CCM I Lecture Notes PDF
CCM I Lecture Notes PDF
Mechanical Engineering
Prepared by
Ismet Demirdzic
University of Sarajevo, Bosnia and Herzegovina
Alojz Ivankovic
University College Dublin, Ireland
Noel O’Dowd
University of Limerick, Ireland
Resume
Continuum is a mathematical idealisation (representation) of real materials such as solids,
fluids and gases. Continuum mechanics is concerned with the behaviour of such continua. It
is based on laws of conservation (of mass, energy, momentum, moment of momentum and
entropy), and is aided by a mathematical apparatus specifically developed for it. With the
advent of computer technology and accompanying computational methods, Computational
Continuum Mechanics (CCM) has become an extremely powerful tool for analysing
engineering problems.
A number of numerical methods have been developed for analysing continuum mechanics
problems. These utilise conventional numerical methods such as Finite Element (FE), Finite
Volume (FV), Boundary Integral, Finite Difference, …. The FE method has been the most
widely used for the numerical analysis of solid mechanics problems, while the FV method is
the ‘number one’ method in computational fluid mechanics. The current trend is that both
methods are being developed not only for one particular discipline of continuum mechanics
but also for interdisciplinary applications.
The CCM course is designed to introduce basics of continuum mechanics, and FV and FE
methods. By doing so, it is aimed to demonstrate how fundamental laws of physics can be
turned into powerful numerical tools.
The CCM course is also planned to provide theoretical basis for practical applications of FV
and FE methods.
i
Contents
1 Introduction 1
3 Stresses 27
3.1 Body and surface forces 27
3.2 Stress tensor 28
3.3 Principal stresses. Stress tensor’s invariants 31
3.3 Spherical and deviatoric stresses 32
6 Constitutive relationships 73
6.1 Introduction 73
6.2 Classical constitutive relations and equations of state 75
6.3 Elastic solids 78
6.4 Ideal and Newtonian fluids 81
6.5 Fourier’s law and the caloric equation of state 83
ii
7 Mathematical models 85
7.1 Introduction 85
7.2 Linear elastic solids 85
7.3 Newtonian fluids 88
7.4 Initial and boundary conditions 90
7.5 Summary 95
iii
1. INTRODUCTION
Chapter 1
Introduction
It is important to recall that much of the technical knowledge results from only a few basic laws
of physics. They form the basis for specialised engineering disciplines such as solid mechanics,
fluid mechanics, heat transfer, etc.. Continuum Mechanics deals with these principles which are
common for practical engineering materials.
A number of numerical methods have been developed for analysing continuum mechanics
problems. These utilise conventional numerical methods such as Finite Element (FE), Finite
Volume (FV), Boundary Integral, Finite Difference, etc. The FE method has been the most
widely used for the numerical analysis of solid mechanics problems, while the FV method is well
known in computational fluid mechanics. Computational Continuum Mechanics (CCM) course is
designed to introduce basics of continuum mechanics. Is not intended to be a review of numerical
methods but to briefly introduce basic principles of FV and FE methods. By doing so, it is aimed
to demonstrate how fundamental laws of physics can be turned into powerful numerical tools.
Solution of complex real-life engineering problems has become an everyday reality. Demands in
this field are on ever-increasing path. Although quite powerful, the present day numerical
methods are far from being perfect, for a number of reasons, and much effort is being put into
improving the existing methods and developing new ones. The current trend is that both FV and
FE methods are being developed not only for one particular discipline of continuum mechanics
but for interdisciplinary (‘multiphysics’) applications. The direct reliance of the FV method on
the basic physical principles, and its inherent iterativeness makes it particularly suitable for these
applications.
1
1. INTRODUCTION
There are several steps that have to be taken, and a number of important decisions to be made
along the way, in order to solve a practical engineering problem (Fig. 1.1). Each of them
inevitably involves a number of assumptions and approximations, which all together influence
the final solution of the problem. We shall now try to briefly address the essential ones.
1) The first step is a definition of the problem at hand. Its success is often limited by our ability
to grasp the full complexity of the problem, to identify the type of materials involved, to
accurately specify initial and boundary conditions etc.
2) As the next step, one has to choose a mathematical model, which can simulate the behaviour
of the problem. There, we can distinguish between:
2.1) Problems for which an adequate mathematical description can be written (e.g. heat
conduction, elastic stress analysis, laminar fluid flow), and
2.2) Problems for which an adequate mathematical description has not yet been found (non-
linear stress analysis, turbulent fluid flow).
Of course, the group into which a given problem falls will depend on what we are prepared
to consider as an 'adequate' description.
world
(1) Real
SYSTEM OF
Time and space ALGEBRAIC Equations
discretisation
(3) Numerical
discretisation (DISCRETISED)
method
EQUATIONS
SOLUTION
2
1. INTRODUCTION
4) The resulting set of algebraic equations is then solved by (approximate) iterative methods, by
necessity if the equations are non-linear or the number of equations is very large and the
direct solution exceeds available computer resources, or by choice if the system is linear and
not too large. Here, we should not forget the finite arithmetics of the computers and
accompanying round-off error propagation which makes theoretically exact methods at best
approximate and at worst useless.
5) Finally, the numerical method has to be translated into a computer language and computer
program assembled. In doing this a great deal of care has to be taken regarding the suitable
data structure, the program modularity etc. Also, pre- and post- processing facilities are
important components governing utility of stress analysis software.
3
2. BASIC CONCEPTS AND DEFINITIONS
Chapter 2
The word continuum comes from mathematics; e.g. the system of real numbers is continuous
since there is always a real number between any two different real numbers (this means that the
system of real numbers is infinite). Time can be described by one system of real numbers, while
for 3D space three systems are required.
The concept of continuum applied to a matter is best illustrated on considering the mass density.
Let’s assume that a material occupies the volume V0 (Fig. 2.1). Let’s also consider a point P in
V0 and in a series of progressively smaller sub-volumes
V1, V2 , ...(Vn 1 Vn , P Vn , n = 1,2,...) all containing point P. If volume Vn contains mass
M n , then the mass density in P is defined as:
4
2. BASIC CONCEPTS AND DEFINITIONS
Mn
(P) = lim . (2.1)
n Vn
Vn 0
If the mass density is defined in every point of V0 and in every time instant, then the mass is
said to be continually distributed, and
= (x, y,z, t) , (2.2)
is a single-valued function of space and time. Here, x, y and z are Cartesian co-ordinates and t is
time. In a similar way one can define energy density, momentum density etc.
V2
V1
V0
However, the real substances are made of elementary particles (molecules) and the continuum
idealisation is only applicable under certain conditions. If, for example, one attempts to define
the mass density of water the problems will occur when the size of Vn becomes comparable to
the molecular dimension of water. When Vn1 and Vn differ by one molecular dimension, then
the difference between Mn 1 / Vn 1 and Mn / Vn can be finite (or even infinite). As the particles
moves in space, the ratio Mn / Vn either does not exist or it fluctuates in time and space (Fig.
2.2).
M
3 V
5
2. BASIC CONCEPTS AND DEFINITIONS
To solve the problem, let’s allow Vn to decrease but to always remain big enough so it contains
a large number of particles. If under this condition the ratio Mn / Vn tends to a final value (P),
then this value is taken as the mass density at point P. It is usually taken that:
Mn
(P) = lim 3 , (2.3)
Vn Vn
where is of the order of the mean free molecular path. If this value is extrapolated to Vn 0,
then we have defined ‘mathematical’ continuum with mass density (P). In practice, this
extrapolation is usually never required. For example, the molecular dimension of water is
approximately 10-10 m, and water can be considered as a continuum in all flow problems where
any characteristic size is greater than 10-10 m. The mean free molecular path in air is ~ 5 10-8 m.
As shown above, the concept of continuum requires that the mean free molecular path (a mean
distance travelled by a molecule in between two collisions) is small in comparison with the
smallest characteristic length of the body. It means it can be applied to problems where the
particular nature of the material microstructure can be neglected. This condition is, however,
fulfilled for the majority of practical materials. The matter on which the concept of continuum
can be applied is called continuum. The part of physics dealing with motion of continuum is
called continuum mechanics.
For example, atmospheric air is continuous but not homogeneous as its density varies with
altitude. Wood can be considered as continuous and homogeneous but is not isotropic. The first
assumption is a requirement for the use of continuum mechanics. The other two are not.
6
2. BASIC CONCEPTS AND DEFINITIONS
Quantity F may be a scalar, a vector or, in general a tensor, and the respective fields are called
scalar, vector or tensor fields. A scalar is a quantity defined by 30=1 number, vector is defined
by 31=3 numbers, while quantity requiring 3n numbers is called a tensor, where n is tensor order
(also called rank). In fact, scalars and vectors can be regarded as tensors of the zero and first
order, respectively. In what follows, we shall mainly consider tensors of the second order
(defined by 9 numbers), and unless stated otherwise, the word tensor will be used to denote a
tensor of the second order.
where x, y and z are the Cartesian coordinates and i, j and k are the unit base vectors in direction
of x, y and z coordinate axes, respectively. Thus, the expression (2.4) will also be written in the
following form
7
2. BASIC CONCEPTS AND DEFINITIONS
Each vector field can be associated with the following three scalar fields
vx = vx (x,y,z,t),
vy = vy (x,y,z,t), (2.13)
vz = vz (x,y,z,t),
where vx , vy and vz are Cartesian components of vector v and
v = vx i + vy j +vz k. (2.14)
Similarly, each tensor field T can be associated with nine scalar fields
where Txx , Txy, ..., Tzz are Cartesian components of tensor T and
8
2. BASIC CONCEPTS AND DEFINITIONS
Quantities ii, ij, ..., kk, which are sometimes written as ii, ij, ..., kk, are nine linearly
independent tensorial or dyadic products of Cartesian unit vectors and have the following
properties
ij ji
In doing this the standard (Einstein) summation convention is employed: whenever an index
appears twice in a monomial, it implies a summation over the range of its values (from 1 to 3 in
the present case).
Vectors and tensors are often identified by their components, and referred to as vector vi and
tensor Tij. This way of writing is called tensor or index notation. According to the convention,
an index may appear in a monomial once or twice. If an index appears once, it may take any one
of the three values 1, 2 or 3 and is called unrepeated or free index. Number of free indices
defines the quantity; 0 – scalar, 1 – vector and 2 – tensor. If an index appears twice in a
monomial, than the summation convention is applied. Such an index is referred to as a dummy
index since it must assume all three values (if replaced by another index, the meaning of the
expression remains unchanged).
9
2. BASIC CONCEPTS AND DEFINITIONS
v x
vT = [v]T = v y . (2.21)
v z
We shall only consider the transformation of Cartesian coordinates, and so-called Cartesian
tensors. In case of arbitrary curvilinear coordinate transformation, the corresponding tensors are
called general tensors. In what follows and unless stated otherwise, the word tensor will be used
to denote a Cartesian tensor.
Coordinates’ transformation
' ' '
Let’s consider two Cartesian coordinate systems Ox1x 2 x 3 and Ox1x 2 x 3 having the same pole
O. Cosines of angles between the coordinate axes of the two systems are usually denoted as aij
'
meaning aij = cos(xi , x j ) according to:
x1 x 2 x 3
' a11 a12 a13
x1 a11 a12 a13
A= ' or A = a21 a22 a23 . (2.23)
x 2 a21 a22 a23 a31
' a32 a33
x 3 a31 a32 a33
Matrix A is called rotation matrix and it can be shown it is orthogonal, i.e. A-1=AT. Also,
det(A)=1. Transformation (2.25) is called orthogonal transformation.
10
2. BASIC CONCEPTS AND DEFINITIONS
Since the coordinates of an arbitrary vector can be defined by the coordinates of its starting and
ending points, then:
v 'i = aij v j vi = a ji v'j or v' = Av v = AT v' . (2.26)
The expression (2.27) can be generalised giving an equation for the transformation of tensor T
' ' T
Tij = aip a jq Tpq or T = A TA . (2.28)
The following formal mathematical definitions can be given by comparing the transformation
laws for tensors of the order 0, 1 and 2 with the coordinate transformation law (2.25):
1) A scalar is a quantity which has only one component invariant of the coordinate system
transformation.
2) A vector is a quantity defined by three components which transform in the same way as the
coordinates of the position vector (compare (2.25) and (2.26)).
3) A tensor (of the second order) is a quantity defined by nine components which transform
according to (2.28). Higher order tensors can be defined in a similar way.
An alternative way of defining a second order tensor is by using it as a linear vector operator,
which relates a vector u with vector v in the following way
v = Tu or vi = Tij u j or v = Tu . (2.29)
Let’s demonstrate that these two definitions are identical. It follows from (2.26) and (2.29) that
' ' '
vi = aipv p = aipTpq uq = aipTpq a jqu j = aipa jqTpq u j , (2.30)
11
2. BASIC CONCEPTS AND DEFINITIONS
Scalar multiplication
12
2. BASIC CONCEPTS AND DEFINITIONS
v (ab) = vi ai b ji j , (2.42)
v T = eijkv iT jl i k i l . (2.44)
13
2. BASIC CONCEPTS AND DEFINITIONS
Identity tensor
while the symbol det denotes the determinant of the tensor T, which is defined as
Txx Txy Txz
det (T) = det (Tij) = Tyx Tyy Tyz . (2.52)
Tzx Tzy Tzz
14
2. BASIC CONCEPTS AND DEFINITIONS
A (symmetric) tensor, however, has three invariants. They are related to the so-called
characteristic equation
det (T – T I) = det (Tij – ij) = 0 (2.59)
which can be written in the (Hamilton-Cayleyev) form
3 – IT2 – IIT – IIIT = 0, (2.60)
where
IT = tr T= Tii,
1 1
IIT = (T : T tr 2 T) = [Tij Tji (Tii )2 ], (2.61)
2 2
1
IIIT = det (T )= ijk pqr TipTjq Tkr
6
are the first, second and third invariant of the tensor T.
15
2. BASIC CONCEPTS AND DEFINITIONS
has non-trivial roots (l1(k) ,l 2(k ) ,l3(k) ) which represent components of the three principal
directions. If coordinate axes are taken along the principal directions, the tensor T becomes a
diagonal tensor
T(1) 0 0
T = 0 T(2) 0 . (2.63)
0 0 T(3)
is called inverse tensor of tensor T and is denoted as S = T-1. Tensor T is called direct tensor of
tensor T-1. The sufficient condition for the existence of an inverse tensor is det(T)0. It can be
shown that the principal components of an inverse tensor are equal to the inverse of the principal
components of the direct tensor, while their principal directions coincide.
Scalar multiplication of gradient by unit vector lo = l/|l| gives derivative of the scalar field s in
direction l defined by vector l
s
= grad s lo. (2.67)
l
Divergence, rotor and gradient of a vector field
Each vector field can be associated with a scalar field of its divergence
v j v x v y vz
div v = =
x j x
+
y
+
z
( )
= v j,j , (2.68)
16
2. BASIC CONCEPTS AND DEFINITIONS
i j k
v
curl v = ijk k ii = , (2.69)
x j x y z
vx vy vz
17
2. BASIC CONCEPTS AND DEFINITIONS
1
(...) = lim
V 0 V
n(...)dS , (2.73)
S
where V is the volume bounded by a surface S and n is the outward normal to the surface S. The
quantity (…) stands for either scalar, vector or tensor and is multiplied with n in a certain way.
For the Cartesian coordinates:
(...) (...) (...) (...)
(...) = ij = i+ j+ k. (2.74)
x j x y z
This operator has the properties of both a vector and a linear differential operator. It follows
that:
grad s = s, grad v = v,
curl v = x v.
Especially:
div x = 3
curl x = 0
(2.77)
grad x = I
grad x = x
where x is a position vector.
where n is the unit vector of the outer normal to the surface S, and circulation of the vector field
along a closed curve C as
= v dx, where dx = dxi + dyj+ dzk , (2.79)
C
18
2. BASIC CONCEPTS AND DEFINITIONS
In connection with the above quantities one can prove the following two very important
theorems:
1) Gauss' theorem
Gauss' (Divergence) theorem relates the source of a field in a region V with the flux of that
field through the boundary surface of the region S
2) Stokes' theorem
According to Stokes' theorem, the circulation of a vector field along a regular closed curve C
is equal to the flux of the curl of that vector field through an arbitrary simply connected
(one-sided) surface S bounded by the curve C
v dx = curl v n dS . (2.81)
C S
Field s is called the potential of the field v. A potential field is also called non-circulatory as
curl v = curl (grad s) = (s) = 0. (2.83)
19
2. BASIC CONCEPTS AND DEFINITIONS
6) Vector field v which is potential and solenoidal at the same time is called Laplace or
harmonic field. For such a field:
v = grad s
2s
div v = div (grad s) = (s) = 2 s = 0 or =0,
x i x i
or in expanded form:
2s 2s 2s
+ + = 0. (2.85)
x 2 y 2 z 2
The function s which satisfies this so called Laplace equation (2.85) is known as the Laplace
or harmonic function.
2.5.1 Definitions
1) A rectangular array of 'scalars' aij
a11 a12 ... a1n
a
21 a22 ... a2n
A = [aij]mxn = (2.86)
... ... ... ...
am1 am2 ... amn
is called an m x n matrix, or matrix of the format m x n, or simply a matrix. aij is the matrix
element situated in row i and column j, m is the number of rows, and n is the number of
columns. In general, number of rows m and columns n may not be the same.
2) n x 1 matrices are called column matrices or column vectors, and 1x n matrices are called
row matrices or row vectors. Their elements are usually written with one subscript only, e.g.
20
2. BASIC CONCEPTS AND DEFINITIONS
x1
x
2
.
x = [xi]nx1 = , xT = [xi]1xn = [x1, x2, ..., xn]. (2.87)
.
.
x n
I = [ij] = {1,0, if i= j
if i j (2.88)
21
2. BASIC CONCEPTS AND DEFINITIONS
– Also, the trace of a square n x n matrix A is defined as the sum of its diagonal
elements
n
tr (A) = aii , (2.93)
i =1
8) The minor Dij of the element aij of a square matrix is the determinant of the submatrix
obtained by deleting the row and the column in which aij lies. The cofactor Aij of the
element aij is (–1)i+jDij.
9) The matrix A = [aij]mxn is said to have rank r if it contains at least one square submatrix of A
of the order r with a non-zero determinant, whereas the determinant of any submatrix of A
of the order r+1 or more is zero. The matrix A is non-singular or regular if and only if m = n
= r, i.e. if it is square and det (A) 0. Otherwise it is singular.
Equality
Two m x n matrices A = [aij]mxn and B = [bij]mxn are equal (A = B) if and only if aij = bij for all i
and j.
Addition and subtraction
The sum or difference of two m x n matrices A = [aij]mxn and B = [bij]mxn is the m x n matrix
A ± B = [aij ± bij]mxn. (2.95)
Scalar multiplication
The product of the m x n matrix A = [aij]mxn by the scalar s is the m x n matrix
sA = [saij]mxn. (2.96)
22
2. BASIC CONCEPTS AND DEFINITIONS
Matrix Multiplication
The product of the m x n matrix A = [aij]mxn and the n x p matrix B = [bjk]nxp is the C = [cik]mxp
matrix
n
AB = [c ik ]m p = aij b jk = [ai1b1k + ai 2b2k + ... + ainbnk ]m p . (2.97)
j =1
m p
It follows from the above definition that two matrices A and B can be multiplied in the order AB
if the number of columns in A is equal to the number of rows in B, i.e. if A and B are
conformable in the order AB. The existence of AB implies that of BA only if A and B are square
matrices. In general BA AB.
Matrix Inversion
The inverse of the non-singular (square) n x n matrix A is the n x n matrix A–1 such that
A–1A = AA–1 = I, (2.98)
where I is the identity (unit) matrix. It can be shown that
[ ]n n = det1(A) [ A ji ]n n ,
1 1
A = aij (2.99)
where Aij is the cofactor of the element aij in the determinant det (A). The above definition
implies that an inverse of a singular matrix does not exist. If A–1 = AT, matrix A is said to be
orthogonal.
Useful relationships
It can be shown that the following relationships involving matrix operations hold
A + B = B + A, A + (B + C) = (A + B) + C
s (A + B) = sA + sB
A (B + C) = AB + AC
IB = B, CI = C (2.100)
A + 0 = A, 0A = 0, A + (–A) = A – A = 0
(AB)–1 = B–1A–1, (sA)–1 = s–1A–1, (A–1)–1 = A
(AB)T = BTAT, (AT)T = A
||A + B|| ||A|| + ||B||, ||sA|| = s ||A||, ||AB|| ||A|| ||B||,
where A, B and C are matrices of any format that allows the indicated operations, I and 0 are
identity and null matrices, respectively, and s is an arbitrary scalar.
23
2. BASIC CONCEPTS AND DEFINITIONS
1 1
A= (A + A T )+ (A A T ),
2 2
1 1
where, it should be noticed, (A + A T )is a symmetric and (A A T )an asymmetric matrix.
2 2
where aij are given numbers, which are called the coefficients of the system and bi are also
given numbers, can be written in a matrix form as
Ax = b, (2.102b)
where the coefficient matrix A = [aij]mxn, the solution vector x = [xi]nx1 and the right hand side
vector b = {bi}mx1 are
x1 b1
x b
a11 a12 ... a1n 2 2
a . .
21 a22 ... a2n
A=
... , x = , b = . (2.103)
... ... ... . .
. .
am1 am2 ... amn
x n bm
The matrix
a11 a12 ... a1n b1
a b2
21 a22 ... a2n
B = (2.104)
... ... ... ... ...
am1 am2 ... amn bm
is called the augmented matrix of the system (2.102).
Existence and general properties of solutions
System (2.102) may have no solutions, precisely one solution or infinitely many solutions.
Regarding the existence and the general properties of solutions of system (2.102), it can be
shown that:
1) System (2.102) has solutions if and only if the coefficient matrix A and the augmented
matrix B have the same rank r. If r = n, there is precisely one solution. If r < n, there are
infinitely many solutions which are obtained by determining r suitable unknowns (whose
24
2. BASIC CONCEPTS AND DEFINITIONS
submatrix of coefficients has rank r) in terms of the remaining n – r unknowns, which can
take arbitrary values.
2) If system (2.102) is homogeneous (i.e. if b = 0), it always has the trivial solution x = 0. Non-
trivial solution exists only if rank of A is less than n.
Systems with square matrix
In a special case when the coefficient matrix A is a square n x n matrix, i.e. when the system
(2.102) has the same number of unknowns as equations, then
x = A–1b (2.105)
if and only if matrix A is non-singular, i.e. if det(A) 0 (which is the necessary and
sufficient condition for the existence of A–1),
2) The system of homogeneous equations has a non-trivial solutions if and only if the matrix A
is singular, i.e. if det (A) = 0.
Over-determined systems
Systems with more equations than unknowns (m > n) are said to be over-determined and, in
general, can be solved only approximately. One of the methods for solution of such systems is
the least square method. It consists in requiring that the sum of the squares of differences
between the right and the left hand side of each equation in (2.102a)
m
2
f(x1, x2,...,xn) = aij x j bi =
2
(aij x j bi )2 (2.106)
i =1
has a minimum (note the use of the summation convention), i.e. that
m
f
= 2 aik (aij x j bi ) = 0 (k = 1, 2, . . ., n), (2.107)
x k
i =1
which results in the following system of n linear algebraic equations with n unknowns
aikaijxj = aikbi (k = 1, 2, . . ., n) (2.108a)
(where i takes values from 1 to m, and j from 1 to n, m > n), or in the compact (easy to
remember) matrix form
ATA x = ATb. (2.108b)
25
2. BASIC CONCEPTS AND DEFINITIONS
A non-trivial solution will exist if and only if the determinant of the system vanishes
P() = det (A – I) = 0. (2.110)
The above non-linear algebraic equation is named the characteristic equation of A, and the
values of which satisfy that equation are called the eignevalues of A. P() is called the
characteristic polynomial of A.
