Sei sulla pagina 1di 127

M.

Mirzaei, Elasticity

-1-

Theory of

Elasticity
by Majid Mirzaei
These lecture notes have been prepared as supplementary material for a graduate level course on Theory of Elasticity. You are welcome to read or print them for your own personal use. All other rights are reserved. No originality is claimed for these notes other than selection, organization, and presentation of the material.

The major references that have been used for preparation of the lecture notes are as follows:

Timoshenko & Goodier, "Theory of Elasticity" Boresi & Chong, "Elasticity in Engineering Mechanics" Phillip L. Gould, "Introduction to Linear Elasticity" Provan, J.W., "Stress Analysis Lecture Notes"

M. Mirzaei, Elasticity

-2-

Theory of Elasticity Lecture Notes


Majid Mirzaei, PhD Associate Professor Dept. of Mechanical Eng., TMU mmirzaei@modares.ac.ir http://www.modares.ac.ir/eng/mmirzaei

Chapter 1 Field Equations of Linear Elasticity


In many engineering applications a pointwise description of the field quantities, such as stress and strain, is required throughout the body. The distinctive feature of the theory of elasticity, compared to the alternative approaches like the strength of materials, is that it provides a consistent set of equations which may be solved (using the advanced techniques of applied mathematics) to obtain a unique pointwise description of the distribution of the stresses, strains, and displacements for a particular loading and geometry.

Tensors, Indicial Notation


In the theory of elasticity we usually deal with physical quantities which are independent of any particular coordinate system that may be used to describe them. At the same time, these quantities (tensors) are usually specified in a particular coordinate system by their components. Specifying the components of a tensor in one coordinate system determines the components in any other system. The physical laws of Elasticity are in turn expressed by tensor equations. Because tensor transformations are linear and homogeneous, such tensor equations, if they are valid in one coordinate system, are valid in any other coordinate system. This invariance of tensor equations under a coordinate transformation is one of the principal reasons for the usefulness of tensor methods in the theory of Elasticity. In a three-dimensional Euclidean space (such as ordinary physical space) the number of components of a tensor is 3N, where N is the order of the tensor. Accordingly a tensor of order zero is specified in any coordinate system in three-dimensional space by one component. Tensors of order zero are called scalars. Physical quantities having magnitude only are represented by scalars. Tensors of order one have three coordinate components in physical space and are known as vectors. Quantities possessing both magnitude and direction are represented by vectors. Second-order tensors are called dyadics. Higher order tensors such as triadics, or tensors of order three, and tetradics, or tensors of order four can also be defined.

M. Mirzaei, Elasticity

-3-

In these notes we use the indicial notation for presentation of tensorial quantities and equations. Under the rules of indicial notation, a letter index may occur either once or twice in a given term. When an index occurs unrepeated in a term, that index is understood to take on the values 1,2,.. .,N where N is a specified integer that determines the range of the index. Unrepeated indices are known as free indices. The tensorial rank of a given term is equal to the number of free indices appearing in that term. Also, correctly written tensor equations have the same letters as free indices in every term. In ordinary physical space a basis is composed of three, non-coplanar vectors, and so any vector in this space is completely specified by its three components. Therefore the range on the index of ai (which represents a vector in physical space) is 1,2,3. Accordingly the symbol ai is understood to represent the three components a1, a2, a3. For a range of three on both indices, the symbol Aij represents nine components (of the second-order tensor (dyadic) A). When an index appears twice in a term, that index is understood to take on all the values of its range, and the resulting terms summed. In this so-called summation convention, repeated indices are often referred to as dummy indices because their replacement by any other letter not appearing as a free index does not change the meaning of the term in which they occur. In general, no index occurs more than twice in a properly written term.

Kinematics of Deformable Solids


Fig. 1.1 shows the undeformed configuration of a material continuum at time t = t0 together with the deformed configuration at a later time t = t. It is useful to refer the initial and final configurations to separate coordinate systems as shown in the figure.

Deformed Configuration

t=t
undeformed Configuration

t=t0
C0

u
x2 x1 b x3

X2

R0

X3

X1
Fig. 1.1

M. Mirzaei, Elasticity

-4-

Suppose a material point is at position X in the undeformed solid, and moves to a position x when the solid is loaded. We may describe the deformation and motion of a solid by a mapping in the following form:

x = ( X, t )

(1-1)

Now we consider the two coordinate systems to be coincident, as shown in Fig. 1.2. The displacement of a material point is:

u(t ) = x(t ) X

(1-2)

C0

u R0 X
X1 x1 t
undeformed Configuration

x
b C

X2 x2

X3 x3

Deformed Configuration

Fig. 1.2

Next, consider two straight parallel lines on the reference configuration of a solid (see Fig. 1.3). If the deformation of the solid is homogeneous, the two lines remain straight in the deformed configuration, and the lines remain parallel.

dS1 dS2 ds1 ds 2

undeformed Configuration

Deformed Configuration

Fig. 1.3

M. Mirzaei, Elasticity

-5-

Furthermore, the lines stretch by the same amount, i.e.:


ds1 ds2 = dS1 dS2

(1-3)

The difference (ds ) 2 (dS ) 2 is used to define a measure of deformation, which occurs in the vicinity of the particles between the initial and final configurations.
(ds ) 2 = dx12 + dx22 + dx32 = dxi dxi = dx dx (dS ) 2 = dX 12 + dX 22 + dX 32 = dX i dX i = dX dX

(1-4)

u+du
dS

ds

undeformed Configuration

Deformed Configuration

Fig. 1.4 Referring to Fig. 1.4, we may further write:


dS + u + du = u + ds, du = ds dS

(1-5)

We also have,

(ds) 2 (dS ) 2 = dxi dxi dX i dX i and,


xi = xi ( X 1 , X 2 , X 3 ) dxi = xi x x dX 1 + i dX 2 + i dX 3 X 1 X 2 X 3

(1-6)

(1-7)

dxi = xi , j dX j = F dX

M. Mirzaei, Elasticity

-6-

where the comma denotes differentiation with respect to a spatial coordinate. In the above F is known as deformation gradient tensor. Next, we define
u = u1e x1 + u2e x2 + u3e x3 = ui e xi

(1-8)

and proceed as follows:


(ds ) 2 (dS ) 2 = xi , j dX j xi , k dX k dX i dX i = (F dX) (F dX) dX dX = ( xi , j xi , k ij ik ) dX j dX k = ( X i + ui ), j ( X i + ui ), k jk dX j dX k = ( ij + ui , j ) ( ik + ui , k ) jk dX j dX k = ( u j ,k + uk , j + ui , j ui ,k ) dX j dX k = jk + u j , k + uk , j + ui , j ui ,k jk dX j dX k

(1-9)

Rewriting the indices in a more proper form, we get


(ds ) 2 (dS ) 2 = ( ui , j + u j ,i + uk ,i uk , j ) dX i dX j
L = 2 ij dX i dX j

(1-10)

L in which we distinguish the Lagrangian strain tensor ij for characterization of the

deformation near a point.

ijL =

1 ( ui, j + u j ,i + uk ,iuk , j ) 2

(1-11)

Alternatively we may write


X i = X i ( x1 , x2 , x3 ) dX i = X i X X dx1 + i dx2 + i dx3 x1 x2 x3 (1-12)

dX i = X i , j dx j = F dx and proceed as follows:

M. Mirzaei, Elasticity

-7-

(ds ) 2 (dS ) 2 = dxi dxi X i , j dx j X i ,k dxk = dx dx (F dx) (F dx) = ( ij ik X i , j X i ,k ) dx j dxk = jk ( xi ui ), j ( xi ui ),k dx j dxk = jk ( ij ui , j ) ( ik ui , k ) dx j dxk = jk jk + uk , j + u j , k ui , j ui ,k dx j dxk = ( u j ,k + uk , j ui , j ui ,k ) dx j dxk

(1-13)

Replacing the dummy indices i and k, we have


(ds ) 2 (dS ) 2 = ( ui , j + u j ,i uk ,i uk , j ) dxi dx j
E = 2 ij dxi dx j

(1-14)

L This time, we define the Eulerian strain tensor ij to characterize the deformation near a

point.

ijE =

1 ( ui, j + u j ,i uk ,iuk , j ) 2

(1-15)

In practice, we need to make a number of assumptions to simplify the equations of linear elasticity. A major one is to assume that deformations are infinitesimal. In most practical circumstances it is sufficient to assume uk ,i uk , j << 1 We use this to define a linear measure of deformation in linear elasticity, and define the infinitesimal strain tensor as:

L ij = ijE = ij =

1 ( ui, j + u j ,i ) 2

(1 16)

In order to check the relationship between the above definition of strain and the conventional definition, we consider a one-dimensional case as follows:
L (ds ) 2 (dS ) 2 = 2 ij dX i dX j = 2 22 dX i dX i = 2 22 (dS ) 2

(ds ) 2 (dS ) 2 (ds + dS )(ds dS ) ds dS 22 = = = 2(dS ) 2 dS (ds + dS ) dS

(1-17)

M. Mirzaei, Elasticity

-8-

The linear strain tensor in Cylindrical and Spherical Coordinate Systems


Cylindrical
ur rr = r 2 r = 1 ur u u + r r r 1 u ur = + r r ur u z + z r u 1 u z 2 z = + z r u zz = z z 2 rz =

(1-18)

Spherical
ur rr = r 2 r = 1 ur u u + r r r 1 u ur = + r r 1 ur u u + r sin r r 1 u 1 u u cot + 2 = r sin r r 1 u ur u cot + + = r sin r r 2 r =

(1-19)

Compatibility
Conditions of compatibility, imposed on the components of strain, are necessary and sufficient to insure a continuous single-valued displacement field. The procedure is to eliminate the displacements between kinematic equations to produce equations with only strain as unknowns. For instance we may write:

ll ,mm = ul ,lmm mm ,ll = um ,mll


1 lm ,lm = (ul ,mlm + um ,llm ) 2 1 = (ul ,lmm + um ,mll ) 2 2 lm,lm = ll ,mm + mm ,ll (1-20)

M. Mirzaei, Elasticity

-9-

In general, we have

ij ,kl + kl ,ij = ik , jl + jl ,ik


which represents 81 equations, only six of them are independent.

