by the best-fit algorithm as ξb, the Barlow points (best-
fit). Note that ξa are the points established by Barlow [2.3] and MacNeal [2.4],
while ξb will correspond to the points given in References 2.6 and 2.7. Note
that the natural coordinate system ξ is being used here for convenience.

Thus, for Scenarios a and b, this leads to

ò δε (ε − ε ) dV = 0
T
(2.11)

whereas for Scenario c, it becomes

ò δε
T
D (ξ ) (ε − ε ) dV = 0 (2.12)

Note that we can consider Equation (2.11) as a special case of the orthogonality
condition in Equation (2.12) with the weight function D(ξ)=1. This case
corresponds to one in which a straightforward application of Legendre
polynomials can be made. This point was observed very early by Moan [2.5]. In
this case, one can determine the points where ε = ε as those corresponding to
points, which are the zeros of the Legendre polynomials. See Table 2.2 for a
list of unnormalised Legendre polynomials. We shall show below that in Equation

19
(2.11), the points of minimum error are the sampling points of the Gauss
Legendre integration rule only if ε is exactly one polynomial order lower than
ε. And in Equation (2.12), the optimal points no longer depend on the nature of
the Legendre polynomials, making it difficult to identify the optimal points.

Scenario a

We shall consider fem solutions using a linear (two-node), a quadratic (three-


node) and a cubic (four-node) element. The true displacement field is taken to
be one order higher than the discretized field in each case.

Linear element (p = 1)

u = quadratic = bo + b1 ξ + b2 ξ
2
(2.13)

p =1
ε = linear = u,ξ = b1 + 2b2 ξ = å ε s Ps (ξ ) (2.14)
s=0

Note that we have written ε in terms of the Legendre polynomials for future
convenience. Note also that we have simplified the algebra by assuming that
strains can be written as derivatives in the natural co-ordinate system. It is
now necessary to work out how the algebra differs for the aliasing and best-fit
approaches.

Aliasing: At ξi = ± 1 , uia = ui ; then points where ε a = ε are given by ξa=0. (The


algebra is very elementary and is left to the reader to work out). Thus, the
Barlow point (aliasing) is ξa = 0, for this case.

Best-fit: u =linear, is undetermined at first. Let ε = c0 , as the element is


capable of representing only a constant strain. Equation (2.11) will now give
ε = c0 = b1 . Thus, the optimal point is ξb = 0, the point where the Legendre
polynomial P1(ξ) = ξ vanishes. Therefore, the Barlow point (best-fit) for this
example is ξb = 0.

Table 2.2 The Legendre polynomials Pi

Order of polynomial i Polynomial


Pi
0 1

1 ξ

(1-3ξ )
2 2

3
(3ξ-5ξ )
3

4
(3-30ξ +35ξ
2 4
)

20
Quadratic element (p = 2)

u = cubic = b0 + b1 ξ + b2 ξ + b3 ξ
2 3

ε = quadratic = u,ξ (2.15)

= (b1 + b3 ) + 2b2 ξ − b3 æç 1 − 3ξ 2 ö÷ = å ε sPs (ξ ) (2.16)


è ø s=0

Aliasing: At ξi = 0, ± 1, uia = ui ; ; then points where ε a = ε are given by ξ a = ± 1 3 .


(Again, the algebra is left to the reader to work out). Thus, the Barlow points
(aliasing) are ξ a = ± 1 3 , for this case.

Best-fit: u = ε = c0 + c1ξ as this element is capable of


quadratic. Let
representing a linear strain. Equation (2.11) will now give ε = (b 1 + b 3 ) + 2b 2 ξ .
Thus, the optimal points are ξb = ± 1 3, the points where the Legendre
(
polynomial P2 (ξ ) = 1 − 3ξ 2 ) vanishes. Therefore, the Barlow points (best-fit) for
this example are ξ b = ± 1 3.

Note that in these two examples, i.e. for the linear and quadratic elements, the
Barlow points from both schemes coincide with the Gauss points (the points where
the corresponding Legendre polynomials vanish). In our next example we will find
that this will not be so.

Cubic element (p = 3)

u = quartic = b0 + b1ξ + b2ξ 2 + b3 ξ 3 + b4ξ 4 (2.17)

ε = cubic = u,ξ

( )
= (b1 + b3 ) + (2b2 + 12b4 5 )ξ − b3 1 − 3ξ 2 − 4b4 5 3ξ − 5ξ 3 ( )
p =3
= å ε sPs (ξ ) (2.18)
s=0

Aliasing: At ξi = ± 1 3 , ± 1, uia = ui ; then points where εa = ε are given by


ξ a = 0, ± 5 3 . Thus, the Barlow points (aliasing) are ξ a = 0, ± 5 3 for this case.
Note that the points where the Legendre polynomial P3 (ξ ) = 3ξ − 5ξ 3 ( ) vanishes are
ξ0 = 0, 3 5 !

(
Best-fit: u = cubic. Let ε = c0 + c1ξ + c2 1 − 3ξ 2 , as this element is capable of )
representing a quadratic strain. Equation (2.11) will now give
( )
ε = (b1 + b3 ) + (2b2 + 12b4 5 ) ξ − b3 1 − 3ξ 2 . Thus, the Barlow points (best-fit) for

21
this example are ξ b = 0, 3 5 , the points where the Legendre polynomial
(
P3 (ξ ) = 3ξ − 5ξ 3
) vanishes.

Therefore, we have an example where the aliasing paradigm does not give
the correct picture about the way the finite element process computes strains.
However, the best-fit paradigm shows that as long as the discretized strain is
one order lower than the true strain, the corresponding Gauss points are the
optimal points. Table 2.1 summarizes the results obtained so far.

The experience of this writer and many of his colleagues is that the best
-fit model, is the one that corresponds to reality. If one were to actually
solve a problem where the true strain varies cubically using a 4-noded element,
which offers a discretized strain which is of quadratic order, the points of
optimal strain actually coincide with the Gauss points.

Scenario b:

So far, we have examined simple scenarios where the true strain is exactly one
polynomial order higher than the discretized strain with which it is replaced.
If P p (ξ ) , denoting the Legendre polynomial of order p, describes the order by
which the true strain exceeds the discretized strain, the simple rule is that he
optimal points are obtained as P p (ξ b ) = 0 . These are therefore the set of p Gauss
points at which the Legendre polynomial of order p vanishes. Consider now a case
where the true strain is two orders higher than the discretized strain - e.g. a
quadratic element (p = 2) modeling a region where the strain and stress field
vary cubically. Thus, we have,

( ) (
ε = (b1 + b3 ) + (2b2 + 12b4 5 )ξ − b3 1 − 3ξ 2 − 4b4 5 3ξ − 5ξ 3 )
ε = c0 + c1ξ (2.19)

Equation (2.11) allows us to determine the coefficient ci in terms of bi; it


turns out that

ε = (b1 + b3 ) + (2b2 + 12b4 5 ) ξ , (2.20)

a representation made very easy by the fact that the Legendre polynomials area
orthogonal and that therefore ε can be obtained from ε by simple inspection.
It is not however a simple matter to determine whether the optimal points
coincide with other well-known points like the Gauss points. In this example, we
have to seek the zeros of

(
b3 1 − 3 ξ 2 ) + 4b 5 (3ξ − 5ξ )
4
3
(2.21)

Since b3 and b4 are arbitrary, depending on the problem, it is not possible to


seek universally valid points where this would vanish, unlike in the case of
Scenario a earlier. Therefore, in such cases, it is not worthwhile to seek
points of optimal accuracy. It is sufficient to acknowledge that the finite

22
element procedure yields strains ε , which are the most reasonable one can
obtain in the circumstances.

Scenario c

So far, we have confined attention to problems where σ is related to ε through a


simple constant rigidity term. Consider an exercise where (the one-dimensional
bar again) the rigidity varies because the cross-sectional area varies or
because the modulus of elasticity varies or both, i.e. σ = D(ξ)ε. The
orthogonality condition that governs this case is given by Equation (2.12).
Thus, it may not be possible to determine universally valid Barlow points a
priori if D(ξ) varies considerably.

2.6 The `best-fit' rule from a variational theorem

Our investigation here will be complete in all respects if the basis for the
best-fit paradigm can be derived logically from a basic principle. In fact, some
recent work [2.6,2.7] shows how by taking an enlarged view of the variational
basis for the displacement type fem approach we will be actually led to the
conclusion that strains or stresses are always sought in the best-fit manner.

The `best-fit' manner in which finite elements compute strains can be


shown to follow from an interpretation uses the Hu-Washizu theorem. To see how
we progress from the continuum domain to the discretized domain, we will find it
most convenient to develop the theory from the generalized Hu-Washizu theorem
[2.8] rather than the minimum potential theorem. As we have seen earlier, these
theorems are the most basic statements that can be made about the laws of
nature. The minimum potential theorem corresponds to the conventional energy
theorem. However, for applications to problems in structural and solid
mechanics, Hu proposed a generalized theorem, which had somewhat more
flexibility [2.8]. Its usefulness came to be recognized when one had to grapple
with some of the problems raised by finite element modeling. One such puzzle is
the rationale for the "best-fit" paradigm.

Let the continuum linear elastic problem have an exact solution described
by the displacement field u, strain field ε and stress field σ. (We project that
the strain field ε is derived from the displacement field through the strain-
displacement gradient operators of the theory of elasticity and that the stress
field is derived from the strain field through the constitutive laws.) Let us
now replace the continuum domain by a discretized domain and describe the
computed state to be defined by the quantities u , ε and σ , where again, we
take that the strain fields and stress fields are computed from the strain-
displacement and constitutive relationships. It is clear that ε is an
approximation of the true strain field ε. What the Hu-Washizu theorem does,
following the interpretation given by Fraejis de Veubeke [2.9], is to introduce
a dislocation potential to augment the usual total potential. This dislocation
potential is based on a third independent stress field σ which can be
considered to be the Lagrange multiplier removing the lack of compatibility
appearing between the true strain field ε and the discretized strain field ε .

23
Note that, σ is now an approximation of σ . The three-field Hu-Washizu theorem
can be stated as

δπ = 0 (2.22)
where

ì T T ü
π = ò íσ ε 2 + σ (ε − ε ) + P ý dv (2.23)
î þ

and P is the potential energy of the prescribed loads.

In the simpler minimum total potential principle, which is the basis for the
derivation of the displacement type finite element formulation in most
textbooks, only one field (i.e., the displacement field u) is subject to
variation. However, in this more general three field approach, all the three
fields are subject to variation and leads to three sets of equations which can
be grouped and classified as follows:

Variation on Nature Equation

u Equilibrium ∇ σ + terms from P = 0 (2.24a)

T
σ Orthogonality ò δσ (ε − ε ) dv = 0 (2.24b)
(Compatibility )

æ σ − σ ö dv = 0
ε ò δε
T
Orthogonality ç ÷ (2.24c)
(Equilibrium ) è ø

Equation (2.24a) shows that the variation on the displacement field u requires
that the independent stress field σ must satisfy the equilibrium equations (∇
signifies the operators that describe the equilibrium condition. Equation
(2.24c) is a variational condition to restore the equilibrium imbalance between
σ and σ . In the displacement type formulation, we choose σ = σ . This satisfies
the orthogonality condition seen in Equation (2.24c) identically. This leaves us
with the orthogonality condition in Equation (2.24b). We can now argue that this
tries to restore the compatibility imbalance between the exact strain field ε
and the discretized strain field ε . In the displacement type formulation this
can be stated as,

ò δσ
T
(ε − ε ) dV =0 (2.25)

Thus we see very clearly, that the strains computed by the finite element
procedure are a variationally correct (in a sense, a least squares correct)
`best approximation' of the true state of strain.

24
2.7 What does the finite element method do?

To understand the motivation for the aliasing paradigm, it will help to


remember that it was widely believed that the finite element method sought
approximations to the displacement fields and that the strains/stresses were
computed by differentiating these fields. Thus, elements were believed to be
"capable of representing the nodal displacements in the field to a good degree
of accuracy." Each finite element samples the displacements at the nodes, and
internally, within the element, the displacement field is interpolated using the
basis functions. The strain fields are computed from these using a process that
involves differentiation. It is argued further that as a result, displacements
are more accurately computed than the strain and stress field. This follows from
the generally accepted axiom that derivatives of functions are less accurate
than the original functions. It is also argued that strains/stresses are usually
most inaccurate at the nodes and that they are of greater accuracy near the
element centers - this, it is thought, is a consequence of the mean value
theorem for derivatives.

However, we have demonstrated convincingly that in fact, the Ritz


approximation process, and the displacement type fem which can be interpreted as
a piecewise Ritz procedure, do exactly the opposite - it is the strain fields
which are computed, almost independently as it were within each element. This
can be derived in a formal way. Many attempts have been made to give expression
to this idea, (e.g., Barlow [2.1] and Moan [2.7]), but it appears that the most
intellectually satisfying proof can be arrived at by starting with a mixed
principle known as the Hu-Washizu theorem [2.8]. This proof has been taken up in
greater detail in Reference 2.3 and was reviewed only briefly in the preceding
section. Having said that the Ritz type procedures determine strains, it follows
that the displacement fields are then constructed from this in an integral
sense. The stiffness equation actually reflecting this integration process and
the continuity of fields across element boundaries and suppression of the field
values at domain edges being reflected by the imposition of boundary conditions.
It must therefore be argued that displacements are more accurate than strains
because integrals of smooth functions are generally more accurate than the
original data. We have thus turned the whole argument on its head.

2.8 Conclusions

In this chapter, we postulated a few models to explain how the displacements


type fem works. We worked out a series of simple problems of increasing
complexity to establish that our conjecture that strains and stresses appear in
a "best-fit" sense could be verified (falsified, in the Popperian sense) by
carefully designed numerical experiments.

An important part of this exercise depended on our careful choice and use
of various stress terms. Thus terms like σ and σf was actual physical states
that were sought to be modeled. The stress terms σ r and σ f were the quantities
that emerged in what we can call the "first-order tradition" analysis in the
language of Sir Karl Popper [2.10] - where the RR or fem operations are
mechanically carried out using functional approximation and finite element
stiffness equations respectively. We noticed certain features that seemed to
relate these computed stresses to the true system they were modeling in a
predictable or repeatable manner. We then proposed a mechanism to explain how

25
this could have taken place. Our bold conjecture, after examining these
numerical experiments, was to propose that it is effectively seeking a best-fit
state.

To confirm that this conjecture is scientifically coherent and complete,


we had to enter into a "second-order tradition" exercise. We assumed that this
is indeed the mechanism that is operating behind the scenes and derived
predicted quantities σ r that will result from the best-fit paradigm when this
was applied to the true state of stress. These predicted quantities turned out
to be exactly the same as the quantities computed by the RR and fem procedures.
In this manner, we could satisfy ourselves that the "best-fit" paradigm had
successfully survived a falsification test.

Another important step we took was to prove that the "best-fit" paradigm
was neither gratuitous nor fortuitous. In fact, we could also establish that
this could be derived from more basic principles - in this regard, the
generalized theorem of Hu was found valuable to determine that the best-fit
paradigm had a rational basis. In subsequent chapters, we shall use this
foundation to examine other features of the finite element method.

One important conclusion we can derive from the best-fit paradigm is that,
an interpolation field for the stresses σ (or stress resultants as the case may
be) which is of higher order than the strain fields ε. On which it must `do
work' in the energy or virtual work principle is actually self-defeating because
the higher order terms cannot be `sensed'. This is precisely the basis for de
Veubeke's famous limitation principle [2.9], that ‘it is useless to look for a
better solution by injecting additional degrees of freedom in the stresses.’ We
can see that one cannot get stresses, which are of higher order than are
reflected in the strain expressions.

2.9 References

2.1. T. S. Kuhn, The Structure of Scientific Revolution, University of Chicago


Press, 1962.

2.2. S. Dasgupta, Understanding design: Artificial intelligence as an


explanatory paradigm, SADHANA, 19, 5-21, 1994.

2.3. J. Barlow, Optimal stress locations in finite element models, Int. J. Num.
Meth. Engng. 10, 243-251 (1976).

2.4. R. H. MacNeal, Finite Elements: Their Design and Performance, Marcel


Dekker, NY, 1994.

2.5 T. Moan, On the local distribution of errors by finite element


approximations, Theory and Practice in Finite Element Structural Analysis.
Proceedings of the 1973 Tokyo Seminar on Finite Element Analysis, Tokyo,
43-60, 1973.

2.6. G. Prathap, The Finite Element Method in Structural Mechanics, Kluwer


Academic Press, Dordrecht, 1993.

26
2.7. G. Prathap, A variational basis for the Barlow points, Comput. Struct. 49,
381-383, 1993.

2.8. H. C. Hu, On some variational principles in the theory of elasticity and


the theory of plasticity, Scientia Sinica, 4, 33-54 (1955).

2.9. B. F. de Veubeke, Displacement and equilibrium models in the finite element


method, in Stress Analysis, Ellis Horwood, England, 1980.

2.10. B. Magee, Popper, Fontana Press, London, 1988.

27
Chapter 3

Completeness and continuity: How to choose shape functions?


3.1 Introduction

We can see from our study so far that the quality of approximation we can
achieve by the RR or fem approach depends on the admissible assumed trial, field
or shape functions that we use. These functions can be chosen in many ways. An
obvious choice, and the one most universally preferred is the use of simple
polynomials. It is possible to use other functions like trigonometric functions
but we can see that the least squares or best-fit nature of stress prediction
that the finite element process seeks in an instinctive way motivates us to
prefer the use of polynomials. When this is done, interpretations using Legendre
polynomials become very convenient. We shall see this happy state of affairs
appearing frequently in our study here.

An important question that will immediately come to mind is - How does the
accuracy or efficiency of the approximate solution depends on our choice of
shape functions. Accuracy is represented by a quality called convergence. By
convergence, we mean that as we add more terms to the RR series, or as we add
more nodes and elements into the mesh that replaces the original structure, the
sequence of trial solutions must approach the exact solution. We want quantities
such as displacements, stresses and strain energies to be exactly recovered,
surely, and if possible, quickly. This leads to the questions: What kind of
assumed pattern is best for our trial functions? What are the minimum qualities,
or essences or requirements that the finite element shape functions must show or
meet so that convergence is assured. Two have been accepted for a long time:
continuity (or compatibility) and completeness.

3.2 Continuity

This is very easy to visualize and therefore very easy to understand. A


structure is sub-divided into sub-regions or elements. The overall deformation
of the structure is built-up from the values of the displacements at the nodes
that form the net or grid and the shape functions within elements. Within each
element, compatibility of deformation is assured by the fact that the choice of
simple polynomial functions for interpolation allows continuous representation
of the displacement field. However, this does not ensure that the displacements
are compatible between element edges. So special care must be taken otherwise,
in the process of representation, gaps or overlaps will develop.

The specification of continuity also depends on how strains are defined in


terms of derivatives of the displacement fields. We know that a physical problem
can be described by the stationary condition δΠ=0, where Π=Π(φ) is the
functional. If Π contains derivatives of the field variable φ to the order m,
then we obviously require that within each element, φ , which is the approximate
field chosen as trial function, must contain a complete polynomial of degree m.
So that φ is continuous within elements and the completeness requirements (see
below) are also met. However, the more important requirement now is that
continuity of field variable φ must be maintained across element boundaries -

28
this requires that the trial function φ and its derivatives through order m-1
must be continuous across element edges. In most natural formulations in solid
mechanics, strains are defined by first derivatives of the displacement fields.
In such cases, a simple continuity of the displacement fields across element
0
edges suffices - this is called C continuity. Compatibility between adjacent
0
elements for problems which require only C -continuity can be easily assured if
the displacements along the side of an element depend only on the displacements
specified at all that nodes placed on that edge.

There are problems, as in the classical Kirchhoff-Love theories of plates


and shells, where strains are based on second derivatives of displacement
fields. In this case, continuity of first derivatives of displacement fields
1
across element edges is demanded; this is known as C -continuity. Elements,
which satisfy the continuity conditions, are called conforming or compatible
elements.

We shall however find a large class of problems (Timoshenko beams, Mindlin


plates and shells, plane stress/strain flexure, incompressible elasticity) where
this simplistic view of continuity does not assure reasonable (i.e. practical)
rates of convergence. It has been noticed that formulations that take liberties,
e.g. the non-conforming or incompatible approaches, significantly improve
convergence. This phenomenon will be taken up for close examination in a later
chapter.

3.3 Completeness

We have understood so far that in the finite element method, or for that matter,
in any approximate method, we are trying to replace an unknown function φ(x),
which is the exact solution to a boundary value problem over a domain enclosed
by a boundary by an approximate function φ (x ) which is constituted from a set of
trial, shape or basis functions. We desire a trial function set that will ensure
that the approximation approaches the exact solution as the number of trial
functions is increased. It can be argued that the convergence of the trial
function set to the exact solution will take place if φ (x ) will be sufficient to
represent any well behaved function such as φ(x) as closely as possible as the
number of functions used becomes indefinitely large. This is called the
completeness requirement.

In the finite element context, where the total domain is sub-divided into
smaller sub-regions, completeness must be assured for the shape functions used
within each domain. The continuity requirements then provide the compatibility
of the functions across element edges.

We have also seen that what we seek is a best-fit arrangement in some


sense between φ (x ) and φ(x). From Chapter 2, we also have the insight that this
best-fit can be gainfully interpreted as taking place between strain or stress
quantities. This has important implications in further narrowing the focus of
the completeness requirement for finite element applications in particular.

By bringing in the new perspective of a best-fit strain or stress


paradigm, we are able to look at the completeness requirements entirely from

29
what we desire the strains or stresses to be like. Of paramount importance now
is the idea that the approximate strains derived from the set of trial functions
chosen must be capable in the limit of approximation (i.e. as the number of
terms becomes very large) to describe the true strain or stress fields exactly.
This becomes very clear from the orthogonality condition derived in Equation
2.25.

From the foregoing discussion, we are now in a position to make a more


meaningful statement about what we mean by completeness. We would like
displacement functions to be so chosen that no straining within an element takes
place when nodal displacements equivalent to a rigid body motion of the whole
element are applied. This is called the strain free rigid body motion condition.
In addition, it will be necessary that each element must be able to reproduce a
state of constant strain; i.e. if nodal displacements applied to the element are
compatible with a constant strain state, this must be reflected in the strains
computed within the element. There are simple rules that allow these conditions
to be met and these are called the completeness requirements. If polynomial
trial functions are used, then a simple assurance that the polynomial functions
0
contain the constant and linear terms, etc. (e.g. 1, x in a one-dimensional C
0
problem;) 1, x, y in a two-dimensional C problem so that each element is
certain to be able to recover a constant state of strain, will meet this
requirement.

3.3.1 Some properties of shape functions for C0 elements

The displacement field is approximated by using shape functions Ni within each


element and these are linked to nodal degrees of freedom ui. It is assumed that
these shape functions allow elements to be connected together without generating
gaps or overlaps, i.e. the shape functions satisfy the continuity requirement
described above. Most finite element formulations in practical use in general
0
purpose application require only what is called C continuity - i.e. a field
quantity φ must have interelement continuity but does not require the continuity
of all of its derivatives across element boundaries.
0
C shape functions have some interesting properties that derive from the
completeness requirements imposed by rigid body motion considerations. Consider
a field φ (usually a displacement) which is interpolated for an n-noded element
according to
n
φ= å N iφi (3.1)
i=1

where the Ni are assumed to be interpolated using natural co-ordinates (e.g. ξ,


η for a 2-dimensional problem). It is simple to argue that one property the
shape functions must have is that Ni must define the distribution of φ within
the element domain when the degree of freedom φi at node i has unit value and
all other nodal φ’s are zero. Thus

i) Ni = 1 when x = xi (or ξ = ξi) and Ni = 0 when x = xj (or ξ = ξj) for i ≠ j.

30
Consider now, the case of the element being subjected to a rigid body motion so
that at each node a displacement ui=1 is applied. It is obvious from physical
considerations that every point in the element must have displacements u=1. Thus
from Eq 3.1 we have

ii) Σ Ni = 1.

This offers a very simple check for shape functions. Another useful check comes
again from considerations of rigid body motion - that a condition of zero strain
must be produced when rigid body displacements are applied to the nodes. The
reader is encouraged to show that this entails conditions such as

iii) Σ Ni,ξ = 0, Σ Ni,η = 0

for a 2-dimensional problem.

Note that for C elements, where derivatives of φ are also used as nodal degrees
1

of freedom, these rules apply only to those Ni which are associated with the
translational degrees of freedom.