It can be shown that if matrix A is symmetric, all its eigenvalues are real, and if any two
eigenvalues are distinct, e.g. 1 2, then the two associated eigenvectors x1 and x2 are
orthogonal, i.e. x1T x 2 = 0 .
The set of eigenvalues of A is called the spectrum of A. The largest of the absolute values of the
eigenvalues of A, (A) = ||max, is called the spectral radius of A, while the ratio ||max/ ||min is
termed the condition number of A.
26
3. STRESSES
Chapter 3
Stresses
3.1 Body and surface forces
Continuum mechanics is mainly concerned with forces acting on a continuum element and the
resulting motion. Those forces are conventionally classified into two categories: body (mass or
volume) and surface (traction) forces.
Body forces act on the total mass of the continuum element and are expressed as force per unit
mass of the element. These are, for example, gravitational force, inertial force, magnetic or
electro-magnetic force etc.
Surface forces are a consequence of the action of the immediate environment onto the
continuum element through their direct contact and are expressed as forces per unit area of the
element. These are, for example contact forces, forces of internal friction in a body etc. Surface
forces are also called stresses.
Different to particle mechanics, continuum mechanics is not concerned with forces themselves,
but with the density of their distribution in space (in a volume, or on a surface). Observe a
small volume of continuum no smaller than 3 (where is of the order of magnitude of free
path of molecules) and call it a continuum element (Fig. 3.1).
27
3. STRESSES
It is important to note an essential difference between the mass force fb and the surface force
fs. While the vector fb is a single-valued function of position vector and time, and consequently
forms a vector field, the vector fs has an infinite number of values in every point in space,
depending on the orientation of the surface on which it acts (Fig. 3.1), and, therefore, it does
not form any vector field.
a) b)
f sn
1
n1 f sn n2
2
S 1
P S2 P
fb fb
V V
S S
28
3. STRESSES
Let’s consider a continuum element of a tetrahedral shape shown in Fig. 3.2. It is oriented in
such a way that three of its sides are in coordinate planes, while the position of the fourth side
is arbitrary and can be defined by its unit normal:
n = (nx, ny, nz) = (cosx, cosy, cosz), (3.6)
where:
nx = n i = cos x , n y = n j = cos y , nz = n k = cos z . (3.7)
fs(-i)
n f sn
fs(-j)
y
S
fs(-k)
x
Figure 3.2 Cauchy tetrahedral
If fb is the resultant body force acting on the element and fsn, fs(-i), fs(-j), fs(-k) are surface forces
acting on the tetrahedral element sides, by applying the Newton second law of motion one
gets:
Va = Vfb + fsn S + fs (i) S x + fs ( j ) S y + fs (k) Sz , (3.8)
where a is the acceleration, V the volume and S, Sx, Sy, Sz are surfaces of the element
(Fig. 3.2). The term on the left hand side of eq. (3.8) and the first term on the right hand side
are proportional to the element volume. Being small quantities of the higher order they can be
neglected compared to other terms which are proportional to the surface, and eq.(3.8) becomes
fsn S = fsi Sx + fsj S y + fsk Sz . (3.9)
Since
S x Sy Sz
nx = cos x = , n y = cos = , nz = cos z = , (3.10)
S S S
then
fsn = n x fsi + n y fsj + n zfsk . (3.11)
By expressing the surface forces in terms of their components in Cartesian coordinate system:
29
3. STRESSES
one gets:
nx = nx xx + ny yx + nz zx ,
ny = nx xy + ny yy + nz zy , (3.13)
nz = nx xz + ny yz + nz zz .
Expressions (3.13) show that components of fsn, which acts on an arbitrary surface, are
expressed as linear functions of the components of surface forces fsi, fsj, fsk, acting on
coordinate planes, i.e. in terms of nine quantities which can be regarded as components of a
tensor with matrix:
xx xy xz
= yx yy yz . (3.14)
zx zy zz
Tensor is called (Cauchy) stress tensor. Consequently, eq. (3.11) can be written in the form:
fsn = n , or fsn,i = njji . (3.15)
Thus, in every point of a continuum exists an infinite number of vectors fsn, whose direction
depends on the orientation n of the plane on which they act (e.g. Fig. 3.1), but only one stress
tensor . The stress tensor, therefore, forms a (tensor) field. It is obvious that the components
of the stress tensor , like the components of the body force vector fb, depend on the choice of
the coordinate system, but the tensor , like the vector fb is an invariant (independent of the
coordinate system) physical quantity.
It will be shown later that stress tensor is symmetric, i.e. that:
= T, or ij = ji . (3.16)
It is worth noticing that the first index in the expression for the surface force components (eqs.
(3.12) and (3.13)) denotes the direction of the normal to the surface on which the force acts,
while the second index denotes the direction of the axis to which the force is projected. For
example, zx denotes ‘x’ component of the force acting on the plane with normal in z direction.
The components with the same indices xx, yy, zz represent the projections of the force in the
direction of the surface normal and are called normal (or direct) stresses, while the projections
in the surface plane directions xy, xz, zy, … are called tangential (or shear) stresses (Fig.
3.3). Stresses can be positive and negative. The usual convention is that the stress is positive if
the outer normal and the force projection are both oriented in either positive or negative
30
3. STRESSES
direction of the corresponding coordinate axes. Otherwise, the stress is negative. Positive
normal stresses cause tension while negative cause compression. All stresses in Fig. 3.3 are
positive.
z
zz k
zy j
zx i yz k
-yx i
- yy j xz k yy j
y
yx i
- yz k xy j
xxi
In continuum mechanics, the stress components are often denoted with for normal and for
tangential stresses, i.e.
xx xy xz
= yx yy yz . (3.17)
zx zy zz
If n is a unit normal to the principal plane (in the direction of the principal axis), then the
surface force acting on the principal plane is fp = n, where is the principal stress. According
to (3.15) fp = n . Since is a symmetric tensor and n = In , then:
( I) n = 0 or (ij - ij) nj = 0. (3.18)
31
3. STRESSES
This system of three homogeneous linear algebraic equation has non-trivial roots (ni 0) if its
determinant is equal to zero, i.e. if:
det (ij - ij) = 0, (3.19)
where:
) = ii ,
I1 = tr(
1
[ ]1
I2 = : tr 2 () = [ ij ij ( ii ) 2 ] ,
2 2
(3.21)
1
I3 = det( ) = ijk pqr ip jq kr ,
6
are the first, second and third invariants of the stress tensor.
If 1, 2 and 3 are the roots of eq. (3.20), and therefore principal stresses, then eq. (3.20) can
be written in the form:
( -1) ( -2) ( -3)=0, (3.22)
For each i the system (3.18) has a non-trivial solution which, using the condition nini = 1,
gives the components of one principal direction.
If principal directions are chosen as coordinate axes, stress tensor becomes diagonal (cannonic
form):
1 0 0
= 0 2 0 . (3.24)
0 0 3
32
3. STRESSES
1 I
p = -p= ii = 1 ., (3.25)
3 3
where conventionally pressure p has a negative sign since the stresses of compression
(pressure) are negative. Then, the stress tensor can be divided into spherical and deviatoric
parts, i.e. into the sum of hydrostatic or spherical stress which causes the change of volume,
and deviatoric stress which causes the change of shape (in isotropic materials):
= pI + d = [pij + (ij – pij)] iiij . (3.26)
The principal directions of and d are the same since they represent the normals to the
planes with zero tangential stresses, and their principal values are:
i = p + di (i = 1, 2, 3). (3.27)
33
4. DEFORMATION AND FLOW
Chapter 4
Deformation is the change in the dimensions and shape of a body between the initial (reference,
undeformed) and final (deformed) configurations. Generally, only initial and final
configurations are considered, without paying attention to intermediate stages of deformations.
4.1.2 Configurations
Let’s assume that at a reference time t = t 0 , which is usually taken as zero and is called the
initial time, a material body B occupies the part (region) of the space R0. Particles of body B
continuously fill the space R0, and there is a one-to-one correspondence with particles of body B
and points in R0. The position of each particle P0 of the body can then be defined with the
position vector (Fig. 4.1), which in a fixed coordinate system O0123 is given as:
= 1i1(0) + 2i (0) (0)
2 + 3i 3 . (4.1)
34
4. DEFORMATION AND FLOW
In the next time instant t the body B continuously occupies the space R, and all particles of B
uniquely correspond to points in R. Then, a new position of the particle P0, denoted as P in the
new configuration R, can be specified in a coordinate system Ox1x2x3 by a position vector:
x = x1i1 + x 2i 2 + x 3i 3 . (4.2)
3
x3
P0
P
(0)
x
i3 i3
i2
O0 O x2
i1
i1
(0)
i2
(0)
2
1
x1
Initial (reference, undeformed) Current (deformed)
configuration configuration
This expression gives the current location x of the particle which was at point at time t = 0.
Equation (4.3) can be interpreted as a mapping of the initial configuration into the current
configuration. If the Jacobian of this transformation
x
J = det i
0 , (4.4)
j
A typical particle which moves according to (4.3) must have its reference (initial) position in
R0 and, therefore, point is its reference point. Regardless of the particle motion, it can only
change its current position and it will always have the same reference position. Coordinates i
can therefore be regarded as fixed and are called material or Lagrangian coordinates. The
35
4. DEFORMATION AND FLOW
system of all particles at t = t 0 , i.e. region R0, is called the reference (initial or undeformed)
configuration.
On the other hand, the position vector x denotes the point in space which can be ‘filled’ with
different particles at different instants of time. Coordinates xi are called spatial or Eulerian
coordinates, and the region R occupied by the body at time t is called the current or deformed
configuration.
Continuum motion can be described by using either material i or spatial xi coordinates. The
current position x of a particle at any time t can be found if its reference position is known, or
the reference position can be found if the current position x is known. If the motion is
desribed using eq. (4.3) such a description is called a material or Lagrangian description, and
the description by (4.5) is called spatial or Eulearian.
When using the Lagrangian description, a particle is followed in time. Eulerian description
gives information of the motion in a given spatial point of different particles which ‘fill’ that
point at different times. The choice of the description will depend on the problem considered,
but usually the Lagrangian description is used for solid body motion (and deformation), and the
Eulerian for fluid body motion.
In continuum mechanics, the coordinate systems O0123 and Ox1x2x3 are normally chosen to
coincide to each other, and then:
u= x . (4.8)
These expressions have different physical interpretation. Expression (4.9) represents the
displacement at time t of the particle with the position vector , while expression (4.10)
represents the displacement of any particle which is in the position x at time t.
36
4. DEFORMATION AND FLOW
4.2 Deformation
For small deformations all of these measures are practically equivalent. The choice of a
particular measure of deformation will depend on the relationship between the stress and the
deformation, i.e. on the constitutive relationship.
4.2.2 Gradients
Deformation gradients
Let’s observe an infinitesimal distance between two particles of a continuum body subjected to
deformation (Fig. 4.2). Then, for example:
x
dx = F d or dx i = i d j , (4.13)
j
and:
i
d = H dx or di = dx , (4.14)
x j j
where
37
4. DEFORMATION AND FLOW
xi
F = (grad x )T or Fij = , (4.15)
j
Q
x3 3 u+du
Qo
+d dx
d
u
Po P
x
x2
x1 2
1
Displacement gradients
If the relative displacement du = dx - d is taken as the measure of the deformation, one gets:
du = dx d = F d - I d = (F I) d = J d , (4.19)
or
du = dx d = I dx - H dx = (I H) dx = K dx , (4.20)
where:
u x
J = (grad u )T = [grad (x ) ]T = F I or Jij = i = i ij , (4.21)
j j
38
4. DEFORMATION AND FLOW
or
x k x k
(dx) 2 = dT G d or (dx)2 = d d , (4.24)
i j i j
where
T x k x k
G = F F or Gij = (4.25)
i j
or
k k
(d) 2 = dxT C dx or (d)2 = dx dx , (4.27)
xi x j i j
where
T k k
C = H H or Cij = (4.28)
xi x j
is the Cauchy deformation tensor. Similar to tensors F and H, tensors G and C are inverse of
each other, i.e. C = G-1.
39
4. DEFORMATION AND FLOW
where:
1 1 1 x x
where:
1 1 1
E= (I C) = (I HT H) or Eij = ij k k , (4.32)
2 2 2 x i x j
L=
1
2 ( ) (
1
( J + I)T (J + I) I = JT + J + J JT , or
2 )
L=
1
2
( T
) ( T
)
grad u + grad u + gradu grad u , or
(4.33)
1 u i u j u k u k
Lij = + +
2 j i i j
.
E=
1
2 ( 1
) (
I (I K )T (I K ) = KT + K K K T , or
2 )
E=
1
2
( [ T T
]
grad x u) + grad x u grad x u (grad x u) , or (4.34)
1 ui u j uk uk
E ij = +
2 x j xi x i x j
.
Factor 1/2 is introduced so that L and E would reduce to the corresponding small strain tensors
after higher order terms are neglected. It can be noticed that in cases where there is no
deformations (du = 0, dx = d ), L and E become null tensors while G and C are unity tensors.
40
4. DEFORMATION AND FLOW
configurations can be ignored (dx d), and L and E reduce to the well known Cauchy small
(or infinitesimal) strain tensor:
Ls = E s =
1
2 [
(gradu)T + gradu , or ]
1 u u j 1 ui u j
lij = ij = i + = +
2 j i
2 x j xi
, (4.35)
where the grad operator can be applied in either Lagrangian or Eulerian coordinates, and lower
case letters l and are used to denote small strains.
Expressions (4.35) can also be obtained in the following way. If u = u(x,t) is expanded in the
vicinity of P0 by using the Taylor series expansion and neglecting the higher order terms, one
gets:
u
u = u0 + du = u0 + (grad u)0 dx or ui = ui 0 + i dx j .
T
(4.36)
x j 0
The tensor (grad u)0 can be expressed as the sum of its symmetric and anti-symmetric parts
T
giving:
u = u0 +
1
2
[ 0
1
2
] [
(grad u)T + grad u dx + (grad u)T grad u dx
0 ,
] (4.37)
= u0 + E s0 dx + 0 dx
=
1
2[(grad u)T grad u ] 1 u
or ij = i
u j
2 x j xi
, (4.38)
It can be seen that the motion of a body can be decomposed into two parts: rigid body motion
consisting of translation and rotation, and deformation. The first part obviously does not cause
stresses and is therefore neglected in continuum mechanics.
41
4. DEFORMATION AND FLOW
Diagonal elements of the small strain tensor represent the change in length per unit original
length in the direction of the coordinate axes, and are called normal strains. Off-diagonal terms
represent one-half the angle between two line elements originally at right angle to each other,
and are called shear strains. The concept of small (infinitesimal) deformations is used in the
analysis of elastic bodies. It is also useful as a first step in analysing of large deformations.
2[
1 s s 2 s
] 1
2
2
I2 = L :L tr (L ) = [lij lij (lii ) ] = ( l1l2 + l2l3 + l3l1) , (4.44)
1
I3 = det(Ls) = eijke pqr lip l jqlkr = l1l2l3 ,
6
are the invariants , and the roots of (4.43) l1, l2, l3 are principal values of the Lagrangian small
strain tensor. Principal directions and planes correspond to these principal values. Analogue
relationships can be obtained for Eulerian small strain tensor Es.
It is worth noticing that small deformation tensors L, E, G and C are symmetric and therefore
have three real principal values with corresponding three (orthogonal) principal directions.
Since L and G, and E and C differ by the constant, their corresponding principal directions
coincide.
42
4. DEFORMATION AND FLOW
xi
F = R U = T R or Fij = = RikU kj = Tik Rkj , (4.45)
j
where U is the right and T is the left stretch tensors. This decomposition can be interpreted
using expression (4.13) and
dx = F d = R U d , (4.46)
as the body stretch by U followed with the rigid body rotation by R (Fig. 4.3), or as the rigid
rotation by R followed by the stretch T.
F
d dx
U R
,
dx
Polar decomposition is theoretically very important, but the stretch tensors are not practical
since their components are complex functions of deformation gradients. From the practical
calculations, components of the deformation tensors G and C are more useful.
[ ( )]
Ls = Ls,s + Ls,d = v I + Ls,d = v ij + lij v ij i ii j . (4.48)
The spherical component represents the change in volume and it can be shown that:
V0
l1 + l2 + l3 = I1 = 3 v , (4.49)
V0
where V0 is the initial volume. The deviatoric component represents the change in the shape of a
body.
43
4. DEFORMATION AND FLOW
If, for example, the components of Cauchy small strain tensor are expressed explicitly in terms
of coordinates, then six independent equations (4.35)
1 u u j
ij = i +
2 x j xi
(4.50)
can be regarded as six partial differential equations with three unknown displacement function
ui. This system is obviously over-determined and, in general, its solution does not exist for
arbitrary ij, or it is not compatible, and additional conditions are needed.
1 3ui uj
2 3
ij
= + , (4.51)
x k x l 2 x jx k x l xi xk x l
2 jl 1 uj 3ul
3
= + , (4.52)
x ix k 2 x l xi x k x j xix k
44
4. DEFORMATION AND FLOW
This is the necessary and satisfactory compatibility conditions, known as St. Venant
compatibility equations. There are 81 equations but only 6 are independent (due to symmetry of
Es):
211 2 22 212
+ = 2 ,
x 22 x12 x1x 2
2 22 2 33 2 23
+ =2 ,
x32 x 22 x2x 3
2 2
33 11 2 31
+ = 2 ,
x12 x 32 x3x1
(4.54)
23 31 12 211
+ +
= ,
x1 x1 x 2 x3 x2x 3
2
23 31 12 22
+
= ,
x 2 x1 x 2 x3 x3x1
23 31 12 233
+ +
= .
x 3 x1 x 2 x3 x1x 2
In case of finite strains, the compatibility conditions are considerably more complex.
The Lagrangian description follows the continuum particle and is conventionally used in solid
mechanics where the deformation is of particular interest. On the other hand, the Eulerian
approach is used in fluid mechanics where the velocity field is the main information required.
While the Lagrangian approach focuses on continuum particles (defined by their positions at t =
0), the Eulerian approach is concerned with a particular region of space (control volume).
Since the functions (4.3) and (4.5) are inverse of each other, any physical (scalar, vector or
tensor) continuum quantity F expressed for a continuum particle (in Lagrangian or material
description) with a function
F = F1(,t) = F1(i ,t) , (4.55)
can also be expressed in reference to a particular location in space occupied by the particle
(Eulerian or spatial description) with a function
F = F1((x,t),t) = F2 (x,t) = F2 (xi ,t), (4.56)
45
4. DEFORMATION AND FLOW
where indices 1 and 2 in F1 and F2 are employed to demonstrate that a function in Eulerian
coordinates is not in general the same as in Lagrangian coordinates.
The instantaneous position vector of a given particle is a function of time, and its material
derivative is the velocity of that particle and, at the same time, represents the velocity of
continuum in a given point of space, i.e.
dx Dx
= = v(x,t). (4.57)
dt Dt
This is an important relation, especially in fluid mechanics, where the function = (x,t) , i.e.
the initial positions of fluid elements passing through a given domain of space (control volume)
is neither of interest nor important.
In Lagrangian coordinates
DF DF1 (,t) F1 (,t)
= = , (4.58)
Dt Dt t
while in Eulerian coordinates (notice that x is a function of t with material derivative Dx/Dt)
DF DF2 (x,t) F2 Dx
= = + gradF2 , (4.59)
Dt Dt t Dt
or by using (4.57)
DF F
= + v(x,t) gradF . (4.60)
Dt t
Local Material
change change
The first term on the right hand side of expression (4.60) represents the temporal rate of change
in a given point of space (local change) and is also called spatial time derivative. Time
derivatives with spatial coordinate x held fixed are called spatial time derivatives. The second
term is a consequence of the fact that the continuum changes the position in space (material
change) and is called convective or transport term.
46
4. DEFORMATION AND FLOW
Based on (4.60), one can define operators for material derivative in Lagrangian coordinates:
D(...) (...)
= . (4.61)
Dt t
or in Eulerian coordinates:
D(...) (...)
= + v grad(...) . (4.62)
Dt t
Material derivative of velocity is acceleration. According to (4.61) and (4.64), the acceleration
in Lagrangian coordinates is given as:
Dv v 2x 2u
a(, t) = = = = , (4.67)
Dt t t 2 t 2
In relation to Lagrangian description, one can define particle trajectory as the system of all
spatial points passed through by the particle. With reference to Eulerian description, a
streamline is defined as the line of the velocity vector field, i.e. at every time instant velocity
vectors are in the direction of tangent to the line. In a given time instant, a velocity vector is
tangent to both trajectory and streamline. The trajectory equation is obtained by solving the
following vector equation:
47
4. DEFORMATION AND FLOW
dx
= v(x,t). (4.69)
dt
with the initial condition x = at t = t0, while the vector equation for the streamline at t = t0 is:
dx = v(x,t 0 ). (4.70)
where is a parameter. If the velocity is not function of time, the corresponding streamline and
trajectory coincide.
where V = V(t) is the material volume (i.e. the volume which moves with the continuum and no
mass flux occurs across the volume boundary), is also well defined function of space and time.
By substituting variables in (4.71) and using (4.3), one gets:
where J = dV/dV0 is the Jacobian of the transformation (4.3) defined by expression (4.4), and V0
= V(t0) is the material volume at t = t0. Based on (4.72), the material derivative of the integral
(4.71) can be expressed as:
D D
Dt F (x,t)dV =
Dt F[x(,t),t]JdV0 . (4.73)
V V0
Noticing that V0 does not depend on time, which means that the order of differentiation and
integration can be swapped, and using Euler expansion formula:
DJ
= Jdivv . (4.74)
Dt
one gets:
D DF DJ DF
Dt Dt Dt
FdV = J +F dV = + Fdivv JdV0 . (4.75)
Dt 0
V V0 V0
48
4. DEFORMATION AND FLOW
where S is the area bounding the material volume V. The equation (4.77) represents the
Reynolds transport theorem which states that the rate of change of integral of function F over
the material volume, is equal to the sum of the rate of change of that integral over the fixed
volume in space which coincide with the material volume at a given time, and the flux of F
through the bounding surface S.
x3 P
v
dx dv
x
Po
vo
xo
x1 x2
If the function v(x, t) is expanded in the vicinity of P0 by using the Taylor series expansion and
higher order terms are neglected, one gets:
v
v = v 0 + dv = v0 + (grad v)0 dx or vi = vi 0 + i dx j ,
T
(4.78)
x j 0
where
v i
Y = (grad v)
T
or Yij = , (4.79)
x j
is the spatial velocity gradient.
49
4. DEFORMATION AND FLOW
As before, the tensor (grad v )T0 can be decomposed into the sum of its symmetric and anti-
symmetric parts giving:
v = v0 +
1
2
[ 0
1
2
] [
(grad v)T + grad v dx + (grad v)T grad v dx
0 ,
] (4.80)
= v 0 + D0 dx + V0 dx
where
D=
1
2[(grad v)T + grad v ] 1 v
or Dij =
i +
v j
2 x j xi
, (4.81)
V=
1
2[( grad v) grad v
T
] or Vij =
1 vi v j
2 x j xi
, (4.82)
The rate of deformation tensor D is not limited to small deformations although it has the same
form as the small strain tensor Es (or Ls). It can be seen that, analogously to deformation, the
velocity of a body can be decomposed into two parts: the first corresponding to the velocity of
rigid body motion (translation and rotation), and the second corresponding to the rate of
deformation. The first part does not cause stresses and is normally neglected in continuum
mechanics.
[ ]
s
DL DE s 1
= (grad v) + grad v = D ,
T
= (4.86)
Dt Dt 2
and
D 1
[
= (grad v) grad v = V .