(1-21)

11,22 + 22,11 = 212,12 22,33 + 33,22 = 2 23,23 33,11 + 11,33 = 2 31,31 12,13 + 13,12 23,11 = 11,23 23,21 + 21,23 31,22 = 22,31 31,32 + 32,31 12,33 = 33,12
(1-22)

Assignment 1:
Noting that = ( ),i ei and using indicial notation, prove the identities or find the right

hand side:
2 = = ,ii ( v ) = v + v 2 ( ) = 2 + 2 + 2 3 ( ) = 2 ( ) = ? ( ) = ?

M. Mirzaei, Elasticity

- 10 -

Kinetics of Deformable Solids


In general, two types of forces can be applied to a solid body. (i) External forces applied to its boundary (e.g. forces arising from contact with another solid or fluid pressure). (ii) Externally applied forces that are distributed throughout a body (e.g. gravitational, magnetic, and inertial forces).

C0

X2

R0
X1 t
undeformed Configuration

R
b C
Deformed Configuration

X3

Fig. 1.5 The stress vector or surface traction t at a point represents the force acting on the surface per unit area and can be defined as:

t = lim

dP dA0 dA

(1-23)

in which dA is an element of area on a surface subjected to a force dP (see Fig. 1.6).

dP

dA
Fig. 1.6 The body force vector denotes the external force acting on the interior of a solid per unit volume and can be defined as: b = lim
dP dV 0 dV

(1-24)

M. Mirzaei, Elasticity

- 11 -

in which dV denotes an infinitesimal volume element subjected to a force dP (Fig. 1.7).

dP

dV
Fig. 1.7

Internal forces induced by external loading


The solid body shown in Fig. 1.8 is subjected to a balanced system of externally applied forces. As a result, internal forces are developed in order to keep the body together. Suppose that the body is cut into two parts. The force Pn represents the resultant force that acts on the two faces in order to keep the two parts in load equilibrium.

n
t Pn t t

R
b C e2

R
-n
e1

e3
Fig. 1.8
The internal traction vector Tn represents the force per unit area acting on a plane with normal vector n inside the deformed solid an can be defined as:
Tn = lim
dPn dA

(1-25)

dA0

M. Mirzaei, Elasticity

- 12 -

in which dA is an element of area in the interior of the solid, with normal n (see Fig. 1.9).

dPn n

dA
Fig. 1.9

The components of Cauchy stress in a given basis can be visualized as the tractions acting on planes with normals parallel to each basis vector, as depicted in Fig. 1.10.

e2 T2
22 21 12 32

23

T3
11 33 31 13

e3 T1

e1

Fig. 1.10

Here, we may write:


T1 = 11e1 + 12e 2 + 13e3 T2 = 21e1 + 22e 2 + 23e3 T3 = 31e1 + 32e 2 + 33e3 or Ti = ij e j

(1-26)

M. Mirzaei, Elasticity

- 13 -

In order to find the components of the traction vector on an arbitrary plane (represented by n) we impose the equilibrium on a tetrahedral element, as shown in Fig. 1.11.

e2 dAn dA1 -e1 T(-e1) dA1 e3 n Tn dAn e1

Fig. 1.11 1 Tn dAn Ti dAi + f hdAn = 0, 3 h 0 Tn dAn Ti dAi = 0 dAi = dAn Cos(n, ei ) Tn dAn Ti dAnn ei = 0 Tn = Ti ei , Ti = ij e j , n ei = ni Ti ei ij e j ni = 0 , Ti ei = ij e j ni = ji ei n j n

(1-27)

Accordingly, we obtain the following expressions which are called: the stress boundary equations
Ti = ji n j

(1-28)

Principle Stresses
For practical purposes, it is convenient to break down the traction vector into normal and shear components as follows:

nn = Tn n = Ti ei n = Ti ni = ji n j ni

(1-29)

M. Mirzaei, Elasticity

- 14 -

and,

ns = (TTi nn 2 ) 2 i

(1-30)

In order to find the extremum values of the normal components of the stress tensor (the principle stresses) we may write:

nn = ji n j ni

(1-31)

= ji n j ni (ni ni 1) where is the lagrangian multiplier. Next, we write:


nn = ji n j ni = 0 ni = ji n j ij n j = 0 = ( ij ij )n j = 0 n

(1-32)

which implies that

11 12 13 21 22 23 = 0 31 31 33
The characteristic polynomial is

(1-33)

3 I I 2 + I II I III = 0

(1-34)

The roots of the above characteristic polynomial are the eigenvalues of our problem, or the principle stresses. The invariants of the stress tensor are defined as: I I = ii 1 I II = ( ii jj ij ji ) 2 I III = ij

(1-35)

Finally, for each eigenvalue, we may find the eigenvectors which are the principle stress directions.

M. Mirzaei, Elasticity
1 1 1 ( ij ij )n j = 0 n1 , n1 , n3 2 2 2 2 ( ij ij )n j = 0 n12 , n2 , n3

- 15 -

(1-36)

3 ( ij ij )n j = 0 n , n , n
3 1 3 2

3 3

Assignment 2:
For the given stress tensor, determine the maximum shear stress and show that it acts in the plane which bisects the maximum and minimum stress planes.
0 5 0 ij = 6 12 1

Conservation of Linear Momentum


In order to find the governing equations for distribution of the stress tensor within a body, we start by considering an arbitrary material region with volume V and surface S. Based on the general concept of conservation of linear momentum, we may write:
&& TdA + fdV = udV
S V V

(1-37)

which, using (1-28), can be written as

ji

&& n j dA + f i dV = ui dV
V V

(1-38)

Now we use the Divergence theorem of Gauss and write

G ndA = GdV
S V

ji , j

&& dV + f i dV ( ji , j + f i ) dV = ui dV
V V V

(1-39)

Since we are dealing with an arbitrary volume, the three equations of motion can be derived from (1-39) as

&& ji , j + fi = ui

(1 40)

M. Mirzaei, Elasticity

- 16 -

The static form of the above equations (usually known as the equilibrium equations) is

ji , j + f i = 0

(1-41)

The Equations of Motion in Cylindrical and Spherical Coordinate Systems can be written as:
rr 1 r rz rr && + + + + f r = ur r r r z r 1 z 2 r && + + + + f = u r r r z rz 1 z zz rz && + + + + f z = uz r r r z and
rr 1 r 1 r 1 && + + + (2 rr r cot ) + f r = ur r r sin r r r 1 1 1 && + + + [3 r + ( ) cot ] + f = u r r sin r r r 1 1 1 && + + + (3 r + 2 cot ) + f = u r r sin r r

(1-42)

(1-43)

Conservation of Angular Momentum


Once again, let us consider a material region with volume V and surface S. Based on the general concept of conservation of angular momentum we may write:
&& (X T)dA + (X f )dV = (X u)dV
S V V

ijk

X iT j dA + ijk X i f j dV = 0
V

(1-44)

which, using (1-28) and the Divergence theorem of Gauss, can be written as:

M. Mirzaei, Elasticity

- 17 -

S V

ijk

X i ( lj nl )dA + ijk X i f j dV = 0
V

ijk

( X i lj ),l dV + ijk X i f j dV = 0
V

ijk [( X i lj ),l + X i f j ]dV = 0


ijk

(1-45)

[ X i ,l lj + X i lj ,l + X i f j ]dV = 0

Next, we use (1-41) to simplify the above as:


ijk X i,l lj + X i lj ,l + f j dV = 0
=0

ijk il lj dV = 0

(1-46)

Since we are dealing with an arbitrary volume, we can write:

ijk ij = 0
ij = ji

(1-47)

which shows that the stress tensor is symmetric.