3.4 A numerical experiment regarding completeness

Let us now return to the uniform bar carrying an axial load P as shown in
Fig. 2.1. The exact solution for this problem is given by a state of constant
strain. Thus, the polynomial series u (x ) = a0 + a1x (see Section 2.3) will contain
enough terms to ensure completeness. We could at the outset arrange for a0 to be
zero to satisfy the essential boundary condition. With u1 (x ) = a1x we were able
to obtain the exact solution to the problem. We can now investigate what will
happen if we omit the a1x term and start out with the trial function set
u2 (x ) = a2 x 2 . It is left to the reader to show that this results in a computed
stress variation σ 2 = 3Px 2EAL ; i.e. a linear variation is seen instead of the
actual constant state of stress. The reader is now encouraged to experiment with
RR solutions based on other incomplete sets like u3 (x ) = a3 x 3 and
u2 +3 (x ) = a2 x 2
+ a3 x . Figure 3.1 shows how erroneous the computed stresses σ2, σ3
3

and σ2+3 are.

The reason for this is now apparent. The term that we have omitted is the
one needed to cover the state of constant strain. In this case, this very term
also provides the exact answer. The omission of this term has led to the loss of
completeness. In fact, the reader can verify that as long as this term is left
out of any trial set containing polynomials of higher order, there will not be
any sort of clear convergence to the true state of strain. In the present
example, because of the absence of the linear term in u (the constant strain
term therefore vanishes), the computed strain vanishes at the origin!

31
Fig. 3.1 Stresses from RR solutions based on incomplete polynomials

3.5 Concluding remarks

We have now laid down some specific qualities (continuity and completeness) that
trial functions must satisfy to ensure that approximate solutions converge to
correct results. It is also hoped that if these requirements are met, solutions
from coarse meshes will be reasonably accurate and that convergence will be
acceptably rapid with mesh refinement. This was the received wisdom or reigning
paradigm around 1977. In the next chapter, we shall look at the twin issues of
errors and convergence from this point of view.

Unfortunately, elements derived rigorously from only these basic paradigms


can behave in unreasonably erratic ways in many important situations. These
difficulties are most pronounced in the lowest order finite elements. The
problems encountered were called locking, parasitic shear, etc. Some of the
problems may have gone unrecorded with no adequate framework or terminology to
classify them. As a very good example, for a very long time, it was believed
that curved beam and shell elements performed poorly because they could not meet
the strain-free rigid body motion condition. However, more recently, the correct
error inducing mechanism has been discovered and these problems have come to be
called membrane locking. An enlargement of the paradigm is now inevitable to
take these strange features into account. This shall be taken up from Chapter 5.

32
Chapter 4

Convergence and Errors


4.1 Introduction

Error analysis is that aspect of fem knowledge, which can be said to belong to
the "second-order tradition" of knowledge. To understand its proper place, let
us briefly review what has gone before in this text. We have seen structural
engineering appear first as art and technology, starting as the imitation of
structural form by human fabrication. The science of solid and structural
mechanics codified this knowledge in more precise terms. Mathematical models
were then created from this to describe the structural behavior of simple and
complex systems analytically and quantitatively. With the emergence of cheap
computational power, techniques for computational simulation emerged. The
finite element method is one such device. It too grew first as art, by
systematic practice and refinement. In epistemological terms, we can see this
as a first-order tradition or level or inner loop of fem art and practice.
Here, we know what to do and how to do it. However, the need for an outer loop
or second-order tradition of enquiry becomes obvious. How do we know why we
should do it this way? If the first stage is the stage of action, this second
stage is now of reflection. Error analysis therefore belongs to this tradition
of epistemological enquiry.

Let us now translate what we mean by error analysis into simpler terms in
the present context. We must first understand that the fem starts with the
basic premises of its paradigms, its definitions and its operational procedures
to provide numerical results to physical problems which are already described
by mathematical models that make analytical quantification possible. The answer
may be wrong because the mathematical model wrongly described the physical
model. We must take care to understand that this is not the issue here. If the
mathematical model does not explain the actual physical system as observed
through carefully designed empirical investigations, then the models must be
refined or revised so that closer agreement with the experiment is obtained.
This is one stage of the learning process and it is now assumed that over the
last few centuries we have gone through this stage successfully enough to
accept our mathematical models without further doubt or uncertainty.

Since our computational models are now created and manipulated using
digital computers, there are errors which occur due to the fact that
information in the form of numbers can be stored only to a finite precision
(word length as it is called) at every stage of the computation. These are
called round-off errors. We shall assume here that in most problems we deal
with, word length is sufficient so that round-off error is not a major
headache.

The real issue that is left for us to grapple with is that the
computational model prepared to simulate the mathematical model may be faulty
and can lead to errors. In the process of replacing the continuum region by
finite elements, errors originate in many ways. From physical intuition, we can
argue that this will depend on the type and shape of elements we use, the
number of elements used and the grading or density of the mesh used, the way
distributed loads are assigned to nodes, the manner in which boundary

33
conditions are modeled by specification of nodal degrees of freedom, etc. These
are the discretization errors that can occur.

Most of such errors are difficult to quantify analytically or determine


in a logically coherent way. We can only rely on heuristic judgement to
understand how best to minimize errors. However, we shall now look only at that
category of discretization error that appears because the computational or
discretized model uses trial functions, which are an approximation of the true
solution to the mathematical model. It seems possible that to some extent,
analytical quantification of these errors is possible.

We can recognize two kinds of discretization error belonging to this


category. The first kind is that which arises because a model replaces a
problem with an infinitely large number of degrees of freedom with a finite
number of degrees of freedom. Therefore, except in very rare cases, the
governing differential equations and boundary conditions are satisfied only
approximately. The second kind of error appears due to the fact that by
overlooking certain essential requirements beyond that specified by continuity
and completeness, the mathematical model can alter the physics of the problem.
In both cases, we must be able to satisfy ourselves that the discretization
process, which led to the computational model, has introduced a certain
predictable degree of error and that this converges at a predictable rate, i.e.
the error is removed in a predictable manner as the discretization is improved
in terms of mesh refinement. Error analysis is the attempt to make such
predictions a priori, or rationalize the errors in a logical way, a posteriori,
after the errors are found.

To carry out error analysis, new procedures have to be invented. These


must be set apart from the first-order tradition procedures that carry out the
discretization (creating the computational model from the mathematical model)
and solution (computing the approximate results). Thus, we must design
auxiliary procedures that can trace errors in an a priori fashion from basic
paradigms (conjectures or guesses). These error estimates or predictions can be
seen as consequences computed from our guesses about how the fem works. These
errors must now be verified by constructing simple digital computation
exercises. This is what we seek to do now. If this cycle can be completed, then
we can assure ourselves that we have carved out a scientific basis for error
analysis. This is a very crucial element of our study. The fem, or for that
matter, any body of engineering knowledge, or engineering methodology, can be
said to have acquired a scientific basis only when it has incorporated within
itself, these auxiliary procedures that permit its own self-criticism.
Therefore, error analysis, instead of being only a posteriori or post mortem
studies, as it is usually practised, must ideally be founded on a priori
projections computed from intelligent paradigms which can be verified
(falsified) by digital computation.

In this chapter, we shall first take stock of the conventional wisdom


regarding convergence. This is based on the old paradigm that fem seeks to
approximate displacements accurately. We next take note of the newly
established paradigm that the Ritz-type and fem approximations seek
strains/stresses in a `best-fit' manner. From such an interpretation, we
examine if it is possible to argue that errors, whether in displacements,
stresses or energies, due to finite element discretization must diminish
2
rapidly, at least in a (l/L) manner or better, where a large structure (domain)
of dimension L is sub-divided into elements (sub-domains) of dimension l. Thus,

34
with ten elements in a one-dimensional structure, errors must not be more than
a few percent. This is the usual range of problems where the continuity and
completeness paradigms explain completely the performance of finite elements.
In subsequent chapters, we shall discover however that a class of problems
exist where errors are much larger - the discretization fail in a dramatic
fashion. Convergence and error analysis must now be founded on a more complex
conceptual framework - new paradigms need to be introduced and falsified. This
will be postponed to subsequent chapters.

4.2 Traditional order of error analysis


The order of error analysis approach that is traditionally used is inherited
from finite difference approaches. This seems reasonable because quite often
simple finite difference and finite element approximations result in identical
equations. It is also possible with suitable interpretations to cast finite
difference approximations as particular cases of weighted residual
approximations using finite element type trial functions. However, there are
some inherent limitations in this approach that will become clear later.

It is usual to describe the magnitude of error in terms of the mesh size.


Thus, if a series of approximate solutions using grids whose mesh sizes are
uniformly reduced is available, it may be possible to obtain more information
about the exact answer by some form of extrapolation, provided there is some
means to establish the rate at which errors are removed with mesh size.

The order of error analysis proceeds from the understanding that the
finite element method seeks approximations for the displacement fields. The
errors are therefore interpreted using Taylor Series expansions for the true
displacement fields and truncations of these to represent the discretized
fields. The simple example below will highlight the essential features of this
approach.

4.2.1 Error analysis of the axially loaded bar problem

Let us now go back to the case of the cantilever bar subjected to a uniformly
distributed axial load of intensity q0 (Section 2.3). The equilibrium equation
for this problem is
u,xx + q 0 AE = 0 (4.1)

We shall use the two-node linear bar elements to model this problem. We have
seen that this leads to a solution, which gives exact displacements at the
nodes. It was also clear to us that this did not mean that an exact solution
had been obtained; in fact while the true solution required a quadratic
variation of the displacement field u(x), the finite element solution u (x) was
piecewise linear. Thus, within each element, there is some error at locations
between the nodes.

Fig. 4.1 shows the displacement and strain error in the region of an
element of length 2l placed with its centroid at xi in a cantilever bar of
length L. If for convenience we choose q 0 AE = 1 , then we can show that

u(x) = Lx − x 2 2 (4.2a)
u (x) = Lx −(l 2
+ 2xxi − xi2 ) 2 (4.2b)

35
u

∈ (x )

∈ (x )

Fig. 4.1 Displacement and strain error in a uniform bar under distributed axial
load q 0

ε(x)=L–x (4.2c)

ε(x) = L − xi (4.2d)

If we denote the errors in the displacement field and strain field by e(x) and
e’(x) respectively, we can show that

e (x ) = (x − xi + l ) (x − xi − l ) 2 (4.3a)

e ′(x ) = (x − xi ) (4.3b)

From this, we can argue that in this case, the strain error vanishes at the
element centroid, x=xi, and that it is a maximum at the element nodes. This is
clear from Fig. 4.1. It is also clear that for this problem, the displacement
error is a maximum at x=xi. What is more important to us is to derive measures
for these errors in terms of the element size l, or more usefully, in terms of
the parameter h = l L , which is the dimensionless quantity that indicates the
mesh division relative to the size of the structure. It will also be useful to
have these errors normalized with respect to typical displacements and strains
occurring in the problem. From Equations (4.2) and (4.3), we can estimate the
maximum normalized errors to be of the following orders of magnitude, where “O”
stands for "order",
e u max = O h 2( ) (4.4a)

36
e′ ε max
= O (h ) (4.4b)

Note that the order of error we have estimated for the displacement field is
that for a location inside the element as in this problem, the nodal
deflections are exact. The strain error is O(h) while the displacement error is
2
O(h ). At element centroids, the error in strain vanishes. It is not clear to us
in this analysis that the discretized strain ε (x ) is a "best-fit" of the true
strain ε (x ) . If this is accepted, then it is possible to show that the error in
2
the strain energy stored in such a situation is O(h ) as well.

It is possible to generalize these findings in an approximate or


tentative way for more complex problems discretized using elements of greater
precision. Thus, if an element with q nodes is used, the trial functions are of
degree q-1. If the exact solution requires a polynomial field of degree q at
q
least, then the computed displacement field will have an error O(h ). If strains
th
for the problem are obtained as the r derivative of the displacement field,
q-r
then the error in the strain or stress is O(h ). Of course, these measures are
tentative, for errors will also depend on how the loads are lumped at the nodes
and so on.

4.3 Errors and convergence from "best-fit" paradigm

We have argued earlier that the finite element method actually proceeds to
compute strains/stresses in a `best-fit' manner within each element. We shall
now use this argument to show how convergence and errors in a typical finite
element model of a simple problem can be interpreted. We shall see that a much
better insight into error and convergence analysis emerges if we base our
treatment on the `best-fit' paradigm.

We must now choose a suitable example to demonstrate in a very simple


fashion how the convergence of the solution can be related to the best-fit
paradigm. The bar under axial load, an example we have used quite frequently so
far, is not suitable for this purpose. A case such as the uniform bar with
linearly varying axial load modeled with two-node elements gives nodal
deflections, which are exact, even though the true displacement field is cubic
but between nodes, in each element, only a linear approximate trial function is
possible. It can actually be proved that this arises from the special
mathematical nature of this problem. We should therefore look for an example
where nodal displacements are not exact.

The example that is ideally suited for this purpose is a uniform beam
with a tip load as shown in Fig. 4.2. We shall model it with linear (two-node)
Timoshenko beam elements which represent the bending moment within each element
by a constant. Since the bending moment varies linearly over the beam for this
problem, the finite element will replace this with a stairstep approximation.
Thus, with increase in number of elements, the stress pattern will approach the
true solution more closely and therefore the computed strain energy due to
bending will also converge. Since the applied load is at the tip, it is very
easy to associate this load with the deflection under the load using
Castigliano's theorem. It will then be possible to discern the convergence for
the tip deflection. Our challenge is therefore to see if the best-fit paradigm
can be used to predict the convergence rate for this example from first
principles.

37
Fig. 4.2 Cantilever beam under tip load

4.3.1 Cantilever beam idealized with linear Timoshenko beam elements

The dimensions of a cantilever under tip load (Fig. 4.2) are chosen such that
the tip deflection under the load will be w=4.0. The example chosen represents
a thin beam so that the influence of shear deformation and shear strain energy
is negligible.

We shall now discretize the beam using equal length linear elements based
on Timoshenko theory. We use this element instead of the classical beam element
for several reasons. This element serves to demonstrate the features of shear
locking which arise from an inconsistent definition of the shear strains (It is
therefore useful to take up this element for study later in Chapter 6). After
correcting for the inconsistent shear strain, this element permits constant
bending and shears strain accuracy within each element - the simplest
representation possible under the circumstances and therefore an advantage in
seeing how it works in this problem.

We shall restrict attention to the bending moment variation as we assume


that the potential energy stored is mainly due to bending strain and that we
can neglect the transverse shear strain energy for the dimensions chosen.

Figure 4.3 shows the bending moment diagrams for a 1, 2 and 4 element
idealizations of the present problem using the linear element. The true bending
moment (shown by the solid line) varies linearly. The computed (i.e.
discretized) bending moments are distributed in a piecewise constant manner as
shown by the broken lines. In each case, the elements pick up the bending
moment at the centroid correctly - i.e. it is doing so in a `best-fit' manner.
What we shall now attempt to do for this problem is to relate this to the
accuracy of results. We shall now interpret accuracy in the conventional
sense, as that of the deflection at the tip under the load. Table 4.1 shows the

Table 4.1 Normalized tip deflections of a thin cantilever beam, L/t = 100

No. of elements 1 2 3
Predicted rate .750 .938 .984
Element without RBF .750 .938 .984
Element with RBF 1.000 1.000 1.000

38
normalized tip deflection with increasing idealizations (neglecting a very
small amount due to shear deformation). An interesting pattern emerges. If
error is measured by the norm {w − w (fem ) } w , it turns out that this is given
exactly by the formula 1 4N 2 where N is the number of elements. It can now be
seen that this relationship can be established by arguing that this feature
emerges from the fact that strains are sought in the `best-fit' manner shown in
Fig. 4.3.

Consider a beam element of length 2l (Fig. 4.4). Let the moment and shear force
at the centroid be M and V. Thus the true bending moment over the element
region for our problem can be taken to vary as M+Vx. The discretized bending
moment sensed by our linear element would therefore be M. We shall now compute
the actual bending energy in the element region (i.e. from a continuum
analysis) and that given by the finite element (discretized) model. We can show
that

Fig. 4.3 Bending moment diagrams for a one-, two- and four-element dealizations
of a cantilever beam under tip load.

39
Energy in continuum model = (l EI) (M 2 + V 2l 2 3) (4.5)
Energy in discretized model = (l EI) (M 2 ) (4.6)

Thus, as a result of the discretization process involved in replacing each


continuum segment of length 2l by a linear Timoshenko beam element which can
give only a constant value M for the bending moment, there is a reduction
2 2
(error) in energy in each element equal to (l/EI) (V l /3). It is simple now to
show from this that for the cantilever beam of length L with a tip load P, the
2
total reduction in strain energy of the discretized model for the beam is U/4N
2 3
where U=P L /6EI is the energy of the beam under tip load.

We are interested now to discover how this error in strain energy


translates into an error in the deflection under load P. This can be very
easily deduced using Castigliano's second theorem. It is left to the reader to
show that the tip deflections of the continuum and discretized model will
differ as {w − w (fem ) } w = 1 4N 2 .

Table 4.1 shows this predicted rate of convergence. Our foregoing


analysis shows that this follows from the fact that if any linear variation is
approximated in a piecewise manner by constant values as seen in Fig. 4.3, this
is the manner in which the square of the error in the stresses/strains (or,
equivalently, the difference in work or energy) will converge with
idealization. Of course, in a problem where the bending moment is constant, the
rate of convergence will be better than this (in fact, exact) and in the case
where the bending moment is varying quadratically or at a higher rate, the rate
of convergence will be decidedly less.

Fig. 4.4 Bending moment variation in a linear beam element.

40
We also notice that convergence in this instance is from below. This can
be deduced from the fact that the discretized potential energy U is less than
the actual potential energy U for this problem. It is frequently believed that
the finite element displacement approach always underestimates the potential
energy and a displacement solution is consequently described as a lower bound
solution. However, this is not a universally valid generalization. We can see
briefly below (the reader is in fact encouraged to work out the case in detail)
where the cantilever beam has a uniformly distributed load acting on it using
the same linear Timoshenko beam element for discretization that this is not the
case. It turns out that tip rotations converge from above (in a 1 2N 2 rate)
while the tip deflections are fortuitously exact. The lower bound solution
nature has been disturbed because of the necessity of altering the load system
at the nodes of the finite element mesh under the `consistent load' lumping
procedure.

As promised above, we now extend the concept of "best-fit" and


variationally correct rate of convergence to the case of uniformly distributed
load of intensity q on the cantilever with a little more effort. Now, when a
finite element model is made, two levels of discretization error are
introduced. Firstly, the uniformly distributed load is replaced by consistent
loads, which are concentrated at element nodes. Thus, the first level of
discretization error is due to the replacement of the quadratically varying
bending moment in the actual beam with a linear bending moment within each beam
element. Over the entire beam model, this variation is piecewise linear. The
next level of error is due to the approximation implied in developing the
stiffness matrix which we had considered above this effectively senses a "best-
approximated" constant value of the bending moment within each element of the
linear bending moment appearing to act after load discretization.

With these assumptions, it is a simple exercise using Castigliano's


theorem and fictitious tip force and moment P and M respectively to demonstrate
that the finite element model of such a problem using two-noded beam elements
will yield a fortuitously correct tip deflection (
w = qL4 8EI) for all
idealizations (i.e. even with one element!) and a tip rotation that converges
( )
at the rate 1 2N 2 from above to the exact value θ = qL3 6EI . Thus, as far as
tip deflections are concerned, the two levels of discretization errors have
nicely cancelled each other to give correct answers. This can deceive an unwary
analyst into believing that an exact solution has been reached. Inspection of
the tip rotation confirms that the solution is approximate and converging.

We see from the foregoing analysis that using linear Timoshenko beam
elements for the tip loaded cantilever, the energy for this problem converges
2
as O(h ) where h=2l=L/N is the element size. We also see that this order of
convergence carries over to the estimate of the tip deflections for this
problem. Many text-books are confused over such relationships, especially those
that proceed on the order of error analysis. These approaches arrive at
conclusions such as strain error is proportional to element size, i.e. O(h) and
2
displacement error proportional to the square of the element size, i.e. O(h )
for this problem. We can see that for this problem (see Fig. 4.3) this estimate
is meaningful if we consider the maximum error in strain to occur at element
nodes (at centroids the errors are zero as these are optimal strain points). We
also see that with element discretization, these errors in strain vanish as

41
O(h). We can also see that the strain energies are now converging at the rate
2
of O(h ) and this emerges directly from the consideration that the discretized
strains are `best-fits' of the actual strain. This conclusion is not so readily
arrived at in the order of error analysis methods, which often argue that the
2
strains are accurate to O(h), then strain energies are accurate to O(h ) because
strain energy expressions contain squares of the strain. This conclusion is
valid only for cases where the discretized strains are `best-fit'
approximations of the actual strains, as observed in the present example. If
the `best-fit' paradigm did not apply, the only valid conclusion that could be
drawn is that the strains that have O(h) error will produce errors in strain
energy that are O(2h).

4.4 The variationally correct rate of convergence

It is possible now to postulate that if finite elements are developed in a


clean and strictly "legal" manner, without violating the basic energy or
variational principles, there is a certain rate at which solutions will
converge reflecting the fidelity that the approximate solution maintains with
the exact solution. We can call this the variationally correct rate of
convergence. However, this prescription of strictly legal formulation is not
always followed. It is not uncommon to encounter extra-variational devices that
are brought in to enhance performance.

4.5 Residual bending flexibility correction

The residual bending flexibility (RBF) correction [4.1,4.2] is a device used to


enhance the performance of 2-node beam and 4-node rectangular plate elements
(these are linear elements) so that they achieve results equivalent to that
0
obtained by elements of quadratic order. In these C elements (Timoshenko theory
for beam and Mindlin theory for plate) there is provision for transverse shear
strain energy to be computed in addition to the bending energy (which is the
1
only energy present in the C Euler-Bernoulli beam and Kirchhoff plate
formulations). The RBF correction is a deceptively simple device that enhances
performance of the linear elements by modifying the shear energy term to
compensate for the deficient bending energy term so that superior performance
is obtained. To understand that this is strictly an extra-variational trick (or
"crime"), it is necessary to understand that a variationally correct procedure
will yield linear elements that have a predictable and well-defined rate of
convergence. This has been carried out earlier in this chapter. It is possible
to improve performance beyond this limit only by resorting to some kind of
extra-variational manipulation. By a variationally correct procedure, we mean
that the stiffness matrix is derived strictly using a BTDB triple product, where
B and D are the strain-displacement and stress-strain matrices respectively.

4.5.1 The mechanics of the RBF correction.

MacNeal [4.2] describes the mathematics of the RBF correction using an original
cubic lateral displacement field (corresponding to a linearly varying bending
moment field) and what is called a aliased linearly interpolated field (giving
a constant bending moment variation for the discretized field). We might recall
that we examined the aliasing paradigm in Section 2.5 earlier. We find that a
quicker and more physically insightful picture can be obtained by using
Equations (4.5) and (4.6) above. Consider a case where only one element of

42
length 2l is used to model a cantilever beam of length L with a tip load P.
Applying Castigliano's theorem, we can show very easily that the continuum and
discretized solutions will differ by,

(
w L = P L2 3EI + 1 kAG ) (4.7)

w L = P (L2
4EI + 1 kAG ) (4.8)
Note the difference in bending flexibilities. This describes the inherent
approximation involved in the discretization process if all variational norms
are strictly followed. The RBF correction proposes to manipulate the shear
flexibility in the discretized case so that it compensates for the deficiency
*
in the bending flexibility for this problem. Thus if k is the compensated shear
correction factor, from Eq (4.7) and (4.8), we have

L2 3EI + 1 kAG = L2 4EI + 1 k* AG

1 k* AG = 1 kAG + l 2 3EI (4.9)

It is not accurate to say here that the correction term is derived from the
bending flexibility. The bending flexibility of an element that can represent
only a constant bending moment (e.g. the two-noded beam element used here) is
L2 4EI . The missing L2 12EI (or l 2 3EI ) is now brought in as a compensation
*
through the shear flexibility, i.e. k is changed to k . This "fudge factor"
therefore enhances the performance of the element by giving it a rate of
convergence that is not variationally permissible. Two wrongs make a right
here; or do they? Table 4.1 shows how the convergence in the problem shown in
Fig. 4.2 is changed by this procedure.

4.6 Concluding remarks.

We have shown that the best-fit paradigm is a useful starting point for
deriving estimates about errors and convergence. Using a simple example and
this simple concept, we could establish that there is a variationally correct
rate of convergence for each element. This can be improved only by taking
liberties with the formulation, i.e. by introducing extra-variational steps.
The RBF is one such extra-variational device.