Dt 2
T
] (4.87)
50
4. DEFORMATION AND FLOW
where components Dij are known as the true (natural) strain increments.
showing that the rate of volumetric dilatation is equal to the sum of diagonal elements of the
rate of deformation tensor.
If the mean rate of volumetric dilatation is defined as:
I 1
= 1 = Dii , (4.93)
3 3
then the deformation rate tensor can be decomposed into the sum of the spherical (dilatational)
part, which corresponds to the rate of the volume change, and the deviatoric part which
represents the rate of change of shape:
s d d
(
D = D + D = I + D or Dij = ij + Dij ij . ) (4.94)
51
5. FUNDAMENTAL LAWS
Chapter 5
Apart from the fundamental laws, which are independent of the type of continuum under
consideration, additional laws or constitutive relationships will also be used. These are
specific for each continuum, such as Hooke’s elasticity law, Stoke’s viscosity law, Fourier’s
law of heat transfer, etc.
52
5. FUNDAMENTAL LAWS
For a system:
If denotes a global property of a system such as mass, or momentum and = (x, t)
represents the density of that property, then the relationship between them is given by a
material integral (over a given material domain instead of a given spatial domain) as:
H= dV , (5.1)
V
where is the mass density and V is the system volume (material domain). All fundamental
laws are concerned with the rate of change of due to the interaction between the system
and its environment. These changes can be proportional to the mass of the system or to the
surface bounding the system as:
DH
Dt
= m dV + S
ndS . (5.2)
V S
This expression can be regarded as the general conservation law or the general transport
equation. By using Gauss theorem (2.80) one gets:
( m + div ) dV ,
DH
= (5.3)
Dt S
V
and due to (5.1) [and (5.20)], the differential form of the general conservation law becomes:
D( )
= m + div S . (5.4)
Dt
Volumetric (body, mass) term m and surface term s will be given in the next section for each
fundamental law.
where V and S represent the volume and the bounding surface of the control volume,
respectively. This is the general conservation law for the control volume in the integral form.
53
5. FUNDAMENTAL LAWS
t
( )
+ div (v) dV = m + div S dV . (5.6)
V V
The differential form of the mass conservation law or the continuity equation can be obtained
using the following form of expression (5.8):
where V is the volume occupied by a material at time t that was occupying volume V0 at t0,
and 0 = ( , t0) and = (x, t) are the densities at t0 and t, respectively. By substituting
variables (x = x ( , t)) in the integral on right hand side of (5.9), one gets:
54
5. FUNDAMENTAL LAWS
since dV = J dV0 and dV0 = const. This equation is Lagrangian or material form of the
continuity equation. From (5.13) it follows that:
D D D D
Dt
( dV ) = ( dV ) +
Dt Dt
dV =
Dt
dV , (5.14)
=0
which has an important implication for other conservation laws.
t dV + v ndS = 0 , (5.15)
V S
Change of mass Mass flux
in the CV through CS
Based on (5.7), the differential continuity equation in the spatial form can be written in the
form:
+ div( v) = 0 . (5.16)
t
Local Convective
change change
or:
D
+ v grad + div v = + div v = 0. (5.17)
t Dt
55
5. FUNDAMENTAL LAWS
D( ) D ( )
= = + div (v) = + v grad . (5.19)
Dt Dt t t
Strong conservation Weak conservation
form form
This is a consequence of Reynold’s transport theorem and mass conservation, and is useful
expression in continuum mechanics. For example, one can see that (5.8) and (5.13) follow
directly from (5.22).
5.2.3 Conservation of linear momentum (1st Euler’s law, 2nd Newton’s law)
According to (5.4) and (5.18), the differential form of the momentum equation takes the form:
Dv
= fb + div . (5.24)
Dt
Rate of change of Body Surface
linear momentum force force
56
5. FUNDAMENTAL LAWS
(v )
t
dV + v( v n)dS = fbdV + ndS , (5.25)
V S V S
Change of mom. Flux of mom. Total body Total surface
in CV across CS force force
Equations (5.24) and (5.26) can be expressed in the index form in Cartesian coordinate
system as:
Dvi ji
= fb,i + , (5.27)
Dt x j
According to (5.4), the differential form of the moment of momentum principle takes the
form:
57
5. FUNDAMENTAL LAWS
D( x v)
= x fb + div (x ) . (5.30)
Dt
Rate of change of Moment of Moment of
moment of momentum body forces surface forces
This is also known as the second Cauchy law. The first term on the left and the second term
on the right hand side of (5.30) can be expressed in the following way:
D( x v) Dv Dx Dv Dv
=x +v =x +v v= x ,
Dt Dt Dt Dt Dt (5.31)
div( x ) = x div + grad x .
hence:
T = or ji = ij . (5.34)
This means that the second Cauchy law can also be formulated in the following way: the
necessary and sufficient condition for the moment of momentum principle to be satisfied is
the symmetry of the Cauchy stress tensor.
[ ( x v)]
t
dV + (x v)( v n)dS = ( x fb )dV + (x ) ndS . (5.35)
V S V S
Change of mom. of Flux of mom. of Total moment of Total moment of
momentum in CV momentum across CS body forces surface forces
58
5. FUNDAMENTAL LAWS
Here, only thermodynamical processes where the sources of energy are mechanical work and
heat are considered. The principle of conservation of energy, i.e. the energy balance
principle, states that the rate of change of total energy is equal to the work done by surface
and body forces plus the heat delivered to the body by the heat flux and other heat sources.
The integral form of the energy balance principle takes the form:
D
Dt
edV = q ndS + hdV + fb vdV + ( v) ndS , (5.37)
V S V V S
Rate of change of Heat Heat Work done (power)
total energy delivered source
where the sign of the heat flux term is negative since positive heat flow is out of the body.
Here, q is the heat flux, h is the heat source, and e is the total specific energy. In continuum
mechanics, e is often taken as the sum of internal energy u and kinetic energy, i.e.:
1 v2
e = u + vv= u + . (5.38)
2 2
According to (5.4) and (5.18), the differential form of the energy balance equation takes the
form:
De
= divq + h + fb v ( v) .
+ div (5.39)
Dt
Rate of change Heat Heat Work by Work by
of energy flux source body forces surface forces
(5.41)
(e)
+ div (ev) = divq + h + fb v + ( v)
div
t
Local Convective Heat Heat Work by Work by
change change flux source body forces surface forces
59
5. FUNDAMENTAL LAWS
D v
2
= fb v + (div ) v .
Dt 2
(5.43)
It is important to notice that this equation is a consequence of the momentum equation and
not the energy equation.
Heat equation
Using (5.38), the energy equation (5.39) [or (5.41)] can be written in the form:
Du Dv
+ v ( ) v + :gradv ,
= divq + h + f b v + div (5.44)
Dt Dt
giving:
Du Dv
+ v fb div() = divq + h + :gradv . (5.45)
Dt Dt
This equation is known as the heat equation. By summing the equation of mechanical energy
and the heat equation one gets the equation of total energy (5.39) [or (5.41)]. Equation (5.46)
in the index notation takes the form:
Du q j v j
= + h + ij , (5.47)
Dt x j xi
60
5. FUNDAMENTAL LAWS
Du q qy qz v x v v
=
x + + + h + xx + yx x + zx x +
Dt x y z x y z
v y v y v y
xy + yy + zy + (5.48)
x y z
v v v
xz z + yz z + zz z
x y z
While the motoric work affects the mechanical energy change, the deformation work is
converted into heat and increases the internal energy in the control volume. By decomposing
the stress tensor into its spherical and deviatoric parts (3.26), one gets:
( d
:gradv = pI + :gradv = ) pdivv + d :gradv. (5.50)
Deformation Reversible Irreversible
work part part
It can be seen that the deformation work consists of the reversible part (which is a
consequence of the spherical stress component) which can be converted back into the
mechanical energy, and of the irreversible part (which is a consequence of the deviatoric
stress component) which cannot be converted back into the mechanical energy.
When the heat flux and heat sources vanish, i.e. in a purely mechanical process, the energy
equation becomes [see (4.81) and (4.82)]:
Du
= :gradv = :(D V) = :D = ij D ji , (5.51)
Dt
which is no longer a partial differential equation. It can be seen that the internal power (the
internal energy rate) is given by the contraction (scalar, inner or double-dot product) of the
rate of deformation and the Cauchy stress. We therefore say that the rate of deformation and
the Cauchy stress are conjugate in power (or in work or in energy).
61
5. FUNDAMENTAL LAWS
quantitative measure of microscopic randomness and disorder. The entropy of a system can
change either through interactions with surroundings, or by changes that take place within the
system, i.e.:
ds = dse + dsi , (5.52)
where ds is the increase in specific entropy, dse is the increase due to interaction with the
exterior, and dsi is the internal increase. The change dsi is never negative, it is zero for a
reversible process, and positive for an irreversible process:
dsi > 0 (irreversible process)
. (5.53)
dsi = 0 (reversible process)
In a reversible process, if dqR denotes the heat supplied per unit mass to the system, the
change dse is given as:
dq R
dse = (reversible process), (5.54)
T
where T is the absolute temperature.
According to the second law, the rate of increase of total entropy of the system is never less
than the sum of the entropy influx through the surface plus the entropy produced internally by
body sources. Mathematically, this entropy principle may be expressed in integral from as the
Clausius-Duhem inequality:
D 1
Dt sdV q ndS +
T dV , (5.55)
V S V
where is the local entropy source per unit mass. is sometimes given as = r / T , where
r is internal heat supply per unit mass and time (e.g. from radiation). The equality in (5.55)
holds for reversible processes, while the inequality applies to irreversible processes. The
difference between the rate of change of entropy and the rate of entropy input in a body
determines the total production of entropy , i.e.:
D 1
Dt
= sdV + q ndS dV 0. (5.56)
T
V S V
This relation describes the direction of the energy transfer and postulates irreversibility of
various thermodynamic processes (observations show that heat flows from the warmer to the
colder regions of a body; mechanical energy can be transformed into heat by friction, and this
can never be converted back into mechanical energy). The situation < 0 never occurs; it
would mean that molecules organise themselves globally. A thermodynamic process is called
reversible if it is not accompanied by any entropy production ( = 0). Real processes are
62
5. FUNDAMENTAL LAWS
always irreversible and characterised by > 0. These are always associated with energy
dissipation.
t T
By using (2.76):
q 1 1
div = divq 2 q gradT , (5.59)
T T T
giving:
Ds 1 1
divq + 2 q gradT + . (5.60)
Dt T T
1
Since the heat flows from hot to cold, we have 2 q gradT 0 . Expression (5.58) gives
T
the local form of the entropy production rate per unit mass as:
Ds 1 q
+ div 0. (5.61)
Dt T
This inequality must be satisfied for any process, and for this reason it plays an important role
in imposing restrictions upon constitutive relations discussed later.
From Gibbs relation (for an ideal gas) we have that:
1
du = Tds pdv = Tds pd , (5.62)
where v is specific volume, and this can be written in the following form:
Ds Du D(1/ )
T = p . (5.63)
Dt Dt Dt
Using the continuity equation (5.17), the last term in the above equation becomes:
D(1 / ) 1 D 1
= 2 = divv. (5.64)
Dt Dt
Now, with the help of (5.64) and (5.46), expression (5.63) takes the form:
63
5. FUNDAMENTAL LAWS
Ds
T = divq + h + :gradv + pdivv, (5.65)
Dt
and by using (5.50) and the fact that p = p (3.25), one gets:
Ds
T = divq + h pdivv + d :gradv + pdivv = divq + h + d :gradv . (5.66)
Dt
From (5.59) it follows that:
1 q 1
divq = div + 2 q gradT , (5.67)
T T T
and finally, (5.66) takes the form:
Ds q 1 h 1 d
= div q gradT + + :gradv . (5.68)
Dt T T2 T T
Total increase Reversible Increase in entropy Increase in entropy
of entropy heat exchange because of irreversible because of irreversible
heat exchange due to exchange of mechanical
gradT work to heat
v y v y v y
(
xy x + yy + p ) y
+ zy +
z
v z v z v z
xz x + yz y + ( zz + p) z
For an irreversible adiabatic process Ds/Dt>0 by the second law, and in much of continuum
mechanics, it can be shown from relations derived similarly to (5.68) that the dissipation
d
function :gradv must be positive, since both and T are positive.
64
5. FUNDAMENTAL LAWS
5.2.7 Discussion
Five fundamental laws of continuum mechanics presented above, result in a system of 9
balance equations in Cartesian coordinate system: 1 continuity, 3 momentum, 1 energy, 3
moment of momentum, and 1 entropy equation. This system, however, is an open system as it
contains 19 unknown scalar functions: , vj, ij, qi, u, T and s (i = 1, 2, 3 and j = 1, 2, 3). To
close the system, 10 additional equations are required, which will link the 19 dependent
variables. For this purpose, the following constitutive equations are used:
1) Relationship between stress tensor and deformation tensor and/or rate of deformation
tensor (6 equations),
2) Relationship between the heat flux and temperature gradient (3 equations),
3) State equation (1 equation).
Basic constitutive equations will be given in the next chapter.
65
5. FUNDAMENTAL LAWS
there are three additional stress tensors, also called pseudo-stress tenors, defined as:
1. Kirchoff stress tensor (1) obtained by scaling the Cauchy tensor:
(1)
= 0 = J . (5.72)
2. First Piola-Kirchoff stress tensor (1), gives the current force acting on the deformed
configuration per unit of the undeformed area dS0:
(1)
df = n0 dS0 . (5.73)
3. Second Piola-Kirchoff stress tensor (2), gives the current force transformed to the
reference configuration df0 per unit of the undeformed area, where df = F df0 :
where the relationship between the deformed area dS = ndS and undeformed area
dS0 = n0dS0 is given by Nanson’s relation:
ndS = 0 n0 F 1dS0 = Jn0 F 1dS0 . (5.75)
Forces and surfaces involved in defining Piola-Kirchoff tensors are illustrated in Fig. 5.1.
df 0
n0 df
dS 0
df
dS
n
S V
S0 V0
Figure 5.1 Surface forces and surface elements involved in defining Piola-Kirchoff tensors
Using expressions (5.71), to (5.75), one can relate Cauchy with Piola-Kirchoff stress tensors.
First, the current force acting on the current surface can be expressed as:
66
5. FUNDAMENTAL LAWS
(1) 1 (2) T
= 0 F = F , (5.78)
( ) ( )
T T
(2) = 0 F 1 F 1 = (1) F 1 . (5.79)
These expressions show that second Piola-Kirchoff stress tensor is symmetric, while the first
Piola-Kirchoff stress tensor is not, and is therefore inconvenient to use.
Integral form:
If substitution x = x(,t) is introduced in the equation of motion (5.23), i.e. the integrals are
expressed in the reference configuration, then:
D
Dt vJdV0 = fb0 JdV0 + n0 (1)dS0 . (5.80)
V0 V0 S0
Taking into account that 0 = J (5.12), one gets the momentum equation for the reference
configuration in terms of the first Piola-Kirchoff stress:
D
0vdV0 = 0fb0dV0 + n0
(1)
dS0 , (5.81)
Dt
V0 V0 S0
where fb0 (,t) = fb[ x(,t)] . This expression is completely analogous to (5.23) written for the
current configuration with Cauchy stress tensor. Since the first Piola-Kirchoff stress tensor is
not symmetric, the second Piola-Kirchoff stress is more often used in the momentum
equations for the reference configuration:
D
Dt 0vdV0 = 0fb0dV0 + n0 (2) FT dS0 , (5.82)
V0 V0 S0
67
5. FUNDAMENTAL LAWS
Differential form:
The differential form of equations (5.81) and (5.82) is:
Dv (1)
0 = 0fb0 + div , (5.83)
Dt
0
Dv
Dt (
= 0fb0 + div (2) F T , ) (5.84)
while the corresponding equations in the index form in Cartesian coordinate system are:
Dvi (1)
ji
0 = 0 fb 0,i + , (5.85)
Dt j
Dvi (2) x i
0 = 0 fb 0,i +
j
jk k
. (5.86)
Dt
Since:
DF D x j Dx j v j v j xi
=
i ji k =
i ji k = i j ik = i i = Y F, (5.88)
Dt Dt k k Dt k xi k j k
68
5. FUNDAMENTAL LAWS
If the deformation work is expressed by V :D dV and the second Piola-Kirchoff stress
tensor is used, then with the help of (5.77) and (5.12) one gets:
Pdef = :D dV =
0( ) (
F (2) FT :D JdV0 = (2): F T D F dV0 , ) (5.90)
V V0 V0
DL T
and since = F D F , where L is defined by (4.30), we have that:
Dt
DL 1 (2) DG
Pdef = (2) : dV0 = : dV0 . (5.91)
Dt 2 Dt
V0 V0
The total heat delivered through the current area S can be written in terms of the undeformed
surface S0 by using Nanson’s relation (5.75):
(
q ndS = q Jn0 F 1dS0 = JF 1 q n0dS0 , ) (5.92)
S S0 S0
where the term in brackets can be regarded as the heat flux per unit reference area. Finally,
the heat equation in the integral form can be written for the reference configuration with the
first Piola-Kirchoff stress tensor as:
DFT
D
Dt ( )
0 udV0 = JF 1 q n0dS0 + h0dV0 + (1) : Dt
dV0 , (5.93)
V0 S0 V0 V0
69
5. FUNDAMENTAL LAWS
Boundary conditions
Let’s consider a mechanical problem in region V with boundary conditions prescribed on S.
The problem is said to be well posed if at every point on S and for every direction of a
coordinate system (e.g. local coordinate system with one axis normal to the surface S), there
is one and only one condition prescribed via either displacement components ui (or velocity
vi) or surface tractions fsi:
ui = ubi or f si = fbi (i = 1,2,3) on S . (5.95)
Virtual displacements
s
Virtual strain tensor E corresponds to the virtual displacement u , and its components are
given as:
ij = +
( )
.
1 (ui ) u j
(5.97)
2 x j x i
Virtual forces
When certain external (body and surface) forces are acting on a body, a stress state is said to
be statically admissible if stress tensor
1. Satisfies the equilibrium equations [eqs. (5.24) or (5.26) with Dv/Dt = 0] in the interior V
of the body:
70
5. FUNDAMENTAL LAWS
Principle
If a problem is well posed, virtual displacements kinematically admissible and virtual forces
statically admissible, then the virtual work of external forces is equal to the virtual work of
internal forces:
Proof of equivalency
Let us show that the virtual work principle (5.100) follows from the equilibrium equation
(5.98). Since forces remain unchanged during virtual displacement, then the dot product of
the total external force and virtual displacement u , gives the virtual work of external forces:
Wext = fb u dV + n u dS . (5.101)
V S
By applying the Gauss theorem to the surface integral in (5.101), one gets:
Wext = [:gradu + ( fb + div) u ]dV = :E s dS , (5.102)
V S
because the term in the curly bracket is zero (equilibrium equation), symmetric part of
:grad u is equal to the virtual strain tensor Es , and the double dot product of symmetric
stress tensor and anti-symmetric part of the tensor :grad u is zero.
It can be shown that the virtual work principle (5.100) in reference configuration expressed in
terms of the first Piola-Kirchoff stress tensor takes the form:
(1) T (2)
From the above two expressions, one can see that and F , and and L are work
conjugate pairs.
71
5. FUNDAMENTAL LAWS
Generalisation
It can also be shown that the virtual work principle (5.100) is valid if the inertial force is
added to the body force. Namely, by taking that fb + div = Dv / Dt in (5.104), one gets:
Dv
fb u dV + n u dS = :E
s
+ u
dV , (5.105)
Dt
V S V
or
1
fb u dV + n u dS = :E dV + v 2 dV .
s
(5.106)
2
V S V V
Work by external forces Deformation Change of kinetic
Work energy
where K is the total kinetic energy, and Wdef is the total work of deformation. By inserting
(5.108) into the energy balance (first law of thermodynamics)
U + K = Q + Wext , (5.109)
=E
U = Q + W def , (5.110)
where E is the total energy and U is the total internal energy. Therefore, the virtual work
principle can be considered as an energy principle.
72
6. CONSTITUTIVE RELATIONS
Chapter 6
Constitutive relationships
6.1 Introduction
There is a large number of different materials, hence there is a large number of corresponding
constitutive equations. However, it is somewhat surprising that two constitutive relationships,
one for linear elastic solid and another for linear viscous fluid, can well describe behaviour of
many real materials for a range mechanical and thermal loading. Of course, real materials are
always more or less different than the two simple ideal materials, and when the differences
73
6. CONSTITUTIVE RELATIONS
become considerable, then one is dealing with non-Newtonian fluids, viscoelastic, plastic or
viscoplastic solids, and corresponding constitutive relations are much more complex.
The ideal elastic (Hookean) solid has the property that it returns to its initial (undeformed)
state after the removal of all mechanical and/or thermal loads. Its constitutive equation is given
by a linear relation between the stress tensor and strain (deformation) tensor of the form:
Stress tensor = f(Deformation tensor).
The ideal Newtonian fluid differs from a solid in that it cannot support a shear stress; in a fluid,
shear deformation (flow) will continue as long as any shear stress is applied. Its constitutive
equations is expressed by a linear relation between the stress and the rate of deformation
(strain rate) tensor of the form:
Stress tensor = f(Rate of deformation tensor).
Unlike the elastic material, viscous fluid has no memory of its initial state; the stress at the
point depends exclusively on the instantaneous rate of deformation at that point.
Not all material can be separated into these two categories (where the stress is only a function
of either deformation or the rate of deformation). The elastic solid and viscous fluid are the
two extreme cases of a wide range of materials including, viscoelastic solids, elastic solids that
have rate dependence but still possess well defined initial undeformed state to which they
eventually return when the loads are removed, or viscoplastic materials, which remind more of
a thick fluid but are still exhibiting some aspects of a solid, etc.
74
6. CONSTITUTIVE RELATIONS
where , μ and are constants which have the same value in all coordinate systems. If cijkl is
symmetric in either ij or kl, the last bracketed term in (6.1) is zero and we have:
( )
cijkl = ij kl + μ ik jl + il jk , (6.2)
which means that any isotropic fourth-order tensor cijkl symmetric in ij or kl is described with
only two constants (scalars) and μ.
These laws describe the simplest, so called linear materials. They are later generalised to give
constitutive equations for arbitrary states of stress and deformation (or flow), and are able to
describe reasonably well the behaviour of a large number of real materials.
where
is the normal stress, is the strain (relative extension), and a proportionality constant
E is so called Young’s modulus.
F S
F
75
6. CONSTITUTIVE RELATIONS
where is the tangential (shear) stress, dv/dy is the velocity gradient, and the proportionality
coefficient μ is the coefficient of (dynamic) viscosity.
v F
h
y
Most solids subjected to relatively small stresses and low temperatures behave according to the
Hooke’s law. However, this is no longer the case under substantial loads and/or high
76
6. CONSTITUTIVE RELATIONS
temperatures, where materials behave in inelastic manner. This inelasticity can be manifested
via plastic flow, hardening, creep, relaxation, etc.
Similarly, almost all gasses and low-density fluids behave according to the Newton’s law.
However, there are some important fluids with non-linear relation between the stresses and rate
of deformation. These are called non-Newtonian fluids.
Inelastic solids and non-Newtonian fluids are also classified as non-linear materials.
dT
q=k , (6.9)
dx
where k is the coefficient of thermal conductivity. It is important to notice that this law is
applicable for all materials, unlike the previous ones which are specific for solids or fluids.
where cv is the specific heat at constant volume, while for an elastic solid we have:
u
(
u = u T, ij ) du = dT = c E dT ,
T
(6.11)
ij
77
6. CONSTITUTIVE RELATIONS
where cE is the specific heat at constant deformation. In the theory of linear elasticity, either cv
and cE can be used in (6.11), as their values are close to each other.