Constitutive Equations of Linear Elasticity


For an elastic body which is gradually strained at constant temperature, the components of stress can be derived from the strain energy density , which is a quadratic function of the strain components.

ij =

eij

(1-48)

Accordingly, we may write the most general form of the Hookes Law as:

ij = Eijkl ekl

(1-49)

M. Mirzaei, Elasticity

- 18 -

in which Eijkl represents 81 components. Due to the symmetry of the stress and strain tensors and also the elastic coefficient matrix, the number of independent elastic constants reduces to 21.
11 E1111 22 33 = 12 23 31 E1122 E2222 E1133 E2233 E3333

2 E1112 2 E2212 2 E3312 2 E1212

2 E1123 2 E2223 2 E3323 2 E1223 2 E2323

2 E1131 11 2 E2231 22 2 E3331 33 2 E1231 12 2 E2331 23 2 E3131 31

(1-50)

which can be simplified to:


11 C11 C12 C22 22 33 = 12 23 31 C13 C14 C23 C24 C33 C34 C44 C15 C25 C35 C45 C55 C16 11 C26 22 C36 33 C46 12 C56 23 C66 31

(1-51)

The above expression represents the constitutive equations for an anisotropic elastic material. Most engineering materials show some degree of isotropic behavior as they possess properties of symmetry with respect to different planes or axes. We start with the plane x2x3 as the plane of symmetry, which implies that the x1 axis can be reversed as shown in Fig. 1.12.

x3

x3 x1
x2

x2

x1
Fig. 1.12: single plane symmetry

M. Mirzaei, Elasticity

- 19 -

This corresponds to a coordinate transformation with the direction cosines shown in table 1.1.
axes

x1 x2 x3

x1 -1

x2
1

x3

Table 1.1 Using the transformation ij = ik jl kl , in which ik and jl are the direction cosines, we find that: 11 = 11 12 = 12 22 = 22 23 = 23 33 = 33 31 = 31 (1-52)

Similarly, we can show that: 11 = 11 12 = 12 22 = 22 23 = 23 33 = 33 31 = 31 (1-53)

Hence, we may write: 11 11 22 22 33 33 = = 12 12 23 23 31 31 =


which results in:

11 22 33 12 23 31 11 22 33 12 23 31
(1-54)

M. Mirzaei, Elasticity

- 20 -

11 C11 C12 C22 22 33 = 12 23 31

C13 C23 C33

C14 C24 C34 C44

C15 C25 C35 C45 C55

C16 11 C26 22 C36 33 C46 12 C56 23 C66 31

(1-55)

As the constants should not change with the transformation, we must have C14, C16, C24, C26, C34, C36, C45, and C56 = 0. Such materials are called monoclinic and need 13 constants to describe their elastic properties. The double symmetry about the x1 and x2 axes will lead to further reduction of the constants down to 9, which represents the orthotropic material (Note that for 2D problems we need 4 elastic constants to describe the behavior of an orthotropic material). axes

x1 x2 x3

x1 -1

x2
-1

x3

We may have an additional simplification by considering the directional independence in elastic behavior by interchanging x2 with x3, and x1 with x2. Such materials are called cubic with three independent constants. axes

x1 x2 x3

x1 1
0 0

x2 0
0 1

x3 0
1 0

axes

x1 x2 x3

x1 0
1 0

x2 1
0 0

x3 0
0 1

Finally, we may consider rotational independence in material property and define the constitutive equations for an isotropic elastic material with only two independent constants: axes

x1 x2 x3

x1 1
0 0

x2
0 cos - sin

x3 0
sin cos

M. Mirzaei, Elasticity
11 2 + 22 33 = 12 23 31

- 21 0 0 0 2 0 0 0 0 2 0 11 0 22 0 33 0 12 0 23 2 31

2 + 2 +

(1-56)

in which and are called the Lam constants. The indicial form can be written as:

ij = 2 ij + ij kk
We may invert the above equation to express the strains in terms of stresses:

(1-57)

ij =

1 ij ij kk 2 2 (2 + 3 )

(1-58)

The Lam constants are quite suitable from mathematical point of view, but they should be related to the Engineering elastic constants obtained in the laboratory (like E and ) as well. Table 1.2 shows the relationships between different elastic constants.

E , E (1 + )(1 2 )
E 2 (1 + ) E 3 (1 2 )

, 2 (1 2 )

E, ( E 2 ) 3 E

k , 3k (1 + )

+ ( 3 + 2 ) ( + ) 2 ( + )
2 3

2 (1 + ) 3 (1 2 )
2 (1 + )

E 3 ( 3 E )
E
E 1 2

3k (1 2 ) 2 (1 + )
k
3k (1 2 )

Table 1.2

M. Mirzaei, Elasticity

- 22 -

Formulation and Solution Methods of Elasticity Problems


In order to find the fifteen unknowns (three components of displacement, six components of strain, and six components of stress) for an elasticity problem, we need fifteen equations as follows:

ij =

1 ( ui, j + u j ,i ) 2

(I) (II)
or

ji , j + f i = 0

ij = 2 ij + ij kk ij =

1 ij ij kk 2 2 (2 + 3 )

(III)

along with the compatibility and stress boundary equations:

ij ,kl + kl ,ij = ik , jl + jl ,ik


Ti = ji n j

(IV) (V)

Classical Displacement Formulation


We start by substituting Eqs.(I) into the first form of Eqs.(III) to eliminate the strains:

ij = ij uk ,k + (ui , j + u j ,i )

(1-59)

Next, we substitute the above expression for the stresses in Eqs.(II) which resulys in:

ij uk ,ki + (ui , ji + u j ,ii ) + f j = 0 uk ,kj + (ui , ji + u j ,ii ) + f j = 0


Finally, using i for the dummy index in all terms we write: (1-60)

u j ,ii + ( + )ui ,ij + f j = 0

(1-61)

M. Mirzaei, Elasticity

- 23 -

or alternatively:

( + G )u j ,ij + Gui , jj + fi = 0

(1-62)

Equations (1-62) are three equilibrium equations in terms of displacements. They are called Navier Equations and constitute the classical displacement formulation.

Navier Displacement Equations in Cylindrical and Spherical Coordinates Cylindrical:


I e 1 z 2G + fr = 0 r z r I ( + 2G ) e 2G r z + f = 0 r r z ( + 2G ) ( + 2G ) where 1 (rur ) 1 u u z + + z r r r u u 1 u z u 2r = ; 2 = r z r z z r 1 (ru ) ur 2 z = r r I e = eii = u = I e (r ) r 2G + fz = 0 z r

(1-63)

M. Mirzaei, Elasticity

- 24 -

Spherical:
( + 2G ) I e 2G ( sin ) + fr = 0 r r sin I 2G 1 (r ) (r ) ( + 2G ) e + f = 0 r r r sin ( + 2G ) I e 2G (r ) (rr ) + f = 0 r sin r r where I e = eii = u = 2r = 1 ( r 2 ur ) 1 (u sin ) 1 u + + r 2 r r sin r sin

(1-64)

1 (u sin ) u 1 ur 1 (ru ) ; ; 2 = r sin r r r sin 1 (ru ) ur 2 = r r

Classical Force Formulation


We start by substituting the RHS of the second form of Eqs. (III) for strains in Eqs. (IV) to produce 81 stress compatibility equations as follows:

ij ,kl + kl ,ij ik , jl jl ,ik =

( 1 +
ij

tt , kl

kl tt ,ij ik tt , jl jl tt ,ik )

(1-65)

However, only six of the above equations are independent:

ij ,kk + kk ,ij ik , jk jk ,ik =

( 1 +
ij

tt , kk

kk tt ,ij ik tt , jk jk tt ,ik )

(1-66)

In general, the above expression represents nine equations with free indices i and j. Now we substitute the equilibrium equations (II) in the above and simplify to have:

2 ij +

1 mm ,ij = fi , j + f j ,i + ij f m,m 1 + 1

(1-67)

M. Mirzaei, Elasticity

- 25 -

which are known as Beltrami-Michell compatibility relations and constitute the classical force formulation.

Assignment 3:
For a hollow sphere under internal and external pressures find the stress, strain and displacement distributions.

