Sunteți pe pagina 1din 47

M16: Materials Modelling J.C.

Tan
Introduction to the Finite Element Method 24.01.2007

Finite Element Method (FEM)


Dr. J.C. Tan
Dr. J.C. Tan
Lent Term 2006

Lent Term 2007

-1-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

Synopsis

Lecture 10 Finite Element Method I


General introduction to FEM, FEM terminologies, Discretisation, Direct Stiffness Method, 1-D
discrete problems, Simple spring structures, Element and global stiffness relations, Stiffness
matrix, Complex spring systems.

Lecture 11 Finite Element Method II


Solution procedures, Gaussian Elimination, Steps in the FE method, 2-D truss analysis, Energy
methods - Virtual work principle, FE analysis of 2-D continuum.

Lecture 12 Finite Element Method III


Constant strain 2-D triangular element, Elastic strains and stresses in triangular element, Elastic
constants matrix, Relation between nodal forces and element stresses, Example problems.

Booklist
Basic Principles of the Finite Element Method, K.M. Entwistle (Alden Press 1999) [Za81]

The Finite Element Method: The Basis, O.C. Zienkiewicz and R.L. Taylor, 5th ed. Vol.1,
(Butterworth-Heinemann 2000) [Za82]

Introduction to the Finite Element Method, N. Ottosen and H. Petersson (Prentice Hall 1992)

i
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

Table of Contents

1. Overview..................................................................................................................................... 3
2. Direct Stiffness Method .............................................................................................................. 5
2.1 Simple Structure Consisting of Springs (1-D problem)...................................................... 5
2.1.1 A Single Spring........................................................................................................... 6
2.1.2 System of Two Springs............................................................................................... 7
2.1.3 Complex Spring Systems............................................................................................ 9
2.1.4 Solution Procedures .................................................................................................. 10
2.1.5 General Steps in FE Method ..................................................................................... 12
2.2 Truss Analysis (2-D problem) .......................................................................................... 12
3. Energy Principles in Finite Element Analysis Virtual Work Principle ................................. 17
4. Finite Element Analysis of 2-D Continuum ............................................................................. 20
4.1 Constant Strain Triangular Element.................................................................................. 21
4.1.1 Elastic Strains in the Element ................................................................................... 23
4.1.2 Stresses in the Element ............................................................................................. 25
4.1.3 Relationship between Element Stresses and Nodal Forces....................................... 27
4.2 Example Problem.............................................................................................................. 32
Question Sheet .................................................................................................................................. 33
Model Answers ................................................................................................................................. 37

-ii-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

1. Overview

All physical phenomena encountered in engineering problems, such as mechanics of materials,


fluid flow, heat transfer, acoustics and electromagnetism can be modelled using differential
equations, with a set of corresponding boundary conditions (in steady-state problems) and initial
conditions (in transient problems), see Fig. 1. However, practical engineering problems are
generally complicated and cannot be solved using analytical approaches. These difficulties may
arise from the complex nature of the governing differential equations addressed, the complex
geometry of the problem, or the difficult boundary conditions (BCs) and initial conditions to be
taken into account. Exact solutions to such problems are often not available. For such problems,
numerical approaches such as finite difference and finite element methods are required to obtain
approximate solutions. In this series of lectures, we will focus on the finite element method
(FEM), sometimes also known as finite element analysis or FEA.

Fig. 1: Modelling of physical phenomenon

The basic idea behind finite element method is to first divide the domain (region to be modelled)
into discrete connected parts, known as finite elements. These elements may be one-, two- or
three-dimensional, depending on the problem at hand (Fig. 2). The points located at the corners
and sometimes along the sides of the elements are called nodes. The process of dividing the
domain into elements is known as discretisation and the collection of all elements within a domain
is called a mesh (Fig. 3).

Fig. 2: Some examples of 1-D, 2-D and 3-D elements

-3-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

Fig. 3 A finite element mesh of an exhaust manifold assemblage. Meshing was accomplished in ABAQUS using a
combination of 3-D brick and prism elements.

The finite element method first seeks piecewise approximations over each element. This
approximation is usually a polynomial (linear, quadratic, cubic, etc.), which describes how the
unknown variable (e.g. displacement, temperature, potential, etc.) changes over the element. It is
assumed that the variable is known at the nodal points. Since the body has been discretised into
smaller parts, rather simple approximations can be adopted to interpolate the variable over the
element. Having determined the response of all individual elements, the elements are patched
together (assembly process) using some specific rules, to form the entire region and eventually
provide an approximate solution for the entire body. To elucidate this, let us consider an example
where we would like to approximate the true temperature distribution, T(x), along a one-
dimensional fin, as shown in Fig. 4(a). This is a continuous system with an infinite number of
unknowns.

Fig. 4: (a) Actual temperature distribution along a one-dimensional fin, (b) approximate temperature distribution using
4 linear elements and (c) approximation obtained using 2 quadratic elements.

We may discretise the fin into four 1-D linear elements (5 DOF) over which the temperature is
assumed to vary linearly between the nodes (Fig. 4(b)). Alternatively, we can choose to use two
1-D quadratic elements, where quadratic equations are used to approximate the temperature
distribution (Fig. 4(c)). Clearly, in order to increase the accuracy of prediction using linear
approximation, we will require a larger number of elements in contrast to quadratic approximation.
Secondly, we note that irrespective of whether the linear or quadratic approximation is used, the
approximate temperature distribution along the fin is known once temperatures at the nodal points
are known, i.e. nodal temperatures are now the unknowns of the problem. The continuous system
has been converted into a discrete system with a finite number of unknowns. We will learn in 2
that the unknowns at the nodal points can be determined by solving a certain system of equations of
the general form [ K ]{u} = { F } , known as the stiffness relation. In this particular case, since we

-4-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

are dealing with a system which consists of 5 degrees of freedom, there will be 5 equations with 5
unknowns. However, real problems generally have at least thousands of unknowns, making them
impossible to solve by hand. So, solutions can be obtained only through the use of computers and
efficient algorithms.

Since the emergence of finite element method in the early 1960s, it has found a wide range of
applications in all fields of engineering and physical sciences. This has also led to the development
of many general-purpose FE programs, two of the most renowned commercial FE packages are
ABAQUS (see www.hks.com, first released in 1978, currently version 6.6) and ANSYS (see
www.ansys.com, first released in 1970, now version 10). These are highly efficient computer
programs capable of simulating many linear and non-linear physical phenomena under static
(steady-state), quasi-static and transient conditions. However, in order to use them effectively to
obtain meaningful and realistic results, having a good knowledge of the underlying theories and
their limitations is essential. Otherwise, there is a danger that these programs may end up being
used to generate solutions (and predictions) that appear neat and attractive when presented in
colourful graphics, but are in fact physically irrelevant!

2. Direct Stiffness Method

2.1 Simple Structure Consisting of Springs (1-D problem)

Fig. 5: A one-dimensional spring element

A coiled spring can be thought of as a simple one-dimensional finite element (Fig. 5). Each spring
element consists of 2 nodes and has 2 degrees of freedom (D.O.F), i.e. two unknown displacements
u1 and u2. Forces F1 and F2 are applied at the nodes. An assembly of springs can be made by
combining several spring elements together, see Fig. 6.

Fig. 6: An assembly of three co-linear springs

The displacements produced when an assembly of springs is loaded, can be analysed based on
3 principles:

i. Condition of compatibility the connected ends of adjacent springs have the same displacements.

ii. Condition of static equilibrium the resultant force at each node is zero.

iii. Constitutive relation which describes how the material behaves. For the case of an elastic
spring, the Hookes law applies. The tension in the spring (F) is

-5-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

proportional to its extension (u), i.e. F = ku , where k (N m ) is the


-1

spring stiffness.

2.1.1 A Single Spring

First, lets consider a single spring which forms part of a linear array (Fig. 7). The nodes are i and j
and the spring element is designated as (e). The stiffness of the spring is k(e). We define forces and
displacements to the right as positive. Ri( e ) and R (j e ) are the reaction forces acting on nodes i and j,
respectively.

Fig. 7: A spring element

Case (a): Assume node j is fixed and node i is displaced by ui. Note: The suffix a designates
case (a).