Under the above conditions it can be assumed that the Cauchy stress tensor is proportional to
the small strain tensor:
= C:E s or ij = cijkl kl , (6.12)
where C is the fourth-order tensor with 81 components. Since and Es are symmetric tensors,
then:
cijkl = c jikl and cijkl = c ijlk , (6.13)
[ ( )]
ij = ij kl + μ ik jl + il jk kl , (6.14)
( )
ij = kkij + 2μ ij or = trE s I + 2μE s , (6.15)
where and μ are Lame constants. They are linked via shear modulus G, modulus of elasticity
(Young’s modulus) E and Poisson coefficient as:
E E
μ = G= and = . (6.16)
2(1 + ) (1 + )(1 2)
If the Hooke’s law is expressed in terms of the deviatoric stress (3.26) and strain tensors
(4.48):
1 1
d = (tr )I and E sd = E s trEs I ,
3 3 ( ) (6.17)
78
6. CONSTITUTIVE RELATIONS
one gets:
2
( ) 1
d = + μ trE s I + 2μE sd (tr )I ,
3 3
(6.18)
( )
where p = 1 / 3(tr ) is pressure, v = 1 / 3 trE is the mean volumetric strain (see Section
s
Function W which satisfies (6.25) is called elastic potential or strain energy function. Since the
reference (zero) level of strain energy function can be chosen arbitrarily, and the zero stress
state corresponds to the zero strain, the simplest form of strain energy function that leads to a
linear stress-strain relation is the quadratic form:
1
W = cijkl kl ij , (6.26)
2
79
6. CONSTITUTIVE RELATIONS
Based on the expressions from the previous section, the strain-energy function (6.26) can be
expressed as:
( )
s 2 E
( )
1 s s 2
W = trE + μE :E = trE s + Es :E . (6.28)
2 2(1+ ) 1 2
As W must always be positive (vanishing only for zero strains), it follows that:
E
or:
2
μ > 0 and + μ = K > 0. (6.30)
3
(since we rule out the cases K=0 and μ=0). This implies that:
1
E > 0 and 1<
< . (6.31)
2
In practice, a material with negative does not exist, while beryllium has value of close to
zero. The value =0.5 implies that μ=E/3 and 1/K=0, or elastic incompressibility,
approximated by some rubbers.
6.3.3 Thermo-elasticity
In general, elastic coefficients in constitutive equation (6.12) are temperature dependent. For
small temperature variations this dependency is often neglected and the coefficients are
assumed constant. However, even then the thermal expansion of solids should be accounted for
if the resulting deformations are comparable to those due to mechanical loads, or if the thermal
deformations are constrained giving rise to thermal stresses. The thermal stresses can be
superimposed onto ‘mechanical’ stresses for linear materials (with linear constitutive
equations).
Constitutive relation which accounts for the temperature changes has the form:
= C:E B (T T0 ) or ij = cijkl kl ij (T T0 ) ,
s
(6.32)
80
6. CONSTITUTIVE RELATIONS
where B is the second-order tensor, and T0 is the reference temperature corresponding to the
unstrained state of the material. Linear theory assumes that ij are independent of deformations
and that cijkl are independent of temperature. Then, for an isotropic solid at temperature T0 cijkl
must have the form (6.2), while for an isotropic material and ij = 0 tensor B must have the
form I, and the constitutive relation (6.32) takes the so called Duhamel-Neumann form of the
Hooke’s law:
( s)
= trE I + 2μE (T T0 )I or
s
, (6.33)
ij = kkij + μ ij (T T0)ij
where is the coefficient of volumetric thermal expansion, and is related to the coefficient of
linear thermal expansion by the following relationship:
E
= = (2μ + 3 ) = 3K . (6.34)
1 2
where p0 is the static pressure equal to the mean normal pressure defined by (3.25), with the
overbar added to distinguish from the thermodynamic pressure p, and with the subscript zero
to indicate that it corresponds to the fluid at rest or in uniform flow.
Thermodynamic pressure is defined by an equation of state called the kinetic equation of state
(to distinguish from the caloric equation of state):
p = p(,T ) , (6.36)
where is the density and T is the absolute temperature. An example of an equation of state is
the ideal-gas law:
p = RT , (6.37)
where R is the gas constant. Notice that definition (6.36) does not guaranty equality between p
and p0 = 1 / 3(tr ) .
81
6. CONSTITUTIVE RELATIONS
where p is the thermodynamic pressure, which satisfies the state equation (6.36) [or more
precisely (6.37)]. No real fluid is inviscid, but the expression (6.38) has allowed a number of
analytical solutions for idealised flows.
such fluid is called Newtonian fluid. Since both the viscous stress tensor S and the rate of
deformation tenor D are symmetric, the constitutive relation for an isotropic fluid, similarly to
the isotropic solid, takes the form:
= pI + (trD)I + 2μD or ij = pij + Dkk ij + 2μDij , (6.40)
where and μ are two independent parameters which characterise fluid viscosity.
Constitutive equation of Newtonian fluids takes an interesting form when expressed in terms
of deviators:
1
d = + pI and Dd = D (trD)I . (6.41)
3
Namely, expression (6.40) becomes:
2
d = ( p p)I + + μ (trD)I + 2μDd , (6.42)
3
Substituting (6.43) in (6.42), the first two terms on the right hand side vanish, and one obtains
the following two relations equivalent to (6.40):
d = 2μD d
d
or ij = 2μDij
1 D , (6.44)
p = p trD = p 3 or p = p Dkk = p + 3
Dt
where p = 1 / 3(tr ) is the mean pressure, =1/3(trD) is the mean rate of volumetric dilatation
[see (4.93)], and
2
= + μ, (6.45)
3
is the volumetric viscosity. The second form of the second equation in (6.44) employs:
1 D
trD = divv = , (6.46)
Dt
82
6. CONSTITUTIVE RELATIONS
The coefficient μ is called the dynamic viscosity. Generally, it is a function of temperature, and
in fluids it drops as the temperature rises, while in gasses it increases with temperature.
where q is the heat flux, and is a second-order tensor. In case of an isotropic material K=kI,
giving:
83
6. CONSTITUTIVE RELATIONS
T
q = kgradT or qi = k , (6.51)
x i
( )
u = u T, ij , (6.52)
which gives the internal energy u as a function of T and ij for solids, and:
u = u(,T ) , (6.53)
giving the internal energy u as a function of and T for fluids (see Section 6.26).
84
7. MATHEMATICAL MODELS
Chapter 7
Mathematical models
7.1 Introduction
In this chapter, mathematical models, i.e. closed systems of equations, are formulated for
simple linear elastic solid and Newtonian fluid by combining the fundamental laws from
Chapter 5, which are valid for any continuum, and constitutive relations from Chapter 6 valid
for linear solids and fluids. Specifying initial (for unsteady problems) and boundary
conditions finally completes the models.
Basic equations
85
7. MATHEMATICAL MODELS
• Constitutive equations:
s
[( ) s
]
= 2μE + trE ( 3 + 2μ ) (T T0 ) I (thermo-elastic solid), (7.3)
Hooke’s law is sometimes expressed in the following form, which is obtained by solving
(7.3) explicitly for Es:
1+
Es = (tr ) T I . (7.5)
E E
• Equation of state:
u = cE T . (7.6)
Mathematical model
Hooke’s law assumes small strains so that the differences between material and spatial
descriptions can be neglected. Now, the basic equations of the linear theory of elasticity can
be written in the following forms:
• Integral form:
t t dV = {μ[(gradu) ] }
u
V S (7.7)
+ fbdV ,
V
(c ET )
dV = kgradT ndS + hdV (3 + 2μ)T0 (divu)dV . (7.8)
t t
V S V V
86
7. MATHEMATICAL MODELS
• Differential form:
u
{[ ]}
= div μ (gradu) + gradu + grad[
divu (3
+ 2μ ) (T T0 )] + fb , (7.9)
t t
T
( c E T )
= div(kgradT ) + h ( 3 + 2μ )T0 (divu) . (7.10)
t t
• Index notation:
ui u u j u
= μ
i + + k ( 3 + 2μ) (T T0 ) + fb,i , (7.11)
t t x j
x j x i xi
x k
( c E T ) T u
= k + h ( 3 + 2μ)T0 k . (7.12)
t x j x j t x k
It can be seen that the resulting system of 4 equations (in either integral or differential forms)
consists of 4 unknown functions of spatial coordinates and time ui(x,t) and T(x,t). Equation
(7.8) [or (7.10)] is called coupled energy equation, since its last term contains displacements
and therefore is coupled to the momentum equation. However, this term can be neglected in
most linear thermo-elastic problems. In such cases, the solution of a thermo-elastic problem
reduces to solving two separate (de-coupled) problems: i) solving of energy equation to
obtain the temperature field, and ii) solving of the momentum equations to obtain the
displacement field and, by using Hooke’s law, the stress field.
If , μ, , , k and cE have constant values, then the equations (7.9) and (7.10) can be written
in the form:
2u
= μ u + ( + μ )grad(divu) (3 + 2μ )gradT + fb ,
2
(7.13)
t 2
T
= k T + h ( 3 + 2μ)T0 (divu) .
2
c E (7.14)
t t
87
7. MATHEMATICAL MODELS
Basic equations
• Balance equations in integral form:
t dV + v ndS = 0,
V S
(v )
dV + v( v n)dS = ndS + fbdV , (7.15)
t
V S S V
( u)
dV + uv ndS = q ndS + hdV + (:gradv)dV .
t
V S S V V
• Constitutive equations:
2
= pI μdivvI + 2μD (Newtonian viscous fluid), (7.17)
3
• Equations of state:
Const. Incompressible fluid
= (p,T ) = ( p) Barotropic fluid (7.18)
p / RT Ideal gas
u = u(,T ).
88
7. MATHEMATICAL MODELS
Using constitutive relation (7.17) and the equations of state for incompressible fluid
(=const., u=cvT), one gets the following mathematical model:
• Integral form:
v ndS = 0 , (7.19)
S
(v )
V
t
S S
{ [
dV + vv ndS = pI + μ (gradv) + gradv ndS + fbdV ,
T
]} V
(7.20)
( c v T )
t
dV + c v Tv ndS = kgradT ndS + hdV +
V S S V
(7.21)
μ[(gradv) ]
T
+ gradv :gradvdV .
V
• Differential form:
divv = 0 , (7.22)
( v)
t {[
+ div( vv) = gradp + div μ (gradv) + gradv + fb ,
T
]} (7.23)
(c v T )
+ div (cv Tv) = div( kgradT ) + h +
t (7.24)
[
μ (gradv) + gradv :gradv .
T
]
• Index notation:
v j
=0, (7.25)
x j
v i (
v iv j ) p v
μ
i +
v j
t
+
x j
= +
x i x j + fb,i , (7.26)
x j xi
(c v T )
+
( )
cv Tv j
=
T v
+ h + μ i +
v j v j
x j k
x j x j . (7.27)
t x j x i x i
89
7. MATHEMATICAL MODELS
One can see that the resulting system (in any of the forms presented above) consist of 5
equations with 5 unknown functions of spatial coordinates and time vi(x,t), p(x,t) and T(x,t).
It can also be noticed that for the incompressible fluids, the continuity (7.22) and momentum
(7.23) equations, which are also called the equations of hydrodynamics, are decoupled from
the energy equation (7.24).
Finally, if μ, k and cv have constant values, then the equations (7.22) to (7.24) can be written
in the following form:
divv = 0 Continuity equation (7.28)
Dv
= gradp + μ 2 v + fb Navier-Stokes equation (7.29)
Dt
DT 2
c v = k T + h + Heat equation (7.30)
Dt
where:
[ ] μ
[ ]
2
= μ (gradv) + gradv :gradv = (gradv)T + gradv = 2μD:D,
T
(7.31)
2
is (always positive) dissipation function, and squaring of the rate of deformation tensor
should be understood as the double dot product.
90
7. MATHEMATICAL MODELS
Hyperbolic problems
C zone of influence D
of point P
P charasteristics
2 3
region of
dependance
of point P
t A B
1
x
Figure 7.1 Region of dependence and zone of influence of point P for a hyperbolic problem
91
7. MATHEMATICAL MODELS
Parabolic problem
Parabolic PDEs have solutions that are smooth in space (may possess singularities at
corners). An example of a parabolic equation is the heat conduction equation. Information
travels at an infinite speed in a parabolic system. For example, if heat source is applied at one
end of a rod, the temperature rises instantaneously along the entire rod according to the heat
conduction equation. Far from the source, the temperature increase may be very small. For a
parabolic problem two characteristics are identical and the region of dependence of point P
reduces to the line segment AB (Fig. 7.2). The zone of influence of point P, on the other hand,
extends to the whole region at one side of the characteristics AB.
2 3
zone of influence
of point P
A
B
P 1
t region of
dependance
of point P
x
Figure 7.2 Region of dependence and zone of influence of point P for a parabolic problem
Elliptic problem
Examples of elliptic PDEs are the Laplace equation. Here the solutions are smooth (even if
boundary conditions are ‘rough’). Also, boundary data at any point affect the entire solution.
The effect of small irregularities in boundary data tend to be confined to the boundary: this is
known as St Venant’s principle. Elliptic problems have no real characteristics and the
solution at P depends on all points in the solution domain , and inversely, the values of a
dependent variable at any other point in is influenced by the value at P. In the other words,
the region of dependence is identical with the zone of influence and both of them coincide
with the whole solution domain (Fig. 7.3).
92
7. MATHEMATICAL MODELS
P
region of dependance
y zone of influence
Figure 7.3 Region of dependence and zone of influence of point P for an elliptic problem
As mentioned above, the classification of PDEs depends on whether lines or surfaces exist
across which the derivatives of the solution are discontinuous. This is equivalent to
examining whether lines exist along which PDEs can be reduced to ordinary differential
equations. Consider a second-order general linear partial differential equation in the domain x
and y:
2u 2u 2u
A( x,y ) + 2B( x,y ) + C ( x,y ) +
x 2 x y y 2
(7.32)
u u
a( x,y ) + b( x, y ) + c ( x, y )u = f (x,y).
x y
D < 0 – Elliptic problem with no real characteristic and with smooth solution.
93
7. MATHEMATICAL MODELS
2. Neumann boundary conditions where the value of a dependent variable gradient at the
boundary is specified (e.g. prescribed traction or given heat flux)
grad (x B ) = f 2 , grad p(x B ) = f 2 p, grad T (x B ) = f 2T , x B N , (7.35)
where D and N are parts of the boundary with prescribed either Dirichlet or
Neumann boundary conditions, respectively. In the case of the momentum equation for
solids, so-called mixed boundary conditions are applicable. Here, for example, the Dirichlet
boundary condition (given displacement) is applied in the direction normal to the boundary
surface, and the Neumann boundary condition (prescribed traction) is applied in the direction
tangential to the boundary surface.
In many practical situations the problem exhibits some kind of symmetry in which case only
a part of the body may be taken as the solution domain and the symmetry boundary
conditions are applied at the interface
t ( x B ) p ( x B ) T ( x B )
n (x B ) = 0, = 0, = 0, = 0, x B s , (7.36)
n n n
where n is direction normal to the symmetry plane s , and n and t are displacement or
velocity vector components in the direction normal and tangential to the symmetry plane,
respectively.
The unsteady momentum equation for solids is hyperbolic or, more appropriately, it is
hyperbolic in time, and to start the calculations displacements and displacement rates at time
t = to have to be given at all points of the solution domain
u (x,t0 )
u( x,t0 ) = u0 (x ), = u̇0 (x ), x . (7.37)
t
The unsteady momentum equation for fluids and the unsteady energy equation are parabolic
or, more appropriately, they are parabolic in time, and require velocity and temperature to be
known at all points of the solution domain at the initial instant of time t = to
94
7. MATHEMATICAL MODELS
7.5 Summary
It is important to note that equations governing momentum and energy transport in a
continuous media (see Sections 7.2 and 7.3 for thermo-elastic solids and Newtonian fluids,
respectively), can all be written in the form of a following generic transport equation:
t
B dV + B v ndS = grad ndS + qS ndS + qV dV , (7.39)
V S S S V
while the continuity equation is combined with momentum equation to obtain an equation for
pressure (in fluids usually). Here, stands for the transported property, i.e. displacement u,
velocity v or temperature T. The meaning of the properties B and are given in Table 7.1.
The term q S contains parts of the heat flux vector or the stress tensor, which are not
included in grad , while q V contains the volumetric source terms.
Table 7.11: The meaning of various terms in the generic transport equation (7.39)
B q V q S
T cT k :gradv + h 0
u
μ(gradu) + [divu ( 3 + 2μ ) (T T0 )]I
T
u μ fb
t
pI + μ (gradv)
T
v v μ fb
In case of thermo-elastic solid, the continuity equation does not have to considered (it serves
for determining density which is constant). In most practical applications it can be assumed
that not only the strains are small but also velocities, thus the (non-linear) convection term
vv , as well as the mechanical energy dissipation term : gradv can be neglected.
95
8. FINITE VOLUME DISCRETISATION
Chapter 8
satisfying eqs. (7.19) and (7.39) [or (7.19) and (7.7), (7.8), (7.20), (7.21)] in a closed form,
except in very few cases. Here, stands for displacement u, velocity v and temperature T.
Therefore an approximate solution has to be sought (Fig. 8.1). If one assumes a certain variation
of dependent variables in time and space, the governing equations may be integrated and if it
happens that the assumed profiles are exact, these equations will be satisfied identically,
together with prescribed initial and boundary conditions. However, in general this is not the
case and the solution of the governing equations (7.19) and (7.39) amounts to guessing the exact
solution. It is not difficult to imagine that this is not an easy task. In order to make these profile
assumptions feasible the solution domain (time and space) is subdivided into a finite number of
time steps and control volumes, such that even a very simple profile assumption (e.g. linear)
will produce plausible results. Then a procedure (numerical method) is devised of utilising this
to obtain the unknown functions (8.1).
All numerical methods consist of transforming the governing equations into a system of (in
general non-linear) algebraic equations, with subsets of these approximating each conservation
equation. In order to achieve this, time, space and equations have to be discretised. The time
discretisation understands a subdivision of a given time interval into a number of smaller
subintervals or time steps. The space discretisation consists of defining a numerical mesh
consisting of a finite number of computational points, thus replacing the continuous functions
(8.1) by their values at those points. Replacement of individual terms in the governing equations
by algebraic expressions connecting nodal values on a numerical mesh is called equation
dicretisation.
It is intuitively expected that the accuracy of the numerical approximation depends directly on
the number of computational points, that is the better the discretised time and space represent
96
8. FINITE VOLUME DISCRETISATION
the continuum, the better the numerical approximation. In other words, the error of a numerical
simulation has to tend to zero when the number of computational points tends to infinity
(discretised time and space approach the actual time and space), and if that is the case the
numerical discretisation is said to be consistent. The speed of this variation is characterised by
the order of the numerical discretisation.
CLOSED-FORM NUMERICAL
SOLUTION SOLUTION
The structure of the algebraic equations resulting from the discretisation procedure leads to two
classes of methods, the explicit and the implicit methods. In explicit methods the unknown
variable can be explicitly expressed in terms of known values. However, this advantage is
seriously offset by the fact that stability of the method imposes severe restrictions on the
maximum admissible time step. While this might not be a limitation for physically unsteady
problems, where the small time step might be needed for the required time accuracy, it leads to
a large number of time steps in order to reach the steady-state solution corresponding to a
97
8. FINITE VOLUME DISCRETISATION
From the practical application point of view a very important property of a numerical method is
its efficiency i.e. computer memory and CPU time required to obtain a solution†. Although this
is influenced by the discretisation procedure, as well as by the physical properties of the system,
the choice of the solution algorithm for the system of algebraic equations resulting from the
discretisation is the most important step for the economy of a numerical method.
Apart from the above numerical aspects, one has to consider physical implications of the
discretisation scheme, like conservation of physically conserved quantities or boundedness of
the solution.
Some aspects of the Finite Volume (FV) method will be discussed in the next Section, where
1D energy and momentum equations are taken as model equations and the decisions regarding
the discretisation practice will be made. Based on those decisions, in Section 8.2 the full
derivation of the discretised equations for 2D situations will be conducted. Finally, FV
discretisation of 3D problems in domains of arbitrary shape will be outlined in Section 8.3.
† For example, Cramer's rule is a perfect mathematical model for the solution of the system of linear algebraic
equations. However, in order to solve a system of 30 equations, ~1033 arithmetic operations have to be performed,
which for the fastest present day computer exceeds by far the present estimate of the life of the Universe. Not to
mention that the solution would be useless due to an excessive round-off error propagation.
98
8. FINITE VOLUME DISCRETISATION
computational node
boundary node
CV
in which case the problem reduces to the solution of energy equation (7.8) [or (7.21)] reduced to
the following form
T
k x n x dS + hdV = 0. (8.3)
S V
Flux term Source term
where, in general, k = k(x), h = h(x) are known functions, with boundary conditions
T (0) = T0, T ( L) = TL . (8.4)
99
8. FINITE VOLUME DISCRETISATION
T0 T(x) TL
x
L
Figure 8.3 Heat conduction in a rod (or channel)
Space discretisation
For simplicity, the solution domain is subdivided into N equal control volumes (Fig. 8.4) with
L
x i = = x (i = 1,2,...,N ) , (8.5)
N
xi
Sw Se
W P E
1 2 i-1 i i+1 N-1 N
xi
Qe Qw HP
The integrals in (8.7) are calculated by employing the Mid-Point rule and the following
approximation is obtained
T T
ke Se kw Sw + hPV = 0 . (8.8)
x e x w
100
8. FINITE VOLUME DISCRETISATION
where the mid-point values of conductivity and heat source are, simply
x x
ke = k xi +
, kw = k xi , hP = h (x i ) . (8.9)
2 2
a) b)
T
T
w e w e
i-1 i i+1 i-1 i i+1
x x
Figure 8.5 Temperature profile approximation: a) zero order polynomial, b) first order polynomial.
Discretised equations
By employing (8.10) eq. (8.8) becomes
ke k
x
( Ti +1 Ti )S w (Ti Ti 1 )S + hPVi = 0.
x
(8.11)
Qi = Qe Qi–1 = Qw Hi = HP
bi = H i = hPVi
101
8. FINITE VOLUME DISCRETISATION
By writing such an equation for all N control volumes a system of N algebraic equations with N
unknowns (T1, T2,..., TN) is obtained
a11T1 + a12T2 = b1
a21T1 + a22T2 + a23T3 = b2
a32T2 + a33T3 + a34 T4 = b3
......................................... .......
ai,i 1Ti 1 + ai ,iTi + ai,i +1Ti +1 (8.14)
= bi
......................................... .......
aN 1,N 2TN 2 + aN 1,N 1TN 1 + aN 1,N TN = bN 1
aN ,N 1TN 1 + aN ,N TN = bN
If conductivity k and heat source h (i.e. the coefficients) are not functions of T, eqs. (8.14) are
linear. If, however, either k or h is a function of T, eqs. (8.14) are non-linear.
The fact that for the calculation of temperature at any point three grid points are involved, which
renders this discretisation scheme implicit, is illustrated in Fig. 8.6, where the so-called
computational molecule is presented. Since maximum three non-zero coefficients are present in
anyone equation, system (8.14) is said to be sparsely populated or, simply, sparse, or to be more
precise three-diagonal.
1 2 i-1 i i+1 N
Figure 8.6 Computational molecule for steady state heat conduction equation.
Before we proceed to describe some methods for the solution of system of equations (8.13), and
to discuss some important issues linked to the steady heat conduction, namely conservativness
and boundedness, it is useful to write the equation of heat balance for a control volume (8.13) in
the following so-called compass notation form
aP TP = aETE + aW TW + bP , (8.15)
where
S S
aE = ke , aW = kw ,
x x
aP = aE + aW , (8.16)
bP = HP = hPVP .