Beltrami-Michell Equations in Cylindrical Coordinates:


( fr ) 1 rr 1 2 rr 2 rr 1 2 It 1 ( rf r ) 1 ( f ) ( f z ) + + = + + 2 r + 2 2 2 2 r r r r r 1 + r 1 r r z z r 2 2 2 ( f ) 1 1 1 It 1 ( rf r ) 1 ( f ) ( f z ) + + 2 r r + r2 2 + z2 + 1 2 = 1 r r r r + r z 1 zz r r r r 1 z r r r r 1 rz r r r r
2 2 1 2 It 1 zz zz + 2 + + = 2 2 2 1 + z 1 z r 2 2 1 z z + 2 + 2 z 2 r

( fz ) 1 ( rf r ) 1 ( f ) ( f z ) + + 2 r z z r r ( f ) ( f z ) 1 2 It + = + 1 + z z

2 2 ( f z ) ( fr ) 1 2 It 1 rz rz + 2 + + = + 2 2 z 1 + r z z r r ( f r ) ( f ) 1 r 1 2 r 2 r 1 2 It r r + r 2 2 + z 2 + + r = + r r r 1

(1-68)

It = rr + + zz

M. Mirzaei, Elasticity

- 26 -

Beltrami-Michell Equations in Spherical Coordinates:


1 2 rr r r r 2 r rr 1 + 2 sin r sin 2 rr 1 1 2 It + 2 2 + = 2 1 + r 2 r sin

2 ( fr ) 1 ( r fr ) 1 1 ( f ) 2 2 + ( sin f ) + 1 r r r sin r sin r

1 2 It 1 2 1 1 sin + 2 2 = r + 2 + r 2 r r r sin r sin 2 1 + 2 2 ( f ) 1 1 ( f ) 1 ( r fr ) 2 2 + ( sin f ) + r r sin r sin 1 r 2 1 2 1 1 1 2 It + = r + 2 sin + 2 2 r r sin r sin 2 1 + 2 r 2 r


2

2 ( f ) 1 1 ( f ) 1 ( r fr ) 2 2 + ( sin f ) + r r sin r sin 1 r

1 2 r r r 2 r 1 2 r r r r 2 r 1 2 r r r 2 r r

( f ) ( f ) 2 1 1 1 2 It + 2 + = + sin + 2 2 2 r sin 1 + r sin 2 2 ( f r ) ( f ) r r 1 1 1 It + = + + 2 sin + 2 2 2 r sin r 1 + r r sin ( f r ) ( f ) r 2 r 1 1 1 2 It + = + + 2 sin + 2 2 2 1 + r r sin r r sin

(1-69)

It = rr + +

M. Mirzaei, Elasticity

- 27 -

Chapter 2: Some Representative Boundary Value Problems


Deformation of a Rod Standing in a Gravitational Field
As a first example in solving elasticity boundary-value problems, we consider a long rod that stands freely in a gravitational field (see Fig. 2.1).

X
Fig. 2.1
We start with the stress boundary conditions.

Ti = ij n j
On the top surface we have: Ti (0, 0, 0), ni (0, 0,1) T1 = 11n1 + 12 n2 + 13 n3 0 = 0 + 0 + 13 13 = 0 T2 = 21n1 + 22 n2 + 23 n3 0 = 0 + 0 + 23 23 = 0 T3 = 31n1 + 32 n2 + 33 n3 0 = 0 + 0 + 33 33 = 0 On the lateral surfaces we have:

(2-1)

(2-2)

M. Mirzaei, Elasticity Ti (0, 0, 0), ni (n1 , n2 , 0) 0 = 11n1 + 12 n2 + 0 0 = 21n1 + 22 n2 + 0 0 = 31n1 + 32 n2 + 0 On the bottom surface we have: Ti (0, 0, gL), ni (0, 0, 1) 0 = 0 + 0 13 0 = 0 + 0 23

- 28 -

(2-3)

(2-4)

gL = 0 + 0 33 33 = gL
We now turn our attention to the overall equilibrium:

ij , j + fi = 0 11,1 + 12,2 + 13,3 + f1 = 0 21,1 + 22,2 + 23,3 + f 2 = 0 31,1 + 32,2 + 33,3 g = 0


Near the bottom surface we have:

(i ) (ii ) (iii )

(2-5)

13 = 0, 23 = 0,
Thus, using (2-5 iii), we can write:

33,3 g = 0 33 = gz + f (1, 2)
B.C. : at z = 0 we have 33 = gL f (1, 2) = gL 33 = gz gL = g ( z L) (2-6)

We may therefore assume that the distribution of the stress tensor has the following form:
0 0 0 0 ij = 0 0 0 0 g (L z)

(2-7)

M. Mirzaei, Elasticity

- 29 -

Note that this form satisfies all the equilibrium equations and the boundary conditions of the problem. Using the following constitutive equations:

ij =

1 (1 + ) ij ij kk E

(2-8)

we can obtain the distribution of the strain tensor throughout the rod.
g 0 0 E (L z) g ij = 0 (L z) 0 E g 0 0 (L z) E

(2-9)

Since the strain components are linear in Z, the compatibility equations are all satisfied. Now we use the kinematic equations to obtain the displacement field:

ij =

1 ( ui, j + u j ,i ) 2

(2-10)

We start with the normal strain components:

1 ( u1,1 + u1,1 ) = u1,1 u1,1 = g ( L z ) 2 E g ( L z ) x + f1 ( y, z ) u1 = E 1 g 22 = ( u2,2 + u2,2 ) = u2,2 u2,2 = (L z) 2 E g ( L z ) y + f 2 ( x, z ) u2 = E 1 g 33 = ( u3,3 + u3,3 ) = u3,3 u3,3 = (L z) 2 E g 1 2 u3 = Lz z + f 3 ( x, y ) 2 E

11 =

(i)

(ii)

(2-11)

(iii)

Now we may invoke the displacement boundary conditions:

M. Mirzaei, Elasticity at x = 0, u1 = 0, f1 = 0 at y = 0, u2 = 0, f 2 = 0

- 30 -

In order to find f3, we calculate the shear strain components and use (i, iii) of (2-11) to write:

13 =

1 ( u1,3 + u3,1 ) u1,3 + u3,1 = 0 2 g g g 2 x + f 3,1 = 0 f 3,1 = x f3 = x + h1 ( y ) E E 2E

(2-12)

Now, using the expressions (ii, iii) of Eqs. (2-11) along with (2-12), we may write:

23 =

1 ( u2,3 + u3,2 ) u2,3 + u3,2 = 0 2 g g g 2 y + h1,2 = 0 h1,2 = y h1 = y +C E E 2E

(2-13)

Hence, the final form of the displacement equations can be written as:

u1 = u2 =

g g
E E

(L z) x (L z) y
(2-14)

u3 =

1 2 g 2 g 2 x + y +C Lz z + E 2 2E 2E at x = 0, y = 0, z = 0, we have u = 0 C = 0

u3 =

1 2 g 2 g 2 x + y Lz z + E 2 2E 2E

M. Mirzaei, Elasticity

- 31 -

Rotating Discs
In this section we study the stresses induced in rotating discs. Initially, we assume that the thickness of the disc is small in comparison with its radius, so that the variation of radial and circumferential stresses over the thickness can be neglected. We also assume that zz = 0 .

Fig. 2.2
We may start by writing the equations of motion in cylindrical coordinates: rr 1 r rz rr && + + + + f r = ur r r r z r 1 z 2 r && + + + + f = u r r r z rz 1 z zz rz && + + + + f z = uz r r r z

(2-15)

Considering our Basic assumptions, and due to the symmetry of geometry and loading, we are left with only one equation: rr rr + + r 2 = 0 r r which can be rearranged into the following form: d (r rr ) + r 2 2 = 0 dr Now we may set:
r rr = F (r )

(2-16)

(2-17)

dF (r ) + r 2 2 dr

(2-18)

M. Mirzaei, Elasticity in which F(r) is a stress function.

- 32 -

The linear strain tensor in cylindrical coordinates is:


ur rr = r ij = 2 r = 1 ur u u + r r r 1 u ur = + r r ur u z + z r u 1 u z + 2 z = z r u zz = z z 2 rz =

(2-19)

which in this case reduces to: ur rr = r ij =

r = 0
u = r r

z = 0 zz = 0

rz = 0

(2-20)

We may also write the constitutive equations for our problem as:
1 (1 + ) ij ij kk E 1 1 rr = [ (1 + ) rr ( rr + ) ] = [ rr ] E E 1 = [ rr ] E

ij =

(2-21)

At this point, we will eliminate the displacements between our kinematic equations to obtain a compatibility equation.

d d u = r dr dr r

dur ur 1 dur ur 1 2 = = [ rr ] = r dr r r rdr r

(2-22)

Now we substitute the constitutive equation in the above compatibility equation: d 1 11 E ( rr ) = r E ( rr + rr ) dr d d 1 rr = (1 )( rr ) dr dr r

(2-23)

M. Mirzaei, Elasticity

- 33 -

Rewriting the above equation in terms of our stress function F(r), we will have:
r rr = F (r ), dF (r ) + r 2 2 , F (r ) F dr d 2F dF F (1 + ) F dF + 2 2 r + 2 = 2 r 2 2 2 dr r dr r r r dr

(2-24)

d 2F dF +r F = (3 + ) 2 r 3 r 2 dr dr
2

This is a nonhomogeneous, second order linear, ordinary differential equation with variable coefficients and has the following general solution: 1 (3 + ) 2 r 3 F (r ) = A1r + A2 r 8 Accordingly, the resulting stresses will be: (2-25)

rr = A1 +

A2 (3 + ) 2 r 2 8 r2 A2 (1 + 3 ) 2 r 2 = A1 2 8 r

(2-26)

The coefficients A1 and A2 can be found by imposing the appropriate boundary conditions. We consider the solution for the Solid Disc, and a Disc with stress-free circular hole. For the solid disc we must have A2=0 to account for the finite stresses at the center. The second B.C. is:

Ti = ij n j Ti (0, 0, 0), ni (1, 0, 0) at r = a Tr = rr nr + r n + rz nz 0 = rr nr + 0 + 0 rr = 0 A1


(2-27)

( 3 + ) 2 a 2 =
8

Accordingly, the final form of the stresses will be:

M. Mirzaei, Elasticity (3 + ) 2 2 2 rr = (a r ) 8 (3 + ) 2 2 (1 + 3 ) 2 2 = a r 8 8

- 34 -

(2-28)

The displacement field can be obtained as follows: ur = r = = r ( rr ) E

r 2 ( 3 2 2 ) a 2 + ( 2 1) r 2 8E

(2-29)

For a disc with stress-free circular hole, we have the following conditions at the outside radius:

Ti = ij n j Ti (0, 0, 0), ni (1, 0, 0) at r = a Tr = rr nr + r n + rz nz 0 = rr nr + 0 + 0 rr = 0


2 2 A2 ( 3 + ) a A1 = 2 + a 8

(2-30)

At the inside radius we have:

Ti = ij n j Ti (0, 0, 0), ni (1, 0, 0) at r = b Tr = rr nr + r n + rz nz 0 = rr nr + 0 + 0 rr = 0


2 2 A2 ( 3 + ) b A1 = 2 + b 8

(2-31)

Solving (2-30) and (2-31) for A1 and A2 , we will have:


A1

8 ( 3 + ) 2 a 2b 2 A2 = 8

( 3 + ) 2 =

(a

+ b2 )

(2-32)

M. Mirzaei, Elasticity

- 35 -

Hence the stress equations become:

rr =

(3 + ) 2 2 a 2b 2 a + b2 2 r 2 8 r

(3 + ) 2 2 a 2b 2 (1 + 3 )r 2 a + b2 + 2 = 8 r (3 + )

(2-33)

Furthermore, we may write the maximum radial and hoop stresses as:

rr

max

max

(3 + ) 2 2 (a b) 8 (3 + ) 2 2 2(1 )b 2 = 2a + 8 (3 + )

(2-34)

Also notice that for a small hole as b 0 , we will have:

hole b0 solid r 0

(3 + ) 2 = ( 2a 2 ) 8 (3 + ) 2 2 a = 8
hole

(2-35)

b 0

= 2

solid r 0

The conclusion is that a stress concentration factor of two exists for the circular hole in our rotating disc.

Alternative solutions
There are alternative solutions to the rotating disc problem, which are in fact quite simpler than the solution presented above. For instance, we may extract our previous differential equation in terms of the stress function F(r) directly from the BeltramiMichell equations in cylindrical coordinates (see Chapter 1). We may also start with the Navier displacement equations in cylindrical coordinates:
( + 2G ) ( + 2G ) I e 1 z 2G r z r + fr = 0 (2-36)

I e 2G r z + f = 0 r r z I (r ) r ( + 2G ) e 2G + fz = 0 z r

M. Mirzaei, Elasticity In the above equations we have:


1 (rur ) 1 u u z + + r r r z u u 1 u z u 2r = ; 2 = r z r z z r I e = eii = u =

- 36 -

1 (ru ) ur ; 2z = r r

(2-37)

Considering our basic assumptions and the symmetry of geometry and loading, the above equations will simplify to the following single equation: r2 d 2ur du 1 2 + r r u = 2 r 3 2 dr dr E (2-38)

The general solution can be shown to be:


ur = A 1 2 2 2 3 1 r r (1 ) A1r (1 + ) 2 8 E r

(2-39)

which upon substitution into the kinematic and constitutive equations will give the required expressions for the radial and hoop stresses.

Assignment 4:
Find the stresses in a rotating disc of variable thickness h=f(r). Hint: use the following form of the equilibrium equation: d ( hr rr ) h + 2 hr 2 = 0 dr

M. Mirzaei, Elasticity

- 37 -

Saint-Venant Torsion
This section presents a general solution to the torsion of prismatic bars, known as SaintVenant torsion problem. This solution represents a classic example of the so called semiinverse approach, in which a displacement field is initially assumed and then it is shown that it satisfies the equilibrium equations and the boundary conditions.

X2

X2
X1

R C
u r u2 u1

X3
Fig. 2.3

X1

Based on a geometric study of the schematic shown in Fig. 2.3, we may define the two in-plane components of displacement vector for an arbitrary point in the cross section of the bar by:
u1 = r cos( + ) r cos u2 = r sin( + ) r sin u1 = r cos cos r sin sin r cos u2 = r sin cos + r cos sin r sin

(2-40)

For very small values of , we may assume that cos 1 , sin , and rewrite the displacements as:
u1 = r cos r sin r cos u2 = r cos + r sin r sin u1 = r sin u2 = r cos

(2-41)

which using the expressions r cos = x1 , rsin = x2 , results in,

M. Mirzaei, Elasticity

- 38 -

u ~ ( x2 , x1 ) With the definition of as the torsion-angle per unit length, we have:

(2-42)

= x3
Hence, the in-plane displacements can be written as follows: u ~ (x2 x3 , x1 x3 ) Next, we assume that the third component of displacement can be defined as: u3 = ( x1 , x2 )

(2-43)

(2-44)

(2-45)

where is called the Warping Function. Finally, the assumed displacement field can be summarized as: ui ~ (x2 x3 , x1 x3 , ) Accordingly, the strain field becomes: (2-46)

ij =

1 ( ui, j + u j ,i ) 2 1 0 0 ( x2 + ,1 ) 2 0 1 x + ij ~ ( 1 ,2 ) 2 0

(2-47)

and the stresses become:

ij = 2 ij + ij kk
0 0 G (,1 x2 ) ij ~ 0 G (,2 + x1 ) 0 Now we turn our attention to the equilibrium equations: (2-48)

M. Mirzaei, Elasticity

- 39 -

11,1 + 12,2 + 13,3 = 0 21,1 + 22,2 + 23,3 = 0 31,1 + 32,2 + 33,3 = G (,11 + ,22 )
which can be summarized as: 2 = 0 in R Now we will examine the boundary conditions.
T = i ni T (0, 0), ni (n1 , n2 , 0) 0 = 11n1 + 12 n2 + 0 0 = 21n1 + 22 n2 + 0 T3 = 3i ni = 0 0 = 31n1 + 32 n2 + 0 = G (,1 x2 ) n1 + (,2 + x1 ) n2 = 0

(2-49)

(2-50)

(2-51)

,1n1 + ,2 n2 = x2 n1 x1n2

We may simplify the above boundary conditions based on the geometric considerations depicted in Fig. 2.4 as follows:
X2 n t X1

Fig. 2.4

M. Mirzaei, Elasticity dx1 dx = cos , n2 = 2 = sin dn dn dx dx n1 = 2 = cos , n2 = 1 = sin dt dt n1 =

- 40 -

(2-52)

Accordingly, the stress boundary condition defined in (2-51) may be rewritten as:
dx1 dx2 dx dx + = x2 2 + x1 1 dt dt x1 dn x2 dn d 1 dr 2 = dn 2 dt on C.

(2-53)

Note that the above is not yet in a suitable form because we have to deal with two variables, n and t. In order to solve this problem, we introduce the conjugate harmonic function : z = + i Using the following Cauchy-Riemann conditions: (2-54)

,1 = ,2 ,2 = ,1
we will have: 2 = 0 in R and the boundary conditions will be: dx2 dx1 1 dr 2 + = x2 dt x1 dt 2 dt d 1 dr 2 on C. = dt 2 dt Thus, for a simply connected region we have:

(2-55)

(2-56)

(2-57)

= r 2 on C.
and the stresses become:

1 2

(2-58)

M. Mirzaei, Elasticity

- 41 -

13 = G ( ,2 x2 ) 23 = G ( ,1 + x1 )

(2-59)

We may further simplify our boundary value problem by another variable change, as follows:
1 = r2 2 = 0 on C, and 2 = 2 in R.

(2-60)

In fact we may solve the torsion problem for different geometries through finding the appropriate function that satisfies the conditions presented in (2-60). Accordingly, the stresses become:
0 0 G ,2 ij ~ 0 G ,1 0

(2-61)

Circular Cross Section


For a circular cross section, we choose:
2 = k ( x12 + x2 a 2 )

(2-62)

which satisfies Eqs. (2-60). Next, we have:

= 0 at r = a 2 = 4 k 1 1 1 2 k = = ( x12 + x2 ) + a 2 2 2 2 so,
2 = + ( x12 + x2 ) = a 2 = constant

(2-63)

1 1 2 2 = constant = C

(2-64)

which means that the circular cross section does not warp in torsion.

M. Mirzaei, Elasticity

- 42 -

Accordingly, the displacements and stresses are:


ui ~ (x2 x3 , x1 x3 , C ) 0 0 Gx2 ij ~ 0 Gx1 0

(2-65)

Now we may calculate the total torque required to produce a certain amount of twist.