Fig. 8: Displacing node i when node j is assumed fixed

The force Ria( e ) to cause displacement ui is


Ria( e ) = k ( e )ui (2.1)

To keep the spring in equilibrium, the reaction force at node j is


Ria( e ) + R (jae ) = 0 (2.2)
R (e)
ja = R (e)
ia = k ui
(e)
(2.3)

Case (b): Assume node i is fixed and node j is displaced by uj.

Fig. 9: Displacing node j when node i is assumed fixed

The force R (jbe ) required to maintain displacement uj is


R (jbe ) = k ( e )u j (2.4)

The force at node i to maintain the spring equilibrium is


Rib( e ) = R (jbe ) = k ( e )u j (2.5)

-6-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

Now, we let both nodes deform at the same time. Using the principle of superposition, i.e.
summing up the reactions from cases (a) and (b), we can write
Ri( e ) = Ria( e ) + Rib( e ) (2.6)
R (j e ) = R (jae ) + R (jbe ) (2.7)

So the reaction forces are


Ri( e ) = k ( e )ui k ( e )u j (2.8)
R (j e ) = k ( e )ui + k ( e )u j (2.9)

The reaction force at each node is the sum of the forces due to displacements ui and uj. The above
two equations can be written in matrix form:
k ( e ) k ( e ) ui Ri( e )
(e) = (e) (2.10)
k k ( e ) u j R j

which has the general form:


K ( e ) {u ( e ) } = { F ( e ) } (2.11)
where
[K(e)] is the element stiffness matrix, which is symmetrical. Since the element has
2 degrees of freedom, the order of the element stiffness matrix is [22].
{u(e)} is the nodal displacement vector of the element.
{F(e)} or {R(e)} is the element force vector.

Eqn.(2.11) is the element stiffness relation, and it describes the relation between element
displacements and element forces at the nodal points.

2.1.2 System of Two Springs

Fig. 10(a) shows a system which consists of a linear assembly of 2 springs. The nodal
displacements u1, u2 and u3 are the 3 unknowns. The external forces F1, F2 and F3 are applied at
the nodal points.

Fig. 10: (a) Linear assembly of two spring elements (b) Isolating the assembly into free bodies by cutting at locations
A, B, C and D. (c) Actions and reactions at the nodal points.

-7-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

By cutting the structure at points A, B, C and D, we can isolate the assembly into free bodies, see
Fig. 10(b). Subsequently, based on the principle of action and reaction (Fig. 10(c)), we can write
down the conditions that forces at each node are in equilibrium:

Node 1: R1(1) = F1 (2.12)


Node 2: R +R
(1)
2
(2)
2 = F2 (2.13)
Node 3: R3(2) = F3 (2.14)

We relate the nodal reaction forces to nodal displacements of each element, using general
eqns.(2.8) and (2.9), previously obtained employing the principle of superposition.

Element (1): R1(1) = k (1)u1 k (1)u2


R2(1) = k (1)u1 + k (1)u2
(2.15)
Element (2): R(2)
2 = k u2 k u3
(2) (2)

R(2)
3 = k (2)u2 + k (2)u3

We can now express the force equilibrium equations in terms of nodal displacements. While doing
this, we re-order these equations by collecting together the u terms:

k (1)u1 k (1)u2 = F1
k (1)u1 + ( k (1) + k (2) ) u2 k (2)u3 = F2 (2.16)
k (2)u2 + k (2)u3 = F3

Recasting eqns.(2.16) in the matrix form of [ K ]{u} = { F } , we obtain:

k (1) k (1) 0 u1 F1
(1)
k k (1) + k (2) k (2) u2 = F2 (2.17)

0 k (2)
k (2) u3 F3

where [K] is the global stiffness matrix of the structure. We again observe that [K] is symmetric.
Since the assembly has 3 degrees of freedom (DOF) or unknowns, the order of global stiffness
matrix is [33]. The column matrix {u} contains the nodal displacements of the structure and the
column matrix {F} contains the external forces applied at the nodes of the assembly.

A system of equations of the form [ K ]{u} = { F } is typical in finite element calculations and the
symmetry of [K] is also typical.

Evidently, the above approach used to derive the global stiffness matrix [K] would become difficult
for a complex assembly of springs. However, from eqn.(2.17), it appears that the global stiffness
matrix [K] can be obtained directly, by just considering the individual element stiffness matrices
[K(e)]. This is how it can be done:

First, using eqn.(2.10), we write down the element stiffness relations for elements (1) and (2):

-8-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

k (1) k (1) u1 F1
Element (1): (1) = (2.18)
k k (1) u2 F2
k (2) k (2) u2 F2
Element (2): (2) = (2.19)
k k (2) u3 F3

By inserting appropriate rows and columns of zeros, eqns.(2.18) and (2.19) can be expanded, so
that each relates to the three displacements in the assembled structure, i.e. u1, u2 and u3.

k (1) 0 u1 F1
k (1)
(1)
Element (1): k 0 u2 = F2
k (1) (2.20)
0
0 0 u3 F3
0 0 0 u1 F1
0 k (2) k (2) u = F
Element (2): 2 2 (2.21)
0 k (2)
k u3 F3
(2)

They can now be added to give:

k (1) k (1) 0 u1 F1
(1)
k k (1) + k (2) k (2) u2 = F2 (2.22)
0
k (2) k (2) u3 F3

This is indeed identical to eqn.(2.17). The process by which the system of equations (2.22) is
constructed from the response of each element is known as assembling. In the assembling process,
we take into account conditions for compatibility (i.e. the elements are connected) and equilibrium
conditions at the nodal points.

We observe that the global stiffness matrix [K] can be established as the sum of the expanded
element stiffness matrices [K(e)]. The displacement vector {u} contains all nodal displacements in
the assembly and the force vector {F} consists of external forces applied at the nodes.
Understanding these points allows one to establish the system of equations directly, which is useful
for more complicated systems.

2.1.3 Complex Spring Systems

The global stiffness matrix [K] for a complex assembly of springs can be written down directly by
following the 2 rules presented below. But first, let us recall that for a spring element connecting
nodes i and j (Fig.7), the element stiffness matrix [K(e)] is as follows:

k (e) k ( e ) kii kij


K ( e ) = ( e ) = k jj
(2.23)
k k ( e ) k ji

It can be seen that there are direct stiffness terms, kii & kjj on the leading diagonal and terms kij &
kji (=kij) in off-diagonal positions. For the element stiffness matrix, defining the diagonal terms as

-9-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

direct stiffness and the off-diagonal terms as indirect stiffness, we can formulate the following
rules:

The term in location ii consists of the sum of the direct stiffnesses of all the elements
meeting at node i.

The term in location ij consists of the sum of the indirect stiffnesses relating to nodes i and j
of all the elements joining node i to node j.

Fig. 11: A more complex assembly consists of five springs

Fig. 11 shows a structure made up of an assembly of 5 springs. There are 5 nodal points where
external forces F1 to F5 may be applied. The displacements of the nodes (the unknowns) are
designated as u1 to u5 and the stiffnesses of the 5 springs are k(1) to k(5). Since there are 5 DOF, the
order of the global stiffness matrix must be [55]. By applying the above rules, we can write down
the following system of equations directly:

k (1) k (1) 0 0 0 u1 F1
(1)
k k (1) + k (2) + k (4) k (2) k (4) 0 u2 F2
0 k (2) k (2) + k (3) k (3) 0 u3 = F3 (2.24)
(5)
0 k (4)
k (3)
k (3)
+k +k
(4) (5)
k u4 F4

0 k (5) k (5) u5 F5
0 0

We observe that the stiffness matrix [K] is banded, i.e. non-zero terms exist only along the leading
diagonal. This a common feature of finite element equations.

2.1.4 Solution Procedures

In the previous sections, we see that the system of equations formulated in finite element analysis is
always in the form of:

[ K ]{u} = {F } (2.25)

In practice, {F} contains the known applied forces and {u} are the unknown displacements to be
determined. To solve for the unknown displacements, we will need to obtain the inverse of the
global stiffness matrix, [ K ] :
1

-10-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

{u} = [ K ] {F }
1
(2.26)

In order to find [ K ] , the determinant of K (det K) is essential. It can be shown that for the
1

systems described by eqns. (2.10), (2.22) and (2.24):

det [ K ] = 0 (2.27)

i.e. these systems of equations are singular (no unique solution) and hence cannot be solved. This
is the case because in establishing these equations, no boundary conditions (BC) in the form of
displacements (u1, u2 ) are prescribed. Physically, we can think of this as the structure is not
secured to the ground and will lead to rigid body motion when external loadings are applied.
Therefore, the structure is in equilibrium in infinitely many positions and may even move in a
constant velocity! To obtain a unique solution, we need to specify at least one nodal point
displacement (to prevent rigid body motion).