102
8. FINITE VOLUME DISCRETISATION
Let us, for clarity, write system (8.14) by employing the following notation
wi Ti 1 + di Ti + ei Ti +1 = bi (i = 1,2,...,N ) . (8.17)
TDMA uses the Gaussian elimination to put system (8.17) into an upper triangular form by
computing the new diagonal coefficients di and the new right hand side vector components bi as
w w
di = di i ei 1, bi = bi i bi 1 (i = 2,3,...,N ). (8.18)
di 1 di 1
Note that in the expression (8.18) the equals sign means "is replaced by". The values of
unknown temperatures are then calculated from the back substitution
b b eT
TN = N , Ti = i i i +1 (i = N 1,N 2,...,1). (8.19)
dN di
It can be seen that the TDMA requires both computer storage and time proportional to N, rather
than to N2 and N3, respectively, as it is generally the case with the Gaussian elimination. This
feature makes the TDMA a very powerful algorithm for solving large systems of linear
equations with a three-diagonal matrix.
Ti
(k+1)
=
1
(
di i
(k+1) (k)
b wi Ti 1 ei Ti +1 ) (i = 1,2,...,N, k = 0,1,2,...) . (8.20)
where k is the iteration counter, and w1 and eN are taken as zero. The method is extremely
simple, but only converges under certain conditions. It is interesting to note that a sufficient
condition for the convergence is the diagonal dominance of the system coefficient matrix
( aP aE + aW ) which will be shown shortly to be one of the requirements for the boundedness
of the solution.
103
8. FINITE VOLUME DISCRETISATION
8.1.3 Conservativness
If the heat flux at each cell face is represented by the same expression for the two adjacent
control volumes, then the discretisation (numerical method) is conservative, because the fluxes
at the internal cell faces cancel each other and the overall heat balance is satisfied exactly
(except for round-off error) irrespective of the number of CVs (computational points) and not
only in the limit of the large number of computational points [see eq. (8.11)]:
– Qo+ Q1 + H1 = 0
– Q1+ Q2 + H2 = 0
– Q2+ Q3 + H3 = 0
...... .... (8.21)
–Qi–1+ Qi + Hi = 0
...... ....
– QN–1+ QN + HN = 0
---------------------------------------------------------
N
– Qo + QN + Hi = 0
i =1
Although inherent in the FVM, the conservativness can be lost due to careless discretisation. If,
for example, in calculating the heat flux through the cell face between the cells P and E the cell-
centre instead the cell-face conductivity is used:
S
flux calculated from the cell P Qi = kP (Ti +1 Ti )
x
S
the same flux calculated from the cell E Qi 1 = kE (Ti +1 Ti )
x
will not cancel each other, since in general kP kE, and the conservativness is lost.
8.1.4 Boundedness
Another important property of a discretisation scheme is its boundedness, which is closely
linked to the physical plausibility of the numerical solution. For example, in the absence of heat
sources (h = 0) no temperature should lie outside the range established by the boundary
temperatures, i.e. for To TL:
T0 Ti TL (i = 1,2,...,N ). (8.22)
Also, an increase in a boundary temperature (or temperature of any other point, with other
conditions remaining the same) must cause an increase (not decrease) of the adjacent point
temperature.
The discretisation formula (8.15) satisfies these requirements because all the coefficients have
the same sign (in this case positive). In addition to that, the central coefficient aP is equal to the
104
8. FINITE VOLUME DISCRETISATION
sum of the neighbour coefficients aE and aW [eq. (8.16)] which means that if boundary
temperatures were increased by a constant, temperatures at all computational points would
increase by the same value. Also, equal boundary temperatures (To = TL) would produce a
uniform temperature field (Ti = To, i = 1,2,...,N).
There are, however, numerous formulations that may lead to coefficients of different signs. For,
example, if the source term HP is a function of temperature, then in order to enable the solution
of the system (8.14), it is often linearised, i.e. expressed in the following form
HP = C0 + C1TP . (8.23)
where Co and C1 are, in general, functions of temperature. In that case eqs. (8.15) and (8.16)
become
aP TP = aETE + aW TW + C0 ,
S S
aE = ke , aW = kw , (8.24)
x x
aP = aE + aW C1 .
If C1 is negative, aP remains positive, i.e. of the same sign as aE and aW. However, if C1 is
positive, aP may become negative resulting in an unbounded solution (and in this case unstable
numerical algorithm). A positive C1 (in all points) clearly indicates a physically unstable
situation, because an increase in TP leads to the increase in the heat source, which in turn leads
to a larger TP, and so on.
Unbounded solutions are frequently encountered in cases of convection dominated flows. This
is illustrated in the next sub-section. Other situations of a potentially unstable and/or unbounded
numerical solution will be discussed in the Section on stability.
T(x), v(x)
. .
m m
x
L
Figure 8.7 Heat convection-conduction in channel
In this case the energy equation (7.21) reduces to the following form:
105
8. FINITE VOLUME DISCRETISATION
T
cTvn x dS = k x n x dS . (8.25)
S S
Convective term Flux term
vn x dS = 0 . (8.26)
S
Obviously, in case of a channel with constant cross sectional area, v=const. If the uniform space
discretisation (8.5) is used, then eqs. (8.25) and (8.26) become:
T T
cTvdS cTvdS = k x dS k x dS . (8.27)
Se Sw Se Sw
If the peace-wise linear variation of temperature between the grid points (Fig. 8.5b) is
employed, eq. (8.27) can be written as:
TP + TE T + TP k k
(cv )e 2
S ( cv ) W
w 2
S = e (TE TP )S w (TP TW )S
x x . (8.29)
Te Tw
[ ]
aE + aW + c ( vS )e ( vS ) w = aE + aW
ṁe + ṁw = 0
Since the last term in the aP coefficient is equal to zero due to the continuity equation (8.28),
the property aP = aE + aW is preserved. However, the coefficients aE and aW are not
unconditionally positive, and whenever
S c
k v S < 0,
x 2
106
8. FINITE VOLUME DISCRETISATION
i.e.
v x
Pe = > 2, (8.32)
D
where Pe is the Peclet number and D = k / (c) is the thermal diffusivity, aE or aW become
negative and an unbounded solution may occur. This may also cause divergence of the Gauss-
Siedel solver (in the limit of zero conductivity the aP coefficients become zero!). For example,
3
consider air (
=1 kg/m , c=1000 J/kgK, k=0.025 W/mK) flowing with constant speed v=0.25
m/s through a channel with a cross section S=1 m2. If x=0.005 m, then Pe=50 and
aE = 120, aW = 130, aP = 10 .
and
a) for TE = 0 and TW = 0 TP = 1300
.
b) for TE = 100 and TW = 0 TP = 1200
3
which is unbounded and, obviously, unrealistic. Situation with flow of water (
=1000 kg/m ,
c=4200 J/kgK, k=0.5 W/mK) will be much worse, even for very small velocities.
A well known solution for the high Peclet number flows is to use the zero order polynomial
approximation (Fig. 8.5a) for the discretisation of the convection term in the so-called upwind
manner, e.g.
TP if flow is from P to E i.e. ṁe 0
Te = , (8.33)
TE if flow is from E to P i.e. ṁe < 0
and similarly for Tw. To avoid the conditional statement in (8.33), the use of min and max
functions has proven useful, e.g.
ṁeTe = max(ṁe ,0)TP + min(ṁe ,0)TE , (8.34)
or by employing (8.35)
107
8. FINITE VOLUME DISCRETISATION
S S
ke + ke + c [ṁe min(ṁe ,0)] + c[ ṁw min(ṁw ,0)]
TP =
x x
(8.37)
S S
ke x c min(ṁe ,0) TE + k w x c min(ṁw ,0) TW ,
Thus, by using the lower order polynomial approximation of the cell-face temperature in the
convection term, a bounded discretisation ( aE > 0, aW > 0, aP = aE + aW ) is obtained.
However, this is achieved by sacrificing accuracy of the higher order approximation.
To recover (some of) the lost accuracy and at the same time maintain the boundedness, a
number of schemes has been proposed, all of which amount to a blend of one lower order and
one higher order scheme, e.g.
Te = Te
LO
( )
+ e TeHO TeLO , (8.39)
where 0 e 1 is the blending factor, which can either be fixed or the solution dependent. One
such scheme is the hybrid scheme, which takes:
TeHO if Pee 2
Te = . (8.40)
TeLO if Pee > 2
Another scheme, which is preferred here, uses a constant blending factor to blend the first order
polynomial approximation with a small amount of the zero order one, e.g.
T + TE
ṁeTe = [max( ṁe ,0)TP + min( ṁe ,0)TE ] + ṁe P [max( ṁe ,0)TP + min( ṁe ,0)TE ] =
2
and similarly for the west cell face. It is important to mention that the blending scheme is
implemented according to the ‘deferred-correction’ practice, i.e. convective flux is split into an
implicit part containing the contribution from the LO only (part of the coefficient), and an
108
8. FINITE VOLUME DISCRETISATION
explicit part consisting of the difference between HO and LO contributions (part of the right
hand side vector), leading to the equation:
aP TP = aETE + aW TW + b , (8.42)
with
S
aE = ke c min(ṁe ,0),
x
S
aW = k w c min(ṁw ,0),
x (8.43)
aP = aE + aW + c (ṁe + ṁw ) = aE + aW ,
ṁ ṁ
b = c
min(ṁe ,0) e (TE TP ) + c
min(ṁw,0) w (TW TP )
2
2
One can see that the coefficients are unconditionally positive and that aP = aE + aW , which
enables the use of iterative solvers like Gauss-Siedel method, because the troublesome HO
contribution is relegated to the right hand side and can be controlled by choosing an appropriate
. However, this means that the solution procedure has to be iterative even if equation (8.42) is
linear.
8.1.5 Stability
A stable discretisation scheme is one for which errors from any source (round-off, truncation,
mistakes) are not permitted to grow in the sequence of a numerical procedure. Various
mathematical methods for stability analysis are available, the most popular being the Von
Neumann method based on a Fourier analysis. However, we shall limit ourselves to simple
physical considerations.
and the governing equation (7.8) [or (7.21)] will get the form
T
t k x n x dS hdV .
cTdV = + (8.45)
V S V
Transient term (R) Flux term (Q) Source term (H)
109
8. FINITE VOLUME DISCRETISATION
Let us now discretise eq. (8.45) and examine the stability of the resulting equation.
i.e. we shall use the mid point rule to calculate integrals and assume a piece-wise linear spatial
variation of temperature (Fig. 8.4b) for the evaluation of gradients.
As for the temporal variation of the terms on the right side of eq. (8.45), the heat flux and the
source term will be assumed to be a weighted average of their values at the two successive time
levels, i.e. they will be evaluated at the time t = tm + (1 ) tm1 :
Qi = Qim + (1 )Qim1 = QPm + (1 )QPm1,
(i = 1,2,...,N; m = 1,2,...) (8.49)
Hi = H im + (1 )Him1 = H Pm + (1 ) HPm1,
Transient term
By employing the mid point rule the integral in the transient term is approximated as
For temporal variation we shall assume a two-time-level linear variation of temperature with
time and obtain the following approximation:
Ri =
t
V
cTdV =
V
t
( [
cT )i (cT )i
m m1
= ]
(8.51)
t [ ]
V
( cT ) m
P
( )
cT m1
P
(i = 1,2,...,N; m = 1,2,...).
110
8. FINITE VOLUME DISCRETISATION
Discretised equations
Thus by using eqs. (8.49) and (8.51) and results from Section 8.1.2, the discretised form of the
eq. (8.45) can be written in the same form as eq. (8.13)
m m m m m m
ai,i 1Ti 1 + ai ,iTi + ai,i +1Ti +1 = bi (i = 1,2,...,N; m = 1,2,...). (8.52)
where
m = k m S S
ai,i +1 e , aim,i 1 = kwm ,
x x
V m
bP = H Pm + (1 )H Pm1
Different values of temporal weighting factor lead to different time discretisation schemes.
The best-known are the explicit scheme ( = 0), the Crank-Nicolson scheme ( = 0.5) and the
fully implicit scheme ( = 1), whose interpretation is given in Fig. 8.8. The only apparent
difference between the explicit and fully implicit schemes is that the former assumes that the old
m1 m1 m m
value QP and HP and the latter that the new value QP and HP prevail over the whole of
the time step. However, these two schemes have remarkably different properties. The Crank-
Nicolson scheme, on the other hand, assumes a linear variation of QP and HP with time and it
can be expected to be more accurate than the previous two.
111
8. FINITE VOLUME DISCRETISATION
explicit
QPm-1= Qm-1
i
Crank-Nicolson
Q, H
Qm = Q m
P i
fully implicit
m-1 m
t
Figure 8.8 QP and HP variation for explicit, Crank-Nicolson and fully implicit schemes.
Explicit scheme
For = 0 eqs. (8.54) and (8.55) can be written as
m m m1 m1 m1 m1 m1 m1 m1
aP TP = aE TE + aW TW + aP TP + H P . (8.56)
and
m1 m1 S m1 m1 S
aE = ke , aW = kw ,
x x
V
m1
aP =
t
( c ) m1 m1 m1
P aE aW , (8.57)
V
aPm =
t
( c ) m
P
One can immediately see that temperature at any computational point at the time tm is
dependent on temperatures from the previous time tm–1 only, which is illustrated by the
computational molecule in Fig. 8.9. This means that the coefficient matrix of the resulting
system of linear algebraic equations (obtained by linearisation of the non-linear equations, if the
original equations are non-linear) is diagonal, leading to a trivial matrix inversion and hence to a
solution with a minimum number of algebraic operations per time step and minimum computer
m
memory requirements (temperature TP can be directly calculated from eq. (8.56) and there is
no need to store coefficient matrix and right hand side vector). Moreover, if a steady state is
sought, one can still use the unsteady formulation (8.56) and march in time till the steady state is
reached. Any other scheme with 0 [as well as the steady state formulation (8.13)] would be
m m m
implicit, i.e. more than one temperature (e.g. Ti 1, Ti , Ti +1 ) is unknown at the same time and
the matrix to be inverted is not diagonal, hence requiring much more algebraic operations per
time step and much more computer memory.
112
8. FINITE VOLUME DISCRETISATION
m+1
t
m-1
1 2 i-1 i i+1 N
This is the well-known stability criterion for the explicit scheme for parabolic differential
equations, which in many cases imposes severe restrictions on the maximum admissible time
step. Namely, the reduction of x required by the spatial accuracy causes maximum t to
decrease proportionally to x2, which makes the explicit scheme prohibitively expensive for
many practical applications. Since the stability of the explicit scheme depends on the parameter
r, it is classified as conditionally stable scheme.
To illustrate this, let us assume constant physical properties of a material and a situation without
sources and write eq. (8.56) in the following form
m
(
m1 m1
TP = r TE + TW )
+ (1 2r)TPm1 . (8.59)
Suppose that at the time tm–1: TEm1 = TWm1 = 100 K and TPm1 = 0 K (Fig. 8.10a). For r =1
formula (8.59) gives TPm = 200 K , which lies outside the bounds and is physically impossible.
m1
If temperature TP is increased to, for example, 200 K, one would expect to get an even
m m
higher temperature TP . However, eq. (8.59) produces TP = 0 K , which is again implausible. It
m1 m1
is interesting to note that for a uniform temperature field TE = TW = C eq. (8.59) gives the
m
correct answer TP = C .
113
8. FINITE VOLUME DISCRETISATION
a) "Heater' b) 'Refrigerator'
200 0
t m t m
100 0 100 100 200 100
m-1 m-1
W P E W P E
x x
Figure 8.10 Physically unrealistic results produced by explicit time differencing scheme for r = 1.
Crank-Nicolson scheme
It can be shown that the Crank-Nicolson scheme (obtained for = 0.5) is unconditionally
stable, but unbounded (physically unrealistic) solutions are possible. Namely, the coefficient
m1
aP (for uniform conductivity ke = kw = k, and V = Sx ) becomes [eq. (8.55)]
V cx k
m1
aP =
t
( c ) P am1
m1
E
m1
aW = S
t
x
(8.60)
m1
Thus, whenever, for a given space discretisation, the time step is not sufficiently small, aP
becomes negative and the solution may become unbounded. The stability of the scheme,
however, ensures that these unphysical oscillations eventually die out. It is important to note
that at any stage of the calculation 6 grid nodes are involved (3 at the current time tm and 3 at
the time tm–1), as illustrated by the computational molecule in Fig. 8.11. This in general requires
storing or recalculating of coefficients from the previous time level.
m+1
t
m
m-1
1 2 i-1 i i+1 N
114
8. FINITE VOLUME DISCRETISATION
aP TP = aETE + aW TW + bP , (8.61)
and
S S
aE = ke , aW = kw ,
x x
V
aP = aE + aW + (c) P , (8.62)
t
V
bP = (c )m1 TPm1 + HP
t P
where the superscript m denoting the new time interval has been omitted to facilitate the
comparison of eqs. (8.61) and (8.62) with the steady state equations (8.15) and (8.16).
Now it is easy to note one important practical feature of the fully implicit scheme. Namely, for
an infinite time step ( t ) the fully implicit equations reduce to their steady state
counterparts. At the same time the fully implicit scheme involves only one previous time level
point, as can be seen from the computational molecule in Fig. 8.12, hence involving
significantly less computations and/or storage than the Crank-Nicolson scheme (not as evident
in this 1D case, as in 2D or 3D situations).
m+1
t
m-1
1 2 i-1 i i+1 N
Figure 8.12 Computational molecule for fully implicit scheme (o – unknown, • – known).
Additional considerations
Figure 8.13a shows the propagation of the solution outward from the initial conditions line in
the case of an explicit scheme. It can be observed that the value of the dependent variable
(temperature) at point P, three time steps away from the initial data line, is calculated without
any awareness of the boundary conditions. However, theoretical considerations in Section 7.4
indicate that the solution at point P should be influenced by the boundary conditions, since the
parabolic energy equation has the characteristics t = const.
115
8. FINITE VOLUME DISCRETISATION
Thus, for a large t (or a small x, i.e. for r > 1/2) the solution is allowed to 'wander' (Fig. 8.13a)
and inevitably leads to unbounded physically unrealistic, unstable solution. For a smaller t (or
a larger x, i.e. for r < 1/2) the boundary conditions are able 'to catch-up with' the solution (Fig.
8.13b) and hold it inside the physical bounds, resulting in a stable procedure. On the other hand,
the implicit schemes involve boundary values in the calculation of temperature at all
computational points at every instant of time tm (m = 1,2,...) and therefore produce a stable
solution.
a)
P
m=3
t
m=2
m=1
m=0
i-1 i i+1
initial data
x line
b)
P
m=3
t
m=2
m=1
m=0
i-1 i i+1
initial data
x line
116
8. FINITE VOLUME DISCRETISATION
p
x
L
Figure 8.14 1D bar loaded by compressive stress pulse
When a force is applied, its action is not transmitted instantly through all parts of the body. The
deformations produced by the force, and so the information about the instantaneous stress
distribution, travel through the body in form of stress waves. In the absence of thermal stresses,
Hooke's law [eq. (6.33) or (7.3)] takes the form
xx = E xx, (8.63)
and when body forces are omitted, the governing equation of motion (7.7) reduces to
u u
t dV = E n x dS ,
t x (8.64)
V S
Inertial Surface
force ( Fin ) force ( Fx )
where u =u(x,t). The right hand side of eq. (8.64) is discretised in the same way as the first term
on the right hand side of heat conduction equation (8.45). To discretise the inertial part of eq.
(8.64), similar procedure to the one for the transient term of eq. (8.45) is used; i.e. by employing
the mid point rule the volume integral is approximated as
m
u u u
m
u
m
t
dV =
dV
t
=
t
i
V =
t V (i = 1,2,....,N; m = 1,2,....) (8.65)
P
V V i
and a two-time-level linear variation of the deformation velocity with time is adopted giving
117
8. FINITE VOLUME DISCRETISATION
u V u
m u
m1
t t
Fini = dV = =
t
t i t i
V
V
t
[( m
2 ui ui
m1
) (
ui
m1 m2
ui = )]
(8.66)
V
t 2 [ uim 2uim1 + uim2 = ]
V m m1 m2
2 [uP 2uP + uP ] (i = 1,2,....,N; m = 1,2,....).
t
The surface force is evaluated at time t = tm + (1 ) tm1 as:
m m 1 m m 1.
Fxi = F xi + (1 )Fxi = Fx P + (1 )FxP (8.67)
Combining the equations (8.66) and (8.67), discretised counterpart of eq.(8.64) is obtained and is
of the same form as eq. (8.52) where temperature T is replaced by displacement u:
m m m m m m
ai,i 1ui 1 + ai,i ui + ai ,i +1ui +1 = bi (i = 1,2,....,N; m = 1,2,....) , (8.68)
where
m m S m m S
ai,i +1 = Ee , ai,i 1 = E w ,
x x
m
V
ai,i = 2 + aim,i +1 + aim,i 1 ,
t
m1 m1 S m1 m S
ai,i +1 = (1 )E e , ai,i 1 = (1 )E w ,
x x
V (8.69)
m1
ai,i = 2 2 aim1 m1
,i +1 ai,i 1,
t
m1 m1 + a m1 um1 V um2
bi = aim1 m1
,i 1ui 1 + ai ,i ui i,i +1 i +1 i .
t 2
118
8. FINITE VOLUME DISCRETISATION
In case of the explicit discretisation scheme (=0), the stability criterion ( aPm1 > 0) leads to a
following condition on the time step size:
x E t
t < , or r = < 1. (8.72)
E / x
Here, E / defines the speed of the wave propagation along the bar made of a linear elastic
isotropic material (Ee = Ew = E).
The computational molecule for fully implicit scheme is shown in Fig. 8.15. It involves only one
point from each previous time level (m–1, m–2) point, and therefore requires much less
computations and/or storage than the Crank-Nicolson scheme.
m+1
t
m-1
m-2
1 2 i-1 i i+1 N
x
Figure 8.15 Computational molecule for fully implicit scheme (o – unknown, • – known).
8.1.6 Accuracy
In Sections 8.1.2, 8.14 and 8.1.5 both surface and volume integrals are calculated by employing
the Mid-Point rule. In Section 8.1.2 the spatial temperature variation was approximated by a
piecewise-linear function of the spatial coordinate x leading to the expressions (8.10) for the
temperature gradient. Similarly, in Section 8.1.5 the time derivative was approximated by a two-
time-level approximation, which for different values of led to the explicit, the Crank-Nicolson
and the fully implicit time differencing schemes. Finally, in Section 8.14 the cell-face values of
the temperature were obtained by blending first and zero order polynomial approximations.
Intergrals
Let us first establish the accuracy of the Mid-Point rule. If function f(x) is continuous and has at
least three continuous derivatives in [xw, xe], then it can be shown that
119
8. FINITE VOLUME DISCRETISATION
xe
xe x w n x 2 ''
f (x)dx =
n mi
f + ( xe x w ) 24 f ( ), ( xe ,x w ) , (8.73)
xw i =1
Exact value Mid-Point formula Error
where n is the number of subintervals, and fmi is the value of f(x) in the middle of the ith
subinterval. Since the error of the above formula is proportional to x2 [x = (xe – xw)/n], the
Mid-Point rule is second order accurate.
Cell-face gradients
In order to examine the accuracy of gradient approximations consider a continuous, sufficiently
times differentiable function f(x) and its Taylor series expansion (TSE) around the point xi
f 1 2 f 2 1 f
3
f (x) = f (xi ) + ( x x i ) + 2 ( x xi ) + 3 ( x x i ) + ...,
3
x i 2 x 6 x (8.74)
i i
Then the values of f(x) at points xi+1 = xi + x and xi–1 = xi – x can be written as
f x + 1 f x 2 + 1 f x 3 + ...