X2
32

dA
31 X2
0

X1
X1
Fig. 2.5

Bases on geometric considerations depicted in Fig. 2.5, we can write,


M = ( x1 32 x2 31 )dA
A

2 = G ( x12 + x2 ) dA A

= G

r 2 rd dr
2 0

r4 = G 4 0 a4 = G 2 4 Ga 4 = 2

(2-66)

M. Mirzaei, Elasticity

- 43 -

Elliptical Cross Section


For an elliptical cross section, we choose:
x2 x2 = k 12 + 2 1 2 a b

(2-67)

which satisfies Eqs. (2-60). Next, we have,


= 0 on C a 2b 2 1 1 2 = 2 = 2k 2 + 2 k = 2 a b ( a + b2 ) = ab x x 1 ( a 2 + b2 ) a b
2 2 2 1 2 2 2 2

(2-68)

Thus the stresses become:


2Ga 2 x2 0 0 2 b + a2 2Gb 2 ij ~ 0 x 2 2 1 b +a 0

(2-69)

We also have,
2 = + ( x12 + x2 )

1 2 2 2 2 a b x12 x2 1 2 2 1 2 2 + ( x1 + x2 ) = 2 2 (a + b ) a b 2 a 2 b2 b2 a 2 a 2b 2 2 x12 + x2 + 2 2 2 2 2 2 (a + b ) 2(a + b ) 2(a + b )

(2-70)

Accordingly, we may calculate the warping function using the Cauchy-Riemann conditions:

M. Mirzaei, Elasticity b2 a 2 x2 a 2 + b2 b2 a 2 = 2 x1 x2 + f ( x2 ) a + b2 ,2 = ,1 b2 a 2 b2 a 2 2 x1 + f ,2 = 2 x1 a + b2 a + b2 f ,2 = 0 f = C = b2 a 2 x1 x2 + C a 2 + b2

- 44 -

,1 = ,2 =

(2-71)

Accordingly, the out-of-plane displacement is: b2 a 2 u3 = = 2 x1 x2 a + b2 (2-72)

Fig. 2.6 shows how an elliptical cross section warps due to the existence of the out-ofplane displacements.
x2

x1

Fig. 2.6

Assignment 5:
Solve the torsion problem for a quadrilateral cross section.

M. Mirzaei, Elasticity

- 45 -

Cantilever Beam under End Loading


We start this section by verification of the basic equations governing the Cantilever Beam from an elementary point of view, i.e., the Bernoulli-Euler beam theory. Later we will present the theory of elasticity point of view. The Bernoulli-Euler beam theory is built upon the following assumptions:

x1

2b 2h

x2 x3

Fig. 2.7
h << l , b << l 33 << 11 33 0

(2-73)

It is further assumed that Linear Elements which are initially normal to the undeformed middle surface remain straight and normal to the deformed middle surface and suffer no extension, so we may write:

Undeformed Geometry

x1 x3
Deformed Geometry

Fig. 2.8

M. Mirzaei, Elasticity
11 12 ij ~ 22 0 23 0

- 46 -

(2-74)

On the other hand, the stresses are assumed to be:

13 11 0 0 ij ~ 0 33 0
The resultant moment is:
M = 11 x3 dA
A

(2-75)

(2-76)

and the resultant shear stress will be:


V = 13 dA
A

(2-77)

In order to verify the validity of the basic assumptions and the proposed stress and strain fields, we start from the assumed Displacement field:
u1 = x3 ( x1 ) u2 = u2 ( xi )

(2-78)

= u3 = u3 ( xi )
in which is the angle of rotation before and after deformation. The strains can be obtained using the kinematic equations as follows:

1 ( ui, j + u j ,i ) 2 u1,1 = x3 ,1 ij ~

ij =

1 u2,1 2 u2,2

1 ( + ,1 ) 2 11 12 1 ( u2,3 + ,2 ) ij ~ 22 2 ,3

0 23 0

(2-79)

which, when compared to the assumed strain field, gives the following results:

M. Mirzaei, Elasticity

- 47 -

13 = 0 = ,1 33 = 0 = ( x )

(2-80)

Now we obtain the strain field from the initially assumed stresses using the following constitutive equations:

ij =

1 (1 + ) ij ij kk E 11 E 0 13 0 0 ij ~ 33 0 0 (1 + ) 13 E 0 11 E (2-81)

11 ij ~

11

If we compare the resulting strains with the initially assumed strains and those obtained directly from the displacements, we will notice how the troublesome terms in the strain field have been ignored to match the stress and strain fields and obtain a workable theory. u2,1 = 0 11 E = x3 ,1 = x3,11 u2 = u2 ( x2 , x3 ) ij ~ 11 = u2,2 E 1 (1 + ) 13 ( + ,1 ) = E 2 is ignored 1 ( u2,3 + ,2 ) = 0 2 u2,3 = ,2 E 11 = 0 ( is ignored) 11 E

(2-82)

Next, we proceed with the derivation of other equations, and write:


u2,2 =

11 , 11 = Ex3,11

(2-83)

u2 = ,11 x2 x3

We also have:

M. Mirzaei, Elasticity

- 48 -

23 =

1 ( u2,3 + ,2 ) = 0 2 ,2 = u2,3 = ,11 x2

(2-84)

and,
2 M = 11 x3 dA = E ,11 x3 dA = E,11 I A A

M ,11 = EI

(2-85)

Hence, we may rewrite the displacements as: u2 =

M
EI

x2 x3 ; ,2 =

M
EI

x2

(2-86)

Now we check the Equilibrium Equations:

11,1 + 12,2 + 13,3 = 0 11,1 + 13,3 = 0 21,1 + 22,2 + 23,3 = 0 identically satisfied 31,1 + 32,2 + 33,3 = 0 31,1 + 33,3 = 0

(2-87)

Multiplying the first equilibrium equation by x3 and integrating across the cross section, we will have:

b h x3 11,1dA + x3 13,3 dx3 dx2 = 0 A b h

x1

( x dA) +
A

11

x h h dx dx = 0 h 13 3 2 b 3 13 h
b

(2-88)

b M h + x3 13 h dx2 V = 0 b x1

For problems with no surface shear force: M V = 0 x1 Integrating the third equilibrium equation across the cross section, we will have: (2-89)

M. Mirzaei, Elasticity
b h 13 33 dA + dx3 dx2 = 0 A x b h x 1 3

- 49 -

(2-90)

x1

( dA) +
A

13

33 33 dx2 = 0 h h

Now we define the resultant normal surface load as: q = 33 h 33 h dx2 b and the Eq. (2-90) becomes: V +q =0 x1 We may also derive other important equations for the B.E. beam as follows:
M V = 0 x1 2 M V 2M = = q +q=0 x12 x1 x12 M EI q 4 = 4 EI x1
b

(2-91)

(2-92)

,11 =

(2-93)

1 2M q 4 = = 4 2 EI x1 EI x1

Finally, the following equations are considered to form the basis of the B.E. Beam theory. M , EI V + q = 0, x1 4 q = , x14 EI (2-94)

,11 =

M V = 0, x1 2M + q = 0, x12

M. Mirzaei, Elasticity

- 50 -

Solution Based on the Theory of Elasticity


We initially assume that the normal stresses are the same as those of the elementary theory and also 32 is taken as zero, but no stipulation is made regarding the other shear stresses 13 and 12. Accordingly, the three equilibrium equations become:

11,1 + 12,2 + 13,3 = 0 12,2 + 13,3 = 21,1 + 22,2 + 23,3 = 0 21,1 = 0 31,1 + 32,2 + 33,3 = 0 31,1 = 0

Px3 I (i) (ii)


(2-95)

From equations (2-95, i) and (2-95, ii), we may conclude that the shearing stresses do not depend on x1, so they are the same for all cross sections. For this case the two relevant stress compatibility equations become:

1 mm,ij = fi , j + f j ,i + ij f m,m 1 + 1 1 P 2 31 = (i ) 1 + I 2 21 = 0 (ii ) 2 ij +


Now we must ensure that the modified solution satisfies the boundary conditions.

(2-96)

31 ( h, x2 , x1 ) = 0 23 ( x3 , b, x1 ) = 0
At this stage we define our stress components in terms of a stress function:

(2-97)

31 = ,2 21 = ,3

2 Px3 + f ( x2 ) 2I

(2-98)

which satisfies the equilibrium equations. Then our first stress compatibility equation (Eq. (2-96, i)) becomes: P 1 P (,233 + ,222 ) + f ,22 = 1 + I I (2-99) P (,33 + ,22 ),2 = f,22 1 + I P ,33 + ,22 = x2 f,2 + C 1 + I

M. Mirzaei, Elasticity

- 51 -

And the second compatibility equation (Eq. (2-96, ii)) is:

,333 ,322 = 0 (,33 + ,22 ),3 = 0

(2-100)

Note that equation (2-99) also satisfies equation (2-100). In order to find C, we express the rotation about the x1 axis by:

32 =

1 ( u3,2 u2,3 ) 2

(2-101)

and its rate of change by:

32,1 =

1 ( u3,2 u2,3 ),1 2 1 = ( u3,1 + u1,3 ),2 ( u2,1 + u1,2 ),3 2 = 31,2 21,3

(2-102)

which can be written in terms of stresses using the constitutive equations.