Fig. 12: An assembly of two springs with a specified boundary condition.

For example, for the system of 2 springs from 2.1.2, by specifying u1 = 0 as the BC, eqn.(2.22)
becomes

k (1) k (1) 0 0 F1
(1)
k k +k k (2) u2 = F2
(1) (2)
(2.28)

0 k (2) k (2) u3 F3

The reaction force F1 is therefore given by


k (1)u2 = F1 (2.29)

Eqn.(2.28) can also be rewritten in the partitioned form

k (1) + k (2) k (2) u2 F2


= (2.30)
k k (2) u3 F3
(2)

The determinant of this coefficient matrix is k (1) k (2) , which is non-zero, implying a unique solution
given by
1 F 1 1
u2 = (1) ( F2 + F3 ) and u3 = (1)2 + (1) + (2) F3 (2.31)
k k k k

-11-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

2.1.5 General Steps in FE Method

The basic steps of the finite element method can be summarised as follows:

1. Establish the element stiffness matrices [K(e)]


2. Assemble the element stiffness matrices to form the global stiffness matrix [K]
3. Establish system of equations for the whole structure [ K ]{u} = { F }
4. Enforce the boundary conditions (BCs)
5. Solve the system of equations (Gaussian Elimination, Gauss-Seidel Iteration etc.)

2.2 Truss Analysis (2-D problem)

Fig. 13: A truss structure made up of pin-jointed bars.

The one-dimensional spring element (2.1) can also be used to analyse structures consisting of
trusses (Fig. 13), i.e. bars joined by frictionless hinges (pin-jointed) that can sustain only tensile
and compressive forces along the longitudinal axes.

Fig. 14. A bar element loaded in the x-axis.

With reference to the single spring element (Fig. 7), we will now establish the stiffness k for a bar
loaded by forces F, as shown in Fig. 14. The bar has a uniform cross-section of A and length of L.
E is the Youngs modulus of the bar material.

The force in the bar F is related to the normal stress by


F = A (2.32)

For the constitutive relation, we assume the bar is linearly elastic according to Hookes law
= E (2.33)

where is the normal strain, defined by


u u
= 2 1 (2.34)
L

Combining eqns.(2.32)-(2.34), we can write


AE
F = k ( u2 u1 ) ; k= (2.35)
L

-12-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

By comparing eqns.(2.35) and (2.8), we observe that the spring element can be applied to a bar
oriented along the x-axis, but the stiffness k is now AE / L . From eqn.(2.10), The element stiffness
relation can be written as
1 1 u1 F1 AE
k = ; k= (2.36)
1 1 u2 F2 L

Fig. 15: A bar element as depicted in (a) local and (b) global coordinate systems.

However, unlike spring assemblies that are confined to one axis only, trusses are usually 2-D
structures with bar members pin-jointed at various angles to one another (Fig. 13). Fig. 15 shows a
bar element in local ( x y ) and global (x-y) coordinate systems. For a bar element located in a
local system, as shown in Fig. 15(a), eqn.(2.36) has to be rewritten in the expanded form

1 0 1 0 u1x F1x
0
0 0 0 u1 y F1 y
k = (2.37)
1 0 1 0 u2 x F2 x

0 0 0 0 u2 y F2 y

We observe that (small) displacements in the y -axis, i.e. u1 y and u2 y create no forces in the bar,
i.e. F1 y = F2 y = 0 . So, a pin-jointed bar member can carry only axial loads, F1x and F2 x .

In assembling the global stiffness matrix [K] of the entire structure, it is important to note that the
element stiffness matrices [K(e)] need to be expressed in the global coordinate system. It can be
seen that the local system can be transformed into the global system through a rotation given by the
angle . By geometrical arguments, it follows directly from Fig. 15(a) and (b) that

F1x = F1x cos F1 y sin


F1 y = F1x sin + F1 y cos
F2 x = F2 x cos F2 y sin
F2 y = F2 x sin + F2 y cos (2.38)

-13-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

Taking = cos and = sin and rewriting these in matrix form


F1x 0 0 F1x
F
1 y 0 0 F1 y
= (2.39)
F2 x 0 0 F2 x
F2 y 0 0 F2 y

{F } = [ L]T {F } (2.40)

and similarly
{u} = [ L]T {u } (2.41)

where [L] denotes the transformation matrix, which relates the two coordinate systems.

[L] has the following useful properties:


[ L][ L]T = [ L]T [ L] = [ I ] (2.42)
1
[ L] = [ L] T
(i.e. orthogonal matrix) (2.43)

Using eqns.(2.42) and (2.43), the local force and displacement column matrices can be expressed as
{F } = [ L]{F } (2.44)
{u } = [ L]{u} (2.45)

The local element stiffness relation is


K ( e ) {u } = {F } (2.46)

Substituting eqns.(2.44)(2.45) into (2.46) and pre-multiplying by [ L]T


[ L]T K ( e ) [ L]{u} = [ L]T [ L]{F }
[ L]T K ( e ) [ L]{u} = {F } (2.47)

Therefore, the global element stiffness matrix [K(e)] is

K ( e ) = [ L]T K ( e ) [ L] (2.48)

We observe that once the local element stiffness relation is known, the global element stiffness
relation can be obtained by means of the transformation matrix [L].

For a bar element, it can be shown that the explicit expression for the global element stiffness
matrix is given by

2 2

AE 2 2
K =
(e)
(2.49)
L 2 2

2
2

Naturally, [K(e)] has been found to be symmetric.

-14-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

Example:

Fig. 16: A truss assembly consists of 2 bar elements and loaded at node 2.

The simple truss structure in Fig. 16 consists of 2 bar elements, mounted at nodes 1 and 3. The
bars are of the same Youngs modulus E and cross-sectional area A. An external load F is applied
onto node 2. Since there are 3 nodes and each node has 2 DOF, the entire structure has 6 DOF,
which also translates into a [66] global stiffness matrix. The global displacement and force
column vectors are of the orders of [61].

u1x F1x
u F
1y 1y
u F
{u} = u2 x ; {F } = F2 x (2.50)
2y 2y
u3 x F3 x

u3 y F3 y

Element (1) is oriented along the x-axis, so = 0D . By employing eqn.(2.49), its element stiffness
matrix in the global coordinate system can be written as

1 0 1 0 0 0
0 0 0 0 0 0

AE 1 0 1 0 0 0
K =
(1)
(2.51)
L 0 0 0 0 0 0
0 0 0 0 0 0

0 0 0 0 0 0

To obtain the correct order, extra zeros were added at the appropriate columns and rows (expanded
matrix). Similarly, for element (2), taking = 135D , we obtain sin 135D = 1 2 and
cos 135D = 1 2 , its element stiffness matrix in the global coordinate system is

-15-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

0 0 0 0 0 0
0 0 0 0 0 0

1 1 1 1
0
2
0
2 2 2
AE

K (2) = 1 1 1 1
0 0 (2.52)
L 2 2 2 2 2

0 1 1 1 1
0
2 2 2 2
1 1 1 1
0 0
2 2 2 2

We are now in the position to write down the stiffness relation [ K ]{u} = {F } of the entire
structure, expressed in the global coordinates

1 0 1 0 0 0
0 0 0 0 0 0

1 1 1 1 u1x F1x
1 0 1+
2 2 2 2 2 2 2 2 u1 y F1 y
AE 1 1 1 1 u2 x F2 x
L
0 0 = (2.53)
2 2 2 2 2 2 2 2 u2 y F2 y
1 1 1 1 u3 x F3 x
0 0
2 2 2 2 2 2 2 2 u3 y F3 y
1 1 1 1
0 0
2 2 2 2 2 2 2 2

which is the system of equations that controls the behaviour of the entire structure.