2 3
f (x i +1 ) = f (xi ) +
x
i 2 x 2
6 x 3
i i
(8.75)
f 1 2 f 1 3 f
f (x i 1) = f (xi ) x + 2
x 2 3
x 3 + ...
x i 2 x 6 x
i i
By using equations (8.75) different approximations to (f/x)i can be obtained. For example,
from the first equation (8.75) it follows that
f f (xi +1) f (x i ) 1 2 f f f
x = 2
x + ... = i +1 i + O(x ) (8.76)
i x 2 x x
i
Exact Taylor series Truncation TSA TE
value approximation error
f f (xi ) f (xi 1 ) 1 2 f f f i 1
x = + 2
x + ... = i + O(x ) (8.77)
i x 2 x x
i
Exact Taylor series Truncation TSA TE
value approximation error
or by subtracting the second from the first equation (8.75) one gets
f (xi +1) f (x i 1) 1 3 f 2
f
x
i
=
2x 6 x
f f
3
x + ... = i +1 i 1 + O x
2x
2
( ) (8.78)
i
Exact Taylor series Truncation TSA TE
value approximation error
120
8. FINITE VOLUME DISCRETISATION
Equations (8.76) and (8.77) correspond to the so-called forward and backward differencing
formulae, respectively, and are known under the common name one-sided approximations (or
one-sided differencing schemes). They are first order accurate, because their truncation error
(TE) is proportional to x. The last approximation [eq. (8.78)] corresponds to the so-called
central differencing scheme, and is second order accurate because its TE is proportional to x2.
It is interesting to note that, although formula (8.78) uses only two points, as the first-order
formulae (8.76) and (8.77), one gets second order accuracy. This is due to the fact that the
tangent at the point xi to a second order polynomial (parabola) fitted through points xi–1, xi, xi+1
is parallel to the secant through the end points xi–1 and xi+1 (Fig. 8.16). If, however, xi (xi–1 +
xi+1)/2, the second order accuracy is lost.
tangent
f
secant
parabola
2x (x)
x
Figure 8.16 Geometrical interpretation of the second order accuracy of the two-point formula.
This leads to the idea that formulae (8.76) or (8.77) can be regarded as the central differencing
formulae with respect to point xe = xi+1/2 = (xi + xi+1)/2 (values in brackets in Fig. 8.16) or xw =
xi–1/2 = (xi –1+ xi)/2 respectively, giving second order accuracy at these points.
121
8. FINITE VOLUME DISCRETISATION
In fact, integrals and dependent variable gradients could be obtained by involving any number
of adjacent points in space and time with the order of approximation increasing with the number
of points. However, there is always a trade-off between the order of accuracy and the number of
points simultaneously (implicitly) involved in the calculation, which determines the bandwidth
of the system of linear algebraic equations to be solved to obtain Ti or ui (i = 1,2,...,N). In the
present equation discretisation, in case of a steady state or implicit and Crank-Nicolson
schemes, in the calculation of dependent variable at any computational point three grid points
are involved (Figs. 8.6, 8.11, 8.12, 8.15) resulting in a three-diagonal system of linear algebraic
equations of the form (8.13).
Figure 8.17 Effects of dissipation and dispersion on the computation of sharp gradients.
(a) Exact solution.
st
(b) Numerical solution distorted primarily by dissipation errors (typical of the 1 order methods).
nd
(c) Numerical solution distorted primarily by dispersion errors (typical of the 2 order methods).
8.1.7 Consistency
Consistency deals with the extent to which the discretised equations approximate governing
conservation equations. A discrete counterpart of a conservation law is said to be consistent if
the difference between the two vanishes as the time and space discretisation become sufficiently
fine (i.e. if t 0, x 0).
122
8. FINITE VOLUME DISCRETISATION
8.1.8 Convergence
In the context of numerical continuum mechanics the term convergence is given various
meanings. In theoretical considerations convergence is usually related to the numerical method
on the whole which is said to be convergent if the solution of the discretised equations
approaches the exact solution of the original governing equations when t and x tend to zero,
i.e. when the number of computational points tends to infinity. It can be proved that for a well-
posed (linear initial value) problem and a consistent discretisation scheme, stability is the
necessary and sufficient condition for convergence.
In the present course the term convergence will be related to two iterative processes:
• (outer) iterations set up to solve the problem of non-linearity and/or inter-equation coupling,
and
• solver (inner) iterations performed during iterative solution of the system of linear algebraic
equations resulting from disretisation and subsequent linearisation and/or decoupling.
8.1.9 Summary
Let us now summarise the findings from the previous sections and make decisions about the
choices regarding the present FV method.
1. In order to achieve a conservative discretisation the diffusion fluxes have to be uniquely
associated with cell faces, and not cells.
2. The sufficient condition for a bounded (physically plausible) solution (in the absence of
sources) is that all the coefficients of the discretised equations are positive and that the
diagonal coefficient is equal to the sum of all the neighbouring (in both space and time)
coefficients.
3. Integrals are calculated by employing the Mid-Point rule, and the linear spatial variation of
dependent variables is assumed guaranteeing a second order accurate spatial discretisation.
4. The fully implicit temporal variation scheme is preferred on the grounds of its unconditional
stability, boundedness and better efficiency than, for example, the Crank-Nicolson scheme.
5. The truncation error of the adopted discretisation O(x2,t) vanishes proportionally to x2
(on a uniform mesh) and to t, which means that the present discretisation is consistent.
123
8. FINITE VOLUME DISCRETISATION
Thermo-elastic solid
In some practical situations the geometry and the (mechanical and thermal) loading
configuration allows reasonably accurate results to be obtained by reducing the problem to a
two-dimensional one. These situations are known as (Fig. 8.18):
z y
x
y
x
z
In all these cases it is sufficient to consider a two-dimensional domain in x–y plane. In case of
axi-symmetric problems, radial direction is represented by x axis and axial one by y axis, and
corresponding displacement components by u and v, respectively. The momentum equations
[second equation in (7.1)] take the form
u
t t
( )
dV = xx n x + xy n y dS + fbx dV < x
dV >,
V S V V
(8.79)
v
t ( )
dV = yx n x + yy n y dS + fby dV ,
t
V S V
with the stress tensor components
u u v u
124
8. FINITE VOLUME DISCRETISATION
u u v u
xy = yx = μ +
,
y x
where the terms enclosed within < > brackets are included only for axi-symmetric cases, and
those within [ ] brackets account for thermal effects.
Situations involving a long prismatic body whose geometry and loading conditions do not vary
significantly in the longitudinal z direction is referred to as a (generalised) plane strain problem
(Fig. 8.18a). In these cases the dependent variables (displacement components and temperature)
can be assumed to be functions of only x and y coordinates and time t
u = u(x, y,t), v = v(x,y,t), T = T (x,y,t). (8.83)
If a further assumption is made that the displacement w in z direction is zero at every cross
section
w = w(x, y,z, t) = 0, (8.84)
then the strain components
xz ,
yz and
zz vanish and equations (6.33) and (7.7) reduce to
equations (8.79) and (8.80).
From equations (6.33) follows that xz = yz = 0, but the normal stress in the z direction is not
zero and can be easily calculated from (6.33) from the condition
zz = 0
( )
zz = xx + yy [ ET ] (8.85)
In contrast to the plane strain condition, where the longitudinal dimension in z direction is large
compared to the dimensions in x and y directions, cases involving very small thickness and no
125
8. FINITE VOLUME DISCRETISATION
loading in z direction are known as a (generalised) plane stress problem (Fig. 8.18b). It is,
normally, assumed that
zz = xz = yz = 0 (8.86)
and the non-zero stress components xx, yy and xy are averaged over the thickness and
assumed to be independent of z. It can be shown that, after some algebraic manipulation, eqs.
(8.80) remain valid if and 3K are replaced by
E E E E
= ; 3K = . (8.87)
(1 + )(1 2 ) 1 2 1 2 1
It can be noted that the above formulae can formally be obtained by replacing 1 – 2 with 1 –
in respective expressions for and 3K in eqs. (6.16) and (6.21). However, although the strain
components xz and yz vanish, the component zz is given by expressions
1+
zz = ( + yy +
1 xx
)
1
( )
T = xx + yy + T ,
E
(8.88)
Axi-symmetric problem
Many engineering problems are concerned with solids of revolution which are symmetrical in
shape about axial axis z and are subjected to axially symmetrical loading (Fig. 8.18c). It is
convenient to express these problems in terms of cylindrical coordinates r,
and z, shown in
Fig. 8.18c. Due to axial symmetry
u = 0, r = z = 0 and r = z = 0, (8.89)
where u
is circumferential displacement component. The non-zero strain components are
ur u u 1 u u
rr = , zz = z , = r , rz = zr = r + z
,, (8.90)
r z r
2 z r
where ur and uz are displacements in r and z directions respectively. The non-zero stress
components are rr, zz,
and rz. Using eqs. (6.33) and (8.90) they can be expressed as
ur u u u
rr = 2μ + r + z + r
[ 3KT ],
r
r z r
uz u u u
zz = 2μ + r + z + r
[ 3KT ],
z
r z r
(8.91)
u u u u
= 2μ r + r + z + r
[ 3KT ],
r
r z r
u u
rz = zr = μ r + z
,
z r
126
8. FINITE VOLUME DISCRETISATION
fby dV ,
V
• Energy
T T
t cTdV = k
x
nx + k n y dS + hdV .
y
V S V
Eqs. (8.92) form a closed system of 3 equations with 3 unknown functions u, v, T of spatial
coordinates x, y and time t, which makes a mathematical model for 2D thermo-elastic stress
analysis. Note that if the heat source term h in energy equation does not contain the thermo-
elastic dissipation term, the energy equation is decoupled from momentum equations.
Newtonian fluid
There are also situations when fluid flow can be considered to be two-dimensional. These are
a) planar 2D flow (e.g. flow in a rectangular duct, Fig. 8.19 a), and
b) axi-symmetric 2D flow (e.g. flow in a pipe, Fig. 8. 19b).
127
8. FINITE VOLUME DISCRETISATION
a) b)
. . . .
m m1 m1+m 2
.
m2
Fig
ure 8.19 Two-dimensional flow in a) rectangular duct, and b) pipe.
In case of a 2D planar and axi-symmetric fluid flows, the mathematical model for an
incompressible Newtonian fluid [eqs. (7.19) to (7.21)] takes the following form:
• Continuity
( un x + vn y )dS = 0,
S
• Momentum
u
u v
t ( )
udV + u un x + vn y dS = 2μ n x + μ + n y dS
V S x S
y x
pn x dS + fb x dV ,
S V
(8.93)
u v v
t ( )
vdV + v un x + uny dS = μ + n x + 2μ n y dS
y x y
V S S
v
pn y dS + fb y dV + < 2μ y 2 dV >,
S V V
• Energy
T T
t ( )
cTdV + cT un x + vn y dS = k
x
nx + k n y dS +
y
V S S
u 2
v
u v
2 2 2
v
μ 2 +
x
y
y x
+ + + < 2 > dV + hdV .
y
V V
128
8. FINITE VOLUME DISCRETISATION
nn
Sn
n
Se
nw
w CV e
ne
y Sw
s
Ss ns
x
Equation discretisation
According to the findings from Section 8.1, for the temporal variation a fully implicit temporal
differencing scheme is used, i.e. all dependent variable values are related to the current time tm,
except for ones coming from the time derivative term, which are denoted by an appropriate
superscript (m–1 for the time tm–t, and m–2 for the time tm–2t). For the spatial variation a
129
8. FINITE VOLUME DISCRETISATION
central differencing scheme is used to express the cell-face values of gradients of dependent
variables in terms of their neighbouring nodal values. For the cell-face values of the dependent
variables featuring in the convection terms the SO scheme blended with a small amount of the
FO scheme will be implemented in a deferred-correction manner. All integrals are calculated by
employing the Mid-Point rule.
Different from the practice employed in Section 8.1, a non-uniform grid spacing will be used
here (Fig. 8.22), implying that the spatial interpolation factor of 0.5 used for calculation of the
cell-face values of dependent variables and physical properties from the two neighbouring nodal
values will be replaced by the following spatial interpolation factors:
Pe Pw Pn Ps
fe = , fw = , fn = , fs = . (8.95)
PE PW PN PS
NW NE
N
nw n ne
yn
y e = y w =yP
W w P e E
sw s se y s
S
SW SE
x w x e
Thermo-elastic solid
a) Momentum
a.1) u-momentum:
Now, the equilibrium of forces acting in the x direction on the CV shown in Fig. 8.23a [first
equation (8.79)] can be written in the following form
u
t t
( )
dV = xx n x + xy n y dS + fb x dV < x
dV >, (k = e,w,n,s). (8.96)
V k Sk V V
130
8. FINITE VOLUME DISCRETISATION
t dV =
t (2μ + ) x + y + < x > [ 3KT ]dS + μ y + x dS
V Se Sn
Fxin Fxe Fxw (8.97)
u v u
u v
(2μ + ) x + y + < x
> [ 3KT ]dS
μ y + x dS +
Sw Ss
Fxn Fxs
1 u u v
fbx dV < x (2μ + ) x + x + y [3KT ]dV > .
V V
Fbx < Fx >
a) b)
Fyn Qn
Fxn
n
Fby Fye
Fxw Qw
w e
F bx Fxe Qe
Fyw CV CV
s
Fxs
F ys Qs
Note that eq. (8.97) is exact, i.e. no approximation has been introduced so far! Before applying
the discretisation practices from Section 8.1 to approximate the individual terms in the above
equation, it should be noticed that apart from the approximation of the so-called normal
derivatives u/x at the east and west cell faces and u/y at the north and south cell faces, one
has to approximate the so-called cross derivatives v/y at the east and west cell faces and v/x
at the north and south cell faces, for which no counterpart exists in the energy equation. This is
done, for the north cell face for example, by interpolating between the gradients at points P and
N (Fig. 8.22) in the following way
v v v v vw v vw
x = (1 f ) x + fn = (1 f n ) e + fn e , (8.98)
n
n
P x N x P x N
where
ve = (1 f e )v P + fe vE ,
v w = (1 f w )v P + f wvW .
131
8. FINITE VOLUME DISCRETISATION
By employing the above and the discretisation practices described in Section 8.1, the forces
acting on the CV in the x direction can be approximated by:
Fxin
V
t 2
(
uP 2uPm1 + uPm2 , )
S v ue
Fxe (2μ + ) e (uE uP ) + Se
+ < Se > [3K
Te Se ],
x e y e xe
S v u
Fxw (2μ + ) w (uP uW ) + Sw
+ < w Sw > [ 3K
TwSw ],
x w y w xw
Sn v
Fxn μ (u uP )
y n N
+ μS n ,
x n
(8.99)
S v
Fxs μ s (uP uS ) + μS s ,
y s x s
Fbx fbxP V,
1 uP 1 1
< Fx (2μ + ) + ( ue u w ) + ( v n v s ) [ 3K
TP ]V > .
x P xP x P y P
where the superscript m is the time step counter and the compass labelling scheme shown in Fig.
8.22 is used. The material properties μe, e etc. are obtained by linear interpolation between the
two neighbouring computational points, e.g.:
μe = (1 fe )μP + fe μE .
132
8. FINITE VOLUME DISCRETISATION
Note that if thermal stresses are not present (T = 0), then expression in [] brackets is zero, if a
steady state is considered the expressions in {} brackets vanish and therefore are not calculated,
while terms in brackets <> are non-zero only in axi-symmetric cases.
a.2) v–momentum:
Similarly, from the equilibrium of forces acting in the y direction one gets:
S S
aE = μ e ; aN = (2μ + ) n ;
xe y n
S S
aW = μ w ; aS = (2μ + ) s ;
x w y s
u
u
b = μSe μSw +
y e y w
(8.102)
u
u
u us
Sn S s + < n Sn S >
x n x s xn xs s
and
aP v P aK v K = b (K = E,W ,N,S ), (8.103)
K
b) Energy
Discretisation of the energy equation is even simpler. For the CV in Fig. 8.23b eq. (8.81)
becomes
t
cTdV = (q x n x + q y n y )dS + hdV (k = e,w,n,s), (8.104)
V k Sk V
133
8. FINITE VOLUME DISCRETISATION
R
V
t[( cT )P (cT )m1
P ]
,
S S
Qe ke e (TE TP ), Qw k w w (TP TW ),
x e x w (8.106)
S S
Qn k n n (TN TP ), Qs k s s (TP TS ),
y n y s
H hPV,
and after introducing the following notation
S S
aE = ke e ; aN = k n n ;
x e y n
S S
aW = k w w ; aS = ks s ;
x w y s
(8.107)
V m1
b = hPV + ( cT )P ;
t
V
aP = aK + ( c ) P (K = E,W ,N,S ),
K t
the energy equation (8.105) gets the same form as momentum eqs. (8.101) and (8.103):
aP TP aK TK = b (K = E,W ,N,S ), (8.108)
K
To start the calculation, all dependent variables have to be set to their initial values: uo, vo and
To and in the case of the momentum equations the displacement vector components u1 and v1
at time t1 = tot are also required.
The expressions for the evaluation of forces and heat fluxes described above are valid for all
interior cell faces, which are shared by the two adjacent cells. Boundary cell faces, however, are
related to only one cell and require special treatment, which depends on the type of the
boundary conditions.
If, for example, an east cell face of a CV happens to be at the solution domain boundary and the
displacement is known (Fig. 8.24a), the uE and vE in the expressions (8.99) and in the
corresponding expressions for y direction are simply replaced by uB and vB. Similarly, if the
temperature TB at the east boundary is given (Fig. 8.24b), then TE in the expression (8.106) for
the heat flux Qe is replaced by TB.
Note that xe now stands for the distance between the points P and B, xB.
134
8. FINITE VOLUME DISCRETISATION
a) b)
tBy , vB
t B, uB
P P qB
B=e tBx , uB B=e
TB
xB xB
Figure 8.24 Dirichlet and Neumann boundary conditions for momentum a) and energy b).
The values of the displacements uE and vE required for the calculation of gradients (v / y)e ,
(u / y)e etc. are obtained by extrapolation from the interior of the solution domain.
Similarly, if the heat flux qB is given (Fig. 8.24b), the Qe given by eq. (8.106) is to be replaced
by
Qe = qB Se . (8.110)
135
8. FINITE VOLUME DISCRETISATION
a) b) vB
t Bx
t By
B=n yB
P
B=e uB
P
xB
n
vP
w P e
TP uP
yB
Symmetry sw s=B se
plane
TS
S
uS
vS
uS = uP , v S = vP ,
us = uB = uP , vs = v B = 0,
u u v
= , = 0,
x s x P x s (8.111)
and
TS = TP, y s = 2yB ,
Ts = TB = TP,
136
8. FINITE VOLUME DISCRETISATION
S v
Newtonian fluid
Firstly, the convection term will be discretised for the case of 2D planar flow and a Cartesian
control volume shown in Fig. 8.21. After that discretised equations for energy and momentum
will be assembled.
(un x + vn y ) dS ṁee
*
Ce = , (8.114)
Se
where e* stands for the cell-face mean value of velocity components u, v or internal energy cT,
and ṁe is the mass flux through the east cell face
u dS ue Se .
*
ṁe = (8.115)
Se
The way cell-face values of velocity components ue* and ve* and convected variable e* are
calculated has a strong influence on both the accuracy and stability of a numerical method. The
* *
velocity components ue and ve are obtained from a special interpolation practice which assures
*
a stable solution procedure as described in the next Section, while e is calculated by blending
the second-order accurate, symmetric central differencing formula
1
eCD = (1 f e ) P + fe E = ( P + E ), for a uniform mesh fe = 0.5 (8.116)
2
with some small amount of the first order asymmetric (upwind) differencing scheme, which
assumes that the value of at the cell face is calculated as
137
8. FINITE VOLUME DISCRETISATION
Hence
(
e* = eUD + eCD UD
e , ) (8.118)
where is a blending factor with a value between zero and one, thus combining the accuracy of
a second order scheme and the stability of a first order one. It is important to mention that the
CD scheme is implemented according to the 'deferred-correction' practice, i.e. convective flux is
split into an implicit part containing the contribution from the UD only (part of the coefficient)
and an explicit part consisting from the difference between CD and UD contributions (part of the
right hand side vector, see Section 8.14).
b) Discretised equations
After assembling all the terms featuring in eqs. (8.93), they can be written in the form of the
following non-linear algebraic equation which is identical to the one derived for solid body
stress analysis (4.1)
aP P aKK = b (K = E,W ,N,S), (8.119)
K
•b1) momentum
The coefficients for both u and v velocity components are the same
S S
aE = e e min(ṁe,0); aN = n n min(ṁn ,0);
x e yn
S S
aW = w w min(ṁw,0); aS = s s min( ṁs ,0); (8.120)
x w ys
V
aP = aK + (K = E,W ,N,S ),
K t
138
8. FINITE VOLUME DISCRETISATION
b2) energy
The coefficients for the energy equation differ from the ones for the heat conduction in solids
(8.107) by the presence of the convection terms only
S S
aE = ke e c e min(ṁe ,0); aN = k n n c n min(ṁn ,0);
x e yn
S S
aW = k w w c w min(ṁw,0); aS = ks s c s min(ṁs ,0);
x w ys
V
aP = aK + cP (K = E,W ,N,S),
K t
(8.122)
b = T ce [min(ṁe ,0) ṁe fe ](TE TP ) + T cw [min(ṁw,0) ṁw f w](TW TP ) +
T c n [min(ṁn ,0) ṁn f n ](TN TP ) + T c s[min(ṁs ,0) ṁs f s ](TS TP ) +
V (cT ) m1
hPV + (:gradv) + P
t
b3) continuity
In assembling equations (8.119) the continuity equation [see eq.(8.93)] in the discretised form
ṁk = 0 (k = e,n,w,s), (8.123)
k
where the mass fluxes ṁk defined by (8.115) are multiplied by the value of dependent variable
(u, v, cT) at point P, and subtracted from the corresponding transport equation. In this way the
central coefficient remains to be the sum of the neighbouring coefficients (in both space and
time). An independent use of the continuity equation, i.e. its conversion in an equation for
pressure will be dealt with in the next Section.
It can be noted that in the above procedure pressure, featuring in the source terms of the
momentum equations (8.121), has remained unknown, while at the same time no independent
use has been made of the continuity equation (8.123). The problem lies in the fact that pressure
does not feature explicitly in the continuity equation which consequently cannot be considered
139
8. FINITE VOLUME DISCRETISATION
as 'an equation for pressure' and it just comes as an additional constraint on the velocity field.
This constraint can be satisfied only by adjusting the pressure field. However, pressure is not a
conserved property and has no its governing transport equation, so it is not immediately clear
how this adjustment of pressure is to be performed.
At the same time, the pressure source term in the momentum equation is calculated in a way
that amounts to using a second-order space-centred scheme for the calculation of pressure
gradient [e.g. (8.78)]. For example, the pressure force acting on a control volume in the x
direction is calculated as (see Fig. 8.23a)
Fpx = ( pdS pdS) = ( pe Se pw Sw ) =
Se Se
(8.124)
1 1 1
( pP + pE )Se ( pW + pP )Sw = ( pE pW )S (if Se = Sw = S)
2 2 2
where a uniform mesh spacing is assumed (fe = 0.5). One can see that only alternate (and not
adjacent) computational nodes are involved in the pressure force calculation and an oscillatory
(checkerboard) pressure field is interpreted as uniform by the momentum equation (Fig. 8.27).