It is also possible to argue that errors, whether in displacements, stresses or


energies, due to finite element discretization must converge rapidly, at least
2
in a O(h ) manner or better. If a large structure (domain) of dimension L is
sub-divided into elements (sub-domains) of dimension l, one expects errors of
the order of (l L )2 . Thus, with ten elements in a one-dimensional structure,
errors must not be more than a few percent. We shall discover however that a
class of problems exist where errors are much larger - the discretizations fail
in a dramatic fashion, and this cannot be resolved by the classical (pre-1977)
understanding of the finite element method. A preliminary study of the issues
involved will be taken up in the next chapter; the linear Timoshenko beam
element serves to expose the factors clearly. Subsequent chapters will
undertake a comprehensive review of the various manifestations of such errors.
It is felt that this detailed treatment is justified, as an understanding of

43
such errors has been one of the most challenging problems that the finite
element procedure has faced in its history. Most of the early techniques to
overcome these difficulties were ad hoc, more in the form of `art' or `black
magic'. In the subsequent chapters, our task will be to identify the principles
that establish the methodology underlying the finite element procedure using
critical, rational scientific criteria.

4.7 References

4.1. R. H. MacNeal, A simple quadrilateral plate element, Comput. Struct., 8,


175-183, 1978.
4.2. R. H. MacNeal, Finite Elements: Their Design and Performance, Marcel
Dekker, NY, 1994.

44
Chapter 5

The locking phenomena


5.1 Introduction

A pivotal area in finite element methodology is the design of robust and


accurate elements for applications in general purpose packages (GPPs) in
structural analysis. The displacement type approach we have been examining so
far is the technique that is overwhelmingly favored in general purpose software.
We have seen earlier that the finite element method can be interpreted as a
piecewise variation of the Rayleigh-Ritz method and therefore that it seeks
strains/stresses in a `best-fit' manner. From such an interpretation, it is
possible to argue that errors, whether in displacements, stresses or energies,
due to finite element discretization must converge rapidly, at least in a (l L )2
manner or better, where a large structure (domain) of dimension L is sub-divided
into elements (sub-domains) of dimension l. Thus, with ten elements in a one-
dimensional structure, errors must not be more than a few percent.

By and large, the elements work without difficulty. However, there were
spectacular failures as well. These are what are now called the ‘locking’
0
problems in C finite elements - as the element libraries of GPPs stabilized
these elements came to be favored for reasons we shall discuss shortly. By
locking, we mean that finite element solutions vanish quickly to zero (errors
saturating quickly to nearly 100%!) as certain parameters (the penalty
multipliers) become very large. It was not clear why the displacement type
method, as it was understood around 1977, should produce for such problems,
answers that were only a fraction of a per cent of the correct answer with a
practical level of discretization. Studies in recent years have established that
an aspect known as consistency must be taken into account.

The consistency paradigm requires that the interpolation functions chosen


to initiate the discretization process must also ensure that any special
constraints that are anticipated must be allowed for in a consistent way.
Failure to do so causes solutions to lock to erroneous answers. The paradigm
showed how elements can be designed to be free of these errors. It also enabled
error-analysis procedures that allowed errors to be traced to the
inconsistencies in the representation to be developed. We can now develop a
family of such error-free robust elements for applications in structural
mechanics.

In this chapter, we shall first introduce the basic concepts needed to


understand why such discretized descriptions fail while others succeed. We
0
compare the equations of the Timoshenko beam theory (a C theory) to the
1
classical beam theory (a C theory) to show how constraints are generated in
such a model. This permits us to discuss the concept of consistency and the
nature of errors that appear during a Ritz type approximation. These same errors
are responsible for the locking seen in the displacement type finite element
models of similar problems.

45
5.2 From C1 to C0 elements

As the second generation of GPPs started evolving around the late 70s and early
1
80s, their element libraries replaced what were called the C elements with what
0
were known as the C elements. The former were based on well known classical
theories of beams, plates and shells, reflecting the confidence structural
analysts had in such theories for over two centuries - namely the Euler-
Bernoulli beam theory, the Kirchhoff-Love plate theory and the equivalent shell
theories. These theories did not allow for transverse shear strain and permitted
the modeling of such structures by defining deformation in terms of a single
field, w, the transverse deflection of a point on what is called the neutral
axis (in a beam) and neutral surface of a plate or shell. The strains could then
be computed quite simply from the assumption that normal to the neutral surface
remained normal after deformation. One single governing differential equation
resulted, although of a higher order (in comparison to other theories we shall
discuss shortly), and this was considered to be an advantage.

There were some consequences arising from such an assumption both for the
mathematical modeling aspect as well as for the finite element (discretization)
aspect. In the former, it turned out that to capture the physics of deformation
of thick or moderately thick structures, or the behavior of plates and shells
made of newly emerging materials such as high performance laminated composites,
it was necessary to turn to more general theories accounting or transverse shear
deformation as well - these required the definition of rotations of normals
which were different from the slopes of the neutral surface. Some of the
1
contradictions that arose as a result of the old C theories - e.g. the use of
the fiction of the Kirchhoff effective shear reactions, could now be removed,
restoring the more physically meaningful set of three boundary conditions on the
edge of a plate or shell (the Poisson boundary conditions as they are called) to
be used. The orders of the governing equations were correspondingly reduced. A
salutary effect that carried over to finite element modeling was that the
elements could be designed to have nodal degrees of freedom which were the six
basic engineering degrees of freedom - the three translations and the three
rotations at a point. This was ideal from the point of view of the organization
of a general purpose package. Also, elements needed only simple basis functions
requiring only the continuity of the fields across element boundaries - these
0 1
are called the C requirements. In the older C formulations, continuity of
slope was also required and to achieve this in arbitrarily oriented edges, as
would be found in triangular or quadrilateral planforms of a plate bending
element, it was necessary to retain curvature degrees of freedom (w,xx, w,xy,
w,yy) at the nodes and rely on quintic polynomials for the element shape or
basis functions. So, as general purpose packages ideal for production run
0
analyses and design increasingly found favour in industry, the C beam, plate
1
and shell elements slowly began to replace the older C equivalents. It may be
instructive to note that the general two-dimensional (i.e. plane stress, plane
strain and axisymmetric) elements and three-dimensional (solid or brick as they
0
are called) elements were in any case based on C shape functions - thus this
0
development was welcome in that universally valid C shape functions and their
derivatives could be used for a very wide range of structural applications.

46
However, life was not very simple - surprisingly dramatic failures came to
be noticed and the greater part of academic activity in the late seventies, most
of the eighties and even in the nineties was spent in understanding and
eliminating what were called the locking problems.

5.3 Locking, rank and singularity of penalty-linked stiffness matrix,


and consistency of strain-field

When locking was first encountered, efforts were made to associate it with the
rank or non-singularity of the stiffness matrix linked to the penalty term (e.g.
the shear stiffness matrix in a Timoshenko beam element which becomes very large
as the beam becomes very thin, see the discussion below). However, on
reflection, it is obvious that these are symptoms of the problem and not the
cause. The high rank and non-singularity is the outcome of certain assumptions
made (or not made, i.e. leaving certain unanticipated requirements unsatisfied)
during the discretization process. It is therefore necessary to trace this to
the origin. The consistency approach argues that it is necessary in such
problems to discretize the penalty-linked strain fields in a consistent way so
that only physically meaningful constraints appear.

In this section, we would not enter into a formal finite element


discretization (which would be taken up in the next section) but instead,
illustrate the concepts involved using a simple Ritz-type variational method of
approximation of the beam problem via both classical and Timoshenko beam theory
[5.1]. It is possible to show how the Timoshenko beam theory can be reduced to
the classical thin beam theory by using a penalty function interpretation and in
doing so, show how the Ritz approximate solution is very sensitive to the way in
which its terms are chosen. An `inconsistent' choice of parameters in a low
order approximation leads to a full-rank (non-singular) penalty stiffness matrix
that causes the approximate solution to lock. By making it `consistent', locking
can be eliminated. In higher order approximations, `inconsistency' does not lead
to locked solutions but instead, produces poorer convergence than would
otherwise be expected of the higher order of approximation involved. It is again
demonstrated that a Ritz approximation that ensures ab initio consistent
definition will produce the expected rate of convergence - a simple example will
illustrate this.

5.3.1 The classical beam theory

Consider the transverse deflection of a thin cantilever beam of length L under


an uniformly distributed transverse load of intensity q per unit length of the
beam. This should produce a linear shear force distribution increasing from 0 at
the free end to qL at the fixed end and correspondingly, a bending moment that
varies from 0 to qL2 2 . Using what is called the classical or Euler-Bernoulli
theory, we can state this problem in a weak form with a quadratic functional
given by,

Π =
L
ò (1 2 EI w,2xx − qw ) dx (5.1)
0

This theory presupposes a constraint condition, assuming zero transverse


shear strain, and this allows the deformation of a beam to be described entirely

47
in terms of a single dependent variable, the transverse deflection w of points
on the neutral axis of the beam. An application of the principle of minimum
total potential allows the governing differential equations and boundary
conditions to be recovered, but this will not be entered into here. A simple
exercise will establish that the exact solution satisfying the governing
differential equations and boundary conditions is,

( )
w (x ) = qL2 4EI x 2 − (qL 6EI ) x 3 − (q 24EI ) x 4 (5.2a)

EI w,xx = (q 2 ) (L − x )2 (5.2b)

EI w,xxx = − q (L − x ) (5.2c)

5.3.1.1 A two-term Ritz approximation

Let us now consider an approximate Ritz solution based on two terms,


w = b2 x + b3 x . Note that the constant and linear terms are dropped,
2 3

anticipating the boundary conditions at the fixed end. One can easily work out
that the approximate solution will emerge as,

( )
w (x ) = 5qL2 24EI x 2 − (qL 12EI ) x 3 (5.3a)

so that the approximate bending moment and shear force determined in this Ritz
process are,

(
EI w ,xx = 5qL2 12 − (qL 2 ) x ) (5.3b)

EI w ,xxx = − qL 2 (5.3c)

If the expressions in Equations (5.2) and (5.3) are written in terms of the
natural co-ordinate system ξ, where x = (1 + ξ ) L 2 so that the ends of the beam
are represented by ξ = -1 and +1, the exact and approximate solutions can be
expanded as,

( )(
EI w,xx = qL2 8 4 3 − 2ξ − 1 3 (1 − 3ξ 2 ) ) (5.4)

EI w ,xx = qL2 8 ( ) (4 3 − 2ξ ) (5.5)

The approximate solution for the bending moment is seen to be a `best-fit' or


`least-squares fit' of the exact solution, with the points ξ = ± 1 3 , which are
the points where the second order Legendre polynomial vanishes, emerging as
points where the approximate solution coincides with the exact solution. From
Equations (5.2c) and (5.3c), we see that the shear force predicted by the Ritz
approximation is a 'best-fit' of the actual shear force variation. Once again we
confirm that the Ritz method seeks the 'best-approximation' of the actual state
of stress in the region being studied.

48
5.3.2 The Timoshenko beam theory and its penalty function form

The Timoshenko beam theory [5.1] offers a physically more general formulation of
beam flexure by taking into account the transverse shear deformation. The
description of beam behavior is improved by introducing two quantities at each
point on the neutral axis, the transverse deflection w and the face rotation θ
so that the shear strain at each point is given by γ = θ − w,x , the difference
between the face rotation and the slope of the neutral axis.

The total strain energy functional is now constructed from the two
independent functions for w(x) and θ(x), and it will now account for the bending
(flexural) energy and an energy of shear deformation.

Π =
0
ò
L
(
1 2 EI θ ,2x + 1 2 α (θ − w,x )2 − q w ) dx (5.6)

where the curvature κ=θ,x, the shear strain γ=θ-w,x and α=kGA is the shear
rigidity. The factor k accounts for the averaged correction made for the shear
strain distribution through the thickness. For example, for a beam of
rectangular cross-section, this is usually taken as 5/6.

The Timoshenko beam theory will asymptotically recover the elementary beam
theory as the beam becomes very thin, or as the shear rigidity becomes very
large, i.e. α→∞. This requires that the Kirchhoff constraint θ-w,x→0 must
emerge in the limit. For a very large α, these equations lead directly to the
simple fourth order differential equation for w of elementary beam theory. Thus,
this is secured very easily in the infinitesimal theory but it is this very same
point that poses difficulties when a simple Ritz type approximation is made.

5.3.2.1 A two-term Ritz approximation

Consider now a two term Ritz approximation based on θ = a1x and w = b1x . This
permits a constant bending strain (moment) approximation to be made. The shear
strain is now given by

γ = a1x − b1 (5.7)

and it would seem that this can represent a linearly varying shear force. The
Ritz variational procedure now leads to the following set of equations,

ìï éL 0 ù é L3 3 − L2 2 ù üï ìa1 ü ìï 0 üï
íEI ê ú + α ê ú ý í ý=í 2 ý (5.8)
ïî ë0 0 û êë − L2 2 L úû þï î b1 þ ïîqL 2 ïþ

Solving, we get

− 3qL2 qL 1.5qL3
a1 = ; b1 = − − (5.9)
12EI + αL2 2α 12EI + αL2

49
As α→∞, both a1 and b1→0. This is a clear case of a solution `locking'. This
could have been anticipated from a careful examination of Equations (5.6) and
(5.7). The penalty limit α→∞ in Equation (5.6) introduces a penalty condition
on the shear strain and this requires that the shear strain must vanish in the
Ritz approximation process - from Equation (5.7), the conditions emerging are
a1→0 and b1→0. Clearly, a1→0 imposes a zero bending strain θ,x→0 as well -
this spurious constraint produces the locking action on the solution. Thus, a
meaningful approximation can be made for a penalty function formulation only if
the penalty linked approximation field is consistently modeled. We shall see how
this is done next.

5.3.2.2 A three-term consistent Ritz approximation

Consider now a three term approximation chosen to satisfy what we shall call the
consistency condition. Choose θ = a1x and w = b1x + b2 x 2 . Again, only a constant
bending strain is permitted. But we now have shear strain of linear order as,

γ = − b1 + (a1 − 2b2 )x (5.10)

Note that as α→∞, the constraint α1-2b2→0 is consistently balanced and will
not result in a spurious constraint on the bending field. The Ritz variational
procedure leads to the following equation structure:

ì é L3 3 − L2 2 − 2L3 3 ù ü ì 0 ü
ï éL 0 0 ù ê ú ï ìa1 ü
ï ê ú ï ï ï ïï 2 ïï
í EI ê0 0 0 ú + α ê − L 2 L2 ú í b1 ý = íqL 2 ý
2
L ý (5.11)
ï ê ú ï ïb ï ï 3 ï
êë0 0 0 úû ê − 2L 3
3
L2 4L3 3 ú î 2þ ïîqL 3 ïþ
ïî ë û ïþ

It can be easily worked out from this that the approximate solution is given by,

qL2
θ =− x (5.12a)
6EI

qL q 2 qL2
w = x− x + x2 (5.12b)
α 2α 12EI

− EI θ ,x = qL2 6 (5.12c)

( )
α θ − w ,x = q (L − x ) (5.12d)

There is no locking seen at all - the bending moment is now a least squares
correct constant approximation of the exact quadratic variation (this can be
seen by comparing Equation (5.12c) with Equations (5.4) and (5.5) earlier). The
shear force is now correctly captured as a linear variation - the consistently
represented field in Equation (5.12) being able to recover this even as a α→∞!

50
A comparison of the penalty linked matrices in Equations (5.8) and (5.11)
shows that while in the former, we have a non-singular matrix of rank 2, in the
latter we have a singular matrix of rank 2 for a matrix of order 3. It is clear
also that the singularity (or reduced rank) is a result of the consistent
condition represented by (α1-2b2) in the linear part of the shear strain
definition in Equation (5.10) - as a direct consequence, the third row of the
penalty linked matrix in Equation (5.11) is exactly twice the first row. It is
this aspect that led to a lot of speculation on the role the singularity or rank
of the matrix plays in such problems. We can further show that non-singularity
of penalty linked matrix arises in an inconsistent formulation only when the
order of approximation is low, as seen for the two-term Ritz approximation. We
can go for a quadratic inconsistent approximation (with linear bending strain
variation) in the next section to show that there is no locking' of the solution
and that the penalty linked matrix is not non-singular - however the effect of
inconsistency is to reduce the performance of the approximation to a sub-optimal
level.

5.3.2.3 A four-term inconsistent Ritz approximation

We now take up a four term approximation which provides theoretically for a


linear variation in the approximation for bending strain, i.e. θ = a1x + a2 x 2 and
w = b1x + b2 x 2 so that κ = θ ,x = a1 + 2a2 x and the shear strain is

γ = − b1 + (a1 − 2b2 )x + a2 x 2 (5.13)

Note now that the condition α→∞ forces a2→0 this becomes a spurious constraint
on the bending strain field. We shall now see what effect this spurious
constraint has on the approximation process. The Ritz variational procedure
leads to the following set of equations:

ì éL L2 0 0ù
ï ê ú
ï ê 2 ú
ï êL 4L3 3 0 0ú
í EI +
ê ú
ï ê0 0 0 0ú
ï ê ú
ï
î ëê0 0 0 0 ûú

(5.14)

é L3 3 L4 4 − L2 2 − 2L3 3 ù ü ì 0 ü
ê ú ï ìa1 ü
ï ï ï ï
ê L4 4 L5 5 − L3 3 − L4 2 ú ïï ïa2 ï ï 0 ï
α ê ú ý í ý = í 2 ý
ê − L2 2 − L3 3 L L2 ú ï ï b1 ï ïqL 2 ï
ê ú ï ï b2 ï ï 3 ï
î þ
ëê − 2 L 3 − L4 2 îqL 3 þ
3
L2 4L3 3 ûú ïþ

It can be seen that the penalty linked 4×4 matrix is singular as the fourth row
is exactly twice the second row - this arises from the consistent representation
(a1 − 2b2 ) of the linear part of shear strain in Equation (5.13). The rank of the

51
matrix is therefore 3 and the solution should be free of "locking" - however the
inconsistent constraint forces a2 → 0 and this means that the computed bending
strain κ → a1 ; i.e. it will have only a constant bending moment prediction
capability. What it means is that this four term inconsistent approach will
produce answers only as efficiently as the three term consistent Ritz
formulation. Indeed, one can work out from Equation (5.14) that,

qL2 15qL2
a1 = − − (5.15a)
6EI 60EI + αL2

15qL2
a2 = (5.15b)
60EI + αL2

qL 5qL3
b1 = −
α

(
2 60EI + αL2 ) (5.15c)

qL2 q
b2 = − + (5.15d)
6EI 2α

As α→∞, the bending moment and shear force variation are given by,

− EI θ ,x = qL2 6 (5.16a)

( )
α θ − w ,x = q (L − x ) + 2.5qL (6 (x L )2 − 6 (x L ) + 1 ) (5.16b)

The solution can sense only a constant bending moment in the thin beam limit.
There are now violent quadratic oscillations in the shear force and these
oscillations can be shown to vanish at the points (
x L = 1 2 1+1 3 ) and
( )
1 2 1 − 1 3 , or ξ = ± 1 3 , which are the points corresponding to the 2 point
Gauss-Legendre integration rule. The effect of the inconsistent representation
has been to reduce the effectiveness of the approximation. We shall next see how
the effectiveness can be improved by making the approximation consistent before
the variational process is carried out.

5.3.2.4 A four-term consistent Ritz approximation

Let us now take up a Ritz approximation with a consistently represented function


for the shear strain defined as γ . This can be achieved by noting that in
Equation (5.13), the offending term is the quadratic term associated with a2. We
also see from Equation (5.16b) that this leads to a spuriously excited quadratic
form (6x 2
)
− 6Lx + L2 . Our experience with consistent finite element formulations
term in Equation (5.13) with Lx − L2 6 so that,
2
[5.2] allows us to replace the x

52
γ = − b1 + (a1 − 2b2 )x + a2 Lx − L2 6( )
(
= − b1 − a2 L2 6 )+ (a1 + a2 L - 2b2 )x (5.17)

In fact, it can be proved using the generalized (mixed) variational theorem such
as the Hellinger-Reissner and the Hu-Washizu principles [5.2] that the
variationally correct way to determine γ from the inconsistent γ is,

T
ò δ γ æç γ − γ ö÷ dx = 0
è ø

and this will yield precisely the form represented in Equation (5.17). The Ritz
variational procedure now leads to,

ì éL L2 0 0ù
ï ê ú
ï ê 2 ú
ï êL 4L3 3 0 0ú
í EI ê +
ú
ï ê0 0 0 0ú
ï ê ú
ï êë0 0 0 0 úû
î
(5.18)

é L3 3 L4 4 − L2 2 − 2L3 3 ù ü ì 0 ü
ê ú ï ìa1 ü
ï ï ï ï
ê L4 4 7L5 36 − L3 3 − L4 2 ú ïï ïa2 ï ï 0 ï
α ê ú ý í ý = í 2 ý
ê − L2 2 − L3 3 L L2 ú ï ï b1 ï ïqL 2 ï
ê ú ï ï b2 ï ï 3 ï
î þ
ëê − 2 L 3 − L4 2 îqL 3 þ
3
L2 4L3 3 ûú ïþ

It is important to recognize now that since the penalty linked matrix emerges
( )
from the terms - b1 + a2 L2 6 and (a1 + a2 L - 2b2 ) there would only be two linearly
independent rows and therefore the rank of the matrix is now 2. The approximate
solution is then given by,

5 qL
− EI κ = qL2 − x (5.19a)
12 2

α γ = q (L - x ) (5.19b)

Comparing Equation (5.19a) with Equation (5.3b) we see that the approximate
solution to the Timoshenko equation for this problem with a consistent shear
strain assumption gives exactly the same bending moment as the Ritz solution to
the classical beam equation could provide. We also see that Equation (5.19b) is
identical to Equation (5.12b) so that the shear force is now being exactly
computed. In other words, as α→∞, the terms a1, a2, b1 and b1 all yield
physically meaningful conditions representing the state of equilibrium
correctly.

53
5.4 Consistency and C0 displacement type finite elements

We shall see in the chapters to follow that locking, poor convergence and
0
violent stress oscillations seen in C displacement type finite element
formulations are due to a lack of consistent definition of the critical strain
fields when the discretization is made - i.e. of the strain-fields that are
constrained in the penalty regime.

The foregoing analysis showed how the lack of consistency translates into
a non-singular matrix of full rank that causes locking in low-order Ritz
approximations of such problems. It is also seen that in higher order
approximations, the situation is not as dramatic as to be described as locking,
but is damaging as to produce poor convergence and stress oscillations. It is
easy to predict all this by examining the constrained strain-field terms from
the consistency point of view rather than performing a post-mortem examination
of the penalty-linked stiffness matrix from rank or singularity considerations
as is mostly advocated in the literature.

We shall now attempt a very preliminary heuristic definition of these


requirements as seen from the point of view of the developer of a finite element
for an application in a constrained media problem. We shall see later that if a
simple finite element model of the Timoshenko beam is made), the results are in
very great error and that these errors grow without limit as the beam becomes
very thin. This is so even when the shape functions for the w and θ have been
chosen to satisfy the completeness and continuity conditions. We saw in our Ritz
approximation of the Timoshenko beam theory in this section that as the beam
becomes physically thin, the shear strains must vanish and it must begin to
enforce the Kirchhoff constraint and that this is not possible unless the
approximate field can correctly anticipate this. In the finite element statement
of this problem, the shape functions chosen for the displacement fields cannot
do this in a meaningful manner - spurious constraints are generated which cause
locking. The consistency condition demands that the discretized strain field
interpolations must be so constituted that it will enforce only physically true
constraints when the discretized functionals for the strain energy of a finite
element are constrained.