Now, we can enforce the following boundary conditions of the structure. The load BCs are:

F2 y = F ; F2 x = 0 (2.54)

The displacement BCs are:

u1x = u1 y = 0 ; u3 x = u3 y = 0 (2.55)

Applying the above BCs and solving for the unknowns, we obtain

F1x = F ; F1 y = 0 ; F3x = F ; F3 y = F (2.56)


and
FL 2 FL
u2 x = ; u2 y = 1 + (2.57)
AE 2 AE

Finally, a useful test of the validity of a solution is to check the calculated forces applied to the
structure are in static equilibrium as they must be. It can be seen that the resultant horizontal and
vertical forces are in fact zero:

-16-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

F x
= F1x + F2 x + F3 x = F + 0 F = 0
F y
= F1 y + F2 y + F3 y = 0 F + F = 0 (2.58)

3. Energy Principles in Finite Element Analysis Virtual Work Principle

In the previous sections, we have considered finite element analysis using simple 1-D elements.
Stiffness coefficients in the [K] matrix for discrete problems, such as assembly of springs and bars
under axial load (trusses) can be determined directly. Also, in the case of bar elements, the
relationship between the load and the stress in the bar is straightforward. However, the analysis of
stress distribution in elastic continuum using 2-D and 3-D elements is far more complicated. For
these problems, the stress varies continuously from point to point within the loaded body. To
achieve solutions, we will require additional equations derived from energy principles.

The Virtual Work Principle states that: A necessary and sufficient condition for equilibrium of a
particle is that the work done by forces on the particle is zero under any virtual displacement.

To elucidate this, let us consider a discrete system (Fig. 17) consists of a particle I attached to four
other particles (1 to 4) by springs, through which the particles can exert force on each other.
Particles 1 to 4 are fixed. A force FI is applied to particle I in the direction shown, generating
tensile forces in the springs FI1, FI2, FI3 and FI4. Now, displace I in the direction of FI by a small
virtual displacement uI. This virtual displacement does not change the tensile forces in the four
springs because it is so small.

Fig. 17: A discrete system used for establishing the virtual work principle.

By the Principle of Virtual Work, since I is in equilibrium, the total work done by all the forces
acting on it is zero. The energy stored in the springs, U, as the result of the displacement uI is

U = FI 1uI 1 + FI 2 uI 2 + FI 3 u I 3 + FI 4 uI 4 (3.1)

where uI1 is the extension of the spring I-1 resulting from the (small) displacement uI of I, and
correspondingly for the other 3 springs. Here, all energy terms are positive because the virtual
displacement increases the elastic energy of the springs.

On the other hand, the change (decrease) in potential energy, V, when the applied force is
displaced by the virtual displacement can be written as

-17-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

V = mg uI = FI u I (3.2)

In the above, we imagine that the force FI is applied by a string which passes over a frictionless
pulley and carries the weight mg.

The Principle of Virtual Work requires that, since I is in equilibrium, the change in the total energy
of the system during the virtual displacement is zero:

U + V = 0 U = V (3.3)

that is the change in elastic strain energy in the elastic members is equivalent to the reduction of
potential energy of the applied load during a virtual displacement. For the above system, by
equating eqns. (3.1) and (3.2),we obtain

FI 1uI 1 + FI 2 u I 2 + FI 3uI 3 + FI 4 uI 4 = FI uI (3.4)

If 1, 2, 3, 4 and I were nodes in a 2-D continuum element (presumably this is a 5-node 2-D
element), if we fix nodes 1, 2, 3, 4 and then give I a small virtual displacement, we can then use the
relation in eqn.(3.4) to relate the nodal forces on an element to the stresses within the element.
When treating a continuum element, this expression can be presented in the form of

dV = F u
V i
i i (3.5)

where the LHS is the total work done by the stresses over an element of volume V. Of course, this
is equivalent to the increase in stored strain energy U. The RHS is the total reduction in potential
energy V as a result of displacing the nodal forces Fi by the virtual displacements ui. The
subscript i denotes the nodal number. This expression relates the nodal forces to element stresses.
More details can be found in 4.1.3.

Example:

Fig. 18: An assembly of four springs arranged in a linear array. The virtual displacements u are applied in turn at
nodes 2, 3 and 4.

Fig. 18 shows a linear array consists of four spring elements. Nodes 1 and 5 are fixed, u1 = u5 = 0 .
Forces F2, F3 and F4 are applied to nodes 2, 3 and 4 and the corresponding displacements are u2, u3
and u4. We will apply the Virtual Work Principle to determine the three unknown displacements.
Prior to the analysis, we recall that in the analysis of load transfer by nodal displacements (using
superposition, see 2.1.1), all nodes are fixed aside from the one under consideration (i.e. active
node).

-18-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

First, give node 2 a small virtual displacement u2 while keeping nodes 3 and 4 fixed. The
tensions, T, in springs (1) and (2) are

T1 = k (1)u2
T2 = k (2) ( u3 u2 ) (3.6)

The elastic energy stored in the two springs caused by the virtual displacement u2 is

U 2 = T1u2 T2 u2
(3.7)
= k (1)u2 u2 k (2) (u3 u2 )u2

In the above, the tensions in the springs, T1 and T2 are unaffected by the virtual displacement, since
u2  u2 . Also, note that the second term is a negative due to the contraction in spring (2).

Whilst the reduction in potential energy as a result of virtual displacement u2 is

V2 = F2 u2 (3.8)

By the Principle of Virtual Work, we write down the energy balance as

U + V = 0

so, from eqns.(3.7) and (3.8), we obtain

k (1)u2 u2 k (2) (u3 u2 )u2 = F2 u2


k (1)u2 k (2) (u3 u2 ) = F2 (3.9)

Similarly, we give node 3 a small virtual displacement u3 while keeping nodes 2 and 4 fixed. The
energy balance gives us

T2 u3 T3u3 = F3u3
k (2) (u3 u2 ) k (3) (u4 u3 ) = F3 (3.10)

Finally, we give node 4 a small virtual displacement u4 while keeping nodes 2 and 3 fixed. We
obtain

T3 u4 T4 u4 = F4 u4
k (3) (u4 u3 ) + k (4)u4 = F4 (3.11)

Gathering together like terms of u found in eqns.(3.9)(3.11), we obtain the condensed matrix of
the form [ K ]{u} = { F } :

k (1) + k (2) k (2) 0 u2 F2



k k (2) + k (3) k (3) u3 = F3
(2)
(3.12)

0 k (3)
k + k u4 F4
(3) (4)

-19-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

It can be seen that the system of equations in (3.12) gives us 3 equations and 3 unknowns, so u2, u3
and u4 can be readily found, as the forces and spring stiffnesses are known. Of course, the same
condensed matrix can also be derived based on force equilibrium condition at each node, see
2.1.2.

4. Finite Element Analysis of 2-D Continuum

In the previous sections, we have considered simple structures that are assemblies of spring and bar
elements (trusses). For these elements, the relationship between the nodal forces and the stresses
developed within the element is straightforward and can be found directly. In this section, we will
look at more complex problems involving the analysis of 2-D elements that form part of an elastic
continuum. Fig. 19 depicts such a problem, where we would like to determine the stress
distribution in the loaded elastic thin sheet with a central circular hole. Clearly, this is a plane
stress problem. In practice, knowing the stress distribution will then enable us to calculate the
stress concentration factor at different regions around the hole.

Fig. 19: (a) A thin sheet with a circular central hole, loaded in uniaxial tension. (b) Quarter of the sheet meshed in
ABAQUS using 3-node triangular elements, a pressure boundary condition (P) was applied at the top edge to simulate
loading. The predicted stress distribution is presented in Fig. 29.

By exploiting the symmetry of the problem, only a quarter of the sheet has to be modelled, hence
reducing the number of equations to be solved and cutting down computational cost. The region of
interest is discretised using 3-node triangular elements, which are connected at nodes at their
corners. The generated mesh is shown in Fig. 19(b). A more refined mesh was applied to region
located next to the hole since stresses are expected to vary steeply here.

First, we would like to establish the relationship between the forces applied to the individual
elements and the nodal displacements, i.e. the stiffness relation [ K ]{u} = {F } . This involves
setting up the element stiffness matrices [ K ( e ) ] , and later assembling them to obtain the global
stiffness matrix [ K ] for the entire body. The procedure to do this is exactly like the one presented

-20-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

in 2. Since we know the applied forces {F } , the unknown nodal displacements {u} can be found
by inverting the global stiffness matrix. Knowing the nodal displacements, we can now derive the
elastic strains in each element and from these, the stress distribution in the loaded body, see Fig.
19(b). As a matter of fact, unlike discrete structures, relating the forces applied to a continuum
element to the strains and stresses they generate is far from trivial. To achieve this, we need to
apply energy principles, such as the Virtual Work Principle (3) to establish the force-stress
relationship (4.1.3).