32,1 =

1 ( 31,2 21,3 ) 2G 1 = (,22 + ,33 + f,2 ) 2G

(2-103)

Also we can combine (2-103) and (2-99) to write: 2G32,1 = 1 P x2 + C 1 + I (2-104)

As we are considering a rectangular cross section and the bending is symmetrical about the x3 axis, at x2=0 we should have the average rotation equal to zero, resulting in C=0. Hence, equation (2-99) becomes:

,33 + ,22 =

P x2 f ,2 1 + I

(2-105)

which is the governing equation for the stress function, provided that the resulting stresses satisfy the boundary conditions.

M. Mirzaei, Elasticity

- 52 -

However, we may write the boundary conditions directly in terms of the stress function by choosing a proper form for the function f(x2), as follows: Ph 2 2I 31 ( h, x2 , x1 ) = ,2 ( h, x2 , x1 ) = 0 23 ( x3 , b, x1 ) = ,3 ( x3 , b, x1 ) = 0 f ( x2 ) = and the governing equation changes to: (2-106)

,33 + ,22 =

P x2 1 + I

(2-107)

Hence, in order to solve the problem at hand, we should choose a proper form of such that it satisfies the above equation and vanishes on the boundary.

Membrane Analogy
At this stage we may draw an analogy between the above governing equation and the equation of a homogeneous membrane whose outline is the same as the cross section of our beam. First, let us start with studying the deformation of a membrane subjected to a uniform tension at the edges and a uniform lateral pressure (see Fig. 2.9).

X1
S S

X3

dx dy

X1

X2
Fig. 2.9

M. Mirzaei, Elasticity

- 53 -

If q is the pressure per unit area of the membrane and S is the uniform tension per unit length of its boundary, the tensile forces acting on the opposite sides (dy sides) of an infinitesimal element give a resultant in the upward direction: S ( 2 z / x 2 )dxdy . In the same manner, the tensile forces acting on the other two opposite sides (dx sides) of the element give the resultant S ( 2 z / y 2 )dxdy (see Section 11.7, Kreyszig, Advanced Engineering Mathematics) and the equation of equilibrium of the element becomes: qdxdy + S 2 z 2 z dxdy + S 2 dxdy = 0 x 2 y

q 2 z 2 z 2 + 2 = S x y

(2-108)

At the boundary, the deflection of the membrane is zero. Comparing the above equation with the governing equation of our beam problem, it is clear that the solution of the beam problem reduces to the determination of the deflections of the membrane which are produced by a continuous load with the intensity proportional to / (1 + )( Px2 / I ) .
The curve mnp in Fig. 2.10 represents the intersection of the membrane with the x1x2 plane.

X2
b

X3 X2
m n p

X1
Fig. 2.10

M. Mirzaei, Elasticity

- 54 -

Now we rewrite the stresses from equation (2-98) (incorporating f ( x2 ) = Ph 2 / 2 I ):


2 Px3 Ph 2 + 2I 2I 2 2 P(h x3 ) = + ,2 2I 21 = ,3

31 = ,2

(2-109)

From the above equations it is clear that the deviations of the shearing stresses from the elementary theory are proportional to the first derivatives or slope of the membrane at any point. For example, the correction to 31 shows a maximum positive value at the two sides, a maximum negative value in the center, and is zero at the quarter points (See Fig. 2.10). From the condition of loading of the membrane it is clear that is an even function of x3 and an odd function of x2. Hence, it is appropriate to take the stress function in the form of the Fourier series:

m = n = m = 0 n =1

2 m +1, n

cos

( 2m + 1) x3 sin n x2
2h b

(2-110)

Substituting from above expression into our governing equation we may find the coefficient of the Fourier series and obtain the final form of our stress function as:
P 8b3 = 1 + I 4

m = n = m = 0 n =1

( 1)

m + n +1

b 2 b ( 2m + 1) n ( 2m + 1) 2 + n 2 4h
2

cos

( 2m + 1) x3 sin n x2
2h

(2-111)

which can be used to calculate the shearing stresses using equation (2-98).

Note that the above membrane analogy and Eq. (2-108) can also be used to obtain solutions to the TORSION problems discussed before.

M. Mirzaei, Elasticity

- 55 -

Approximate Solutions
If the depth of the beam is large compared with the width h>>b, we may consider a cylindrical shape for the surface of the membrane at the points sufficiently far from the short sides at x3 = h (assume = ( x2 ) ), so we may write:

= ( x2 )
Px2 1 + I P 3 ( x2 ) = ( x2 + C1 x2 + C2 ) 1 + 6 I (0) = 0 C2 = 0 2 (b) = 0 C1 = b P 3 2 ( x2 ) = ( x2 b x2 ) 1 + 6 I P 2 2 31 ( x3 , x2 ) = ( h x3 ) + 2I 1 + ,22 =

(2-112)

2 b2 x2 3

On the other hand, if the width of the beam is large compared to the depth (b>>h), for the points far from the short sides at x2 = b , we may consider the deflections of the membrane as a linear function of x2 and write: Px2 1 + I Px2 x3 ,3 = + C1 1 + I Px2 2 x3 + C1 x3 + C2 ( x3 , x2 ) = 1 + 2 I Px2 h 2 ( h, x2 ) = 0 C2 = , C1 = 0 1 + 2 I Px2 2 2 ( x3 , x2 ) = ( x3 h ) 1 + 2 I 2 2 2 P ( h 2 x3 ) 1 P ( h x3 ) 31 = 1 1 + = 1 + 2I 2I ,33 =

(2-113)

21 =

Px3 x2 1 + 2 I

M. Mirzaei, Elasticity

- 56 -

Anti-Plane Shear Problem


Consider a cylindrical solid with arbitrary cross-section, as shown in the Fig. 2.11. Also assume that the length of the cylinder greatly exceeds any cross sectional dimension. We may have states of anti-plane shear in a solid by an appropriate loading, as we will see in the following example. The governing equations and boundary conditions for anti-plane shear problems are very simple and lead to solving the Laplaces equation (for which we have many powerful solution techniques).
We consider the following boundary value problem, with body force b = b3 ( x1 , x2 )e 3

T1 = T2 = 0 u3 = u ( x1 , x2 ) T3 = T3 ( x1 , x2 )
3

on on on

C Cu CT (2-114)

X2
R1

X1

R2
Anti Plane Shear

X3

C
Fig. 2.11

In addition, for a traction boundary value problem we must ensure the existence of a static equilibrium solution:

( x1 , x2 )ds + b3 ( x1 , x2 )dV = 0 R

(2-115)

M. Mirzaei, Elasticity

- 57 -

To accept any solution with zero resultant force and moment acting on the ends of the cylinder, we set the boundary conditions on R ( = 1, 2) as:

n dA = 0 r n dA = 0
(2-116)

Recall the field equations:

ij = 1 (ui , j +u j ,i ) 2 ij = ij kk + 2 ij ij , j +bi = 0

(2-117)

Since all forces and boundary displacements act in the x3 direction, it is natural to assume that the displacements are in the x3 direction everywhere. Lets assume a solution of the form: u1 = u2 = 0 % u3 = u3 ( x1 , x2 ) The strains and stresses follow as: (2-118)

= 0, = 0,

33 = 0 33 = 0,

% 3 = 1 u3 , 2 % 3 = u3 ,

(2-119)

where Greek subscripts range from 1 to 2. The equilibrium equations reduce to:

3 , +b3 = 0
% u3 , +b3 = 0

% u3 , =

b3

(2-120)

The boundary conditions may be re-written as,


% u3 = u3 ( x1 , x2 )

on

Cu on CT

% T3 = 3 n = u3 , n = T3 ( x1 , x2 )

(2-121)

M. Mirzaei, Elasticity

- 58 -

As mentioned before, the Laplaces equation can be solved in different ways. Here we will use the complex variable method. This approach is based on the fact that both the real and imaginary parts of an analytic function satisfy the Laplaces equation. We start with the following definitions:
f ( z ) = v( x1 , x2 ) + iw( x1 , x2 ) z = x1 + ix2 i = 1

(2-122)

If f ( z ) is analytic, the derivative with respect to z is path independent, which requires: v w = x1 x2 w v = x1 x2

(2-123)

These are known as the Cauchy-Riemann conditions. Hence, we have:


v, 2v 2v w w = 2+ 2 = =0 x1 x2 x1 x2 x2 x1

(2-124)

or, in general: v, = w, = 0 (2-125)

Thus, both the real and imaginary parts of an analytic function satisfy the Laplace equation. Now, let z = x1 + ix2 = rei characterize the position of a point in the plane of the solid of interest. Then let f ( z ) denote any analytic function of position. We can set:
% u3 = e( f ( z )) = v( x1 , x2 )

(2-126)

% % Note that u3 automatically satisfies the equilibrium equation u3, = 0 . We can solve our
elasticity problem by finding an analytic function that satisfies appropriate boundary conditions. Of course, we may also use:

% u3 = Im( f ( z )) = w( x1 , x2 )
Now, we can determine the stresses directly from f(z). Note that f ( z ) = v + iw f ( z ) = v,1 +iw,1 = v,1 iv,2 = w,2 +iw,1

(2-127)

(2-128)

M. Mirzaei, Elasticity

- 59 -

% Suppose we choose u3 = e( f ( z )) , then using (2-119) we have:


% % f ( z ) = u3 ,1 iu3 ,2 f ( z ) = 31 i 32

(2-129)

f ( z ) = 31 + i 32

Where f ( z ) denotes the complex conjugate.