Such a scheme can produce a correct pressure-difference field, although the underlying pressure
field possesses an unphysical oscillatory profile.
p2
1/2(p1+p 2)
p
p1
* V p pw pE pW
ue = ue + v e , (8.125)
aP e x P xe
v
where the overbar denotes a spatially interpolated quantity [formula (8.116)], aP the
corresponding momentum equation central coefficient and xP = xn = xs (Fig. 8.22). The
second term in eq. (8.125) is a third-order pressure diffusion term, whose role is to smooth out
140
8. FINITE VOLUME DISCRETISATION
the oscillatory pressure velocity profile, and at the same time introduce pressure into the
continuity equation in such a manner that a pressure-correction equation can easily be
constructed. This is achieved by employing the predictor-corrector procedure defined by the
SIMPLE algorithm which will now be briefly described.
by expressing the velocity corrections in terms of the pressure corrections by using a truncated
form of eq. (8.125)
' V p pW
' '
ue = v E , (8.127)
aP e xe
and requiring that the double-starred cell-face velocities satisfy the continuity equation (8.123),
an equation of the form (8.119) for pressure correction p' is obtained, with coefficients
V S V Sn
aE = v e , aN = v ,
aP e x e aP n x n
(8.128)
V Sw V S
aW = v , aS = v s ,
aP w x w aP s x s
aP = aK (K = E, N, W , S) ,
K
b = ṁk (k = e, n, w, s) ,
k
where all variables have their predictor stage values, i.e. those that satisfy the (linearised)
momentum equations.
p = ppred + pp',
141
8. FINITE VOLUME DISCRETISATION
where p is an under-relaxation factor (typically p = 0.2 to 0.3), and is necessary because the
approximations introduced in deriving the pressure-correction equation often result in
overestimating the magnitude of p', which in turn leads to slow convergence or divergence of
the solution procedure. Finally, the mass balance satisfying fluxes are calculated as follows, e.g.
'
( '
ṁe = ṁe,pred + ṁe = ṁe,pred aE pE pP
'
) (8.130)
c4) Closure
It should be noted that the right hand side of the pressure-correction equation is the mass
imbalance [see eqs. (8.128)], caused by the fact that the predictor stage velocity components do
not in general satisfy the continuity equation at the initial stage of calculation. When eventually
the continuity equation is satisfied, b is equal to zero at all computational points, which results
in a zero p' field and a converged solution is obtained. In fact, the p' field is of no concern
whatsoever, and these quantities only serve in a temporary way to procure a satisfactory linkage
between the pressure and velocity fields so that they ultimately satisfy both the momentum and
continuity equations. Thus, taking the truncated form of the expression (8.125) (or any other
approximation) during the derivation of the pressure correction equation is justified as long as
the procedure converges.
The nature and significance of the third-order pressure diffusion term in expression (8.125) are
identical to those introduced by expression (8.146) in Section 8.4.1 where the diffusion
transport through an arbitrarily oriented cell face is discussed. Once a converged solution is
obtained, this correction term will vanish if the pressure varies linearly or quadratically in space,
or it will become small in the limit of a fine grid. In that case the velocity at the cell face will be
a simple spatial average, implying that the cell-centre velocities will also satisfy the continuity
equation.
It is important to note that the coefficient matrix of the pressure correction equation is
symmetric, but singular. However, in case of a well posed problem (the global mass balance
satisfied exactly) the resulting set of linear algebraic equations is undetermined, which
practically means that the pressure correction, and consequently the pressure level is
undetermined. This, however, has no consequences on the solution, since only pressure
gradients (differences), and not the pressure itself, feature in fluid momentum equations [see
e.g. (8.124)].
142
8. FINITE VOLUME DISCRETISATION
A number of different types of boundary conditions is encountered in fluid flow problems. The
most frequently used ones will now be discussed.
d1) Inlet
This is an inflow boundary where the distribution of all variables is know, and it is therefore of
a Dirichlet type.
d2) Outlet
This is an outflow boundary where the distribution of dependent variables is not known and a
kind of Neumann boundary condition is applied. Namely, the velocity components are
extrapolated in a way that forces an exact (to the round-off error) overall mass balance, e.g. if
outlet is at the east boundary
uB = uP +u, (8.131)
where
ṁinf uP dS
SB
u = , (8.132)
S B
where ṁinf is the total mass inflow and SB is the total outlet area. The other dependent variables
are obtained by simple (usually zero order) extrapolation.
d3) Wall
For the momentum equation Dirichlet boundary conditions are usually applied, i.e. the boundary
velocity components are set equal to the wall velocity (zero in the case of a stationary wall),
while the boundary conditions for the energy equation can be either of Dirichlet (specified
temperature) or Neumann (specified heat flux) type.
e) Turbulent flow
Observations of fluid flow have revealed that two different flow regimes, laminar and turbulent,
can be distinguished. While in a laminar flow the migration of fluid elements from one layer to
143
8. FINITE VOLUME DISCRETISATION
another does not take place, i.e. the flow can be thought of as a motion in essentially parallel
paths, turbulent flow is characterised by a chaotic motion of fluid elements and their migration
not only to the neighbouring, but also to distant regions and it is inherently three-dimensional
and unsteady. It is also observed that a stable laminar flow occurs only at a very low range of
Reynolds numbers (in pipes Re =
uD/ < 2300) and otherwise is turbulent.
Stokes' constitutive relationship is valid irrespective of the flow regime. However, the small
length and time scales of turbulent flow make its direct solution absolutely impractical, and
some kind of turbulence modelling is often used to overcome that difficulty. The most
frequently employed is the so-called two equation k– model of turbulence, which solves a
transport equation of the form [3rd expression in (7.15) or in (8.93)] for both the turbulent
kinetic energy k and its dissipation rate , and then calculates the so-called turbulent viscosity
tur and turbulent conductivity ktur via simple algebraic relationships involving k and . The
governing transport equations (7.15 or (8.93)) retain their form, but u, v, p and T now stand for
the respective time averaged quantities and and k for their effective values (eff = + tur,
keff = k + ktur).
In conjunction with the k– model the so-called wall functions are used in formulating boundary
conditions at wall boundaries.
144
8. FINITE VOLUME DISCRETISATION
The inter-equation coupling (e.g. v displacement appearing in the u-momentum equation, and
vice-versa) is often another reason for an iterative solution procedure. Namely, instead of
solving a system of equations containing all dependent variables, the decoupled sub-systems for
each dependent variable are obtained by temporarily assuming all other dependent variables to
be known (taken from an initial guess or previous iteration time step). This approach, known as
segregated solution algorithm, is employed by the present FV method and will be presented in
the next Section. A part of that algorithm is the solution of the systems of linear algebraic
equations and the Incomplete Cholesky Conjugate Gradient (ICCG) is used.
Since the Finite Volume methods favour segregated over simultaneous solutions of discretised
algebraic equations, in this Section a segregated algorithm is described.
As a result of FV discretisation of conservation equations, and by integrating over a time
interval t and over all N CVs three in general mutually coupled sets of N non-linear algebraic
equations with three unknowns (u, v, T) of the form (8.119) are obtained for either thermo-
elastic solid or Newtonian fluid
where stands for u, v or T. An iterative procedure for the solution of these equations will now
be presented.
145
8. FINITE VOLUME DISCRETISATION
System of 3 x N
(non)linear alg.
equations
'Genuine'
iterative
methods
outer iterations
no
CONVERGED iterations Iterative Direct
inner
yes
no yes no
SOLUTION Converged CONVERGED
yes
SOLUTION
• First, all dependent variables are given the corresponding initial values (actual initial values
in case of a transient analysis, and an arbitrary initial guess if only the steady state is sought).
Then the boundary conditions which correspond to the first time step are applied and the sets of
equations for each individual dependent variable are linearised and temporarily decoupled by
assuming that coefficients and source terms are known (calculated by using dependent variable
values from the previous iteration or the previous time step) resulting in a system of linear
algebraic equations of the form
A = b (8.133)
for each dependent variable, where A is an N x N matrix, vector contains values of dependent
variable at N nodal points and b is the source vector.
• These systems are then solved sequentially in turn until a converged solution is obtained.
The procedure is assumed converged when the following conditions are satisfied for all three
sets of equations
146
8. FINITE VOLUME DISCRETISATION
N N
aK K + b aP P < p aP P ,
i =1 K i =1 (8.134)
i( n) (i n 1) < q i(n 1) (i = 1,2,...,N)
where p and q are typically of the order of 10–3 and superscripts (n) and (n–1) denote values at
the two successive iterations.
• In the next time step the whole procedure is repeated, except that the initial values are
replaced by the values from the previous time level.
Systems of linear equations [eq. (8.133)] can be solved by a number of methods. Convergence
of those so called inner (solver) iterations is promoted by diagonal dominance of the coefficient
matrix. Since (coefficients and) sources are only approximated (based on the values of
dependent variables from the previous iteration/time step), it does not make sense to solve eq.
(8.133) to a tight tolerance. Normally, reduction of the sum of absolute residuals for one order
of magnitude suffices.
In order to promote convergence of this solution method, an under-relaxation is often necessary.
This has been introduced in the following manner, by replacing A and b in eq. (8.133) with
1 1
A A+ D and b b+ D (n-1) , (8.135)
The segregated solution strategy employed enables reuse of the same storage for matrix A and
vector b for all dependent variables , thus requiring only 6N storage locations in a 2D case (8N
storage locations in a 3D case).
It is also important to recall that the fully implicit time differencing used avoids stability-related
time step restrictions. In principle, it allows any magnitude of the time step to be used, and in
practice it is limited only by the required temporal accuracy. The number of iterations per time
step depends on the size of the time increment t; for smaller t fewer iterations are required to
obtain the solution at the time tm. If only a steady state is sought, the temporal accuracy is not
important and the solution can be obtained by using a single, infinite time step.
147
8. FINITE VOLUME DISCRETISATION
In case of fluid flow calculations it may, however, be useful to mention the following:
(1) Due to the presence of the convective transport fluid flow equations are always non-
linear.
(2) Due to the deferred-correction implementation of the discretised convection terms, which
employs the asymmetric upwind scheme, the coefficient matrix of the linearised momentum and
energy equations (8.133) is always asymmetric and the ICCG method cannot be used (one can
employ the Gauss-Siedel method, or some variant of the incomplete Cholesky preconditioned
conjugate gradient method applicable to systems with asymmetric matrices, e.g. CGSTAB).
(3) The under-relaxation factors are typically: for the momentum equations v = 0.7 – 0.8,
for the pressure correction equation p' 1 – v, and for the energy equation T = 0.9 – 1.
(4) From the point of view of a segregated solution procedure the existence of the
continuity equation is very convenient, because it is a single scalar equation which links all
velocity components and 'takes care' of the inter-equation coupling. Consequently, unlike
thermo-elastic stress analysis where the employment of a multigrid acceleration procedure is
crucial for a good convergence rate, coupling among velocity components is not a bottle-neck
for convergence, and the problem of convergence moves towards resolving non-linearities.
Finally, it may be useful to explicitly state the way the pressure-correction equation is
incorporated into the solution algorithm:
Step 1. Provide initial values of dependent variables (at the time to).
Step 2. Assemble and solve equation (8.133) for the velocity components, employing the
currently available pressures and mass fluxes.
Step 3. Assemble and solve equation (8.133) for the pressure correction and use calculated
values to correct mass fluxes, velocity components and pressure [eqs. (8.129) and (8.130)].
Step 4. Assemble and solve equation (8.133) for internal energy and obtain temperature.
Step 5. Return to Step 2 and repeat until the sum of the absolute residuals for all equations
has fallen by a prescribed number (typically three) orders of magnitude.
Step 6. Advance the time by the time increment t and return to Step 2; repeat until the
prescribed number of time steps is completed.
148
8. FINITE VOLUME DISCRETISATION
8.4 Summary
The main features of the present FV discretisation method are discussed and
• a second order accurate (on a uniform mesh) fully conservative spatial differencing, enabling
reasonably accurate and physically meaningful results to be obtained on a coarse numerical
mesh are selected.
Then, the 2D momentum and energy balance equations are discretised for cases which can be
'meshed' by a non-uniform Cartesian mesh. This resulted in a 5-diagonal system of 3 x N
generally non-linear algebraic equations for each time step.
The adopted procedure of solving large systems of non-linear algebraic equations incorporates:
• a segregated iterative solution algorithm offering a very low memory requirement which
enables large problems to be solved on small computers; in addition, due to its iterative nature,
it is directly applicable to non-linear solids,
• an efficient conjugate gradient linear equations solver which is considered to be one of the
most powerful methods for the solution of linear algebraic equations, which involves little, or,
sometimes, no memory overheads.
149
8. FINITE VOLUME DISCRETISATION
In this Section a second-order accurate numerical method will be presented which enables a
solution domain to be discretised by polyhedral cells of a completely arbitrary shape, thus
enabling geometrically complex problems to be solved accurately and efficiently.
Due to the time constraint, the energy and momentum equations for a thermo-elastic solid will
only be considered. The discretisation of the fluid equations normally requires discretisation of
the convective terms [see Sections 8.1.4 and 8.2.2] and pressure terms (employing continuity
equation, Section 8.2.2), and will not be considered here.
In case of thermo-elastic solids, the generic transport equation (7.39) can be simplified as:
t
B dV = grad ndS + q S ndS + qV dV , (8.136)
V S S V
where stands for the displacement u and temperature T, and all other terms are defined in table
1 [Section 7.5].
a) Vector and tensor components related to a global Cartesian coordinate system are
preferred because they lead to a strong conservation form of momentum equations,
without terms requiring a smooth numerical mesh.
b) The space is discretised by an arbitrary unstructured mesh, because it is by now generally
accepted that only unstructured meshes are capable of describing accurately complicated
3D geometries. The time is subdivided into a number of unequal time intervals. Spatial
gradients are interpolated by a second-order accurate scheme, while for the temporal
discretisation a fully implicit time differencing is employed.
c) All dependent variables are stored at the cell centre, i.e. a collocated variable arrangement
is used, which requires only one set of control volumes to be generated.
150
8. FINITE VOLUME DISCRETISATION
Details about the implementation of the above options into the present method will now be
explained.
n2 S1
S2
P
nn Sn
Nn
Sj
z nj
xP
dj
Nj
y
x
After that eq. (8.136) for each control volume is written as follows
n n
B dV = grad ndS + qS ndS + qV dV ,
t
(8.137)
V j =1 Sj j =1 Sj V
where n is the number of cells which share cell faces with cell P (so called nearest neighbours)
and and B in the case of momentum equations are now related to the Cartesian displacement
components ui (i = 1,2,3).
151
8. FINITE VOLUME DISCRETISATION
Eq. (8.137) has three distinct parts: transient rate of change, diffusion and source. This equation
is exact, i.e. no approximation has been introduced so far. However, in order to evaluate
integrals in the above equation, a way of calculating integrals and a distribution of dependent
variable and physical properties in space and time have to be assumed.
Calculation of integrals
Both volume and surface integrals are calculated by using the mid-point formula.
Spatial distribution
Here, the following linear spatial distribution is employed (although, in principal, any other
distribution can be used)
(x,t ) = P ( t) + gP ( t)( x x P ) (8.138)
where stands for dependent variable , physical properties of continuum or gradient of , xP
is the position vector of point P and gP(t) is an approximation of gradient of at point P.
The unknown vector gP is calculated by ensuring a fit to a set of sampling points usually
consisting of the nearest neighbours of point P, i.e. by solving the following set of equations
d j g P = N j P ( j = 1,...,n) (8.139)
where d j = x N j x P is the distance vector joining point P with its neighbour Nj (Fig. 8.29). To
solve this over determined set of equations the least square method is used, resulting in the
following normal equation
-1
DgP = h i.e. gP = D h (8.140)
where
dTj ( N )
n n
D= dTj d j and h= j
P (8.141)
j =1 j =1
It is important to notice that for all 's and for assumed linear distribution, D is a symmetric 3x3
matrix, whose coefficients depend solely on the mesh geometrical properties and hence, for a
given cell, a single evaluation of D1 suffices.
Formula (8.138) is used to calculate values of at the faces j, necessary for evaluation of the
surface integrals in (8.137), leading to a second-order symmetric formula
j =
1
2 ( 1
) ( ) ( )
P + N j + (grad ) P x j x P + (grad ) N x j x N j
2 j
(8.142)
152
8. FINITE VOLUME DISCRETISATION
Temporal distribution
As far as temporal distribution of is concerned, according to the adopted fully implicit time
discretisation all values of dependent variables are expressed as
( x,t) =
(x,tm ) =
m (8.143)
where m is the time step counter and tm is the current instant of time, except in the transient
term where a linear variation of in time is assumed. Henceforth the time step counter will
normally be omitted, and all values of will refer to the current time tm, unless indicated
otherwise.
Rate of change
where the meaning of B and B m1 for different is given in Table 8.1.
m1
Table 8.1: The meaning of B and B
in eq. (8.144).
m1
B B
m1 m1
T cT c T
ui
1
t m ( m1
ui ui ) 1
t m1 ( m1
ui ui
m2
)
Diffusion flux
Diffusion flux Dj of through the cell face j can be approximated by
Dj =
grad
ndS
j (grad
) j n j S j ,
*
(8.145)
Sj
where
j stands for the cell-face mean value of the diffusivity, obtained by using second-order
formula (8.142), and the value of (grad ) j is constructed in the following manner
*
N
P grad
d
(grad
)*j = (grad
) j + j
j
n , (8.146)
dj dj j
where the overbar means an arithmetic average between nodes P and Nj straddling the cell face j
and value of (grad ) j is calculated by using second-order formula (8.142).
The gradient based on expression (8.142) is second order space-centred and as such cannot
sense the oscillations which have the period twice the characteristic length of the numerical
mesh (see expression 8.125 and Fig. 8.27 in Section 8.2.2). As a result, once induced, the
unphysical oscillatory profile remains superimposed onto the otherwise smooth spatial variation
153
8. FINITE VOLUME DISCRETISATION
of the dependent variable. Because of this, a third-order diffusion term (term in [] brackets) is
introduced into expression (8.146) for the cell-face gradient of variable . This additional
diffusion transport smoothes out these unphysical oscillations and, once a converged solution is
obtained, it will vanish if the spatial variation of is linear or quadratic. Otherwise, its
magnitude is proportional to the truncation error of the second-order scheme, so it does not
affect the second-order behaviour of presented discretisation practice.
The contribution to the diffusion flux coming from the second term in eq. (8.146) is treated
implicitly. The rest of the diffusion flux is due to component of the gradient normal to the
distance vector dj (the 'cross-diffusion') and it vanishes when the grid is orthogonal, and is small
compared to the so called 'normal diffusion' if the grid non-orthogonality is not severe. For this
reason it is treated explicitly. This simplifies the coefficient matrix of the resulting set of linear
algebraic equations since only the contributions of the nearest neighbours of node P are then
treated implicitly.
Source term
The part of the source term coming from the true, volume sources is integrated by assuming a
linear variation of the source over the control volume
QV = qV dV qV VP ,
P
( ) (8.147)
V
while the part coming from the stress tensor in momentum equations is discretised in the same
manner as diffusion term
[μ(gradu) ]
n n
+ [divu (3 + 2μ) (T T0 )]I ii n j S j (8.148)
T
Qu iS = qS ndS
j =1 S j j =1
j
where ii (i = 1,2,3) are the Cartesian base vectors. Note that qTS = 0.
It is often possible to linearise volumetric source term Q V and add a part of it to the central
coefficient of the resulting coefficient matrix which will then increase its diagonal dominance,
thus enhancing the convergence of the resulting linear algebraic system. It may also be
beneficial, from a stability and efficiency point of view, to take some parts of the surface source
Q S into account implicitly, as was done in Section 8.2.
After assembling all the terms featuring in eq. (8.137), it can be written in the form of the
following non-linear algebraic equation which links the value of dependent variable at the
control volume centre with its values at the points in the neighbourhood
n
aP P aj Nj = b (8.149)
j =1
154
8. FINITE VOLUME DISCRETISATION
where
Sj
a
j =
j ,
dj
n E
VP
a
P = a
j + tm
, (8.150)
j =1
n grad
d j
m1 m1
E
VP
b
=
j (grad
) j n j
S + Q
V + Q
S +
j =1
dj
j tm
m1
and E and E
are defined in Table 8.2.
m1
Table 8.2: The meaning of E and E
in eq. (8.150).
m1
E E
T c c m1T m1
ui
t m
m1
t m
ui +
u
tm1 i (
m1 m2
ui )
Eq. (8.149) is of the same form as eqs. (8.101), (8.103) and (8.108), and it is easy to verify that
the coefficients (8.150) reduce to the corresponding ones from Section 8.2 [(8.100), (8.102) and
(8.107)] if a Cartesian space discretisation is employed. The type and implementation of the
initial and boundary conditions is analogous to those described in Section 8.2, while the
segregated solution procedure is entirely applicable to the solution of the system of equations
(8.149) with coefficients and sources (8.150).
8.5.2 Conclusions
The numerical method presented here can be applied to a wide range of problems in continuum
mechanics. Although very general as far as its applicability to different geometrical and
physical situations is concerned, the method is extremely simple. Namely, at first sight an
arbitrary CV that can be used for space discretisation (Fig. 8.29) might suggest that a method
which can use such discretisation must involve very complicated mathematics. The fact is,
however, that all one has to deal with are vectors and tensors in a global Cartesian coordinate
frame (unless someone wishes to do otherwise). As far as physics is concerned, the method
reduces to a straight forward application of the first principles to a control volume.
The use of cells of different topologies that can be combined in an arbitrary way is the most
general approach to space discretisation and greatly facilitates the mesh generation for complex
geometrical configurations when an accurate description of boundaries is desired. This feature
is particularly useful for adaptive mesh refinement when injection and depletion of
computational points is straightforward and at the same time should help automatic mesh
generation, which is rapidly becoming a very important issue in computational continuum
155
8. FINITE VOLUME DISCRETISATION
mechanics. The discretisation presented facilitates the use of such meshes. It is performed in
real space, using global Cartesian coordinates, thus eliminating the need for coordinate
transformation, which makes it easy to understand and implement. The discretisation promotes
the conservation of momentum and energy on a discrete level, which, combined with a fully
implicit time differencing, enables physically meaningful results to be obtained on very coarse
meshes and for large time steps.
156
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
Chapter 9
9.1 Background
As discussed in chapters 1–8, continuum mechanics involves the solution of equations (e.g.
conservation of momentum) defined over a body, subject to initial and boundary conditions.
When the geometry of the body of interest is simple, analytical methods can be used (e.g.
beam theory) and exact solutions may often be obtained. Such methods cannot easily
be applied to bodies of complex shape. However, by dividing the body into small, finite
‘elements’ of regular shape and applying the laws of continuum mechanics to the collection
of elements, an approximate solution to the problem can be obtained. As the number of
elements used is increased (and their size decreased) we approach the exact continuum
mechanics (infinitely small elements) solution.
The development of the finite element method is often attributed to Clough and co-
workers in the 1960s at Boeing in the US, who coined the term finite element, and used
the method to analyse aircraft structures. Argyris at Imperial College was also one of the
pioneers in this field. (A very brief and personal history of the finite element method is
provided by Hughes of Stanford University at www.usacm.org/T.J.Hughesacceptanceremarks.htm).
The solution of a finite element problem can involve many thousands of repetitive
calculations and therefore most problems must be solved by computer. With the increasing
availability and affordability of computers in the 1980s and 1990s, the finite element method
expanded rapidly. Nowadays, most structural design makes use of the finite element method,
as it can significantly decrease the time and cost to market of a product, by reducing the need
for expensive prototype testing. Many commercial finite element codes exist, e.g. ABAQUS
(www.abaqus.com), ANSYS (www.ansys.com), ADINA (www.adina.com) and DYNA (www.lstc.com),
most of which are available within the mechanical engineering department. The largest
finite element problems can involve millions of variables (degrees of freedom) and can take
weeks to solve even on the fastest computers currently available.