4.1 Constant Strain Triangular Element

Let us now analyse a 3-node triangular plane stress element, of uniform thickness t, as shown in
Fig. 20. This element is suitable for modelling plane stress distribution under biaxial loading (Fig.
19). Biaxial stress system is typically encountered in materials in the form of thin sheets, where
there are no stresses in the direction perpendicular to the plane of the sheet (i.e. thickness
direction).

Fig. 20: A 2-D 3-node triangular plane stress element with uniform thickness t.

In Fig. 20, the element is set up in an xy global coordinate space. The 3 nodes are denoted as i, j
and k in an anti-clockwise order, and their coordinates are (xi, yi), (xj, yj) and (xk, yk), respectively.
At each node, the nodal displacements (u and v) and nodal forces (Fu and Fv) are specified by their
components in directions parallel to the x and yaxes. Since there are 2 DOF at each node, the
element has 6 DOF. From this, we can deduce that the order of the element stiffness matrix is
[66]. So, both nodal displacement vector {u} and nodal force vector {F} have orders of [61],
where

-21-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

ui Fui
v F
i vi
u F
{u} = j ; {F } = uj (4.1)
v j Fvj
uk Fuk

vk Fvk

The nodal displacements will distort the element with the result that all points within the element
will be displaced. Let the components of the displacement of a general point P(x, y) within the
element be u and v. Now, we need to make assumption on the way in which u and v varies across
the element. For a 3-node element, we can only assume that u and v vary linearly with x and y,
which implies that:

u = 1 + 2 x + 3 y
and
v = 4 + 5 x + 6 y (4.2)

where the six coefficients 1 to 6 are constants for a particular element. Substituting the 6 nodal
displacement components in eqn.(4.2), we obtain:

ui = 1 + 2 xi + 3 yi
vi = 4 + 5 xi + 6 yi
u j = 1 + 2 x j + 3 y j
(4.3)
v j = 4 + 5 x j + 6 y j
uk = 1 + 2 xk + 3 yk
vk = 4 + 5 xk + 6 yk

which can be gathered in a [66] expanded matrix.

ui 1 xi yi 0 0 0 1
v 0 0 0 1 xi yi 2
i
u j 1 x j yj 0 0 0 3
= (4.4)
v j 0 0 0 1 xj y j 4
uk 1 xk yk 0 0 0 5

vk 0 0 0 1 xk yk 6

Writing eqn.(4.4) in the general form:

{u} = [ A ]{ } (4.5)

Now, we can invert eqn.(4.5) to express the coefficients in terms of the nodal coordinates found
within [A]-1 and the nodal displacements {u}.

{ } = [ A ] {u}
1
(4.6)

-22-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

4.1.1 Elastic Strains in the Element

In the previous section, we have established the way in which the displacements u and v within the
element vary with positions x and y. The next step is to determine the elastic strains, which consist
of direct and shear strains.

Fig. 21: A small element x at a point within the triangular element, (a) before and (b) after an elastic deformation in
the x-direction.

Fig. 21(a) shows a small element of length x lying in the direction of x (and u), at a point in the
finite element. If we displace the left hand end by u and right hand end by u+u, as depicted in Fig.
21(b), the element extension is

(u + u ) u = u

So the direct strain generated along the x-axis is

u
xx = (4.7)
x

Similarly, it can be shown that the direct strain in the y-direction is

v
yy = (4.8)
y

Fig. 22: Shearing a small element of sides x and y at a point within the triangular element.

Fig. 22 shows the effect of shearing a small element of sides x and y. The ends of x and y are
rotated through (small) angles of 1 and 2 . The (engineering) shear strain xy is therefore the
total rotation of the two lines that are initially orthogonal to each other:

u v
xy = 1 + 2 = + (4.9)
y x

-23-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

Eqns.(4.7)(4.9) can be written as a strain vector:

u

xx x
v
{ } = yy = (4.10)
y
xy u v
+
y x

By differentiating eqn.(4.2), we obtain expressions for the strains

xx 2
{ } = yy = 6 (4.11)

xy 3 + 5

At this point, it is important to note that since all coefficients are constants within an element, all
strains are independent of x and y. For this reason, this element is called the constant strain
element. Due to the constant strain pattern, the initially straight edges of the element remain
straight in the deformed state. So, neighbouring elements will stay join together at their corner
nodes after straining and their edges still fit together in the deformed state. This is the state of
compatibility.

Let us rewrite eqn.(4.11) in a [66] expanded matrix form

1

0 1 0 0 0 0
2


{ } = 0 0 0 0 0 1 3 (4.12)

0 0 1 0 1 0 4
5

6
or
{ } = [C]{ } (4.13)

Combining eqns.(4.6) and (4.13):

{ } = [C][ A ] {u}
1
(4.14)

Defining [ B] = [ C][ A ] , then


1

{ } = [ B]{u} (4.15)

The matrix eqn.(4.15) is important, since it relates the elastic strains in the element {} to the nodal
displacements {u} through the matrix [B]. The derivation of matrix [B] involves tedious matrix
manipulations that leads to:

-24-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

b1 0 b2 0 b3 0
1
[ B] = 0 a1 0 a2 0 a3 (4.16)
2
a1 b1 a2 b2 a3 b3

where is the area of the element and

a1 = ( xk x j ) ; a2 = ( xi xk ) ; a3 = ( x j xi )
(4.17)
b1 = ( y j yk ) ; b2 = ( yk yi ) ; b3 = ( yi y j )

From eqn.(4.16), it is clear that [B] contains only the nodal coordinates, which are known
quantities.

4.1.2 Stresses in the Element

Fig. 23: A biaxial stress system at a point within the triangular element, showing direct and shear components.

After establishing the relations of elastic strains (4.1.1), we will now proceed to relate the strains
to the stresses in the element using constitutive relations from the theory of elasticity. The stresses
concerned with a biaxial stress system are shown in Fig. 23.

Fig. 24: (a) Extension due to xx and (b) contraction from Poissons effect due to yy.

First, let us consider the tensile strain in the x-direction, xx . It consists of the extension due to xx
and contraction due to Poissons ratio strain caused by yy , that is

-25-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

xx yy
xx = (4.18)
E E

Similarly, in the y-direction, we obtain

yy xx
yy = (4.19)
E E

Fig. 25: Engineering shear strain xy generated by xy and yx.

The relationship between shear stress and shear strain is shown in Fig. 25. The engineering shear
strain xy is produced by the shear stresses xy and yx , which are related by

xy E
xy = ; G= (4.20)
G 2 (1 + )

where G is the shear modulus. So, we obtain

2 (1 + )
xy = xy (4.21)
E

Rearranging eqns.(4.18)(4.21) gives

E
xx = ( + ) (4.22)
(1 2 ) xx yy
E
yy = ( + ) (4.23)
(1 2 ) yy xx
E (1 )
xy = (4.24)
(1 2 ) 2 xy
Recasting eqns.(4.22)(4.24) into matrix form

xx 1 0 xx
E
yy = 2
1 0 yy (4.25)
(1 ) (1 )
xy 0 0 2 xy

-26-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

The above matrix equation relates the stress and strain for the biaxial stress state. It can be written
in the condensed form as

{ } = [ D]{ } (4.26)

where
1 0
E
[ D] = 1
(1 2 )
0 (4.27)
(1 )
0 0 2

Here, [D] is termed the elastic constants matrix (stiffness tensor) for plane stress.

Substituting eqn.(4.15) into (4.26):

{ } = [ D][ B]{u} (4.28)

Now, we are able to relate the nodal displacements, which are the principal unknowns in the
analysis, to the element stresses. The next step is to introduce the nodal forces, which are known
quantities, into eqn.(4.28). We will now apply energy principles to this end.

4.1.3 Relationship between Element Stresses and Nodal Forces

To complete the analysis, we apply the Virtual Work Principle (3) to establish the relation
between element stresses and nodal forces. We recall that for the triangular constant strain element
of constant thickness t, the nodal force vector {F} is

{F } = {Fui Fvk }
T
Fvi Fuj Fvj Fuk (4.29)

and the nodal displacement vector {u} is

{u} = {ui vk }
T
vi uj vj uk (4.30)

The forces generate stresses in the element. Fig. 26 shows the biaxial stress system in a small
section within the triangular element.

-27-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

Fig. 26: A volume element under biaxial loading within a constant strain triangular element.