We can elaborate on this definition in the following descriptive way. In


the development of a finite element, the field variables are interpolated using
interpolations of a certain order. The number of constants used will depend on
the number of nodal variables and any additional nodeless variables (those
corresponding to bubble functions). From these definitions, one can compute the
strain fields using the strain-displacement relations. These are obtained as
interpolations associated with the constants that were introduced in the field
variable interpolations. Depending on the order of the derivatives of each field
variable appearing in the definition of that strain field (e.g. the shear strain
in a Timoshenko theory will have θ and the first derivative of w), the
coefficients of the strain field interpolations may have constants from all the
contributing field variable interpolations or from only one or some of these. In
some limiting cases of physical behavior, these strain fields can be constrained
to be zero values, e.g. the vanishing shear strain in a thin Timoshenko beam.
Where the strain-field is such that all the terms in it (i.e. constant, linear,
quadratic, etc.) have, associated with it, coefficients from all the independent
interpolations of the field variables that appear in the definition of that

54
strain-field, the constraint that appears in the limit can be correctly
enforced. We shall call such a representation field-consistent. The constraints
thus enforced are true constraints. Where the strain-field has coefficients in
which the contributions from some of the field variables are absent, the
constraints may incorrectly constrain some of these terms. This field-
inconsistent formulation is said to enforce additional spurious constraints. For
simple low order elements, these constraints are severe enough to produce
solutions that rapidly vanish - causing what is often described as locking.

5.5 Concluding remarks

These exercises show us why it is important to maintain consistency of the basis


functions chosen for terms in a functional, which are linked, to penalty
multipliers. The same conditions are true for the various finite element
formulations where locking, poor convergence and stress oscillations are known
to appear. It is also clear why the imposition of the consistency condition into
the formulation allows the correct rank or singularity of the penalty linked
stiffness matrix to be maintained so that the system is free of locking or sub-
optimal convergence. Again, it is worthwhile to observe that non-singularity of
the penalty linked matrix occurs only when the approximate fields are of very
low order as for the two-term Ritz approximation. In higher order inconsistent
formulations, as for the four-term inconsistent Ritz approximation, solutions
are obtained which are sub-optimal to solutions that are possible if the
formulation is carried out with the consistency condition imposed a priori. We
shall see later that the use of devices such as reduced integration permits the
consistency requirement to be introduced when the penalty linked matrix is
computed so that the correct rank which ensures the imposition of the true
constraints only is maintained.

In the next chapter, we shall go to the locking and other over-stiffening


phenomena found commonly in displacement finite elements. The phenomenon of
shear locking is the most well known - we shall investigate this closely. The
Timoshenko beam element allows the problem to be exposed and permits a
mathematically rigorous error model to be devised. The consistency paradigm is
introduced to show why it is essential to remove shear locking. The correctness
concept is then brought in to ensure that the devices used to achieve a
consistent strain interpolation are also variationally correct. This example
serves as the simplest case that could be employed to demonstrate all the
principles involved in the consistency paradigm. It sets the scene for the need
for new paradigms (consistency and correctness) to complement the existing ones
(completeness and continuity) so that the displacement model can be
scientifically founded. The application of these concepts to Mindlin plate
elements is also reviewed.

The membrane locking phenomenon is investigated in Chapter 7. The simple


curved beams are used to introduce this topic. This is followed up with studies
of parasitic shear and incompressible locking.

In Chapters 5 to 7 we investigate the effect of constraints on strains due


to the physical regimes considered, e.g. vanishing shear strains in thin
Timoshenko beams or Mindlin plates, vanishing membrane strains in curved beams
and shells etc. In Chapter 8 we proceed to a few problems where consistency
between strain and stress functions are needed even where no constraints on the

55
strains appear. These examples show the universal need of the consistency aspect
and also the power of the general Hu-Washizu theorem to establish the complete,
correct and consistent variational basis for the displacement type finite
element method.

5.6 References

5.1 S. P. Timoshenko, On the transverse vibrations of bars of uniform cross-


section, Philosophical Magazine 443, 125-131, 1922.

5.2 G. Prathap, The Finite Element Method in Structural Mechanics, Kluwer


Academic Press, Dordrecht, 1993.

56
Chapter 6

Shear locking
In the preceding chapter, we saw a Ritz approximation of the Timoshenko beam
problem and noted that it was necessary to ensure a certain consistent
relationship between the trial functions to obtain accurate results. We shall
now take up the finite element representation of this problem, which is
essentially a piecewise Ritz approximation. Our conclusions from the preceding
chapter would therefore apply to this as well.

6.1 The linear Timoshenko beam element

An element based on elementary theory needs two nodes with 2 degrees of freedom
at each node, the transverse deflection w and slope dw/dx and uses cubic
1
interpolation functions to meet the C continuity requirements of this theory
(Fig. 6.1). A similar two-noded beam element based on the shear flexible
0
Timoshenko beam theory will need only C continuity and can be based on simple
linear interpolations. It was therefore very attractive for general purpose
applications. However, the element was beset with problems, as we shall
presently see.

6.1.1 The conventional formulation of the linear beam element

The strain energy of a Timoshenko beam element of length 2l can be written as


the sum of its bending and shear components as,

(
ò 1 2 EI χ χ + 1 2 kGA γ γ dx
T T
) (6.1)
where
χ = θ ,x (6.2a)

γ = θ − w,x (6.2b)

In Equations (6.2a) and (6.2b), w is the transverse displacement and θ the


section rotation. E and G are the Young's and shear moduli and the shear
correction factor used in Timoshenko's theory. I and A are the moment of inertia
and the area of cross-section, respectively.

(a) 1 2
w w
w,x w,x

(b) 1 2
w w
θ θ

Fig. 6.1 (a) Classical thin beam and (b) Timoshenko beam elements.

57
In the conventional procedure, linear interpolations are chosen for the
displacement field variables as,

N 1 = (1 − ξ ) 2 (6.3a)
N 2 = (1 + ξ ) 2 (6.3b)

where the dimensionless coordinate ξ=x/l varies from -1 to +1 for an element of


length 2l. This ensures that the element is capable of strain free rigid body
motion and can recover a constant state of strain (completeness requirement) and
that the displacements are continuous within the element and across the element
boundaries (continuity requirement). We can compute the bending and shear
strains directly from these interpolations using the strain gradient operators
given in Equations (6.2a) and (6.2b). These are then introduced into the strain
energy computation in Equation (6.1), and the element stiffness matrix is
calculated in an analytically or numerically exact (a 2 point Gauss Legendre
integration rule) way.

For the beam element shown in Fig. 6.1, for a length h the stiffness
matrix can be split into two parts, a bending related part and a shear related
part, as,

é0 0 0 0ù é 1 h2 −1 h 2ù
ê ú ê ú
ê0 1 0 − 1ú êh 2 h 2 3 - h 2 h2 6 ú
kb = EI h ê ú ks = kGt h ê ú
ê0 0 0 0ú ê -1 - h 2 1 - h 2 úú
ê ú ê
ê0 ê ú
ë − 1 0 1 úû 2
êëh 2 h 6 - h 2 h 2 3 úû

We shall now model a cantilever beam under a tip load using this element,
considering the case of a "thin" beam with E=1000, G=37500000, t=1, L=4, using a
fictitiously large value of G to simulate the "thin" beam condition. Table 6.1
shows that the normalized tip displacements are dramatically in error. In fact
with a classical beam element model, exact answers would have been obtained with
one element for this case. We can carefully examine Table 6.1 to see the trend
as the number of elements is increased. The tip deflections obtained, which are
several orders of magnitude lower than the correct answer, are directly related
to the square of the number of elements used for the idealization. In other
words, the discretization process has introduced an error so large that the
2
resulting answer has a stiffness related to the inverse of N . This is clearly
unrelated to the physics of the Timoshenko beam and also not the usual sort of
discretization errors encountered in the finite element method. It is this very
phenomenon that is known as shear locking.

Table 6.1 - Normalized tip deflections

No. of elements “Thin” beam


0.200 × 10
1 -5

0.800 × 10
-5
2
0.320 × 10
-4
4
0.128 × 10
-3
8
0.512 × 10
-3
16

58
The error in each element must be related to the element length, and
therefore when a beam of overall length L is divided into N elements of equal
length h, the additional stiffening introduced in each element due to shear
2
locking is seen to be proportional to h . In fact, numerical experiments showed
that the locking stiffness progresses without limit as the element depth t
decreases. Thus, we now have to look for a mechanism that can explain how this
2
spurious stiffness of (h/t) can be accounted for by considering the mathematics
of the discretization process.

The magic formula proposed to overcome this locking is the reduced


integration method. The bending component of the strain energy of a Timoshenko
beam element of length 2l shown in Equation (6.1) is integrated with a one-point
Gaussian rule as this is the minimum order of integration required for exact
evaluation of this strain energy. However, a mathematically exact evaluation of
the shear strain energy will demand a two-point Gaussian integration rule. It is
this rule that resulted in the shear stiffness matrix of rank two that locked.
An experiment with a one-point integration of the shear strain energy component
causes the shear related stiffness matrix to change as shown below. The
performance of this element was extremely good, showing no signs of locking at
all (see Table 4.1 for a typical convergence trend with this element).

é0 0 0 0 ù é1 h 2 −1 h 2ù
ê ú ê ú
ê0 1 0 − 1ú êh 2 h 2 4 − h 2 h 2 4 ú
kb = EI h ê ú ks = kGt h ê ú
ê0 0 0 0 ú ê− 1 − h 2 1 − h 2 úú
ê ú ê
ê0 ê ú
− 1 0 1úû
ë ëêh 2 h 4 − h 2 h 4 úû
2 2

6.1.2 The field-consistency paradigm

It is clear from the formulation of the linear Timoshenko beam element using
exact integration (we shall call it the field-inconsistent element) that
ensuring the completeness and continuity conditions are not enough in some
problems. We shall propose a requirement for a consistent interpolation of the
constrained strain fields as the necessary paradigm to make our understanding of
the phenomena complete.

If we start with linear trial functions for w and θ, as we had done in


Equation 6.3 above, we can associate two generalized displacement constants with
each of the interpolations in the following manner

w = a0 + a1 (x l ) (6.4a)

θ = b0 + b1 (x l ) (6.4b)

We can relate such constants to the field-variables obtaining in this


element in a discretized sense; thus, a1/l=w,x at x=0, b0=θ and b1/l=θ,x at x=0.
This denotation would become useful when we try to explain how the
discretization process can alter the infinitesimal description of the problem if
the strain fields are not consistently defined.

59
If the strain-fields are now derived from the displacement fields given in
Equation (6.4), we get

χ = (b1 l ) (6.5a)
γ = (b0 − a1 l ) + b1 (x l ) (6.5b)

An exact evaluation of the strain energies for an element of length h=2l will
now yield the bending and shear strain energy as

U B = 1 2 (EI ) (2l ) {(b1 l )}2 (6.6a)


Us = 1 2 (kGA ) (2l ) {(b0 − a1 l )2 + 1 3 b12 } (6.6b)

It is possible to see from this that in the constraining physical limit of a


very thin beam modeled by elements of length 2l and depth t, the shear strain
energy in Equation (6.6b) must vanish. An examination of the conditions produced
by these requirements shows that the following constraints would emerge in such
a limit

b0 − a1 l → 0 (6.7a)
b1 → 0 (6.7b)

In the new terminology that we had cursorily introduced in Section 5.4,


constraint (6.7a) is field-consistent as it contains constants from both the
contributing displacement interpolations relevant to the description of the
shear strain field. These constraints can then accommodate the true Kirchhoff
constraints in a physically meaningful way, i.e. in an infinitesimal sense, this
is equal to the condition (θ-w,x)→0 at the element centroid. In direct
contrast, constraint (6.7b) contains only a term from the section rotation θ. A
constraint imposed on this will lead to an undesired restriction on θ. In an
infinitesimal sense, this is equal to the condition θ,x→0 at the element
centroid (i.e. no bending is allowed to develop in the element region). This is
the `spurious constraint' that leads to shear locking and violent disturbances
in the shear force prediction over the element, as we shall see presently.

6.1.3 An error model for the field-consistency paradigm

We must now determine that this field-consistency paradigm leads us to an


accurate error prediction. We know that the discretized finite element model
will contain an error which can be recognized when digital computations made
with these elements are compared with analytical solutions where available. The
consistency requirement has been offered as the missing paradigm for the error-
free formulation of the constrained media problems. We must now devise an
operational procedure that will trace the errors due to an inconsistent
representation of the constrained strain field and obtain precise a priori
measures for these. We must then show by actual numerical experiments with the
original elements that the errors are as projected by these a priori error
models. Only such an exercise will complete the scientific validation of the
consistency paradigm. Fortunately, a procedure we shall call the functional re-
constitution technique makes it possible to do this verification.

60
6.1.4 Functional re-constitution

We have postulated that the error of shear locking originates from the spurious
shear constraint in Equation (6.7b). We must now devise an error model for the
case where the inconsistent element is used to model a beam of length L and
depth t. The strain energy for such a beam can be set up as,

Π=
0
ò
L
{1 2 EI θ ,2x + 1 2 kGA (θ − w,x )2 dx} (6.8)

If an element of length 2l is isolated, the discretization process produces


energy for the element of the form given in Equation (6.6). In this equation,
the constants, which were introduced due to the discretization process, can be
replaced by the continuum (i.e. the infinitesimal) description. Thus, we note
that in each element, the constants in Equations (6.6a) and (6.6b) can be traced
to the constants in Equations (6.4a) and (6.4b) and can be replaced by the
values of the field variations θ, θ,x and w,x at the centroid of the element.
Thus, the strain energy of deformation in an element is,

π e = 1 2 (EI ) (2l ) (θ ,x )2 + 1 2 (kGA ) (2l ) (θ − w,x )2 + 1 6 (kGAl )2 (θ ,x )2 (6.9)

Thus the constants in the discretized strain energy functional have been re-
constituted into an equivalent continuum or infinitesimal form. From this re-
constituted functional, we can argue that an idealization of a beam region of
length 2l into a linear displacement type finite element would produce a
modified strain energy density within that region of,

(
π e = 1 2 EI + kGAl2 3 ) (θ,x )2 + 1 2 (kGA ) (θ − w,x )2 (6.10)

This strain energy density indicates that the original physical system has been
altered due to the presence of the inconsistent term in the shear strain field.
Thus, we can postulate that a beam of length L modeled by equal elements of
length 2l will have a re-constituted functional

Π =
0
ò {
1 2 (EI + kGAl2 3 ) (θ ,x )2 + 1 2 (kGA ) (θ − w,x )2 }dx
L
(6.11)

We now understand that the discretized beam is stiffer in bending (i.e. its
flexural rigidity) by the factor kGAl2 3EI . For a thin beam, this can be very
large, and produces the additional stiffening effect described as shear locking.

6.1.5 Numerical experiments to verify error prediction

Our functional re-constitution procedure (note that this is an auxiliary


procedure, distinct from the direct finite element procedure that yields the
stiffness matrix) allows us to critically examine the consistency paradigm. It
indicates that an exactly-integrated or field-inconsistent finite element model
tends to behave as a shear flexible beam with a much stiffened flexural rigidity
I’. This can be related to the original rigidity I of the system by comparing
the expressions in Equations (6.8) and (6.11) as,
I ′ I = 1 + kGAL2 3EI (6.12)

61
We must now show through a numerical experiment that this estimate for the
error, which has been established entirely a priori, starting from the
consistency paradigm and introducing the functional re-constitution technique,
anticipates very accurately, the behavior of a field-inconsistent linearly
interpolated shear flexible element in an actual digital computation. Exact
solutions are available for the static deflection W of a Timoshenko cantilever
beam of length L and depth t under a vertical tip load. If W fem is the result
from a numerical experiment involving a finite element digital computation using
elements of length 2l, the additional stiffening can be described by a parameter
as,

efem = W Wfem − 1 (6.13)

From Equation (6.12), we already have an a priori prediction for this factor as,

e = I ′ I − 1 = kGAl2 3EI (6.14)

We can now re-interpret the results shown in Table 6.1 for the thin beam
case. Using Equations (6.13) and (6.14), we can argue a priori that the
inconsistent element will produce normalized tip deflections (W fem W ) = 1 (1 + e ) .
Since e>>1, we have

( )
W fem W = N 2 5 × 10 −5 (6.15)

for the thin beam. Table 6.2 shows how the predictions made thus compare with
the results obtained from an actual finite element computation using the field-
inconsistent element.

This has shown us that the consistency paradigm can be scientifically verified.
Traditional procedures such as counting constraint indices, or computing the
rank or condition number of the stiffness matrices could offer only a heuristic
picture of how and why locking sets in.

It will be instructive to note here that conventional error analysis norms in


the finite element method are based on the percentage error or equivalent in
some computed value as compared to the theoretically predicted value. We have
seen now that the error of shear locking can be exaggerated without limit, as
the structural parameter that acts as a penalty multiplier becomes indefinitely

Table 6.2 - Normalized tip deflections for the thin beam (Case 2) computed from
fem model and predicted from error model (Equation (6.15)).

N Computed (fem) Predicted


0.200 × 0.200 × 10
-4 -4
1 10
0.800 × 0.800 × 10
-4 -4
2 10
0.320 × 0.320 × 10
-3 -3
4 10
0.128 × 0.128 × 10
-3 -3
8 10
0.512 × 0.512 × 10
-3 -3
16 10

62
large. The percentage error norms therefore saturate quickly to a value
approaching 100% and do not sensibly reflect the relationship between error and
the structural parameter even on a logarithmic plot. A new error norm called the
additional stiffening parameter, e can be introduced to recognize the manner in
which the errors of locking kind can be blown out of proportion by a large
variation in the structural parameter. Essentially, this takes into account, the
fact that the spurious constraints give rise to a spurious energy term and
consequently alters the rigidity of the system being modeled. In many other
examples (e.g. Mindlin plates, curved beams etc.) it was seen that the rigidity,
I, of the field consistent system and the rigidity, I’, of the inconsistent
system, were related to the structural parameters in the form, I’/I = α(l/t)
2

where l is an element dimension and t is the element thickness. Thus, if w is


the deflection of a reference point as predicted by an analytical solution to
the theoretical description of the problem and wfem is the fem deflection
predicted by a field inconsistent finite element model, we would expect the
relationship described by Equation 6.14. A logarithmic plot of the new error
norm against the parameter (l/t) will show a quadratic relationship that will
continue indefinitely as (l/t) is increased. This was found to be true of the
many constrained media problems. By way of illustration of the distinction made

2 2
Fig. 6.2 Variation of error norms e, E with structural parameter kGL /Et for a
cantilever beam under tip shear force.

63
by this definition, we shall anticipate again, the results above. If we
represent the conventional error norm in the form E = (W − W fem ) W , and plot both
E and the new error norm e from the results for the same problem using 4 FI
2
elements against the penalty multiplier (l/t) on a logarithmic scale, the
dependence is as shown in Fig. 6.2. It can be seen that E saturates quickly to a
value approaching 100% and cannot show meaningfully how the error propagates as
the penalty multiplier increases indefinitely. On the other hand, e captures
this relationship, very accurately.

6.1.6 Shear Force Oscillations

A feature of inconsistently modeled constrained media problems is the presence


of spurious violently oscillating strains and stresses. It was not understood
for a very long time that in many cases, stress oscillations originated from the
inconsistent constraints. For a cantilever beam under constant bending moment
modeled using linear Timoshenko beam elements, the shear force (stresses)
displays a saw-tooth pattern (we shall see later that a plane stress model using
4-node elements will also give an identical pattern on the neutral bending
surface). We can arrive at a prediction for these oscillations by applying the
functional re-constitution technique.

If V is the shear force predicted by a field-consistent shear strain


field (we shall see soon how the field-consistent element can be designed) and V
the shear force obtained from the original shear strain field, we can write from
Equation (6.5b),

V = kGA (b0 − a1 l ) (6.16a)

V = V + kGA b1 (x l ) (6.16b)

We see that V has a linear term that relates directly to the constant that
appeared in the spurious constraint, Equation (6.7b). We shall see below from
Equation (6.17) that b1 will not be zero, in fact it is a measure of the bending
moment at the centroid of the element. Thus, in a field-inconsistent
formulation, this constant will activate a violent linear shear force variation
when the shear forces are evaluated directly from the shear strain field given
in Equation (6.5b). The oscillation is self-equilibrating and does not
contribute to the force equilibrium over the element. However, it contributes a
finite energy in Equation (6.9) and in the modeling of very slender beams, this
spurious energy is so large as to completely dominate the behavior of the beam
and cause a locking effect.

Figure 6.3 shows the shear force oscillations in a typical problem - a


straight cantilever beam with a concentrated moment at the tip. One to ten equal
length field-inconsistent elements were used and shear forces were computed at
the nodes of each element. In each case, only the variation within the element
at the fixed end is shown, as the pattern repeats itself in a saw-tooth manner
over all other elements. At element mid-nodes, the correct shear force i.e. V=0
is reproduced. Over the length of the element, the oscillations are seen to be
linear functions corresponding to the kGA b1 (x/l) term. Also indicated by the
solid lines, is the prediction made by the functional re-constitution exercise.
We shall explore this now.

64
Fig. 6.3 Shear force oscillations in element nearest the root, for N element
models of a cantilever of length L = 60.

Consider a straight cantilever beam with a tip shear force Q at the free
end. This should produce a linearly varying bending moment M and a constant
shear force Q in the beam. An element of length 2l at any station on the beam
will now respond in the following manner. Since, a linear element is used, only
the average of the linearly varying bending moment is expected in each finite
element. If the element is field-consistent, the constant b1 can be associated
after accounting for discretization, to relate to the constant bending moment M0
at the element centroid as,

M 0 = EI b1 l or

b1 = M 0l EI (6.17)

In a field-inconsistent problem, due to shear locking, it is necessary to


consider the modified flexural rigidity I’ (see Equation 6.17) that modifies b1
to b1′ , that is,

b1′ = M 0l EI ′
= {M 0l EI (1 + e )}
= b1 (1 + e ) (6.18)

where e = kGAl2 3EI .

Thus, in a field-inconsistent formulation, the constant b1 gets stiffened


by the factor e; the constant bending moment M0 is also underestimated by the

65
same factor. Also, for a very thin beam where e>>1, the centroidal moment M0
2
predicted by a field-consistent element diminishes in a t rate for a beam of
rectangular cross-section. These observations have been confirmed through
digital computation.

The field-consistent element will respond with V = V0 = Q over the entire


element length 2l. The field-inconsistent shear force V from Equations (6.16)
and (6.18) can be written for a very thin beam (e>>1) as,

V = Q + (3M 0 l ) (x l ) (6.19)

These are the violent shear force linear oscillations within each element, which
originate directly from the field-inconsistency in the shear strain definition.

These oscillations are also seen if field-consistency had been achieved in


the element by using reduced integration for the shear strain energy. Unless the
shear force is sampled at the element centroid (i.e. Gaussian point, x/l=0),
these disturbances will be much more violent than in the exactly integrated
version.

6.1.7 The consistent formulation of the linear element

We can see that reduced integration ensures that the inconsistent constraint
does not appear and so is effective in producing a consistent element, at least
in this instance. We must now satisfy ourselves that such a modification did not
violate any variational theorem.

The field-consistent element, as we now shall call an element version free


of spurious (i.e. inconsistent) constraints, can and has been formulated in
various other ways as well. The `trick' is to evaluate the shear strain energy,
in this instance, in such a way that only the consistent term will contribute to
the shear strain energy. Techniques like addition of bubble modes, hybrid
methods etc. can produce the same results, but in all cases, the need for
consistency of the constrained strain field must be absolutely met.

We explain now why the use of a trick like the reduced integration
technique, or the use of assumed strain methods allows the locking problem to be
overcome. It is obvious that it is not possible to reconcile this within the
ambit of the minimum total potential principle only, which had been the starting
point of the conventional formulation.

We saw in Chapter 2, an excellent example of a situation where it was


necessary to proceed to a more general theorem (one of the so-called mixed
theorems) to explain why the finite element method computed strain and stress
fields in a `best-fit' sense. We can now see that in the case of constrained
media problems, the mixed theorem such as the Hu-Washizu or Hellinger-Reissner
theorem can play a crucial role in proving that by modifying the minimum total
potential based finite element formulation by using an assumed strain field to
replace the kinematically derived constrained field, no energy, or work
principle or variational norms have been violated.