The application of the finite element method was originally limited to static solid me-
chanics problems but has since been extended to the study of many field problems, e.g.
heat transfer, fluid mechanics and impact and dynamics problems. In this part of the course
we focus on the solution of solid mechanics problems, but the methods presented can be
relatively easily extended to deal with more complicated problems.
Some of the content of these notes is based on the textbook by Belytschko, Liu and
Moran: Nonlinear Finite Elements for Continua and Structures, published by Wiley and
sons, 2000. Another useful reference book, which is perhaps easier to follow for the beginner
but uses somewhat different notation from these notes, is The Finite Element Method, by
157
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
Zienkiewicz and Taylor, 2000. This book is in three volumes, the most relevant to this
course being Volume 1: The Basis.
Consider a virtual displacement field, δui . This field may take any value provided it is
continuous and bounded (not infinite) over V and δui = 0 on the displacement boundary,
Γu (see section 5.4.1). Multiply the linear momentum equation, Eq. 9.1(a) by this virtual
field, δui , and integrate over the body V to get:
δui (σji,j + ρfib − ρüi )dV = 0. (9.2)
V
158
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
Using Gauss’s theorem (Eq. 2.80) we can rewrite the first term on the RHS of the above
equation as follows:
∂
(σji δui )dV = · (σδu)dV
V ∂xj V
= (σδu)ndΓ
Γ
Equation 9.4 is known as the weak form of the momentum balance equation or the principle
of virtual work. In Eq. 9.4 δui is the virtual displacement field which may take any value
provided it is zero on Γu .
We have shown that the strong form of the momentum equation, Eq. 9.1 leads to the
weak form, Eq. 9.4. It can also be shown that the inverse applies—if Eq. 9.4 is satisfied for
all virtual displacement fields, δui , then the displacement field ui satisfies Eq. 9.1(a) and
the corresponding stress fields satisfies the traction boundary conditions, Eq. 9.1(b). Note
that the unknown in Eq. 9.4 is the displacement field, ui (σij depends on ui through the
strain εij which is the derivative of ui .) It will be seen that the virtual field δui does not
appear in the final form of the finite element equations.
159
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
We first simplify the weak form equation by assuming equilibrium, üi = 0, taking body
forces as zero, fib = 0, and invoking symmetry of the stress tensor, σji = σij . The equation
becomes:
σij δui,j dV − δui fisn dΓ = 0. (9.5)
V Γt
where Uia are the discrete (unknown) nodal displacements, Na (x) are the (known) shape
functions, which we will discuss further later, and Nn is the total number of nodes in the
model.
The virtual displacements δui are interpolated using the same shape functions:
Nn
δui (x) = Na (x)δUia , (9.7)
a=1
160
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
Nn
∂ ∂
⇒ δui (x) = Na (x)δUia
∂xj x
a=1 j
(9.9)
Nn
= Na,j (x)δUia
a=1
Nn
⇒ δUia σij Na,j dV − Na fisn dΓ =0 (9.10)
a=1 V Γt
The above equation must hold for any virtual displacements δUia . Therefore, we can write,
σij Na,j dV − Na fisn dΓ = 0
V Γt
or
σij Na,j dV = Na fisn dΓ (9.11)
V Γt
The finite element equation, Eq. 9.11, may be further simplified by using the constitu-
tive law to write the stress in terms of the unknown nodal displacements. We specialise to
the case of linear elasticity, though there is not too much difficulty in dealing with non-linear
161
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
material behaviour. (In the latter case the finite element equations become non-linear and
cannot be solved in a single step (iteration)).
For linear elasticity, following Eq. 6.12,
1
σij = cijkl (uk,l + ul,k )
2
= cijkl uk,l (due to symmetry of cijkl ).
Now
Nn
uk = Nb Ukb ,
b=1
Nn
⇒ uk,l = Nb,l Ukb ,
b=1
so
Nn
σij = cijkl Nb,l Ukb . (9.14)
b=1
or
Kiakb Ukb = fia
or
KU = f , (9.16)
where U is the vector of unknown nodal displacements (e.g. U21 is the displacement at node
1 in the 2 (y) direction), f is the nodal force vector and K is the finite element ‘stiffness’
matrix (note the analogy with the equation for a spring, F = kx, where F is the force in
the spring, k the spring stiffness, and x the displacement of the spring).
The solution to Eq. 9.16 is obtained by any appropriate numerical method, which
essentially involves the inversion of the stiffness matrix, i.e.
U = K−1 f .
162
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
Equation 9.16 is the finite element equation for a linear analysis—linear geometry (small
displacements and rotations) and linear material response. The formulation for non-linear
response is similar. It will take the same form as Eq. 9.16 but the stiffness matrix will not
be constant (as in the linear case) but will depend on the applied load or displacement.
The incorporation of dynamic effects (via the ρüi term in the momentum equation) is also
relatively straightforward.
The main task in a finite element analysis is the calculation and inversion of the finite
element stiffness matrix. In a non-linear analysis this operation may be carried out many
thousands of times. The matrix K in equation 9.16 is often referred to as the ‘global’
stiffness matrix. The term Kiakb can be interpreted as the stiffness at node a in the degree
of freedom i due to movement of node b in degree of freedom k.
Each element in the finite element mesh contributes an amount to the global stiffness
matrix and the amount it contributes is called the ‘local’ stiffness matrix. The finite element
procedure operates on the principle that all calculations carried out at an element level require
only information from the element itself, which considerably streamlines the computational
analysis as each element can be treated independently and the calculations carried out on
each element are identical. This feature is achieved through the use of the shape functions,
Na , whereby the value of the variables within an element depend only on the values at the
nodes of that element.
The global shape functions Na (x), used to interpolate the displacement field, are
constructed by adding together the local (element) shape functions, Nae (x). For now we
focus on the latter. The requirements for the element shape functions Nae are that
(i) Nae = 1 at node a
= 0 at all other nodes
N
EL
(ii) Nae = 1 where N EL is the number of nodes per element
a=1
(iii) Nae = 0 outside element e.
Within an element the displacement u is given by
N
EL
ui (x) = Nae (x)Uia . (9.17)
a=1
For the rest of this sub-section the superscript e is dropped for convenience.
163
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
Consider an example of a 1-D element with 2 nodes (see Fig. 9.3). The nodes are
designated to lie at x = −1 and x = +1 respectively. (This is known as the parent element.
When carrying out a finite element analysis, the actual elements are mapped to this parent
element (more on this later).)
U1 u(x)
U2
node 1, x = −1 node 2, x = +1
1
N1 (x) = (1 − x)
2
1
N2 (x) = (1 + x).
2
It is clearly seen that requirement (i) above is satisfied for both shape functions, since
N1 (−1) = 1, N1 (+1) = 0, N2 (−1) = 0, N2 (+1) = 1 and clearly N1 + N2 = 1.
The use of the shape functions in this way is simply a convenient way of defining a
linear interpolation between node 1 and node 2. To find the value of the variable u at any
point within the element,
2
1 1
u(x) = Na (x)U a = (1 − x)U 1 + (1 + x)U 2 ,
a=1
2 2
e.g.
U1 + U2
u(0) = (as expected).
2
Quadratic interpolation over an element can also be used. For this three nodes are
required, with the third node usually lying halfway between the other two (see Fig. 9.4).
This provides more flexibility in describing the variation of u within an element but also
increases the complexity of the problem as an extra node has been introduced. Note that
164
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
the 3 node quadratic element in Fig. 9.4 could be replaced by two 2 noded linear elements.
In general quadratic elements are preferred, though there are situations when the use of
quadratic elements can cause problems. For example, in contact problems linear elements
are recommended.
U1
U3 U2
x(x − 1)
N1 (x) =
2
x(x + 1)
N2 (x) =
2
N3 (x) = 1 − x2
For the more common two dimensional elements we can again describe linear and
quadratic shape functions, though in 2D they are actually bi-linear and bi-quadratic (e.g.
see Fig. 9.5). For example, the linear shape functions are given by,
1
N1 (x, y) = (1 − x)(1 − y)
4
1
N2 (x, y) = (1 + x)(1 − y)
4
1
N3 (x, y) = (1 + x)(1 + y)
4
1
N4 (x, y) = (1 − x)(1 + y)
4
165
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
Shape functions of higher than quadratic order are rarely used except for specialised appli-
cations.
We have seen in section 9.5 that the global stiffness matrix may be written as
Kiakb = cijkl Na,j Nb,l dV .
V
The terms Na and Nb in the above are the global shape functions. The global shape
functions are simply the sum over the body of the local shape functions, i.e.
Ne
Na (x) = Nae (x),
e=1
N2 ( x ) N1e ( x )
1 1 2 2 3 3 4
1 2
166
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
1 2
The cross terms (e.g. N2,j × N1,l ) disappear because of the property that Nae (x) = 0 and
Naf (x) = 0 outside the element e or f respectively.
We can then rewrite the stiffness matrix as
Ne
e
Kiakb = kiakb ,
e=1
e
where kiakb is the element stiffness,
kiakb = cijkl Na,j Nb,l dV , (9.19)
V
So far we have used tensorial notation to denote the stiffness matrix and the finite
element equation. In some cases the equations are more easily represented if we use the
equivalent ‘matrix’ notation (sometimes referred to as Voigt notation).
Consider the constitutive law for an isotropic linear elastic material (Eqs. 6.12 and
6.15),
167
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
σm = Dmn εn , (9.23)
with m, n ranging from 1 to 6. Note the factor of 2 in the shear strain terms in Eq. 9.22
(not all finite element codes follow this convention). This is for convenience and to allow
us to write for example,
σ · ε = σn εn .
N EL
1 1
εij = (ui,j + uj,i ) = Na,j Uia + Na,i Uja . (9.24)
2 2 a=1
For example,
N EL N EL
1 a a
ε11 = (Na,1 U1 + Na,1 U1 ) = Na,1 U1a .
2 a=1 a=1
N EL
1
ε12 = (Na,2 U1a + Na,1 U2a ).
2 a=1
We can expand the sum in the strain-displacement relation above to get (for a 4 noded
element):
168
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
εm = Bmp Up , (9.27),
where B is the matrix of shape function derivatives, as illustrated in Eq. 9.26. In Eq. 9.27
m ranges from 1 to N ST R, where N ST R is the number of strain components (6 in Eq.
9.26) and p ranging from 1 to N EL × N DF , where N EL is the number of nodes per
element (4 in Eq. 9.26) and N DF is the number of degrees of freedom per node (3 in Eq.
9.26).
If we recall the original form of the weak form expression (Eq. 9.5), the contribution
of the stiffness term for a linear elastic material may be written as
σij δui,j dV = cijkl uk,l δui,j dV .
V V
Looking at the integrand of the term on the RHS of the above equation, we can write,
In ‘matrix’ form,
cijkl δεij εkl = Dmn δεm εn ,
169
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
+1 N
QP
I= f (x)dx ≈ wp f (ξp ), (9.30)
−1 p=1
where ξp give the position of the quadrature points (sometimes called gauss points, as the
method is also known as Gaussian quadrature), wp are the ‘weights’ assigned to each gauss
point and N QP is the order of quadrature. If N QP = 1 we have linear quadrature; if
N QP = 2 we have quadratic quadrature etc.
If we have N QP quadrature points it may be shown that we can integrate exactly a
polynomial of order 2 × N QP − 1, i.e. with 1 gauss point we can integrate a linear function
exactly and if N QP = 2 we can integrate a cubic function.
170
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
The coordinates and weights of the gauss points corresponding to a particular order
of integration are given in any standard text book on numerical methods. They may be
evaluated by integrating simple expressions whose results are known and comparing with the
result obtained from Gauss quadrature. This is illustrated below for the case of N QP = 1.
9.7.1 First Order Quadrature
For N QP = 1 we consider integrating a constant function f (x) = 1 and a linear function
f (x) = x over the ‘parent’ element.
1 1
1dx = 2; wp f (ξp ) = w1 × 1 = w1
−1 p=1
⇒ w1 = 2
1 1
xdx = 0; wp f (ξp ) = w1 × ξ1 = 2ξ1
−1 p=1
⇒ ξ1 = 0.
i.e. the midpoint rule. Equation 9.31 will exactly integrate functions of the form
f (x) = a0 + a1 x.
NQP ξ w
1 0 2
2 −1/√3 +1/√3 1 1
3 − 6 / 10 0 − 6 / 10 5/9 8/9 5/9
4
− (3 + 4.8 ) / 7 − (3 − 4.8 ) / 7 (3 − 4.8 ) / 7 (3 + 4.8 ) / 7 1/2 − √30/36 1/2 + √30/36 1/2 + √30/36 1/2 − √30/36
171
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
N QP
1 1 1
I= f (x, y)dxdy = wp f (ξp , y) dy
−1 −1 −1 p=1
N
QP 1
= wp f (ξp , y)dy
p=1 −1
N QP
N QP
= wp wq f (ξp , ξq ) ,
p=1 q=1
i.e.
1 1 N
QP N
QP
f (x, y)dxdy = wp wq f (ξp , ξq ). (9.32)
−1 −1 p=1 q=1
For NQP = 2 we have 2 × 2 integration as shown in Fig. 9.7 and Eq. 9.32 is,
2
2
I= f (ξp , ξq ),
p=1 q=1
× × × × ×
− − −
× × × ×
− √ √
× × × × ×
− − −
172
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
We can now apply Gaussian quadrature to the integration of the element stiffness
matrix. Assuming a finite element mesh with two spatial dimensions,
k= BT DBdV
V
N
QP N
QP (9.33)
T
= wp wq B (ξp , ξq )DB(ξp , ξq ).
p=1 q=1
In the above it is assumed that the material properties, D, (λ and μ for isotropic elasticity)
do not vary with position.
We next consider what order of integration is needed to integrate the above equation
exactly. The D matrix is assumed independent of position and does not affect the order of
integration needed. Since the B matrix is made up of the derivatives of shape functions,
the order of B will depend on the order of shape function used, e.g. linear or quadratic.
If we have a linear 2D element the shape functions are bi-linear. Therefore the B matrix
is linear and the integrand of the stiffness matrix, BT DB is quadratic. Therefore we need
N QP = 2. If we are using a quadratic element, then BT DB is quartic (to the power of
4). Therefore in this case, to integrate exactly, we require N QP = 3 (see Fig. 9.7).
The above analysis assumes that the element is equivalent to the parent element. We
will subsequently generalise our analysis to the case of an arbitrarily shaped element.
To assemble the stiffness matrix using Eq. 9.33 the matrix B is evaluated at the
integration points. Therefore it is relatively simple to evaluate strain at the integration
points using Eq. 9.27 and then stress using Eq. 9.23. Thus finite element codes often
provide as output the stress and strain at the integration points rather than e.g. at the nodal
points (note that the displacement is always obtained at the nodes). However, there is no
difficulty, if required, in extrapolating stress or strain values determined at the integration
points to the nodes or indeed averaging the values over an element, which often provides a
more accurate value for the stress.
As discussed in Section 9.8 we carry out the integration on the parent element using
numerical quadrature. The next step is to map the result to the actual element (see Fig.
9.8). In what follows the notation s and s1 , s2 will be used to denote coordinates on the
parent element and x and x1 , x2 to denote the mapped element (a 1D or 2D analysis is
assumed throughout though the procedure is identical in 3D).
173
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
Figure 9.8, Mapping from the parent element to the finite element for 1D and 2D elements.
Consider integration of a function f (x) over a 1D finite element space,
I= f (x)dx
x
By using a change of variables, we can integrate over the parent element, e.g.,
dx
f (x)dx = f (x(s)) ds,
x s ds
The term dx/ds is known as the jacobian, J, of the mapping and we can write,
f (x)dx = f (s)J(s)ds. (9.34)
X S
To calculate the jacobian we need to know the mapping from s space to x space.
Generally in a finite element analysis, the shape functions are used (i.e. to account for the
element ‘shape’).
(s1, s2)
(x1, x2)
(S 1
2
,S 2
2 )
(X 1
2
, X 22 )
Figure 9.9, Mapping from the s space to the x space.
We write, in the general case, that
N
EL
x(s) = Na (s)Xia , (9.35)
a=1
174
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
where Xia indicate the coordinates of the nodes of the mapped element (see Fig. 9.9).
Elements which use the same shape functions to map the nodal coordinates and the
nodal displacements are known as isoparametric elements—the vast majority of elements
do, though not all.
In 2D we define a jacobian matrix, Jij and
N EL N EL
∂xi ∂
Jij = = Na (s)Xia = Na,j Xia (9.36)
∂sj ∂sj a=1 a=1
and the jacobian of the transformation from s-space to x-space is the determinant of this
matrix,
∂xi
J = det .
∂sj
We can then write that
N
QP N
QP
k= wp wq BT (ξp , ξq )DB(ξp , ξq )J(ξp , ξq ). (9.37)
p=1 q=1
The value of the shape functions at the mapped gauss points is the same as the value
on the parent element (this is a property of the isoparametric transformation) so they can be
evaluated using the parent element definitions. Note, however, that there are derivatives in
the expression for k. These derivatives are with respect to x, i.e. the mapped coordinates,
while the shape functions are expressed in terms of parent coordinates, s. To carry out the
differentiation we need the inverse of the jacobian matrix, i.e.,
∂Na ∂Na ∂si ∂Na −1
= = Jij . (9.38)
∂xj ∂si ∂xj ∂si
It is these mapped derivatives of the shape functions that should be inserted into the
B matrix and used in Eqs. 9.27 and 9.29, if the element of interest is not the same as the
parent element.
9.9.1 Example Calculation of Jacobian for a 4 Noded Element
Consider the 4 noded (linear) element in Fig. 9.10. Writing s1 and s2 as the coordinates
in the parent space, the linear shape functions are given by,
1
N1 (s1 , s2 ) = (1 − s1 )(1 − s2 )
4
1
N2 (s1 , s2 ) = (1 + s1 )(1 − s2 )
4
1
N3 (s1 , s2 ) = (1 + s1 )(1 + s2 )
4
1
N4 (s1 , s2 ) = (1 − s1 )(1 + s2 )
4
175
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
N
EL
xi (s) = Na (s)Xia
a=1
where Xia indicate the coordinates of the nodes of the mapped element. (It is easy to
see, using the linear shape functions, that, for example, node 1 on the parent element,
with coordinates, (−1, −1), is mapped to node 1 on the mapped element with coordinates
(X11 , X21 ).)
The jacobian matrix is then
N EL N EL
∂xi ∂
Jij = = Na (s)Xia = Na,j Xia
∂sj ∂sj a=1 a=1
The jacobian of the transformation from s-space to x-space is the determinant of this matrix,
J = det (J) .
176
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
The derivative of the shape functions may be written down immediately, e.g. for the
linear shape functions,
1 1
N1,1 (s1 , s2 ) = − (1 − s2 ) ; N1,2 (s1 , s2 ) = − (1 − s1 )
4 4
1 1
N2,1 (s1 , s2 ) = (1 − s2 ) ; N2,2 (s1 , s2 ) = − (1 + s1 )
4 4
1 1
N3,1 (s1 , s2 ) = (1 + s2 ) ; N3,2 (s1 , s2 ) = (1 + s1 )
4 4
1 1
N4,1 (s1 , s2 ) = − (1 + s2 ) ; N4,2 (s1 , s2 ) = (1 − s1 )
4 4
Consider a regular square element shown in Fig. 9.12.
N
EL
Jij = Na,j Xia
a=1
Similarly,
As discussed in Section 9.5 the global finite element equation is written as (Eq. 9.16),
KU = f
Here f is the nodal force vector, which, as with the stiffness matrix, is made up of contribu-
tions from the individual elements, and in the same way we define a local force vector and
a global force vector.
It is seen from Eq. 9.15 that with applied tractions fisn on Γt the nodal force vector is
given by
fia = Na fisn dΓ. (9.39)
Γt
If body forces exist these are also included as part of the force vector. It may be shown,
starting from the virtual work expression, Eq. 9.4, that the contribution to the force vector
from body forces is given by,
178
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
fia = Na ρfib dV . (9.40)
V
Typical examples of body forces are weight, where |f | = g and the force acts towards
the ground, and centrifugal force, |f | = ω 2 r, where r is the distance to the axis of rotation
and the force acts along the radius, r. Note that Eq. 9.39 involves integration over the
element surface while Eq. 9.40 involves integration over the element volume. Therefore
the same procedure for integrating the stiffness matrix may be used to integrate the body
force term, but the integration of the traction term, Eq. 9.39, is different. It should also
be pointed out that a different order of integration can be used to integrate the traction
and/or body force term than is used for the element stiffness matrix. For example if the
traction over the element is constant (uniform applied stress) then single point quadrature,
over the element surface is sufficient.
If an applied load may be considered to be a concentrated force, then the force vector
is determined directly, i.e.
fia = Fi . (9.41)
where Fi are the components of the applied force at the node a.
where ε0kl are the initial strains (independent of the deformation). A common example of
initial strains are those generated due to a change in temperature. In this case
where α is the coefficient of thermal expansion, ΔT is the increase in temperature and δkl
is the Kronecker delta.
Inserting Eq. 9.42 into the virtual work equation we get:
0
cijkl (εkl − εkl )Na,j dV = Na fisn dΓ
V Γ
t
sn
⇒ cijkl εkl Na,j dV = Na fi dΓ + cijkl ε0kl Na,j dV ,
V Γt V
179
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
i.e., the equivalent nodal force due to the internal strain is given as,
a
fi = cijkl ε0kl Na,j dV .
V
The next stage in the analysis is the assembly of the global system matrices. Each
element will make a contribution to the global stiffness and force matrices. This is best
illustrated by a simple example. Consider a two element truss illustrated in Fig. 9.13. Each
node has two potential degrees of freedom.
F
⎡ ⎤⎡ 1 ⎤ ⎡ 1 ⎤
K1111 K1121 K1112 K1122 K1113 K1123 U1 f1
⎢ K2111 K2121 K2112 K2122 K2113 K2123 ⎥ ⎢ U21 ⎥ ⎢ f21 ⎥
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎢ K1211 K1221 K1212 K1222 K1213 K1223 ⎥ ⎢ U12 ⎥ ⎢ f12 ⎥
⎢ ⎥⎢ ⎥ = ⎢ ⎥. (9.43)
⎢ K2211 K2221 K2212 K2222 K2213 K2223 ⎥ ⎢ U22 ⎥ ⎢ f22 ⎥
⎣ ⎦⎣ 3 ⎦ ⎣ 3 ⎦
K1311 K1321 K1312 K1322 K1313 K1323 U1 f1
K2311 K2321 K2312 K2322 K2313 K2323 U23 f23
The term Kiakb in the stiffness matrix implies the stiffness at node a in the degree of
freedom i due to movement of node b in degree of freedom k. The stiffness matrix will
be symmetric, Kiakb = Kkaib , and the diagonal terms will be positive or zero (this is true
in general, not just for this example). Since node 2 is common to two elements (see Fig.
9.13) , there will be a contribution from element 1 and element 2 to all terms which have
1 2
a and b = 2, e.g. K1212 = k1212 + k1212 . Note also that since node 3 and node 1 are not
connected, all terms of the form Ki3k1 = Ki1k3 = 0. For this problem assembly of the
global stiffness matrix is relatively trivial, but the principle is the same for all FE meshes.
The assembly of the force vector is also trivial in this case. The only force is a con-
centrated force at node 2 in the 2 direction. Therefore f22 = F and all other nodal forces
180
9. INTRODUCTION TO THE FINITE ELEMENT METHOD
are zero. There are nodal reaction forces at node 1 and 2, but as will be seen in the next
section, these are not included in the final system of equations.
Steps 1, 2, 9 and usually 3 are carried out by the user. Steps 4 to 8 and sometimes 3
(automatic meshing) are carried out by the finite element program.
181