First, we keep all nodal displacements zero except for ui and apply a small virtual displacement ui.
The element stresses are not affected by this small displacement, but the strains have increased by
very small increments xx, yy and xy, leading to a very small increase in the elastic strain
energy stored in the small section of the element depicted in Fig. 26. The magnitude of energy
increase is equivalent to the work done by the stresses when the faces of an element are displaced
by the small strains.

Fig. 27: Volume element dV acted upon by xx.

Start by considering the work done by xx. From Fig. 27, it can be seen that for the small volume
element dx dy t = dV located within the finite element acted on by xx, the work is the product of
force and displacement, that is

( xx tdy ) ( dx xx ) = xx xx ( t dx dy )
(4.31)
= xx xx dV

Similarly, the work done by yy during the virtual displacement yy is

( yy tdx ) ( dy yy ) = yy yy dV (4.32)

It is important to note that the above equations are true because the small virtual displacement does
not change the forces.

-28-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

To obtain the work done by the shear stresses, lets consider Fig. 28(a) and (b).

Fig. 28: Work done by shear components (a) xy and (b) yx.

The total work done by the two shear components ( xy = yx ) is given by

xy 2 dV + yx 1 dV = xy xy dV (4.33)

The total increase of stored strain energy dU in the element of volume dV, as a result of the virtual
displacement ui is the sum of eqns.(4.31)(4.33):

dU = ( xx xx + yy yy + xy xy ) dV (4.34)

We obtain the elastic strain energy increase in the whole triangular element U by integrating
eqn.(4.34) over the total volume, that is

U = ( xx xx + yy yy + xy xy ) dV (4.35)
V

However, for the 3-node triangular element, the stresses and strains are constant for the whole
volume V of the each element,

U = ( xx xx + yy yy + xy xy ) dV
V (4.36)
= ( xx xx + yy yy + xy xy ) V

The above will not be the case for higher order elements (quadratic, cubic etc.) that allow both
stresses and strains to vary across each element. For these elements, eqn.(4.35) has to be evaluated
by applying numerical integration methods, such as Gaussian quadrature etc.

Expressing eqn.(4.36) in matrix form

xx

U = V { xx yy xy } yy (4.37)

xy

which can be written in condensed form:

U = V { } { }
T
(4.38)

-29-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

In 4.1.1, we have derived a relation (eqn.(4.15)) for strain {} and nodal displacements {u}

{ } = [ B]{u}

So, a similar relation must exist between virtual strains {} and virtual nodal displacements {u}

{ } = [ B]{ u} (4.39)
and
{ } = { u} [ B]
T T T
(4.40)

Substituting eqn.(4.40) into (4.38), the elastic strain energy increase in the element becomes

U = V { u} [ B] { }
T T
(4.41)
b1 0 a1
0 a1 b1

V b a2
{ ui vk } 2
0
U = vi u j v j uk { } (4.42)
2 0 a2 b2
b3 0 a3

0 a3 b3

The a and b terms are defined in eqn.(4.17). Since we displace only ui and fixed all the other
degrees of freedom, only ui is non zero. So, multiplying out the RHS of eqn.(4.42):

V
U = ui [b1 0 a1 ]{ } (4.43)
2

The balancing decrease of potential energy of the nodal applied force is Fui ui . By the Virtual
Work Principle, we write down the energy balance (eqn.(3.5)) as

V
ui [b1 0 a1 ]{ } = Fui ui
2
V
[b1 0 a1 ]{ } = Fui (4.44)
2

At this point, we have successfully related the (known) nodal force Fui to the stresses generated by
it {}. Of course, now we need to derive expressions for the stresses produced by the other 5 load
components, by applying these virtual displacements - vi, uj, vj, uk and vk, in turn. The six
equations obtained can then be gathered into a single matrix equation.

-30-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

b1 0 a1 Fui
0 a1 b1 F
vi
V b2 0 a2 Fuj
{ } = (4.45)
2 0 a2 b2 Fvj
b3 0 a3 Fuk

0 a3 b3 Fvk

Written in condensed form

V [ B] { } = { F }
T
(4.46)

Substituting eqn.(4.28) into eqn.(4.46), we obtain

V [ B] [ D ][ B]{u} = { F }
T
(4.47)

This is one of the most important relations in finite element theory. It relates the known nodal
forces to the unknown nodal displacements. It has the well-known general form

[ K ]{u} = {F } (4.48)

When considering a single element, the element stiffness matrix [K(e)] is given by

K ( e ) = V [ B] [ D ][ B]
T
(4.49)

or, when written in full

b1 0 a1
0 a1 b1
1 0 b1 0 b2 0 b3 0
1
2
E b2 a2
0 0 a3
0
K ( e ) = V 1 a1 0 a2 0 (4.50)
2 (1 2 ) 0 a2 b2
(1 ) a b1 a2 b2 a3 b3
b3 0 a3 0 0 2 1

0 a3 b3

From eqn.(4.50), it can be seen that for the constant strain triangular element, we have all the
information we need to evaluate the three matrices that make up the element stiffness matrix [K(e)].
The order of [K(e)] is the same as of the three matrices multiplied together, that is [66] or 36
stiffness coefficients altogether. The element stiffness relation has the following structure

-31-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

k11 k12 k13 k14 k15 k16 ui Fui


k k22 k23 k24 k25 k26 vi Fvi
21
k31 k32 k33 k34 k35 k36 u j Fuj
= (4.51)
k41 k42 k43 k44 k45 k46 v j Fvj
k51 k52 k53 k54 k55 k56 uk Fuk

k61 k62 k63 k64 k65 k66 vk Fvk

In a full FE analysis, we will first set up the stiffness matrices for all the elements [K(e)], and then
assemble them to produce a global stiffness matrix [K]. The technique to do this is exactly the
same as that for a linear array of springs (2.1.2). Finally, to achieve solution, we invert the
stiffness relation [K]{u}={F}, to attain {u} for the entire finite element mesh.

4.2 Example Problem

Shown in Fig. 29 is the finite element solution to the problem discussed in 4. It can be seen that
due to the use of triangular constant strain element, the calculated stresses are also constant across
each element. But, a smoother profile can be attained by averaging the stresses at common nodes.
High stresses were found at the edge of the hole, located along the horizontal axis.

(a) (b)
Fig. 29: Mises stress distribution (in N m-2) in a thin sheet with a circular central hole, loaded in uniaxial tension.
Modelling parameters are given in Fig. 19. (a) The constant stress distribution over each element. (b) A smoother stress
profile obtained by averaging stresses at the nodes.

To improve the accuracy, higher order elements such 6-node triangular element or 8-node
quadrilateral should be used. They employ quadratic approximation and therefore allow stresses
and strains to vary linearly over the element. More details can be found in the references listed
under Booklist (see page ii).

-32-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

Question Sheet

1. In the context of finite element analysis, state the condition of compatibility for a system
consisting of pin-jointed bar elements. For a continuum model meshed using 3-node
triangular elements, explain why compatibility is always satisfied across the interelement
boundaries. Demonstrate why a 3-node triangular element is also known as a constant strain
element.

2. In linear elasticity problems of spring or pin-jointed bar elements, simple relationship exists
between the nodal forces and the element stresses. However, it is less straightforward to
establish such a relation for continuum elements. Explain how the nodal forces can be related
to the stresses within a continuum element.

3. What is meant by rigid body motion in the context of finite element analysis? Shown below
are structures that can be analysed as two-dimensional:

(a) a clamped plate loaded by three forces;


(b) a circular disk loaded by two pairs of concentrated forces.

Identify the symmetry lines and show how each structure can be simplified, include the
relevant boundary conditions on your sketches.

(a) (b)

4. Shown below is a system of linear elastic springs. The spring constants of the elements are
k(1), k(2) and k(3), and F is an external force applied at node 2. It is known that the global
stiffness relation will be in the form of [K]{u}={F}. Using the Direct Stiffness Method,
formulate the global stiffness matrix [K]. Subsequently, write down the global stiffness
relation, taking into account all appropriate boundary conditions.

-33-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

5. Tripos 2006, Q25(b)

Shown below is a plane truss structure made of bar elements (1), (2) and (3). The structure is
mounted at nodes 1 and 3. An external load P is applied at node 2, inclined at 30 from the
horizontal axis. Each bar has cross-sectional area A and Youngs modulus E. How many
degrees of freedom (DOF) are there in this assembly?