66
To eliminate problems such as locking, we look for a consistent
constrained strain field to replace the inconsistent kinematically derived
strain field in the minimum total potential principle. By closely examining the
strain gradient operators, it is possible to identify the order up to which the
consistent strain field must be interpolated. In this case, for the linear
displacement interpolations, Equations (6.5b), (6.7a) and (6.7b) tell us that
the consistent interpolation should be a constant. At this point we shall still
not presume what this constant should be, although past experience suggests it
is the same constant term seen in Equation (6.7a). Instead, we bring in the
Hellinger-Reissner theorem in the following form to see the identity of the
consistent strain field clearly. For now, it is sufficient to note that the
Hellinger-Reissner theorem is a restricted case of the Hu-Washizu theorem. In
this theorem, the functional is stated in the following form,

(
ò − 1 2 EI χ χ + EI χ χ − 1 2 kGA γ
T T T
γ + kGA γ T
)
γ dx (6.20)

where χ and γ are the new strain variables introduced into this multi-field
principle. Since we have difficulty only with the kinematically derived γ we can
have χ = χ and recommend the use of a γ which is of consistent order to replace
γ. A variation of the functional in Equation (6.20) with respect to the as yet
undetermined coefficients in the interpolation for γ yields

ò δγ
T
(γ − γ ) dx = 0 (6.21)

This orthogonality condition now offers a means to constitute the coefficients


of the consistent strain field from the already known coefficients of the
kinematically derived strain field. Thus, for γ given by Equation (6.5b), it is
possible to show that γ = (b0 − α 1 l ) . In this simple instance, the same result is
obtained by sampling the shear strain at the centroid, or by the use of one-
point Gaussian integration. What is important is that, deriving the consistent
strain-field using this orthogonality relation and then using this to compute
the corresponding strain energy will yield a field-consistent element which does
not violate any of the variational norms, i.e. an exact equivalence to the mixed
element exists without having to go through the additional operations in a mixed
or hybrid finite element formulation, at least in this simple instance. We say
that the variational correctness of the procedure is assured. The substitute
strain interpolations derived thus can therefore be easily coded in the form of
strain function subroutines and used directly in the displacement type element
stiffness derivations.

6.1.8 Some concluding remarks on the linear beam element

So far we have seen the linear beam element as an example to demonstrate the
principles involved in the finite element modeling of a constrained media
problem. We have been able to demonstrate that a conceptual framework that
includes a condition that specifies that the strain fields which are to be
constrained must satisfy a consistency criterion is able to provide a complete
scientific basis for the locking problems encountered in conventional
displacement type modeling. We have also shown that a correctness criterion
(which links the assumed strain variation of the displacement type formulation

67
to the mixed variational theorems) allows us to determine the consistent strain
field interpolation in a unique and mathematically satisfying manner.

It will be useful now to see how these concepts work if a quadratic beam
element is to be designed. This is a valuable exercise as later, the quadratic
beam element shall be used to examine problems such as encountered in curved
beam and shell elements and in quadrilateral plate elements due to non-uniform
mapping.

6.2 The quadratic Timoshenko beam element

We shall now very quickly see how the field-consistency rules explain the
behavior of a higher order element. We saw in Chapter 5 that the conventional
formulation with lowest order interpolation functions led to spurious
constraints and a non-singular assembled stiffness matrix, which result in
locking. In a higher order formulation, the matrix was singular but the spurious
constraints resulted in a system that had a higher rank than was felt to be
desirable. This resulted in sub-optimal performance of the approximation. We can
now use the quadratic beam element to demonstrate that this is true in finite
element approximations as well.

6.2.1 The conventional formulation

Consider a quadratic beam element designed according to conventional principles,


i.e. exact integration of all energy terms arising from a minimum total
potential principle. As the beam becomes very thin, the element does not lock;
in fact it produces reasonably meaningful results. Fig. 6.4 shows a typical
comparison between the linear and quadratic beam elements in its application to
a simple problem. A uniform cantilever beam of length 1.0 m, width 0.01 m and
10
depth 0.01 m has a vertical tip load of 100 N applied at the tip. For E=10
N/m and µ=0.3, the engineering theory of beams predicts a tip deflection of
2

w=4.0 m. We shall consider three finite element idealizations of this problem -


with the linear 2-node field-consistent element considered earlier in this
section (2C, on the Figure), the quadratic 3-node field-inconsistent element
being discussed now (3I, on the Figure) and the quadratic 3-node field-
consistent element which we shall derive later (3C). It is seen that for this
simple problem, the 3C element produces exact results, as it is able to simulate
the constant shear and linear bending moment variation along the beam length.
The 3I and 2C elements show identical convergence trends and behave as if they
are exactly alike. The curious aspects that call for further investigation are:
the quadratic element (3I) behaves in exactly the same way as the field-
consistent linear element (2C), giving exactly the same accuracy for the same
number of elements although the system assembled from the former had nearly
twice as many nodes. It also produced moment predictions, which were identical,
i.e., the quadratic beam element, instead of being able to produce linear-
accurate bending moments could now yield only a constant bending moment within
each element, as in the field-consistent linear element. Further, there were now
quadratic oscillations in the shear force predictions for such an element. Note
now that these curious features cannot be explained from the old arguments,
which linked locking to the non-singularity or the large rank or the spectral
condition number of the stiffness matrix. We shall now proceed to explain these
features using the field-consistency paradigm.

68
Fig. 6.4 A uniform cantilever beam with tip shear force -
convergence trends of linear and quadratic elements.

If quadratic isoparametric functions are used for the field-variables w


and θ in the following manner

w = a0 + a1 (x l ) + a2 (x l )2
θ = b0 + b1 (x l ) + b2 (x l )2

the shear strain interpolation will be,

(
γ = (b0 + b2 3 − a1 l ) + (b1 − 2a2 l )ξ − b2 3 1 − 3ξ 2 ) (6.22)

Again, we emphasize the usefulness of expanding the strain field in terms of the
Legendre polynomials. When the strain energies are integrated, because of the
orthogonal nature of the Legendre polynomials the discretized energy expression
becomes the sum of the squares of the coefficients multiplying the Legendre
polynomials. Indeed, the strain energy due to transverse shear strain is,

U s = 1 2 (kGA ) (2l ) {(b


0 + b2 3 − a1 l )2 + 1 3 (b1 − 2a2 l )2 + 4b22 45 } (6.23)

Therefore, when we introduce the penalty limit condition that for a thin
beam the shear strain energies must vanish, we can argue that the coefficients
of the strain field expanded in terms of the Legendre polynomials must vanish
separately. In this case, three constraints emerge:

(b0 + b2 3 − a1 l ) → 0 (6.24a)

69
(b1 − 2a2 l ) → 0 (6.24b)
b2 → 0 (6.24c)

Equations (6.24a) and (6.24b) represent constraints having contributions


from the field interpolations for both w and θ. They can therefore reproduce, in
a consistent manner, true constraints that reflect a physically meaningful
imposition of the thin beam Kirchhoff constraint. This is therefore the field-
consistent part of the shear strain interpolation.

Equation (6.24c) however contains a constant only from the interpolation


for θ. This constraint, when enforced, is an unnecessary restriction on the
freedom of the interpolation for θ, constraining it in fact to behave only as a
linear interpolation as the constraint implies that θ,xx→0 in a discretized
sense over each beam element region. The spurious energy implied by such a
constraint does not contribute directly to the discretized bending energy,
unlike the linear beam element seen earlier. Therefore, field-inconsistency in
this element would not cause the element to lock. However, it will diminish the
rate of convergence of the element and would induce disturbances in the form of
violent quadratic oscillations in the shear force predictions, as we shall see
in the next section.

6.2.2 Functional reconstitution

We can use the functional re-constitution technique to see how the


inconsistent terms in the shear strain interpolation alter the description of
the physics of the original problem (we shall skip most of the details, as the
material is available in greater detail in Ref. 6.1).

The b2 term that appears in the bending energy also makes an appearance in
the shear strain energy, reflecting its origin through the spurious constraint.
We can argue that this accounts for the poor behavior of the field-inconsistent
quadratic beam element (the 3I of Fig. 6.4). Ref. 6.1 derives the effect more
precisely, demonstrating that the following features can be fully accounted for:

i) the displacement predictions of the 3I element are identical to that made by


the 2C element on an element by element basis although it has an additional mid-
node and has been provided with the more accurate quadratic interpolation
functions.
ii) the 3I element can predict only a constant moment within each element,
exactly as the 2C element does.
iii) there are quadratic oscillations in the shear force field within each
element.

We have already discussed earlier that the 3I element (the field-


inconsistent 3-noded quadratic) converges in exactly the same manner as the 2C
element (the field-consistent linear). This has been explained by showing using
the functional re-constitution technique, that the b2 term, which describes the
linear variation in the bending strain and bending moment interpolation, is
"locked" to a vanishingly small value. The 3I element then effectively behaves
as a 2C element in being able to simulate only a constant bending-moment in each
region of a beam, which it replaces.

70
6.2.3 The consistent formulation of the quadratic element

As in the linear element earlier, the field-consistent element (3C) can be


formulated in various ways. Reduced integration of the shear strain energy using
a 2-point Gauss-Legendre formula was the most popular method of deriving the
element so far. Let us now derive this element using the `assumed' strain
approach. We use the inverted commas to denote that the strain is not assumed in
an arbitrary fashion but is actually uniquely determined by the consistency and
the variational correctness requirements. The re-constitution of the field is to
be done in a variationally correct way, i.e. we are required to replace γ in
Equation (6.22) which had been derived from the kinematically admissible
displacement field interpolations using the strain-displacement operators with
an `assumed' strain field γ which contains terms only upto and including the
linear Legendre polynomial in keeping with the consistency requirement. Let us
write this in the form

γ = c0 + c1ξ (6.25)

The orthogonality condition in Equation (6.21) dictates how γ should replace γ


over the length of the element. This determines how c0 and c1 should be
constituted from b0, b1 and b2. Fortunately, the orthogonal nature of the
Legendre polynomials allows this to be done for this example in a very trivial
fashion. The quadratic Legendre polynomial and its coefficient are simply
truncated and c0=b0 and c1=b1 represent the variationally correct field-
consistent `assumed' strain field. The use of such an interpolation subsequently
in the integration of the shear strain energy is identical to the use of reduced
integration or the use of a hybrid assumed strain approach. In a hybrid assumed
strain approach, such a consistent re-constitution is automatically implied in
the choice of assumed strain functions and the operations leading to the
derivation of the flexibility matrix and its inversion leading to the final
stiffness matrix.

6.3 The Mindlin plate elements

A very large part of structural analysis deals with the estimation of stresses
and displacements in thin flexible structures under lateral loads using what is
called plate theory. Thus, plate elements are the most commonly used elements in
general purpose structural analysis. At first, most General Purpose Packages
(GPPs) for structural analysis used plate elements based on what are called the
1
C theories. Such theories had difficulties and limitations and a1so attention
0
turned to what are called the C theories.

The Mindlin plate theory [6.2] is now the most commonly used basis for the
development of plate elements, especially as they can cover applications to
moderately thick and laminated plate and shell constructions. It has been
estimated that in large scale production runs using finite element packages, the
simple four-node quadrilateral plate element (the QUAD4 element) may account for
as much as 80% of all usage. It is therefore important to understand that the
evolution of the current generation of QUAD4 elements from those of yester-year,
over a span of nearly three decades was made difficult by the presence of shear
locking. We shall now see how this takes place.

71
The history behind the discovery of shear locking in plate elements is
quite interesting. It was first recognized when an attempt was made to represent
the behavior of shells using what is called the degenerate shell approach [6.3].
In this the shell behavior is modeled directly after a slight modification of
the 3D equations and shell geometry and domain are represented by a 3D brick
element but its degrees of freedom are condensed to three displacements and two
section rotations at each node. Unlike classical plate or shell theory, the
transverse shear strain and its energy is therefore accounted for in this
formulation. Such an approach was therefore equivalent to a Mindlin theory
formulation. These elements behaved very poorly in representing even the trivial
example of a plate in bending and the errors progressed without limit, as the
plates became thinner. The difficulty was attributed to shear locking. This is
in fact the two-dimensional manifestation of the same problem that we
encountered for the Timoshenko beam element; ironically it was noticed first in
the degenerate shell element and was only later related to the problems in
designing Timoshenko beam and Mindlin plate elements [6.4]. The remedy proposed
at once was the reduced integration of the shear strain energy [6.5,6.6]. This
was only partially successful and many issues remained unresolved. Some of these
were,

i) the 2×2 rule failed to remove shear locking in the 8-node serendipity plate
element,
ii) the 2×2 rule in the 9-node Lagrangian element removed locking but introduced
zero energy modes,
iii) the selective 2×3 and 3×2 rule for the transverse shear strain energies
from γxz and γyz recommended for a 8-node element also failed to remove shear
locking,
iv) the same selective 2×3 and 3×2 rule when applied to a 9-noded element is
optimal for a rectangular form of the element but not when the element was
distorted into a general quadrilateral form,
v) even after reduced integration of the shear energy terms, the degenerate
shell elements performed poorly when trying to represent the bending of curved
shells, due to an additional factor, identified as membrane locking [6.7],
originating now from the need for consistency of the membrane strain
interpolations. We shall consider the membrane-locking phenomenon in another
section.

We shall confine our study now to plate elements without going into the
complexities of the curved shell elements.

In Kirchhoff-Love thin plate theory, the deformation is completely


described by the transverse displacement w of the mid-surface. In such a
description, the transverse shear deformation is ignored. To account for
transverse shear effects, it is necessary to introduce additional degrees of
freedom. We shall now consider Mindlin's approximations, which have permitted
such an improved description of plate behavior. The degenerate shell elements
that we discussed briefly at the beginning of this section can be considered to
correspond to a Mindlin type representation of the transverse shear effects.

In Mindlin's theory [6.2], deformation is described by three quantities,


the section rotations θx and θy (i.e. rotations of lines normal to the
midsurface of the undeformed plate) and the mid-surface deflection w. The
bending strains are now derived from the section rotations and do not cause any

72
difficulty when a finite element model is made. The shear strains are now
computed as the difference between the section rotations and the slopes of the
neutral surfaces, thus,

γ xz = θ x − w,x
γ yz = θ y − w,y (6.26)

The stiffness matrix of a Mindlin plate element will now have terms from the
bending strain energy and the shear strain energy. It is the inconsistent
representation of the latter that causes shear locking.

6.3.1 The 4-node plate element

The 4-node bi-linear element is the simplest element based on Mindlin theory
that could be devised. We shall first investigate the rectangular form of the
element [6.4] as it is in this configuration that the consistency requirements
can be easily understood and enforced. In fact, an optimum integration rule can
be found which ensures consistency if the element is rectangular. It was
established in Ref. 6.4 that an exactly integrated Mindlin plate element would
lock even in its rectangular form. Locking was seen to vanish for the
rectangular element if the bending energy was computed with a 2×2 Gaussian
integration rule while a reduced 1-point rule was used for the shear strain
energy. This rectangular element behaved very well if the plate was thin but the
results deteriorated as the plate became thicker. Also, after distortion to a
quadrilateral form, locking re-appeared. A spectral analysis of the element
stiffness matrix revealed a rank deficiency - there were two zero energy
mechanisms in addition to the usual three rigid body modes required for such an
element. It was the formation of these mechanisms that led to the deterioration
of element performance if the plate was too thick or if it was very loosely
constrained. It was not clear why the quadrilateral form locked even after
reduced integration. We can now demonstrate from our consistency view-point why
the 1-point integration of the shear strain energy is inadequate to retain all
the true Kirchhoff constraints in a rectangular thin plate element. However, we
shall postpone the discussion on why such a strategy cannot preserve consistency
if the element was distorted to a later section.

Following Ref. [6.4], the strain energy for an isotropic, linear elastic
plate element according to Mindlin theory can be constituted from its bending
and shear energies as,

U = UB + US

) {ò ò [θ
Et2
+θ y2 ,y +2ν θ x ,x θ y , y
=
(
2
x ,x
24 1 − ν 2

(
+ (1 − ν ) 2 θ x,y + θ y,x )2 ] dx dy (6.27)

+
6k (1 − ν )
t 2
[
ò ò (θ x − w,x ) + (θ y − w,y ) dx dy
2 2
] }

73
Fig. 6.5 Cartesian and natural coordinate system for a four-node rectangular
plate element.

where x, y are Cartesian co-ordinates (see Fig. 6.5), w is the transverse


displacement, θx and θy are the section rotations, E is the Young's modulus, ν
is the Poisson's ratio, k is the shear correction factor and t is the plate
thickness. The factor k is introduced to compensate for the error in
approximating the shear strain as a constant over the thickness direction of a
Mindlin plate.

Let us now examine the field-consistency requirements for one of the shear
strains, γxz, in the Cartesian system. The admissible displacement field
interpolations required for a 4-node element can be written in terms of the
Cartesian co-ordinates itself as,

w = a0 + a1x + a2 y + a3 xy (6.28a)
θ = b0 + b1x + b2 y + b3 xy (6.28b)

The shear strain field derived from these kinematically admissible shape
functions is,

γ xz = (b0 − a1 ) + (b2 − a3 )y + b1x + b3 xy (6.29)

As the plate thickness is reduced to zero, the shear strains must vanish. The
discretized constraints that are seen, to be enforced as γ xz → 0 in Equation
(6.29) are,
b0 − a1 → 0 (6.30a)
b2 − a3 → 0 (6.30b)
b1 → 0 (6.30c)
b3 → 0 (6.30d)

74
The constraints shown in Equations (6.30a) and (6.30b) are physically meaningful
and represent the Kirchhoff condition in a discretized form. Constraints (6.30c)
and (6.30d) are the cause for concern here - these are the spurious or
`inconsistent' constraints which lead to shear locking. Thus, in a rectangular
element, the requirement for consistency of the interpolations for the shear
strains in the Cartesian co-ordinate system is easily recognized as the
polynomials use only Cartesian co-ordinates. Let us now try to derive the
optimal element and also understand why the simple 1-point strategy of Ref. 6.4
led to zero energy mechanisms.

It is clear from Equations (6.29) and (6.30) that the terms b1x and b3xy
are the inconsistent terms which will contribute to locking in the form of
spurious constraints. Let us now look for optimal integration strategies for
removing shear locking without introducing any zero energy mechanisms. We shall
consider first, the part of the shear strain energy contributed by γxz. We must
integrate exactly, terms such as (b0-a1), (b2-a3)y, b1x, and b3xy. We now
identify terms such as (b0-a1), (b2-a3), b1, and b3 as being equivalent to the
quantities (θx-w,x)0, (θx-w,x),y0, (θx,x)0, and (θx,xy)0 where the subscript ‘0’
denotes the values at the centroid of the element (for simplicity, we let the
centroid of the element lie at the origin of the Cartesian co-ordinate system).

An exact integration, that is a 2×2 Gaussian integration of the shear


strain energy leads to

ò ò γ xzdx dy = 4lh [(θ x − w,x )0 + h (


3 (θ x − w,x )2 , y0 + l 2 3 (θ x ,x )02 + h 2l 2 9 θ x ,xy )02 ]
2 2 2
(6.31)

In the penalty limit of a thin plate, these four quantities act as constraints.
The first two reproduce the true Kirchhoff constraints and the remaining two act
as spurious constraints that cause shear locking by enforcing θx,x→0 and θx,xy→0
in the element.

If a 1×2 Gaussian integration is used, we have,

ò ò γ xzdx dy = 4lh [(θ x − w,x )0 + h 3 (θ x − w,x )2 , y0 ]


2 2 2
(6.32)

Thus, only the true constraints are retained and all spurious constraints are
removed. This strategy can also be seen to be variationally correct in this
case; we shall see later that in a quadrilateral case, it is not possible to
ensure variational correctness exactly. By a very similar argument, we can show
that the part of the shear strain energy from γyz will require a 2×1 Gaussian
integration rule. This element would be the optimal rectangular bi-linear
Mindlin plate element.

Let us now look at the 1-point integration strategy used in Ref. 6.4. This
will give shear energy terms such as,

ò ò γ xz dx dy = 4lh [(θ x − w,x )0 ]


2 2
(6.33)

75
We have now only one true constraint each for the shear energy from γxz
and γyz respectively while the other Kirchhoff constraints (θ x − w,x ), y0 → 0 and
(θ y − w,y ),x0 → 0
are lost. This introduces two zero energy modes and accounts for
the consequent deterioration in performance of the element when the plates are
thick or are very loosely constrained, as shown in Ref. 6.4.

We have seen now that it is a very simple procedure to re-constitute


field-consistent assumed strain fields from the kinematically derived fields
such as shown in Equation (6.29) so that they are also variationally correct.
This is not so simple in a general quadrilateral where the complication arising
from the isoparametric mapping from a natural co-ordinate system to a Cartesian
system makes it very difficult to see the consistent form clearly. We shall see
the difficulties associated with this form in a later section.

6.3.2 The quadratic 8-node and 9-node plate elements

The 4-node plate element described above is based on bi-linear functions. It


would seem that an higher order element based on quadratic functions would be
far more accurate. There are now two possibilities, an 8-node element based on
what are called the serendipity functions and a 9-node element based on the
Lagrangian bi-quadratic functions. There has been a protracted debate on which
version is more useful, both versions having fiercely committed protagonists.
By now, it is well known that the 9-node element in its rectangular form is free
of shear locking even with exact integration of shear energy terms and that its
performance is vastly improved when its shear strain energies are integrated in
a selective sense (2×3 and 3×2 rules for γ xz and γ yz terms respectively). It is
in fact analogous to the quadratic Timoshenko beam element, the field-
inconsistencies not being severe enough to cause locking. This is however not
true for the 8-node element which was derived from the Ahmad shell element [6.3]
and which actually pre-dates the 4-node Mindlin element. An exact integration of
bending and shear strain energies resulted in an element that locked for most
practical boundary suppressions even in its rectangular form. Many ad-hoc
techniques e.g. the reduced and selective integration techniques, hybrid and
mixed methods, etc. failed or succeeded only partially. It was therefore
regarded for some time as an unreliable element as no quadrature rule seemed to
be able to eliminate locking entirely without introducing other deficiencies. It
seems possible to attribute this noticeable difference in the performance of the
8- and 9-node elements to the missing central node in the former. This makes it
more difficult to restore consistency in a simple manner.

6.3.3 Stress recovery from Mindlin plate elements

The most important information a structural analyst looks for in a typical


finite element static analysis is the state of stress in the structure. It is
therefore very important for one to know points of optimal stresses in the
Mindlin plate elements. It is known that the stress recovery at nodes from
displacement elements is unreliable, as the nodes are usually the points where
the strains and stresses are least accurate. It is possible however to determine
points of optimal stress recovery using an interpretation of the displacement
method as a procedure that obtains strains over the finite element domain in a
least-squares accurate sense. In Chapter 2, we saw a basis for this
interpretation. We can apply this rule to determine points of accurate stress

76
recovery in the Mindlin plate elements. For a field-consistent rectangular 4-
node element, the points are very easy to determine [6.8] (note that in a field-
inconsistent 4-node element, there will be violent linear oscillations in the
shear stress resultants corresponding to the inconsistent terms). Thus, Ref. 6.8
shows that bending moments and shear stress resultants Qxz and Qyz are accurate
at the centroid and at the 1×2 and 2×1 Gauss points in a rectangular element for
isotropic or orthotropic material. It is coincidental, and therefore fortuitous,
that the shear stress resultants are most accurate at the same points at which
they must be sampled in a selective integration strategy to remove the field-
inconsistencies! For anisotropic cases, it is safest to sample all stress
resultants (bending and shear) at the centroid.

Such rules can be extended directly to the 9-node rectangular element. The
bending moments are now accurate at the 2×2 Gauss points and the shear stress
resultants in an isotropic or orthotropic problem are optimal at the same 2×3
and 3×2 Gauss points which were used to remove the inconsistencies from the
strain definitions. However, accurate recovery of stresses from the 8-node
element is still a very challenging task because of the difficulty in
formulating a robust element. The most efficient elements known today are
variationally incorrect even after being made field-consistent and need special
filtering techniques before the shear stress resultants can be reliably sampled.

So far, discussion on stress sampling has been confined to rectangular


elements. When the elements are distorted, it is no simple matter to determine
the optimal points for stress recovery - the stress analyst must then exercise
care in applying these rules to seek reliable points for recovering stresses.

6.4 Concluding remarks

We can conclude this section on shear locking by noting that the available
understanding was unable to resolve the phenomena convincingly. The proposed
improvement, which was the consistency paradigm, together with the functional
re-constitution procedure, allowed us to derive an error estimate for a case
under locking and we could show through numerical (digital) experiments that
these estimates were accurate. In this way we are convinced that a theory with
the consistency paradigm is more successful from the falsifiability point of
view than one without.

6.5 References

6.1 G. Prathap and C. R. Babu, Field-consistent strain interpolations for the


quadratic shear flexible beam element, Int. J. Num. Meth. Engng. 23, 1973-
1984, 1986.
6.2 R. D. Mindlin, Influence of rotary inertia and shear on flexural motion of
elastic plates, J. Appl. Mech. 18, 31-38, 1951.
6.3 S. Ahmad, B. M. Irons and O. C. Zienkiewicz, Analysis of thick and thin
shell structures by curved finite elements, Int. J. Num. Meth. Engng. 2,
419-451, 1970.
6.4 T. J. R. Hughes, R. L. Taylor and W. Kanoknukulchal, A simple and efficient
finite element for plate bending, Int. J. Num. Meth. Engng. 411, 1529-
1543, 1977.
6.5 S. F. Pawsey and R. W. Clough, Improved numerical integration of thick shell
finite elements, Int. J. Num. Meth. Engng. 43, 575-586, 1971.

77
6.6 O. C. Zienkiewicz, R. L. Taylor and J. M. Too, Reduced integration technique
in general analysis of plates and shells, Int. J. Num. Meth. Engng. 43, 275-
290, 1971.
6.7 H. Stolarski and T. Belytschko, Membrane locking and reduced ingression for
curved elements, J. Appl. Mech. 49, 172-178, 1982.
6.8 G. Prathap and C. R. Babu, Accurate force evaluation with a simple bi-linear
plate bending element, Comp. Struct. 25, 259-270, 1987.

78
Chapter 7

Membrane locking, parasitic shear and incompressible locking


7.1 Introduction

The shear locking phenomenon was the first of the over-stiffening behavior that
was classified as a locking problem. Other such pathologies were noticed in the
behavior of curved beams and shells, in plane stress modeling and in modeling of
three dimensional elastic behavior at the incompressible limit (ν→0.5). The
field-consistency paradigm now allows all these phenomena to be traced to
inconsistent representations of the constrained strain fields. A unifying
pattern is therefore introduced to the understanding of the locking problems in
constrained media elasticity - whether it is shear locking in a straight beam or
flat plate element, membrane locking in a curved beam or shell element,
parasitic shear in 2D plane stress and 3D elasticity or incompressible locking
in 2D plane strain and 3D elasticity as the Poisson's ratio ν→0.5.

This chapter will now briefly summarize this interpretation. Unlike the
chapter on shear locking, a detailed analysis is omitted as it would be beyond
the scope of the present book and readers should refer to the primary published
literature.

7.2 Membrane locking

Earlier, we argued that the most part of structural analysis deals with the
behavior of thin flexible structures. One popular and efficient form of
construction of such thin flexible structures is the shell - enclosing space
using one or more curved surfaces. A shell is therefore the curved form of a
plate and its structural action is a combination of stretching and bending. It
is possible to perform a finite element analysis of a shell by using what is
called a facet representation - i.e. the shell surface is replaced with flat
triangular and/or quadrilateral plate elements in which a membrane stiffness
(membrane element) is superposed on a bending stiffness (plate bending element).
Such a model is understandably inaccurate in that with very coarse meshes, they
do not capture the bending-stretching coupling of thin shell behavior. Hence,
the motivation for designing elements with mid-surface curvature taken into
account and which can capture the membrane-bending coupling correctly. There are
two ways in which this could be done. One is to use elements based on specific
shell theories (e.g. the Donnell, Flugge, Sanders, Vlasov theories, etc.). There
are considerable controversies regarding the relative merits of these theories
as each has been obtained by carrying out approximations to different degrees
when the 3-dimensional field equations are reduced to the particular class of
shell equations. The second approach is called the degenerate shell approach -
three dimensional solid elements can be reduced (degenerated) into shell
elements having only mid-surface nodal variables - these are no longer dependent
on the various forms of shell theories proposed and should be simple to use.
They are in fact equivalent to a Mindlin type curved shell element.

However, accurate and robust curved shell elements have been extremely
difficult to design - a problem challenging researchers for nearly three
decades. One key issue behind this difficulty is the phenomenon of membrane
locking - the very poor behavior of curved elements, as they become thin. This

79
Fig. 7.1 The classical thin curved beam element.

was neither recognized nor understood for a very long time. It was believed at
first that this was because curved elements could not satisfy the strain-free
rigid body motion condition because of their inherent curved geometry. In this
section, we shall apply the field-consistency paradigm to the problem to
recognize that a consistent representation of membrane strains is desired to
avoid the phenomenon called membrane locking and that it is the need for
consistency rather than the requirement for strain-free rigid body motion which
is the key factor. This is most easily identified by working with the one-
dimensional curved beam (or arch) elements.

In Chapter 6, we studied structural problems where both bending and shear


deformation was present. We saw that simple finite element models for the shear
deformable bending action of a beam or plate behaved very poorly in thin beam or
plate regimes where the shear strains become vanishingly small when compared
with the bending strains. This phenomenon is known as shear locking. We also saw
how the field-consistency paradigm was needed to explain this behavior in a
scientifically satisfying way.

We shall now see that a similar problem appears in curved finite elements
in which both flexural and membrane deformation take place. Such elements
perform very poorly in cases where inextensional bending of the curved structure
was predominant, i.e. the membrane strains become vanishingly small when
compared to the bending strains. This phenomenon has now come to be called
membrane locking.

In this section, we shall examine in turn, the classical thin curved beam
element and the linear and quadratic shear flexible curved beam elements in a
curvilinear co-ordinate system.

7.2.1 The classical thin curved beam element

Figure 7.1 describes the simplest curved beam element of length 2l and radius of
curvature R based on classical thin beam theory. The displacement degrees of
freedom required are the circumferential displacement u and the radial
displacement w. The co-ordinate s follows the middle line of the curved beam.
The membrane strain ε and the bending strain χ are described by the strain-
displacement relations

ε = u,s + w R (7.1a)
χ = u,s R − w,ss (7.1b)

0
A C description for u and a C1 description w is required. Kinematically
admissible displacement interpolations for u and w are

80
Fig. 7.2 Geometry of a circular arch

u = a0 + a1ξ (7.2a)
w = b0 + b1ξ + b2ξ 2
+ b3 ξ 3
(7.2b)

where ξ = s l and a0 to b3 are the generalized degrees of freedom which can be


related to the nodal degrees of freedom u, w and w,s at the two nodes.

The strain field interpolations can be derived as

( ) (
χ = a1 Rl − 2b2 l 2 − 6b3 l 2 ξ ) (7.3a)

ε = (a1 l + b0 R + b2 3R ) + (b1 R + 3b3 5R )ξ − b 2 3R (1 − 3ξ )− b 2


3 (
5R 3ξ − 5ξ 3 ) (7.3b)

If a thin and deep arch (i.e. having a large span to rise ratio so that L/t>>1
and R/H is small, see Fig. 7.2) is modeled by the curved beam element, the
physical response is one known as inextensional bending such that the membrane
strain tends to vanish. From Equation (7.3b) we see that the inextensibility
condition leads to the following constraints:

a1 l + b0 R + b2 3R → 0 (7.4a)

b1 + 3b3 5 → 0 (7.4b)

b2 → 0 (7.4c)

b3 → 0 (7.4d)

Following the classification system we introduced earlier, we can observe that


constraint (7.4a) has terms participating from both the u and w fields. It can
therefore represent the condition ε = u,s + w R → 0 in a physically meaningful way,

81
i.e. it is a true constraint. However, the three remaining constraint (7.4b) to
(7.4d) have no participation from the u field. Let us now examine in turn what
these three constraints imply for the physical problem. From the three
constraints we have the conditions b1 → 0, b2 → 0 and b3 → 0. Each in turn
implies the conditions w,s → 0, w,ss → 0 and w,sss → 0 . These are the spurious
constraints. This element has been studied extensively and the numerical
evidence given there confirms that membrane locking is mainly determined by the
first of these three constraints. Next, we shall briefly describe the functional
re-constitution analysis that will yield semi-quantitative error estimates for
membrane locking in this element.

We can now use the functional re-constitution technique, (which in Chapter


6 revealed how the discretization process associated with the linear Timoshenko
beam element led to large errors known as shear locking) to see how the
inconsistent constraints in the membrane strain field interpolation (Equations
(7.4b) to (7.4d)) modify the physics of the curved beam problem in the
discretization process.

The analysis is a little more difficult than for the shear locking case
and would not be described in full here. For a curved beam of length 2l, moment
of inertia I and area of cross-section A, the strain energy can be written as,

(
π e = ò 1 2 EI χ T χ + 1 2 EA ε T ε ds ) (7.5)

where ε and χ are as defined in Equation (7.1). For simplicity assume that the
cross-section of the beam is rectangular so that if the depth of the beam is t,
then I = At2 12 . After discretisation, and reconstitution of the functional, we
can show that the inconsistent membrane energy terms disturb the bending energy
terms - the principal stiffening term is physically equivalent to a spurious in-
(
plane or membrane stiffening action due to an axial load N = EAl2 3R 2 . )
This action is quite complex and it is not as simple to derive useful
universally valid error estimates as was possible with the linear Timoshenko
beam element. We therefore choose a problem of simple configuration which can
reveal the manner in which the membrane locking action takes place. We consider
a shallow circular arch of length L and radius of curvature R, simply supported
at the ends. The loading system acting on it is assumed to induce nearly
inextensional bending. We also assume that the transverse deflection w can be
approximated in the form w=csin(πs/L). If the entire arch is discretized by
curved beam elements of length 2l each, the strain energy of the discretized
arch (again assuming that χ=-w,ss and neglecting the consistently modeled part
of the membrane energy) one can show that due to membrane locking, the bending
rigidity of the arch has now been altered in the following form:

(EI )′ (EI ) = {1 + (4 )
π 2 (Ll Rt )2 } (7.6)

We can therefore expect membrane locking to depend on the structural parameter


2
of the type (Ll/Rt) .

82
Fig. 7.3 The Mindlin curved beam element.

Numerical experiments with various order integration rules applied to


evaluate the various components of the stiffness matrix indicated that the
inconsistent elements locked exactly as predicted by the pattern describe above.
The consistent element was entirely free of locking.

This classical curved beam element (the cubic-linear element) was


discredited for a very long time because it was not possible to understand why
the element should perform so poorly. We have now found that the explanation for
its very poor behavior could be attributed entirely to the inconsistency in the
membrane strain field interpolations. The belief that the lack of the strain
free rigid body motion in the conventional formulation was the cause of these
errors cannot be substantiated at all. We have also seen how we could restore
the element into a very accurate one (having the accuracy of the straight beam
models) by designing a new membrane strain field which met the consistency
requirements.

7.2.2 The Mindlin curved beam elements

The exercise above showed that the locking behavior of curved beams could be
traced to the inconsistent definition of the membrane strain field. The curved
beam element used to demonstrate this was based on classical thin beam theory
and used a curvilinear co-ordinate description. Of interest to us now is to
understand how the membrane locking phenomenon can take place in a general
curved shell element. Most general shell elements which have now been accepted
into the libraries of general purpose finite element packages are based on what
is called the degenerate shell theory which is equivalent to a shear deformable
theory and use quadratic interpolations (i.e. three nodes on each edge to
capture the curved geometry in a Cartesian system).

To place the context of membrane locking in general shell elements


correctly, it will be useful to investigate the behavior of the Mindlin curved
beam elements. These are now described by the circumferential (tangential)
displacement u, the radial (normal) displacement w and the rotation of a normal
to the midline θ. As in the Timoshenko beam element, this rotation differs from
the rotation of the mid-line w,s by an amount γ = θ − w,s which describes the
shear strain at the section. Fig. 7.3 shows how the Mindlin curved beam element
is specified.

The strain energy in a Mindlin curved beam is U=UM+UB+US where the


respective contributions to U arise from the membrane strain ε, the bending

83
strain χ and the transverse shear strain γ. For an element of length 2l and
radius R,
U M = (1 2 ) (EA ) ò ε 2 ds where ε = u,s + w R (7.7a)

U B = (1 2 ) (EI ) ò χ 2 ds where χ = u,s R − θ ,s (7.7b)

U S = (1 2 ) (kGA ) ò γ 2 ds where γ = θ − w ,s (7.7c)

where EA, EI and kGA represent the respective rigidities.

There are no difficulties with the bending strain. However, depending on


the regime, both shear strains and membrane strains can be constrained to vanish
leading to shear and membrane locking. The arguments concerning shear locking
are identical to those expressed in Chapter 6, i.e. while the inconsistency in
the shear strain field is severe enough to cause locking in the former (a
significant stiffening which caused errors that propagated indefinitely as the
structural parameter kGl2 Et2 became very large), it is much milder in the
latter and is responsible only for a poorer rate of convergence (the performance
of the element having reduced to that of the linear field-consistent element).
However with membrane locking, the inconsistencies at the quadratic level can
cause locking - therefore both the linear and quadratic elements must be
examined in turn to get the complete picture.
0
For the linear Mindlin curved beam element, as a C displacement type
formulation is sufficient, we can choose linear interpolations of the following
form,

u = a0 + a1ξ (7.8a)
w = b0 + b1ξ (7.8b)
θ = c0 + c1ξ (7.8c)

where ξ = s l is the natural co-ordinate along the arch mid-line. These


interpolations lead to the following membrane strain field interpolation:

ε = (a1 l + b0 R ) + (b1 R ) ξ (7.9)

There is now one single spurious constraint in the limit of inextensional


bending of an arch, i.e. b1→0, which implies a constraint of the form w,s→0 at
the centroid of the element. The membrane locking action is similar to but
simpler than that seen for the classical thin curved beam element above. In a
thin beam, in the Kirchhoff limit, θ→w,s, which implies that membrane locking
induces the section rotations θ to lock. It can therefore be seen that a
constraint of this form leads to an in-plane stiffening action of the form
corresponding to the introduction of a spurious energy

(1 2 ) (EAl2 ) òw , 2
3R 2 s ds (7.10)

84
The membrane locking action is therefore identical to that predicted for the
classical thin curved beam. A variationally correct field-consistent element can
easily be derived using the Hu-Washizu theorem. This is a very accurate element.

The quadratic Mindlin curved beam element can also be interpreted in the
light of our studies of the classical curved beam earlier. Spurious constraints
at both the linear (i.e. w,s→0) and quadratic (i.e w,ss→0) levels lead to
errors that propagate at a (Ll Rt )2 rate and l 2 Rt (
rate respectively. We )
2

expect now that in a quadratic Mindlin curved beam element, the latter
constraint is the one that is present and that these errors are large enough to
make the membrane locking effect noticeable. This also explains why the
degenerate shell elements, which are based on quadratic interpolations, suffer
from membrane locking.

A quadratic (three-node) element having nodes at s=-l, 0 and l (see Fig.


7.3) would have a displacement field specified as,

u = a0 + a1ξ + a2ξ 2 (7.11a)


w = b0 + b1ξ + b2ξ 2
(7.11b)
θ = c0 + c1ξ + c2ξ 2 (7.11c)

The membrane strain field can be expressed as:

(
ε = (a1 l + b0 R + b2 3R ) + (2a2 l + b1 R ) ξ − b2 3R 1 − 3ξ 2 ) (7.12)

The spurious constraint is now b2→0 a discretized equivalent of imposing the


constraint w,ss→0 at the centroid of each element. To understand that this can
cause locking, we note that in a thin curved beam, θ ,s → w,ss and therefore a
spurious constraint of this form acts to induce a spurious bending energy. A
functional re-constitution approach will show that the bending energy for a thin
curved beam modifies to,

ò
é
ë
( )

U B = (1 2 ) (EI ) ê1 + (4 15 ) l 2 Rt ú w,ss
û
2
ds (7.13)

This can be interpreted as spurious additional stiffening of the actual bending


( )
rigidity of the structure by the (4 15 ) l 2 Rt 2 term. It is a factor of this form
that caused the very poor performance of the exactly integrated quadratic shell
elements.

A variationally correct field consistent element is obtained by using the


Hu-Washizu theorem or very simply by using a two-point integration rule for the
shear and membrane strain energy energies. This is an extremely accurate
element.

It is interesting to compare the error norm for locking for this element,
i.e. (l 2
Rt2 ) with the corresponding error norm for locking in the inconsistent

85
( )
linear Mindlin curved beam element, which is L l Rt 2 . It is clear that since
L>>l when an arch is discretized with several elements, the locking in the
quadratic element is much less than in the linear element. One way to look at
this is to examine the rate at which locking is relieved as more elements are
added to the mesh, i.e. as l is reduced. In the case of the inconsistent linear
2
element, locking is relieved only at an l rate whereas in the case of the
4
quadratic element, locking is relieved at the much faster l rate. However,
locking in the quadratic element is significant enough to degrade the
performance of the quadratic curved beam and shell elements to a level where it
is impractical to use in its inconsistent form. This is clearly seen from the
fact that to reduce membrane-locking effects to acceptable limits, one must use
element lengths l << Rt . Thus for an arch with R=100 and t=1, one must use
about a 100 elements to achieve an accuracy that can be obtained with two to
four field-consistent elements.

7.2.3 Axial force oscillations

Our experience with the linear and quadratic straight beam elements showed that
inconsistency in the constrained strain-field is accompanied by spurious stress
oscillations corresponding to the inconsistent constraints. We can now
understand that the inconsistent constraint b1→0 will induce linear axial force
oscillations in the linear curved beam element. Similarly, the inconsistent
quadratic constraint will trigger off quadratic oscillations in the quadratic
element.

7.2.4 Concluding remarks

So far, we have examined the phenomenon of membrane locking in what were called
the simple curved beam elements. These were based on curvilinear geometry and
the membrane locking effect was easy to identify and quantify. However, curved
beam and shell elements which are routinely used in general purpose finite
element packages are based on quadratic formulations in a Cartesian system. We
shall call these the general curved beam and shell elements. Membrane locking is
not so easy to anticipate in such a formulation although the predictions made in
this section have been found to be valid for such elements.

7.3 Parasitic shear

Shear locking and membrane locking are phenomena seen in finite element modeling
using structural elements based on specific theories, e.g. beam, plate and shell
theories. These are theories obtained by simplifying a more general continuum
description, e.g. the three dimensional theory of elasticity or the two
dimensional theories of plane stress, plane strain or axisymmetric elasticity.
The simplifications are made by introducing restrictive assumptions on strain
and/or stress conditions like vanishing of shear strains or vanishing of
transverse normal strains, etc. In the beam, plate and shell theories, this
allows structural behavior to be described by mid-surface quantities. However,
these restrictive assumptions impose constraints which make finite element
modeling quite difficult unless the consistency aspects are taken care of. We
shall now examine the behavior of continuum elements designed for modeling
continuum elasticity directly, i.e. 2D elements for plane strain or stress or
axisymmetric elasticity and solid elements for 3D elasticity. It appears that

86
Fig. 7.4 (a) Constant strain triangle, (b) Rectangular bilinear element,
(c) Eight-node rectangle and (d) Nine-node rectangle.

under constraining physical regimes, locking behavior is possible. One such


problem is parasitic shear. We shall see the genesis of this phenomenon and show
that it can be explained using the consistency concepts. The problems of shear
locking, membrane locking and parasitic shear obviously are very similar in
nature and the unifying factor is the concept of consistency.

The finite element modeling of plane stress, plane strain, axisymmetric


and 3D elasticity problems can be made with two-dimensional elements and three-
dimensional elements (sometimes called continuum elements to distinguish them
from structural elements), of which the linear 3-noded triangle and bi-linear 4-
noded rectangle (Fig. 7.4) are the simplest known in the 2D case. In most
applications, these elements are reliable and accurate for determining general
two-dimensional stress distributions and improvements can be obtained by using
the higher order elements, the quadratic 6-noded triangle and 8- or 9-noded
quadrilateral elements (Fig. 7.4).

However, in some applications, e.g. the plane stress modeling of beam flexure or
in plane strain and axisymmetric applications where the Poisson's ratio
approaches 0.5 (for nearly incompressible materials or for materials undergoing
plastic deformation), considerable difficulties are seen. Parasitic shear occurs
in the plane stress and 3D finite element modeling of beam flexure. Under pure
bending loads, for which the shear stress should be zero, the elements yield
large values of shear stress except at certain points (e.g. centroids in linear
elements). In the linear elements, the errors progress in the same fashion as
for shear locking in the linear Timoshenko beam element, i.e. the errors
progress indefinitely as the element aspect ratio increases. Reduced integration
of the shear strain energy or addition of bubble modes for the 4-noded

87
Fig. 7.5 Plane stress model of beam flexure.

rectangular element completely eliminates this problem. The 8- and 9-noded


elements do not lock but improve in performance with reduced integration. These
elements, in their original (i.e. field-inconsistent) form show severe linear
and quadratic shear stress oscillations.

Here, we shall extend the consistency paradigm to examine this phenomenon


in detail, confining attention to the rectangular elements.

7.3.1 Plane stress model of beam flexure

Figure 7.5 shows a two-dimensional plane stress description of the problem of


beam flexure. The beam is of length L, depth T and thickness (in the normal to
the plane direction) b. Two independent field variables are required to describe
the problem, the displacements u and v (Fig. 7.5). The strain energy functional
for this problem consists of energies from the normal strains, UE and shear
strains, UG. For an isotropic case, where E is the Young's modulus, G is the
shear modulus and ν is Poisson's ratio,

U = U E + UG

)( )
L T é ù
Eb
( )
= ò0 ò0 ê
(
êë 2 1 - ν 2
u 2 ,x + v2 ,y +2ν u,x v , y + Gb 2 u, y +v ,x 2 ú dx dy
úû
(7.14)

A one-dimensional beam theory can be obtained by simplifying this problem. For a


very slender beam, the shear strains must vanish and this would then yield the
classical Euler-Bernouilli beam theory. Note that although shear strains vanish,
shear stresses and shear forces would remain finite. We shall see now that it is
this constraint condition that leads to difficulties in plane stress finite
element modeling of such problems.

7.3.2 The 4-node rectangular element

An element of length 2l and depth 2t is chosen. The rectangular form is chosen


so that the issues involved can be identified clearly. There are now 8 degrees
of freedom, u1-u4 and v1-v4 at the four nodes. The field-variables are now
interpolated in the following fashion:

88
u = a0 + a1x + a2 y + a3 xy (7.15a)
v = b0 + b1x + b2 y + b3 xy (7.15b)

We need to examine only the shear strain field that will become
constrained (i.e. become vanishingly small) while modeling the flexural action
of a thin beam. This will be interpolated as,

γ = u,y +v,x = (a2 + b1 ) + a3 x + b3 y (7.16)

and the shear strain energy within the element will be,

( )2 dx dy
t
U e (G ) = ò-l ò Gb 2 a2 + b1 + a3 x + b3 y
l

[(a ]
-t

= (Gb 2 ) (4tl ) 2 + b1 )2 + (a3l )2 3 + (b3t )2 3 (7.17)

Equation (7.17) suggests that in the thin beam limit, the following constraints
will emerge:

a2 + b1 → 0 (7.18a)
a3 → 0 (7.18b)
b3 → 0 (7.18c)

We can see that Equation (7.18a) is a constraint that comprises constants from
both u and v functions. This can enforce the true constraint of vanishing shear
strains in a realistic manner. In contrast, Equations (7.18b) and (7.18c) impose
constraints on terms from one field function alone in each case. These are
undesirable constraints. We shall now see how these lead to the stiffening
effect that is called parasitic shear.

In the plane stress modeling of a slender beam, we would be using elements


which are very long in the x direction, i.e. elements with l>t would be used.
One can then expect from Equation (7.17) that as l >>t , the constraint a3→0
2 2

will be enforced more rapidly than the constraint b3→0 as the beam becomes
thinner. The spurious energies generated from these two terms will also be in a
similar proportion. In Equation (7.17), one can interpret the shear strain field
to comprise a constant value, which is `field-consistent' and two linear
variations which are `field-inconsistent'. We see now that the variation along
the beam length is the critical term. Therefore, let us consider the use of the
shear strain along y=0 as a measure of the averaged shear strain across a
vertical section. This gives

γ = (a2 + b1 ) + a3 x (7.19)

Thus, along the length of the beam element, γ have a constant term that reflects
the averaged shear strain at the centroid, and a linear oscillating term along
the length, which is related to the spurious constraint in Equation (7.18b). As
in the linear beam element in Section 4.1.6, this oscillation is self-
equilibrating and does not contribute to the force equilibrium over the element.
However, it contributes a finite energy in Equation (7.17) and in the modeling

89
of very slender beams, this spurious energy is so large as to completely
dominate the model of the beam behavior and cause a locking effect. We shall now
use the functional re-constitution technique to determine the magnitude of
locking and the extent of the shear stress oscillations triggered off by the
inconsistency in the shear strain field.

We shall simplify the present example and derive the error estimates from
the functional re-constitution exercise. We shall denote by a3 , the coefficient
determined in a solution of the beam problem by using a field-consistent
representation of the total energy in the beam, and by a3 , the value from a
solution using energies based on inconsistent terms. The field-consistent and
field-inconsistent terms will differ by,

(
a3 a3 = 1 + Gl 2 Et 2 ) (7.20)

(
This factor, e L = 1 + Gl2 Et2) is the additional stiffening parameter that
determines the extent of parasitic shear.