Given that [ K ]{u} = { F } is the global stiffness relation, where [K] is the global stiffness
matrix. How many stiffness coefficients are there?

Write down the expressions for displacement vector {u} and force vector {F}

The element stiffness matrix K ( e ) in global coordinate system x-y is given by:

2 2

AE 2 2
K ( e ) =
L 2 2

2 2

where = cos and = sin . Write down the element stiffness matrix of each element.

By applying the Direct Stiffness Method, assemble the element stiffness matrices to form the
global stiffness relation of the entire assembly. By enforcing suitable boundary conditions,
solve for the unknown displacements and element stresses. Based on your analysis, do you
have any recommendations for optimising the existing design?

[Data: P = 10 kN; A = 100 mm2; L = 1 m; E = 200 GPa]

-34-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

6. Tripos 2005, Q25(b)

The figure below shows a composite bar of unit cross-sectional area, insulated on its sides,
consisting of two different materials a and b. The temperatures at its ends are maintained
constant. Associated numerical data are presented in the table below.

Segment Length (m) Thermal conductivity (W m-1 K-1)


a 0.10 100
b 0.55 55

Derive the thermal stiffness matrices for each of the two sections of the bar, taking each
segment as an element of your analysis. From these, derive the global stiffness matrix and
hence calculate the heat flows Q1 and Q3, and the temperature at the junction between the two
segments a and b.

Plot the temperature predicted by the finite element model against distance and explain the
main limitations of this model. How could the analysis be refined?

7. Evaluate the stiffness matrix [K(e)] for the 2-D continuum element depicted below, assuming
plane stress conditions. The three nodes are designated as i, j and k, and the coordinates are
in centimetres. Let E = 210 GPa, = 0.25 and thickness, t = 1 mm.

Assume that the element nodal displacements have been found to be ui = 0, vi = 0.05 mm;
uj = 0.03 mm, vj = 0; uk = 0 and vk = 0.06 mm. Determine the element stresses
( xx , yy and xy ) and the principle stresses.

-35-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

8. Solve this problem using a computer programme, such as ABAQUS.

For the steel plate shown below, determine the maximum principal stresses and their
locations.

[Data: E = 210 GPa, = 0.25 and thickness, t = 10 mm]

9. Solve this problem using a computer programme, such as ABAQUS.

Determine the maximum principal stresses and their locations for a flat dog bone shaped
specimen subjected to tensile forces, as depicted in the figure below. Make use of symmetry
where possible.

[Data: E = 210 GPa, = 0.25 and thickness, t = 2 mm]

-36-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

Model Answers

Q1

Bar elements that are connected to one another have the same displacements at shared nodal points
(hinges).

Compatibility is guaranteed in a 3-node triangular element because the initially straight edges of
the element remain straight in the deformed state. So, neighbouring elements will stay join
together at their corner nodes after straining and their edges always fit together in the deformed
state.

The displacements (u, v) along an edge of a 3-node triangular element are linear functions of the
coordinates (x, y):
u = 1 + 2 x + 3 y
v = 4 + 5 x + 6 y

where the six coefficients 1 to 6 are constants. Therefore, the strain vector is given by:
u

xx x 2
v
{ } = yy = = 6

y +
xy u v 3 5

+
y x

Since coefficients are constants within an element, the strains are independent of x and y,
therefore the name - constant strain element.

Q2

For continuum elements, establishing the relationship between nodal forces and element stresses
requires the use of energy principles, such as Virtual Work Principle (or the Minimisation of Total
Potential). The Virtual Work Principle states that a particle remains in equilibrium under any small
virtual displacement ui, since the increase in the stored elastic strain energy U will be balanced
by decrease in potential energy V when the applied force is displaced by the virtual displacement,
i.e.
U = V

When treating a continuum element, this expression can be expressed as

dV = F u
V i
i i

where the LHS is the total work done by the stresses over an element of volume V, which is
equivalent to the increase in stored strain energy U. The RHS is the total reduction in potential
energy V, as a result of displacing the nodal forces Fi by the virtual displacements ui. The

-37-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

subscript i denotes the nodal numbers. This expression therefore relates the nodal forces to element
stresses.

Q3

In finite element analysis, rigid body motion occurs when the system of equations [ K ]{u} = { F }
(stiffness relation) is singular, i.e. det [ K ] = 0 . This is the case because in establishing these
equations, no boundary conditions in the form of displacements have been prescribed. Physically,
we can think of this as the structure is not secured to the ground and is free to move when external
loadings are applied. Such a structure can achieve equilibrium in infinitely many positions, so
there is no unique solution.

(a) Only half of the structure needs to be modelled.

(b) Only a quarter of the structure needs to be considered.

Q4

The system has 4 nodes and each node has 1 DOF, so the global stiffness matrix will be of the
order of [44]. First, write down each of the element matrices [K(1)], [K(2)] and [K(3)].

The element stiffness matrix for spring element (1), connecting nodes 1 & 2:

k (1) k (1) 0 0
(1)
k (1)
k (1) k k (1) 0 0
K (1) = (1) (1)
, and its expanded form
k k 0 0 0 0

0 0 0 0

-38-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

For spring element (2), connecting nodes 2 & 3:


0 0 0 0
k k (2) 0
(2)
k (2)
0 k (2)
K (2) = (2) , and its expanded form
k k (2) 0 k (2) k (2) 0

0 0 0 0

For spring element (3), connecting nodes 2 & 4:


0 0 0 0
k (3) 0 k (3)
k (3) 0 k (3)
K (3) = (3) , and its expanded form
k k (3) 0 0 0 0

0 k 0 k (3)
(3)

The global stiffness matrix [K] is obtained by assembling the 3 element stiffnesses

k (1) k (1) 0 0
(1)
k k (1) + k (2) + k (3) k (2) k (3)
[ K ] =
0 k (2) k (2) 0

0 k (3) 0 k (3)

The displacement boundary conditions (BCs) are: u1x = u3 x = u4 x = 0

The load boundary condition: F2x = F

Enforcing the BCs and writing down the global stiffness relation:

k (1) k (1) 0 0 0 F1
(1)
k k (1) + k (2) + k (3) k (2) k (3) u2 F
=
0 k (2) k (2) 0 0 F3

0 k (3) 0 k (3) 0 F4

Q5

The truss structure consists of 3 pin-jointed bar elements. There are 3 nodes and each node has 2
DOF (degrees of freedom), so the entire structure has 6 DOF, which translates into a [66] global
stiffness matrix. There are altogether 36 stiffness coefficients.

-39-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

The global displacement and force column vectors are of the orders of [61]:

u1x F1x
u F
1y 1y
u F
{u} = u2 x ; {F } = F2 x
2y 2y
u3 x F3 x

u3 y F3 y

To obtain the correct order for the expanded matrices, i.e. [66], extra zeros are added at the
appropriate rows and columns. The element stiffness matrices are as follows.

Element (1) is oriented at = 0D , so = 1 and =0:


1 0 1 0 0 0
0 0 0 0 0 0
1 0 1 0

AE 0 0 0 0 AE 1 0 1 0 0 0
K = L 1 0 1 0 L 0
(1)
Expanded form
0 0 0 0 0
0
0 0 0 0 0 0 0 0 0

0 0 0 0 0 0

Element (2) is oriented at = 90D , so = 0 and =1:


0 0 0 0 00
0 0
0 0 0 0 0 0 0 0

AE 0 1 0 1 AE 0 0 0 0 00
K =
(2)

L 0 0 0 0 L 0 0 0 1 0 1
0
0 1 0 1 0 0 0 0 0

0 0 0 1 0 1

1
Element (3) is oriented at = 45D , so = = :
2

1 1 1 1
2 2 0 0
1 1 2 2 2 2 2 2
1

1

2 1 1 1 1

2 2 2
2 2 0 0
1 2 2 2 2 2 2

1 1 1

AE 2 2 2 2 AE 0 0 0 0 0 0
K (3) =
L 2 1 1 1 1 L 0 0 0 0 0 0

2 2 2 2 1 1 1 1
1 1 1 1 2 2 0 0
2 2
2 2 2 2
2 2 2 2 1 1 1 1
2 2 0 0
2 2
2 2 2 2

-40-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

The global stiffness relation is given by

[ K ]{u} = {F }
By assembling the above expanded element matrices, we obtain

1 1 1 1
1 + 2 2 2 2
1 0
2 2 2 2


1 1 1 1 u1x F1x
2 2 0 0
2 2 2 2 2 2 u1 y F1 y

AE 1 0 1 0 0 0 u2 x F2 x
=
L 0 0 0 1 0 1 u2 y F2 y

1 1 1 1 u3 x F3 x
2 2
2 2
0 0
2 2

2 2 u3 y F3 y

1 1 1 1
2 2 0 1 1+
2 2
2 2 2 2

The displacement boundary conditions are:

At node 1: u1x = u1 y = 0
At node 3: u3 x = u3 y = 0

The load boundary conditions are:

At node 2: F2 x = P cos 30D = (10, 000 ) ( cos 30D ) = +8660.25 N


F2 y = P sin 30D = ( 10, 000 ) ( sin 30D ) = 5000 N

By enforcing the above boundary conditions, we obtain

1 1 1 1
1 + 2 2 2 2
1 0
2 2 2 2


1 1 1 1 0 F1x
2 2 0 0
2 2 2 2 2 2 0 F1 y

AE 1 0 1 0 0 0 u2 x 8660.25
=
L 0 0 0 1 0 1 u2 y 5000

1 1 1 1 0 F3 x
2 2
2 2
0 0
2 2

2 2 0 F3 y


1 1 1 1
2 2 0 1 1+
2 2
2 2 2 2

Solving for the unknown displacements at node 2, i.e. u2x and u2y:

AE 1 0 u2 x 8660.25
=
L 0 1 u2 y 5000

-41-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

AE (8660.25)(1)
u2 x = 8660.25 N u2 x = = 0.433 mm
L (100 106 )( 200 109 )
AE ( 5000 )(1)
u2 y = 5000 N u2 y = = 0.25 mm
L (100 106 )( 200 109 )
Solving for the unknown reactions at nodes 1 and 3.
At node 1:
AE 1 0 u2 x F1x
=
L 0 0 u2 y F1 y

AE (100 106 )( 200 109 )



L
( u2 x ) = F1x F1x =
1
( 0.433 103 ) = 8660 N

F1 y = 0 N

At node 3:

AE 0 0 u2 x F3 x
=
L 0 1 u2 y F3 y
F3 x = 0 N
AE (100 106 )( 200 109 )

L
( u2 y ) = F3 y F3 y =
1
( 0.25 103 ) = 5000 N

Calculate the stresses, ( e ) , in each element:

8660
Element (1): (1) = = 86.6 MPa (in tension)
100 106
5000
Element (2): (2) = = 50 MPa (in tension)
100 106
Element (3): (3) = 0

Clearly, for the current loading configuration, bar element (3) does not carry any axial load and can
be removed from the current design for weight reduction, without affecting the performance of the
truss assembly.

-42-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

Q6

The 1-D steady-state heat flow problem is in fact an analogue to the 1-D bar element analysis (see
page 12). So, the global stiffness relation [ K ]{u} = { F } can also be expressed as [ K ]{T } = {Q} ,
where K is now the thermal stiffness matrix, T is the unknown temperature and Q is the heat energy
(J s-1 or W).

Fouriers law describes 1-D steady-state heat flow as

dT
Q = kA
dx

For a general 1-D thermal element as depicted above, consisting of nodes 1 and 2, we obtain
kA
Q1 = (T2 T1 )
L

Using continuity relation, we have Q1 + Q2 = 0 , and hence


kA
Q2 = (T2 T1 )
L

Recasting the above equations as [ K ]{T } = {Q}


kA 1 1 T1 Q1
=
L 1 1 T2 Q2

Here, we note that the thermal stiffness matrix of a single element is in fact given by
k ( e ) A( e ) 1 1
K ( e ) =
L( e ) 1 1

Element (a), connecting nodes 1 & 2:


1 1 0
k ( a ) A 1 1 k (a) A
K (a)
= ( a ) , and its expanded form ( a ) 1 1 0
L 1 1 L
0 0 0

Element (b), connecting nodes 2 & 3:


0 0 0
k ( b ) A 1 1 k (b ) A
K = ( b )
(b )
, and its expanded form ( b ) 0 1 1
L 1 1 L
0 1 1
k (a) A k (b ) A
We note that ( a ) = 1000 and ( b ) = 100 , where A is 1 unit2.
L L

Assembling and writing down the global stiffness relation:

-43-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

[ K ]{T } = {Q}
1000 1000 0 T1 Q1
1000 1100 100 T = Q
2 2
0 100 100 T3 Q3

Enforcing the BCs: T1 = 400 C, T3 = 100 C, Q1 = Q3

1000 ( 400 ) 1000T2 = Q1 (1)


1000 ( 400 ) + 1100T2 100(100) = Q2 (2)
100T2 + 100 (100 ) = Q1 (3)

Solving the simultaneous equations, we obtain

T2 = 372.7 C
Q1 = Q3 = 399,373 W

The temperature profile across the bar is shown below. Since the prediction was carried out using
only two linear elements, the profile is not smooth and hence will be less accurate. This analysis
can be improved either (i) by increasing the number of elements (use finer mesh size), or (ii) by
employing a higher order approximation function, e.g. quadratic element (see page 4 for a similar
discussion). Of course, both (i) and (ii) can also be implemented together to improve accuracy.

-44-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

Q7

The element stiffness matrix [K(e)] is given by eqn.(4.49):


K ( e ) = V [ B] [ D ][ B]
T
where Volume of element, V = t = 210-7 m3

First, lets determine matrix [B] (eqn.(4.16)):


b1 0 b2 0 b3 0
[ B] = 0 a1 0 a2 0 a3
1
2
a1 b1 a2 b2 a3 b3

The area of the element, = 2 2 cm2 = 2104 m2

The coordinates of the 3 nodes are as follows (in cm):


xi = 0 ; yi = 1
xj = 2 ; yj = 0
xk = 0 ; yk = 1

Obtaining the as and bs (eqn.(4.17)), in cm


a1 = ( xk x j ) = 2 ; a2 = ( xi xk ) = 0 ; a3 = ( x j xi ) = 2
b1 = ( y j yk ) = 1 ; b2 = ( yk yi ) = 2 ; b3 = ( yi y j ) = 1
So
1 0 2 0 1 0
[ B] = 25 0 2 0 0 0 2 m1
2 1 0 2 2 1

For plane stress conditions, the [D] matrix is given by eqn.(4.27)


1 0
E
[ D] = 2
1 0
(1 ) 0 0 (1 )
2

1 0.25 0
210 109
= 0.25 1 0 N m 2
( 0.9375) 0 0.375
0

We are now in position to determine K ( e ) = V [ B] [ D ][ B]


T

1 0 2
0 2 1
1 0.25 0 1 0 2 0 1 0
210 109
K ( ) = ( 2 107 ) 25 0 25 0 2 0 0 0 2

e 2 0 0
0.25 1
0 0 2 ( 0.9375 ) 0 0 0.375 2 1 0 2 2 1
1 0 2

0 2 1

-45-
M16: Materials Modelling J.C. Tan
Introduction to the Finite Element Method 24.01.2007

Performing the matrix triple product, we obtain the [66] matrix

2.5 1.25 2 1.5 0.5 0.25


1.25 4.375 1 0.75 0.25 3.625

2 1 4 0 2 1
( e)
N m1
K = 2.8 10 1.5 0.75 0
7

1.5 1.5 0.75
0.5 0.25 2 1.5 2.5 1.25

0.25 3.625 1 0.75 1.25 4.375

To compute the element stresses, we employ eqn.(4.28) which relates the element stresses to the
nodal displacements:

{ } = [ D][ B]{u}
0
0.05
xx 1 0.25 0 1 0 2 0 1 0
210 109 0.03
yy = 0.25 1 0 25 0 2 0 0 0 2 (1 10 )
3

( 0.9375 ) 0 0.375 2 1 0 2 2 1 0
xy 0
0

0.06
Performing the matrix triple product, we obtain
xx 364

yy = 196 MPa
231
xy

Now, we can calculate the principle stresses

xx + yy xx yy
2

1,2 = + xy
2

2 2
364 + 196 364 196
2

1,2 = + ( 231)
2

2 2
1 = 525.8 MPa and 2 = 34.2 MPa

at principal angles:
2 xy
tan 2 p =
xx yy
p = 35

Q8 and Q9

Solutions will be given in the Example Class.

-46-

S-ar putea să vă placă și