We can relate a3 to the physical parameters relevant to the problem of a


beam in flexure to show that if e L >> 1 , as in a very slender beam, there are
shear force oscillations,

V = V0 + (3M 0 l ) (x l ) (7.21)

The oscillations are proportional to the centroidal moment, and for e L >> 1 , the
oscillations are inversely proportional to the element length, i.e. they become
more severe for smaller lengths of the plane stress element. It is important to
note that a field-inconsistent two node beam element based on linear
interpolations for w and θ (see Equation (7.19)) produces an identical stress
oscillation.

7.3.3 The field-consistent elements

It is well known that the plane stress elements behave reliably in general 2D
applications where no constraints are imposed on the normal or shear strain
fields. The need for field-consistency becomes important only when a strain
field is constrained. Earlier, we saw that in the plane stress modeling of thin
beam flexure, the shear strain field is constrained and that these constraints
are enforced by a penalty multiplier Gl2 Et2 . The larger this term is, the more
severely constrained the shear strain becomes. Of the many strategies that are
available for removing the spurious constraints on the shear strain field, the
use of a variationally correct re-distribution is recommended. This requires the
field-consistent γ to be determined from γ using the orthogonality condition

ò δγ
T
(γ − γ ) dx dy = 0 (7.22)

Thus instead of Equation (7.16), we can show that the shear strain-field
for the 4-node rectangular plane stress element will be,

90
γ = (a2 + b1 ) (7.23)

It is seen that the same result is achieved if reduced integration using a 1


Point Gaussian integration is applied to shear strain energy evaluation. In a
very similar fashion, the consistent shear strain field can be derived for the
quadratic elements.

7.3.4 Concluding remarks on the rectangular plane stress elements

The field-consistency principle and the functional re-constitution technique can


be applied to make accurate error analyses of conventional 4-node and 8-node
plane stress elements. These a priori estimates have been confirmed through
numerical experiments.

A field-consistent re-distribution strategy allows these elements to be


free of locking and spurious stress-oscillations in modeling flexure. These
options can be built into a shape function sub-routine package in a simple
manner so that where field-consistency requirements are paramount, the
appropriate smoothed shape functions will be used.

7.3.5 Solid element model of beam/plate bending

So far, we have dealt with the finite element modeling of problems which were
simplified to one-dimensional or two-dimensional descriptions. There remains a
large class of problems which need to be addressed directly as three dimensional
states of stress. Solid or three dimensional elements are needed to carry out
the finite element modeling of such cases. A variety of 3D solid elements exist,
e.g. tetrahedral, triangular prism or hexahedral elements. In general cases of
3D stress analysis, no problems are encountered with the use of any of these
elements as long as a sufficiently large number of elements are used and no
constrained media limits are approached. However, under such limits the
tetrahedral and triangular prism elements cannot be easily modified to avoid
these difficulties. It is possible however to improve the hexahedral 8-node and
27-node elements so that they are free of locking and of other ill-effects like
stress oscillations. One problem encountered is that of parasitic shear or shear
locking when solid elements are used to model regions where bending action is
predominant. In fact, in such models, parasitic shear and shear locking merge
indistinguishably into each other. We briefly examine this below.

The 8-noded brick element is based on the standard tri-linear


interpolation functions and is the three dimensional equivalent of the bi-linear
plane stress element. It would appear now that the stiffness matrix for the
element can be derived very simply by carrying out the usual finite element
operations, i.e. with strain-displacement matrix, stress-strain matrix and
numerically exact integration (i.e. 2×2×2 Gaussian integration). Although this
element performs very well in general 3D stress analysis, it failed to represent
cases of pure bending (parasitic shear) and cases of near incompressibility
(incompressible locking).

The element is unable to represent the shear strains in a physically


meaningful form when the shear strains are to be constrained, as in a case of
pure bending. The phenomenon is the 3D analogue of the problem we noticed for

91
the bi-linear plane stress element above. The problem of designing a useful 8-
node brick element has therefore received much attention and typical attempts to
alleviate this have included techniques such as reduced integration, addition of
bubble functions and assumed strain hybrid formulations. We shall now look at
this from the point of view of consistency of shear strain interpolation.

We can observe here that the shear strain energy is computed from
Cartesian shear strains γ xy , γ xz and γ yz . In the isoparametric formulation,
these are to be computed from terms involving the derivatives of u, v, w with
respect to the natural co-ordinates ξ, η, ζ and the terms from the inverse of
the three dimensional Jacobian matrix. It is obvious that it will be a very
difficult, if not impossible task, to assure the consistency of the
interpolations for the Cartesian shear strains in a distorted hexahedral.
Therefore attention is confined to a regular hexahedron so that consistency
requirements can be easily examined.

To facilitate understanding, we shall restrict our analysis to a rectangular


prismatic element so that the (x, y, z) and (ξ, η, ζ) systems can be used
interchangeably. We can consider the tri-linear interpolations to be expanded in
the following form, i.e.

u = a1 + a2 x + a3 y + a4z + a5 xy + a6 yz + a7 xz + a8 xyz (7.24a)


v = b1 + b2 x + b3 y + b4z + b5 xy + b6 yz + b7 xz + b8 xyz (7.24b)
w = c1 + c2 x + c3 y + c4z + c5 xy + c6 yz + c7 xz + c8 xyz (7.24c)

where a1 to a8 and b1 to b8 are related to the nodal degrees of freedom u1 to u8


and v1 to v8 respectively. We shall now consider a case of a brick element
undergoing a pure bending response requiring a constrained strain field
γ xy = u,y +v ,x → 0 . From Equations (7.24a) and (7.24b) we have,

γ xy = (a3 + b2 ) + (a6 + b7 )z + a5 x + b5 y + a8 xz + b8 yz (7.25)

When Equation (7.25) is constrained, the following spurious constraint


conditions appear: a5→0; b5→0; a8→0 and b8→0.

Our arguments projecting these as the cause of parasitic shear would be


proved beyond doubt if we can derive a suitable error model. These error models
are now much more difficult to construct, but have been achieved using a
judicious mix of a priori and a posteriori knowledge of zero strain and stress
conditions to simplify the problem. These error models have been convincingly
verified by numerical computations [7.1].

A field-consistent representation will be one that ensures that the


interpolations for the shear fields will contain only the consistent terms. A
simple way to achieve this is to retain only the consistent terms from the
original interpolations. Thus, for γxy we have,

γ xy = (a3 + b2 ) + (a6 + b7 )z (7.26)

92
Similarly, the other shear strain fields can be modified. These strain fields
should now be able to respond to bending modes in all three planes without
locking.

7.4 Incompressible locking

Conventional displacement formulations of 2D plane strain and 3D elasticity fail


when Poisson's ratio ν→0.5, i.e. as the material becomes incompressible. Such
situations can arise in modeling of material such as solid rocket propellants,
saturated cohesive soils, plastics, elastomers and rubber like materials and in
materials that flow, e.g. incompressible fluids or in plasticity. Displacement
fields lock and highly oscillatory stresses are seen - drawing an analogy with
shear and membrane locking, we can describe the phenomenon as incompressible
locking. In fact, a penalty term (bulk modulus→∞ as ν→0.5) is easily
identifiable as enforcing incompressibility. Since this multiplies the
volumetric strain, it is easy to argue that a lack of consistency in the
definition of the volumetric strain will cause locking and will appear as
oscillatory mean stresses or pressures.

7.4.1 A simple one-dimensional case - an incompressible hollow sphere

It is instructive to demonstrate how field-consistency operates in the finite


element approximation of a problem in incompressible (or nearly-incompressible)
elasticity by taking the simplest one-dimensional case possible. A simple
problem of a pressurized elastic hollow sphere serves to illustrate this.

Consider an elastic hollow sphere of outer radius b and inner radius a


with an internal pressure P. If the shear modulus G is 1.0, the radial
displacement field for this case is given by,

é 1
u=ρ ê +
(1 − 2ν ) r ù
ú (7.27)
ë 4r 2 2 (1 + ν ) û

[(
where ρ = Pba3 G b 3 − a3 )] and r=R/b non-dimensional radius and ν is the Poisson's
ratio.

A finite element approximation approaches the problem from the minimum


total potential principle. Thus, a radial displacement u leads to the following
strains:

ε r = u,r εθ = u r εφ = u r

and the volumetric strain is

ε = ε r + ε θ + ε φ = u,r + 2u r

The elastic energy stored in the sphere (when G=1) is

(ε )
1
U = ò
2
r + ε θ2 + ε φ2 + 1 2 α ε 2 r 2 dr (7.28)
a

93
and α = 2ν (1 − 2ν ) when G=1 and a = a b is the dimensionless inner radius. The
displacement type finite element representation of this very simple problem runs
into a lot of difficulties when ν→0.5, i.e., when α→∞ in the functional
represented by Equation (7.28).

These arguments now point to the volumetric strain field in Equation


(7.28) as the troublesome term. When α→∞, this strain must vanish. In a
physical situation, and in an infinitesimal interpretation of Equation (7.28),
this means that ε→0. However, when a finite element discretization is
introduced, this does not happen so simply. To see this, let us examine what
will happen when a linear element is used to model the problem.

7.4.2 Formulation of a linear element

The displacement field u and the radial distance r are interpolated by


linear isoparametric functions as

r = 0.5 (r1 + r2 ) + 0.5 (r2 − r1 )ξ (7.29a)


u = 0.5 (u1 + u2 ) + 0.5 (u2 − u1 )ξ (7.29b)

where the nodes of the element are at r1 and r2.

The troublesome term in the potential energy functional is the volumetric


strain and this contribution can be written as,

òε r dr = ò (ru,r +2u )
2 2 2
dr (7.30)

It is the discretized representation of this term (ru,r +2u ) that will


determine whether there will be locking when α→∞. Thus, unlike in previous
sections, the constraint is not on the volumetric strain ε but on rε because of
the use of spherical coordinates to integrate the total strain energy. It is
therefore meaningful to define this as a pseudo-strain quantity γ=rε. To see how
this is inconsistently represented, let us now write the interpolation for this
derived kinematically from the displacement fields as,

é (r1 + r2 ) ù
(ru,r +2u ) = ê (u2 − u1 ) + (u1 + u2 )ú + [3 2 (u2 − u1 )] ξ (7.31)
ë 2 (r2 − r1 ) û

Consider now the constraints that are enforced when this discretized field
is used to develop the strain energy arising from the volumetric strain as shown
in Equation (7.31) when α→∞.

(r1 + r2 ) (
u2 − u1 ) + (u1 + u2 ) → 0 (7.32a)
2 (r2 − r1 )

(u2 − u1 ) → 0 (7.32b)

94
Condition (7.32a) represents a meaningful discretized version of the
condition of vanishing volumetric strain and is called a true constraint in our
field-consistency terminology. However, condition (7.32b) leads to an undesired
restriction on the u field as it implies u,r → 0 . This is therefore the spurious
constraint that leads to locking and spurious pressure oscillations.

7.4.3 Incompressible 3D elasticity

We now go directly to modeling with 3D elasticity. We re-write the strain energy


for 3D elasticity in the following form:

U =1 2G ò ε d D ε d dV + 1 2 λ ò ε n ε n dV
T T
(7.33)

In Equation (7.33), εd is the distortional strain and εn is the volumetric


strain. In this form, it is convenient to describe the elasticity matrix in
terms of the shear modulus, G and the bulk modulus K, where G = E [2 (1 + ν )] ,
K = E [3 (1 - 2ν )] and λ = (K − 2G 3 ) = E ν [(1 + ν ) (1 - 2ν )] .
Consider the fields given in Equations (7.24) for the 8-noded brick
element. An elastic response for incompressible or nearly-incompressible
materials will require a constrained strain field,

ε n = u,x +v , y +w,z → 0 (7.34)

where εn is the volumetric strain. From Equations (7.24) and (7.34) we have,

ε n = (a2 + b3 + c4 ) + (b5 + c7 )x + (a5 + c6 )y + (a7 + b6 )z + (a8 yz + bz xz + c8 xy ) (7.35)

Equation (7.35) can be constrained to zero only when the coefficients of each of
its terms vanish, giving rise to the constraint conditions,

a2 + b3 + c4 → 0 (7.36a)
b5 + c7 → 0 (7.36b)
a5 + c6 → 0 (7.36c)
a7 + b6 → 0 (7.36d)
a8 → 0 (7.36e)
b8 → 0 (7.36f)
c8 → 0 (7.36g)

Equation (7.36b) to (7.36) are the inconsistent terms that cause locking. To
ensure a field-consistent representation, only the consistent terms should be
retained, giving

ε n = a2 + b3 + c4 (7.37)

This field will now be able to respond to the incompressible or nearly


incompressible strain states and the element should be free of locking.

95
7.4.4 Concluding remarks

It is clear from this section that incompressible locking is analogous to shear


locking, etc. A robust formulation must therefore ensure consistent
representations of volumetric strain in cases where incompressibility or near
incompressibility is expected.

7.5 References

7.1 G. Prathap, The Finite Element Method in Structural Mechanics, Kluwer


Academic Press, Dordrecht, 1993.

96
Chapter 8

Stress consistency
8.1 Introduction

In Chapters 6 to 7 we examined the difficulties experienced by the displacement


approach to the finite element formulation of problems in which some strain
fields are constrained. To obtain accurate solutions at reasonable levels of
discretization, it was necessary to modify these strain fields and use these in
computing the stiffness matrix and also in stress recovery. The criterion
governing the relationship between the various terms in the modified strain
field interpolation was described as consistency - i.e. the strain field
interpolations must maintain a consistent balance internally of its contributing
terms. This allows the constraints that emerge after discretization to remain
physically faithful to the continuum problem. This was the rule that guided the
locking-free design of all elements discussed so far.

In this chapter, we take a look at a class of problems where no


constraints are imposed on the strains but there is a need to relax the
satisfaction of the constitutive relationship linking discretised stress to
discretised strain so that again a certain degree of consistency is maintained.
Such situations develop where there are structural regions in which the rigidity
varies spatially due to varying elastic moduli or cross-sectional area and also
in initial strain problems, of which the thermal strain problem is the most
commonly encountered.

In structural regions with varying rigidity the spatial variation of


strain-fields and stress or stress-resultant fields will not match. In the
discretization of such cases, it is necessary to consider a form of external
consistency requirement between the discretized strain fields and the
discretized stress or stress-resultant fields. This is necessary so that a
correct interpretation of computed stresses and stress resultants is possible;
otherwise, oscillations will be seen.

In initial strain problems, the initial strain variation and the total
strain variations may be of different order. This is a familiar problem in
finite element thermal stress analysis. Here, it is necessary to obtain
kinematically equivalent thermal nodal forces from temperature fields and also
ensure that the discretised thermal strains corresponding to this are consistent
with the discretised description of total strains. Again, one must carefully
identify the conflicting requirements on the order of discretised functions to
be used.

8.2 Variable moduli problems

8.2.1 A tapered bar element

We shall now conduct a simple numerical experiment with a tapered bar element to
show that the force field computed directly from strains in a structural element
of varying sectional rigidities has extraneous oscillations. A linear element
would have sufficed to demonstrate the basic principles involved. However, we
use a quadratic tapered bar element so that the extraneous oscillations, which

97
for a general case can be of cubic form, are not only vividly seen but also need
special care to be reduced to its consistent form. In a linear element, this
exercise becomes very trivial, as sampling at the centroid of the element gives
the correct stress resultant.

We shall consider an isoparametric formulation for a quadratic bar element


of length 2l with mid-node exactly at the mid-point of the element. Then the
interpolations for the axial displacement u and the cross sectional area A in
terms of their respective nodal values are,

u = u2 + (u3 − u1 ) 2 ξ + (u1 − 2u2 + u3 ) 2 ξ 2


A = A 2 + (A 3 − A1 ) 2 ξ + (A1 − 2A 2 + A3 ) 2 ξ2

We shall first examine how the element stiffness is formulated when the minimum
total potential principle is used. We start with a functional written as,

π = 1 2 ò N T ε dx − W
where,

ε = du dx the axial strain,


N = E A (x ) ε the kinematically constituted axial force,
W = ò pu dx the potential of external forces,
p distributed axial load,
E Young's modulus of elasticity

After discretization, ε will be a linear function of ξ but N will be a


cubic function of ξ. The strain energy of deformation is then expressed as,

U =1 2 òN ε dx
T

From this product the terms of the stiffness matrix emerge. Due to the
orthogonal nature of the Legendre polynomials, terms from N which are linked
with the quadratic and cubic Legendre polynomials, N3 and N4 respectively, will
not contribute to the energy and therefore will not provide terms to the
stiffness matrix! It is clear that, the displacements recovered from such a
formulation cannot recognize the presence of the quadratic and cubic terms N3
and N4 in the stress field N as these have not been accounted for when the
stiffness matrix was computed. Hence, in a displacement type finite element
formulation, the stresses recovered from the displacement vector will have
extraneous oscillations if N3 and N4 are not eliminated from the stress field
during stress recovery. We shall designate by BAR3.0, the conventional element
using N for stiffness matrix evaluation and recovery of force resultant.

Next, we must see how the consistent representation of the force field
denoted by N must be made. N should comprise only the terms that will
contribute to the stiffness and strain energy and the simplest way to do this is
to expand the kinematically determined N in terms of Legendre polynomials, and
retain only terms that will meaningfully contribute to the energy in (N ε ) .
T

98
Thus, N must be consistent with ε, i.e. in this case, retain only up to linear
terms:
N = N 1 + N 2ξ (8.1)

Such an element is denoted by BAR3.1. It uses N for stiffness matrix evaluation


and recovery of forces resultant. To see the variational basis for the procedure
adopted so far, the problem is re-formulated according to the Hu-Washizu
principle which allows independent fields for assumed strain and assumed stress
functions.

8.2.2 Numerical experiments

We shall perform the computational exercises with the two versions of the
element; note that in both cases, the stiffness matrices and computed
displacements are identical. Figure 8.1 shows a tapered bar clamped at node-1
and subjected to an axial force P at node-3. The taper is defined by the
parameters,
α = (A 3 − A1 ) 2A 2 and β = (A1 − 2A 2 + A 3 ) 2A 2

Finite element results from a computational exercise using the two versions
described above for a bar with cross section tapering linearly from the root to
the tip for which β=0 are obtained. Thus α is given by,

α = (A 3 − A1 ) (A1 + A3 )

Thus α can vary from 0 to –1.0. Fig. 8.2 shows the axial force patterns
obtained from the finite element digital computation for a case with α=-0.9802
(with A3=0.01 A1) It can be noted here that the results are accurate at
ξ = ±1 3.

Fig. 8.1 A cantilever bar modeled with a single element.

99
Fig. 8.2 Axial force pattern for linearly tapered bar (α=0.9802 and β=0.0) with
A3=0.01 A1.

A general case of taper is examined next where A1=1.0, A2=0.36 and


A3=0.04, and the area ratios are α=-4/3 and β=4/9. Fig. 8.3 shows the results
from the finite element computations. Due to the presence of both quadratic and
cubic oscillations, there are no points which can be easily identified for
accurate force recovery! Therefore it is necessary to perform a re-constitution
of the force resultant fields on a consistency basis using the orthogonality
principle as done here before reliable force recovery can be made.

8.2.3 Reconstitution of the stress-resultant field using the Hu-Washizu


principle

In forming the Hu-Washizu functional for the total potential, an as yet


undetermined assumed force function N is introduced but the assumed strain
field ε can be safely retained as ε (note that in a constrained media problem
it will be required to introduce a field-consistent ε that will be different
from the kinematically derived and therefore usually field-inconsistent ε) - the
functional now becomes

π = ò {1 2 (EAε ) }
ε + N T (ε − ε ) dx − W
T

100
Fig. 8.3 Axial force pattern for bar with combined linear and quadratic taper
(α=-4/3 and β=4/9).

A variation of the Hu-Washizu energy functional with respect to the


kinematically admissible degree of freedom u, gives the equilibrium equation,

ò δu {dN dx − p} dx = 0
T

Variation with respect to the assumed strain field ε gives rise to a


constitutive relation

ò δε
T
{- N + EA ε } dx = 0 (8.2)

and variation with respect to the assumed force field N gives rise to the
condition

ò δN
T
{ε - ε } dx = 0 (8.3)

Now Equations (8.2) and (8.3) are the orthogonality conditions required
for reconstituting the assumed fields for the stress resultant and the strain.
The consistency paradigm suggests that the assumed stress resultant field N

101
should be of the same order as the assumed strain field ε . Then Equation (8.3)
gives the orthogonality condition for strain field-redistribution.

On the other hand, orthogonality condition (8.2) can be used to


reconstitute the assumed stress-resultant field N from the kinematically
derived field N. Now, Equation (8.2) can be written as,

ò δε
T
{N - N } dx = 0 (8.4)

Thus, if N is expanded in terms of Legendre polynomials, it can be proved


that N which is consistent and orthogonally satisfies Equation (8.4) is
obtained very simply by retaining all the Legendre polynomial terms that are
consistent with ε , i.e. as shown in Equation (8.1). Thus the procedure adopted
in the previous section has variational legitimacy.

8.3 Initial strain/stress problems

Finite element thermal stress analysis requires the formulation of what is


called an initial strain problem. The temperature fields which are imposed must
be converted to discretised thermal (initial) strains and from this
kinematically equivalent thermal nodal forces must be computed. The usual
practice in general purpose codes is to use the same shape functions to
interpolate the temperature fields and the displacement fields. Thermal stresses
computed directly from stress-strain and strain-displacement matrices after the
finite element analysis is performed thus show large oscillating errors. This
can be traced to the fact that the total strains (which are derived from
displacement fields) are one order higher than the thermal strains (derived from
temperature fields). Some useful rules that are adopted to overcome this
difficulty are that the temperature field used for thermal stress analysis
should have the same consistency as the element strain fields and that if
element stresses are based on Gauss points, the thermal stresses should also be
based on these Gauss point values. This strategy emerged from the understanding
that the unreliable stress predictions originate from the mismatch between the
element strain ε and the initial strain due to temperature ε0. We shall now show
that this is due to the lack of consistency of their respective interpolations
within the element.

Earlier in this chapter, we saw that stress resultant fields computed from
strain fields in a displacement type finite element description of a domain with
varying sectional rigidities showed extraneous oscillations. This was traced to
the fact that these stress resultant fields were of higher interpolation order
than the strain fields and that the higher degree stress resultant terms did not
participate in the stiffness matrix computations. In this section, we show that
much the same behavior carries over to the problem of thermal stress
computations.

8.3.1 Description of the problem

With the introduction of initial strains ε0 due to thermal loading, the stress
to strain relationship has to be written as

102
σ = D (ε - ε0 ) = Dε m (8.5)

The strain terms now need to be carefully identified. {ε} is the total strain
and {εm} is the mechanical or elastic strain. The free expansion of material
produces initial strains
ε0 = α T (8.6)

where T is the temperature relative to a reference value at which the body is


free of stress and α is the coefficient of thermal expansion. The total strains
(i.e. the kinematically derived strains) are defined by the strain-displacement
matrix,
{ε}=[B]{d} (8.7)

where {d} is the vector of nodal displacements. In a finite element description,


the displacements and temperatures are interpolated within the domain of the
element using the same interpolation functions. The calculation of the total
strains {ε} involves the differentiation of the displacement fields and the
strain field functions will therefore be of lower order than the shape
functions. The initial strain fields (see Equation (8.6)) involve the
temperature fields directly and this is seen to result in an interpolation field
based on the full shape functions. The initial strain matrix is of higher degree
of approximation than the kinematically derived strain fields if the temperature
fields can vary significantly over the domain and are interpolated by the same
isoparametric functions as the displacement fields. It is this lack of
consistency that leads to the difficulties seen in thermal stress prediction.
This originates from the fact that the thermal load vector is derived from a
part of the functional of the form,

ò δ {ε } D {ε0 } dV
T
(8.8)

Again, the problem is that the `higher order' components of the thermal (or
initial) stress vector are not sensed by the total strain interpolations in the
integral shown above. In other words, the total strain terms "do work" only on
the consistent part of the thermal stress terms in the energy or virtual work
integral. Thus, a thermal load vector is created which corresponds to a initial
strain (and stress) vector that is `consistent' with the total strain vector.
The finite element displacement and total strain fields which are obtained in
the finite element computation then reflect only this consistent part of the
thermal loading. Therefore, only the consistent part of the thermal stress
should be computed when stress recovery is made from the nodal displacements;
the inclusion of the inconsistent part, as was done earlier, results in thermal
stress oscillations.

We demonstrate these concepts using a simple bar element.

8.3.2 The linear bar element - Derivation of stiffness matrix and


thermal load vector
Consider a linear bar element of length 2l. The axial displacement u, the total
strain ε, the initial strain ε0 and stress σ are interpolated as follows:

u = (u1 + u2 ) 2 + ξ (u2 − u1 ) 2 (8.9a)

103
Fig. 8.4 Linear bar element.

ε = u,x = (u2 − u1 ) 2l (8.9b)

ε 0 = α {(T1 + T2 ) 2 + ξ (T2 − T1 ) 2} (8.9c)

σ = E (u2 − u1 ) 2l − Eα {(T1 + T2 ) 2 + ξ (T2 − T1 ) 2} (8.9d)

where ξ =x l (see Fig.8.4), E is the modulus of elasticity, A the area of


cross-section of the bar and α the coefficient of expansion.

To form the element stiffness and thermal load vector, an integral of the
form

ò δε σ dx
T
(8.10)

is to be evaluated. This leads to a matrix equation for each element of the


form,

EA é 1 − 1ù ìu1 ü (T1 + T2 ) ì − 1ü ìF1 ü


ê ú í ý − EAα í ý=í ý (8.11)
2l ë− 1 1û î 2þ
u 2 î 1 þ îF2 þ

where F1 and F2 are the consistently distributed nodal loads arising from the
distributed external loading. By observing the components of the interpolation
fields in Equations (8.9b) to (8.9d) carefully (i.e. constant and linear terms)
and tracing the way they participate in the `work' integral in Equation (8.10),
it is clear that the (T2 − T1 ) 2 term associated with the linear (i.e. ξ) term in
σ (originating from ε0) cannot do work on the constant term in δεT and therefore
vanishes from the thermal load vector; see Equation (8.11). Thus the equilibrium
equations that result from the assembly of the element equilibrium equations
represented by Equation (8.11) will only respond to the consistent part of the
initial strain and will give displacements corresponding to this part only.

If these computed displacements are to be used to recover the initial


strains or thermal stresses, only the consistent part of these fields should be
used. The use of the original initial strain or stress fields will result in
oscillations corresponding to the inconsistent part. We shall work these out by

104
Fig. 8.5 Clamped bar subject to varying temperature

hand using a simple example below and compare it with the analytical solution.

8.3.3 Example problem

Figure 8.5 shows a bar of length L=4l clamped at both ends and subjected to a
varying temperature field. We shall consider a case where two conventionally
derived linear elements are used, so that we require the nodal temperatures T1,
T2 and T3 as input. The nodal reactions are F1 and F3 (corresponding to the
clamped conditions u1=u3=0). We have the assembled equations as,

é 1 −1 ù ì 0 ü ìF1 ü ì − (T1 + T2 )ü
EA ê ú ï ï ï ï EAα ï ï
−1 2 −1 ú íu2 ý = í 0 ý + í (T1 − T3 ) ý (8.12)
2l ê ï 0 ï ïF ï 2 ï (T + T )ï
êë −1 1úû î þ î 3þ î 2 3 þ

From this, we can compute the displacements and nodal reactions as,

u2 = αl (T1 − T3 ) 2 and F1 = −F3 = EAα (T1 + 2T2 + T3 ) 4 (8.13)

and these are the correct answers one can expect with such an idealization. If
the stress in the bar is computed from the nodal reactions, one gets a constant
stress σ = −Eα (T1 + 2T2 + T3 ) 4 , in both elements, which is again the correct answer
one can expect for this idealization. It is a very trivial exercise to show
analytically that a problem in which the temperature varies linearly from 0 to T
at both ends will give a constant stress field σ = − EαT 2 , which the above model
recovers exactly.

Problems however appear when the nodal displacement computed from


Equations (8.12) is used in Equations (8.9b) to (8.9d) to compute the initial
strains and stresses in each element. We would obtain now, the stresses as
(subscripts 12 and 23 denote the two elements)

σ 12 = −Eα (T1 + 2T2 + T3 ) 4 − ξ Eα (T2 − T1 ) 2 (8.14a)

σ 23 = −Eα (T1 + 2T2 + T3 ) 4 − ξ Eα (T3 − T2 ) 2 (8.14b)

It is now very clear that a linear oscillation is introduced into each element
and this corresponds to that inconsistent part of the initial strain or stress
interpolation which was not sensed by the total strain term. This offers us a
straightforward definition of what consistency is in this problem - retain only
that part of the stress field that will do work on the strain term in the

105
functional. To see how this part can be derived in a variationally correct
manner, we must proceed to the Hu-Washizu theorem.

8.3.4 A note on the quadratic bar element

We may note now that if a quadratic bar element had been the basis for the
finite element idealization, the total strains would have been interpolated to a
linear order; the initial strain and thermal stress field will now have a
quadratic representation (provided the temperature field has a quadratic
( )
variation) and the inconsistency will now be of the 1 − 3ξ 2 type; thus thermal
stresses derived using a formal theoretical basis will show these quadratic
oscillations which will vanish to give correct answers at the points
corresponding to ξ = ± 1 3 ; i.e. the points corresponding to the 2-point Gauss
integration rule.

8.3.5 Re-constitution of the thermal strain/stress field using the Hu-


Washizu principle

We now seek to find a variational basis for the use of the re-constituted
consistent initial strain and stress interpolations in the Hu-Washizu principle.

The minimum total potential principle states the present problem as, find
the minimum of the functional,

Π MTP = ò {σ }
ε m 2 + P dV
T
(8.15)

where σ and εm are as defined in (8.5) to (8.7) and the displacement and strain
fields are interpolated from the nodal displacements using the element shape
functions and their derivatives and P is the potential energy of the prescribed
loads.

We now know that the discretized stress field thus derived, σ, is


inconsistent to the extent that the initial strain field ε0 is not of the same
order as the total strain field ε. It is therefore necessary to reconstitute the
discretized stress field into a consistent stress field σ without violating any
variational norm. In the examples above, we had seen a simple way in which this
was effected.

To see how we progress from the inconsistent discretized domain (i.e.


involving ε0 and σ) to the consistent discretized domain (i.e. introducing ε0
and σ , it is again convenient to develop the theory from the generalized Hu-
Washizu mixed theorem. We shall present the Hu-Washizu theorem from the point of
view of the need to re-constitute the inconsistent ε0 to a consistent ε0 without
violating any variational norms. We proceed thus:

Let the continuum linear elastic problem have a discretized solution based
on the minimum total potential principle described by the displacement field u,
strain field ε and stress field σ (we project that the strain field ε is derived
from the displacement field through the strain-displacement gradient operators

106
of the theory of elasticity and that the stress field σ is derived from the
strain field ε through the constitutive laws as shown in (8.2). Let us now
replace the discretized domain by another discretized domain corresponding to
the application of the Hu-Washizu principle and describe the computed state to
be defined by the quantities ε, ε0 and σ , where again, we take that the stress
fields σ are computed using the constitutive relationships, i.e. σ = D (ε − ε0 ) .
It is clear that ε0 is an approximation of the strain field ε0. Note that we
also argue that we can use ε = ε as there is no need to introduce such a
distinction here (in a constrained media elasticity problem it is paramount that
ε be derived as the consistent substitute for ε.)

What the Hu-Washizu theorem does, following the interpretation given by de


Veubeke, is to introduce a "dislocation potential" to augment the usual total
potential. This dislocation potential is based on a third independent stress
field σ which can be considered to be the Lagrange multiplier removing the lack
of compatibility appearing between the kinematically derived strain field ε0 and
the independent strain field ε0 . The three-field Hu-Washizu theorem can be
stated as,

δΠ HW = 0 (8.16)

where

Π HW = ìíσ T ε m 2 + σ
ò
T
(ε m − ε m ) + P üý dV (8.17)
î þ

where ε m = (ε − ε0 ) . At this stage we do not know what σ or ε0 are except that


they are to be of consistent order with ε.

In the simpler minimum total potential principle, which is the basis for
the derivation of the displacement type finite element formulation in most
textbooks, only one field (i.e. the displacement field u), is subject to
variation. However, in this more general three field approach, all three fields
are subjected to variation and leads to three sets of equations which can be
grouped and classified as follows:

Variation on Nature Equation,

u Equilibrium ∇ σ + terms from P=0 (8.18a)

σ Orthogonality òδσ
T
(ε0 − ε0 ) dV =0 (8.18b)
(Compatibility)
T
æ σ − σ ö dV = 0
ε0 Orthogonality ò δε0 ç
è
÷
ø
(8.18c)
(Equilibrium)

107
Fig. 8.6 (a) Clamped bar under linear temperature variation and its (b) Bar
element model, (c) Plane stress model.

Let us first examine the orthogonality condition in (8.18c). We can


interpret this as a variational condition to restore the equilibrium imbalance
between σ and σ . In this instance this condition reduces to σ = σ . Note that
in a problem where the rigidity modulus D can vary significantly over the
element volume, this condition allows σ to be reconstituted from σ in a
consistent way.

The orthogonality condition in (8.18b) is now very easily interpreted.


Since we have σ = σ consistent with ε, this condition shows us how to smooth ε0
to ε0 to maintain the same level of consistency as ε. This equation therefore
provides the variationally correct rule or procedure to determine consistent
thermal strains and stresses from the inconsistent definitions.

Therefore, we now see how the consistent initial strain field ε0 can be
derived from the inconsistent initial strain field ε0 without violating any
variational norm. We thus have a variationally correct procedure for re-
constituting the consistent initial strain field - for the bar element above,
this can be very trivially done by using an expansion of the strain fields in
terms of the orthogonal Legendre polynomials, as done in the previous section
for the consistent stress-resultants.

108
8.3.6 Numerical examples

In this section we take up the same example of a bar with fixed ends subjected
to linearly varying temperature along its axis and model it with the bar element
and a plane stress element for demonstrating the extraneous stress oscillations
resulting from the lack of consistency in initial strain definition.

A bar of length 8 units, depth 2 units subjected to a linear temperature


distribution as shown in Fig 8.6a is modeled first with two bar elements (see
Fig 4. 18.6b). Let BAR.0 and BAR.1 represent the bar element versions with
inconsistent and consistent thermal stress fields. As already predicted, both
give accurate displacements (-0.002 units at the mid point), see Equation
(9.13)); BAR.1 gives exact stress throughout while BAR.0 shows linear
oscillations (see Equations (8.14a) and (8.14b)) as shown in Fig 8.7.

Fig.8.7 Consistent and inconsistent thermal stress recovery for a clamped bar
problem.

109
8.3.7 Concluding remarks

Before we close this section, it will be worthwhile to discuss the "correctness"


of the various approaches from the variational point of view. Note that Equation
(9.2) and (9.3) would have been fulfilled if an assumed stress-resultant field
of the same order as N had been used in place of the N of consistent order.
This would result in the Hu-Washizu formulation yielding the same stress
predictions as the minimum potential principle. Thus, both N and N are equally
"correct" in the Hu-Washizu sense, but only N is "correct" with the minimum
potential energy formulation. Where they differ is in the consistency aspect -
i.e. N is consistent with ε whereas N isn't. It follows that from the
theoretical point of view of the variational or virtual work methods, the
potential energy formulation and the Hu-Washizu formulation describe the
continuum physics in exactly the same way. Both the potential energy and Hu-
Washizu formulations are equally valid as far as variational correctness is
concerned. However, when approximations are introduced in the finite element
method, the Hu-Washizu formulation gives decidedly better results than the
potential energy formulation because of the flexibility it provides to modify
stress fields and strain fields to satisfy the consistency requirements in
situations where they play a prominent role. Note that the consistency paradigm,
in this case that used to justify the truncation of N (obtained directly from
the displacement approach) to a consistent N on the argument that the truncated
terms are not sensed in the discretized strain energy computations, lies clearly
outside the variational correctness paradigm. Similarly, in the earlier chapters
of this book, we saw another variation of the consistency paradigm - the need to
ensure a proper balance in the discretized representation of constrained strain
fields, which again lies outside the scope of the variational correctness
paradigm. It is very important to understand therefore that both consistency and
correctness paradigms are mutually exclusive but are needed together to ensure
that the finite element formulations are correctly and consistently done.

The stress field-consistency paradigm introduced here also applies to


finite element applications where certain terms in either the stress or strain
fields do not participate in the displacement recovery due to their inconsistent
representation e.g. initial stress/strain problems, problems with varying
material properties within an element etc. In the next section, we shall extend
this paradigm to extract consistent thermal stress and/or strains from a
displacement type formulation.

In this section, we have demonstrated another variation of the consistency


paradigm - in applications to evaluation of stresses and stress resultants in
finite element thermal stress computations. It is shown that such stresses must
be computed in a consistent way, recognizing that the displacement type finite
element procedure can sense only stresses which have the same consistency as the
total strain field interpolations used within each element. Thus, in an element
where the temperature field interpolations do not have the same consistency as
the strain fields, it is necessary to derive a consistent initial strain-field
for purposes of recovery of stresses from the computed displacement field. A
simple and variationally correct way to do this has been established from the
Hu-Washizu theorem.

110
Chapter 9

Conclusion

9.1 Introduction

Technology and Science go hand in hand now, each fertilizing the other and
providing an explosive synergy. This was not so in the beginning; as there was
Technology much before there was Science. In fact, there was technology before
there were humans and it was technology which enabled the human race to evolve;
science came later. The FEM is also an excellent example of a body of knowledge
that began as Art and Technology; Science was identified more slowly. This
chapter will therefore sum up the ideas that constitute the science of the
finite element method and also point to the future course of the technology.

9.2 The C-concepts: An epistemological summing-up

One would ideally like to find a set of cardinal or first principles that govern
the entire finite element discretisation procedure, ensuring the quality of
accuracy and efficiency of the method. At one time, it was believed that two
rules, namely continuity and completeness, provided a necessary and sufficient
basis for the choice of trial functions to initiate the discretisation process.
Chapter 3 of this book reviewed the understanding on these two aspects. However,
finite element practitioners were to be confronted with a class of problems
where the strict implementation of the continuity condition led to conflicts
where multiple strain fields were required. This was the locking problem which
we examined carefully in chapters 5 to 7. Let us recapitulate our fresh insight
in an epistemological framework below.

9.2.1 Continuity conflict and the consistency paradigm

We saw in chapters 5 to 7 that in a class of problems where some strain fields


need to be constrained, the implementation of the continuity rules on the
displacement trial functions blindly led to a conflict whereby excessive
constraining of some strain terms caused a pathological condition called
locking.

Let us again take up the simple problem of a Timoshenko beam element. Two
strain energy components are important; the bending energy which is constituted
from the bending strain as κ = θ ,x and the shear strain energy which is based on
the shear strain γ = θ − w,x . The simplistic understanding that prevailed for a
long time was that the existence of the θ ,x and w,x terms in the energy
expression demanded that the trial functions selected for the θ and w fields
must be C continuous; i.e. continuity of θ and w at nodes (or across edges in a
0

2D problem, etc.) was necessary and sufficient.

However, this led to locking situations. To resolve this conflict, it


became necessary to treat the two energy components on separate terms. The shear
strain energy was constrained in the regime of thin beam bending to become
vanishingly small. This required that the shear strains must vanish without

111
producing spurious constraints. The idea of field consistent formulation was to
anticipate these conditions on the constrained strain field. Thus for θ − w,x to
vanish without producing gratuitous constraints, the trial functions used for θ
and w in the shear strain field alone must be complete to exactly the same order
- this is the consistency paradigm. We can also interpret this as a multiple
continuity requirement. Note that although the bending strain requires θ to be
C continuous, in the shear strain component, we require θ not to have C
0 0
-1
continuity, i.e. say a C continuity, is. Thus, such a conflict is resolved by
operating with consistently represented constrained strain fields.

9.2.2 The correctness requirement

The consistency requirement now demands that when trial functions are chosen for
displacement fields, these fields (such as θ in γ = θ − w,x ) which need to be
constrained must have a reconstituted consistent definition. The minimum total
potential principle, being a single field variational principle, does not
provide a clue as to how this reconstitution of the strain fields can be
performed without violating the variational rules.

In the course of this book, we have come to understand that a multi-field


variational principle like the generalized Hu-Washizu theorem provided exactly
that flexibility to resolve the conflicting demands between consistency and
continuity. This, we called the correctness requirement - that the re-
constitution of the strain field which needed a reduced-continuity satisfying
displacement field from an inconsistent strain field kinematically derived from
the original continuous displacement fields must be done strictly according to
the orthogonality condition arising from the HW theorem.

9.2.3 The correspondence principle

The study so far with consistency and correctness showed the primacy of the Hu-
Washizu theorem in explaining how the finite element method worked. It also
became very apparent that if the internal working of the discretisation
procedure is examined more carefully, it turns out that it is strain and stress
fields which are being approximated in a "best-fit" sense and not the
displacement fields as was thought earlier. This brings us to the stress
correspondence paradigm, and the discovery that this paradigm can be axiomatised
from the Hu-Washizu theorem, as shown in chapter 2.

9.3 How do we learn? How must we teach? In a chronological sense or in a


logical sense?

In an epistemological sense, often, it is phenomena which first confronts us and


is systematically recognized and classified. The underlying first principles
that explain the phenomena are identified very much later, if not at all. The
clear understanding of first principles can also lead to the unraveling of new
phenomena that was often overlooked.

Thus, if and when the first principles are available, then the logical
sequence in which one teaches a subject will be to start with first principles
and derive the various facets of the phenomena from these principles. In
pedagogical practice, we can therefore have a choice from two courses of action:

112
teach in a chronological sense, explaining how the ideas unfolded themselves to
us, or teach in the logical sense, from first principles to observable facts. It
is not clear that one is superior to the other.

In my first book [9.1], I choose the chronological order. The conflicts


created by the locking phenomena were resolved by inventing the consistency
paradigm. This led to the correctness principle and the recognition that the Hu-
Washizu principle formed a coherent basis for understanding the fem procedure.
From these, one could axiomatise the correspondence rule - that the fem approach
manipulates stresses and strains directly in a best-approximation sense. Thus,
chapter 12 of Reference 9.1 sums up the first principles.

In the present book, I have preferred a logical sequence - the HW and


correspondence rules are taken up first and then the various phenomena like
locking, etc.

9.4 Finite element technology: Motivations and future needs

Advances in computer science and technology have had a profound influence on


structural engineering, leading to the emergence of this new discipline we call
computational structural mechanics (CSM). Along with it a huge software industry
has grown. CSM has brought together ideas and practices from several disciplines
- solid and structural mechanics, functional analysis, numerical analysis,
computer science, and approximation theory.

CSM has virtually grown out of the finite element method (FEM). Algorithms
for the use of the finite element method in a wide variety of structural and
thermomechanical applications are now incorporated in powerful general purpose
software packages. The use of these packages in the Computer-Aided-
Design/Computer-Aided-Manufacturing cycle forms a key element in new
manufacturing technologies such as the Flexible Manufacturing Systems. These
allow for unprecedented opportunities for increase in productivity and quality
of engineering by automating the use of structural analysis techniques to check
designs quickly for safety, integrity, reliability and economy. Very large
structural calculations can be performed to account for complex geometry,
loading history and material behavior. Such calculations are now routinely
performed in aerospace, automotive, civil engineering, mechanical engineering,
oil and nuclear industries.

Modern software packages, called general purpose programmes, couple FEM


software with powerful graphics software and the complete cycle of operations
involving pre-processing (automatic description of geometry and subsequent sub-
division of the structure) and post-processing (projecting derived information
from FEM analysis on to the geometry for color coded displays to simplify
interpretation and make decision making that much easier). Already artificial
intelligence in the form of knowledge based expert systems and expert advisers
and optimization procedures are being coupled to FEM packages to reduce human
intervention in structural design to a bare minimum.

It is not difficult to delineate many compelling reasons for the vigorous


development of CSM [9.2]. These are:

113
1. There are a large number of unsolved practical problems of current interest
which still await experimental and/or numerical solutions. Some of these
demand large computational power. Some of the examples described in reference
9.2 are: simulation of response of transportation vehicles to
multidirectional crash impact forces, dynamics of large flexible structures
taking into account joint nonlinearities and nonproportional damping, study
of thermoviscoelatic response of structural components used in advanced
propulsion systems etc. In many structural problems, the fundamental
mechanics concepts are still being studied (e.g. in metal forming, adequate
characterization of finite strain inelasticity is still needed).

2. Computer simulation is often required to reduce the dependence on extensive


and expensive testing; in certain mission critical areas in space, computer
modeling may have to replace tests. Thus, for large space structures (e.g.
large antennas, large solar arrays, the space station), it may not be
possible for ground-test technology in 1-g environment to permit confident
testing in view of the large size of the structures, their low natural
frequencies, light weight and the presence of many joints.

3. Emerging and future computer systems are expected to provide enormous power
and potential to solve very large scale structural problems. To realize this
potential fully, it is necessary to develop new formulations, computational
strategies, and numerical algorithms that exploit the capabilities of these
new machines (e.g. parallelism, vectorization, artificial intelligence).

Noor and Atluri [9.2] also expect that high-performance structures will
demand the following technical breakthroughs:

1. Expansion of the scope of engineering problems modeled, such as:

a) examination of more complex phenomena (e.g. damage tolerance of structural


components made of new material systems);

b) study of the mechanics of high-performance modern materials, such as metal-


matrix composites and high-temperature ceramic composites;

c) study of structure/media interaction phenomena (e.g. hydrodynamic/structural


coupling in deep sea mining, thermal/control/structural coupling in space
exploration, material/aerodynamic/structural coupling in composite wing
design, electromagnetic/thermal/structural coupling in microelectronic
devices);

d) use of stochastic models to account for associated with loads, environment,


and material variability

e) development of efficient high-frequency nonlinear dynamic modeling


capabilities (with applications to impulsive loading, high-energy impact,
structural penetration, and vehicle crash-worthiness);

f) improved representation of structural details such as damping and flexible


hysteritic joints:

g) development of reliable life-prediction methodology for structures made of


new materials, such as stochastic mechanisms of fatigue, etc.

114
h) analysis and design of intelligent structures with active and/or passive
adaptive control of dynamic deformations, e.g. in flight vehicles, large
space structures, earthquake-resistant structures

i) Computer simulation of manufacturing processes such as solidification,


interface mechanics, superplastic forming.

2. Development of practical measures for assessing the reliability of the


computational models and estimating the errors in the predictions of the
major response quantities.

3. Continued reduction of cost and/or time for obtaining solutions to


engineering design/analysis problems.

Special hardware and software requirements must become available to meet the
needs described above. These include:

1. Distributed computing environment having high-performance computers for large


scale calculations, a wide range of intelligent engineering workstations for
interactive user interface/control and moderate scale calculations.

2. User-friendly engineering workstations with high-resolution and high speed


graphics, high speed long distance communication etc.

3. Artificial intelligence-based expert systems, incorporating the experience


and expertise of practitioners, to aid in the modeling of the structure, the
adaptive refinement of the model and the selection of the appropriate
algorithm and procedure used in the solution.

4. Computerized symbolic manipulation capability to automate analytic


calculations and increase their reliability.

5. Special and general purpose application software systems that have advanced
modeling and analysis capabilities and are easy to learn and use.

9.5 Future directions

It is expected that CSM will continue to grow in importance. Three areas which
are expected to receive increasing attention are: 1) modeling of complex
structures; 2) predata and postdata processing and 3) integration of analysis
programs into CAD/CAM systems.

The accurate analysis of a complex structure requires the proper selection


of mathematical and computational models. There is therefore a need to develop
automatic model generation facilities. Complex structures will require an
enormous amount of data to be prepared. These can be easily performed by using
predata and postdata processing packages using high resolution, high throughput
graphic devices. The generation of three dimensional color movies can help to
visualize the dynamic behavior of complex structural systems.

115
9.6 Concluding remarks

In this concluding chapter, we have summarized the ideas that should inform the
rational development of robust finite elements and we have also described the
current trends in the use of structural modeling and analysis software in
engineering design. It has been seen that CSM has greatly improved our
capabilities for accurate structural modeling of complex systems. It has the
potential not only to be a tool for research into the basic behavior of
materials and structural systems but also as a means for design of engineering
structures.

9.7 References

9.1 G. Prathap, The Finite Element Method in Structural Mechanics, Kluwer


Academic Press, Dordrecht, 1993.
9.2 A. K. Noor and S. N. Atluri, Advances and trends in computational structural
mechanics, AIAA J, 25, 977-995 (1987).

116

S-ar putea să vă placă și