% Similar expressions can be determined for u = m( f ( z )) . In this case % % f ( z ) = u3 ,2 +iu3 ,1 f ( z ) = 32 + i 31


(2-130)

As an example we consider the fundamental solution for an infinite solid and find the displacement and stress fields induced by a line load F = Fe 3 acting at the origin. We start by generating the solution from the following function:

f ( z ) = C log( z )

(2-131)

X2

r
X1

Fig. 2.12

This is analytic everywhere except the origin. Next we choose:

% u3 = e( f ( z ))

(2-132)

M. Mirzaei, Elasticity

- 60 -

and compute the resultant force acting on a circular arc enclosing the origin:
F3 = ( 31n1 + 32 n2 )ds =
A B
2

(
0

31

cos + 23 sin )rd

(2-133)

Here we may write:

31 cos + 32 sin = e ( 31 + i 32 ) (cos i sin )


= e( f '( z )e i ) which using (2-131), can be written as:

(2-134)

31 cos + 32 sin = e C

e i z e i = e C i = C re r

(2-135)

Substituting (2-135) into the right hand side of (2-133), results in:
F3 = 2 C C = F3 2 (2-136)

Therefore, the solution can be found and simplified as follows: F3 e {log( z )} 2 F F % u3 = 3 e[log(rei )] = 3 log(r ) 2 2

% u3 =

(2-137)

r = x x

Assignment 6: Find the stress field for a Screw Dislocation problem. b % Hint: Assume u = m {log( z )} 2

M. Mirzaei, Elasticity

- 61 -

The Mode III Crack Problem


The analysis of Mode III crack problem is relatively simple because we may assume that u1=u2=0 and u3 = u3(x1 , x2), which clearly represent an Anti-Plane Shear problem.

III

Fig. 2.13

Referring to the material covered in the above sections, and ignoring the body forces in our crack problem, we may write:

% % u3 , = 2u3 = 0
Here, we consider a solution in the following form:

(2-138)

% u3 =

f ( z) + f ( z)

(2-139)

Hence, the strains become:

31 =

1 f ( z ) + f ( z ) 2 i f ( z ) f ( z ) 32 = 2

(2-140)

Using the constitutive equations we find that:

31 i 32 = 2 f ( z )

(2-141)

Now, let the origin of the coordinate system be located at the tip of a crack lying along the negative x1 axis as shown in Fig. 2.14.

M. Mirzaei, Elasticity

- 62 -

X2

X1

Fig. 2.14

Now, we will focus our attention upon a small region (denoted as D) containing the crack tip and consider the holomorphic function: f ( z ) = Cz +1 , C = A + iB (2-142)

where A, B, and are real undetermined constants. For finite displacements at the crack tip we must have: ( z = r = 0), f 1 ). Then we will have:

31 i 32 = 2( + 1)Cz = 2( + 1)r ( A + iB)(cos + i sin )


where,

(2-143)

31 = 2( + 1)r ( A cos B sin ) 32 = 2( + 1)r ( A sin + B cos )

(2-144)

The boundary condition that the crack surfaces be traction free requires that 32 = 0 on

= . Consequently we have:
A sin + B cos = 0 A sin B cos = 0 To avoid the trivial solution, the determinant of the coefficients of the above equations must vanish. This leads to:
sin 2 = 0

(2-145)

(2-146)

which for > 1 has the following roots:

= , n / 2, n = 0,1, 2,...

1 2

(2-147)

M. Mirzaei, Elasticity

- 63 -

Of the infinite set of functions of the form of the above equation that yield traction-free crack surfaces within D, the function with = 1/2 for which A = 0, provides the most significant contribution to the crack-tip fields. For this case the stresses and displacements become, respectively:
31 K III = 1/ 2 32 (2 r ) sin( / 2) cos( / 2)

(2-148)

and,
2 K III r 2 % u3 = sin( / 2) 2
1

(2-149)

where B has been chosen such that: K III = lim {(2 r )1/ 2 32
r 0

=0

(2-150)

The quantity KIII is referred to as the Mode III stress intensity factor, which is established by the far field boundary conditions and is a function of the applied loading and the geometry of the cracked body. Whereas the stresses associated with the other values of are finite at the crack tip, the stress components associated with = 1/2 have an inverse square root singularity at the crack tip. It is clear that the latter components will dominate as the crack tip is approached and represent the asymptotic forms of the elastic stress and displacement fields.

M. Mirzaei, Elasticity

- 64 -

Chapter 3
2D Static Boundary Value Problems: Plane Elasticity
Chinese Proverb - It is better to ask a question and look like a fool for five minutes, than not to ask a question at all and be a fool for the rest of your life.

Two Dimensional Elastostatic Problems


Plane Strain Formulation
Consider a long prismatic bar loaded with surface tractions and body forces normal to the x3 axis as depicted in Fig 3.1.

X2

f
T

X1

f
T
Plane stress

R
Plane strain

X3

C
Fig. 3.1

Also assume that the magnitude of loadings does not vary along this axis. Under these conditions we may write:

M. Mirzaei, Elasticity

- 65 -

u = u ( x , x ) u = u ( x ) 1 1 1 2 , = 1, 2 u2 = u2 ( x1 , x2 ) u3 = constant, generalized plane strain u3 = 0, plane strain Accordingly, the strain field becomes:
11 12 1 ij = ( ui , j + u j ,i ) ~ 22 2 0 0 0

(3.1)

(3.2)

In order to calculate the stress field we use the following constitutive equations:

ij = kk ij + 2G ij
Partitioning the stress tensor, we have

(3.3)

ij ~

3 33

(3.4)

Here we have three constitutive equations:

= + 2G 3 = 3 = 0 33 = = ( 11 + 22 )

(i) (ii)

(3.5)

Contracting on in equation (3.5i) we have:

= 2 + 2G
= 1 2 ( + G ) (3.6)

Using the above equation and substituting for in equation (3.5ii) we can write:

33 =

= 2( + G)

(3.7)

which shows that in plane strain condition the stress in the x3 direction is a dependent quantity. Accordingly, the strains can be written as:

M. Mirzaei, Elasticity

- 66 -

1 2G 1 = 2G 2 ( + G ) 1 = 2G

(3.8)

It is important to note that the above equation cannot be obtained from the original constitutive equation by simply replacing i,j by ,. 1 2G 11 ( 11 + 22 ) ~ 1 22 ( 11 + 22 ) 2G 1 12 2G

(3.9)

The equilibrium equations in the plane strain case become:

, + f = 0
11,1 + 12,2 + f1 = 0 12,1 + 22,2 + f 2 = 0 (3.10)

Now, we assume that the body force f can be derivable from a potential, V. Hence, we may write:
f = V, , V, = 0 ( 11 V ),1 + 12,2 = 0 ( 22 V ),2 + 12,1 = 0

(3.11)

The above equations can be satisfied by assuming the existence of two functions A(x) and B(x), such that:

11 V = A,2 ; 12 = A,1 22 V = B,1 ; 12 = B,2


B,2 + A,1 = 0 The latter is satisfied by assuming the existence of a stress function such that: B = ,1 ; A = ,2 (3.13) (3.12)

M. Mirzaei, Elasticity

- 67 -

Accordingly, the stresses can be written in terms of the Airy stress function as:

11 V = ,22 22 V = ,11 12 = 21 = ,12


The stress boundary conditions for the plane strain reduce to: T = n

(3.14)

(3.15)

From the six stress compatibility equations we have only one important non-trivial equation, which upon substitution of the stresses by the stress function becomes:
2 33 + 2 = ( 1 2V mm,33 = 1 + 1 2 2 ), = 2 + 2 x1 x2

2 =

2V

1 2V 1 1 2 (,22 + ,11 ) = 2 2V 1 1 2 2 22 = V 1 2 (V + ,22 + V + ,11 ) = 1 2 4 = 1 2 V

(3.16)

The above is the governing partial differential equation for the Airy Stress Function. If the potential function is harmonic, i.e., 2V = 0 , then satisfies the homogeneous biharmonic equation:
4 = 0 4 = 2 2 = (

),

4 4 4 +2 2 2 + 4 x14 x1 x2 x2

(3.17)

M. Mirzaei, Elasticity

- 68 -

Plane Stress Formulation


Let us consider a short prismatic body loaded with surface tractions and body forces normal to the x3 axis as depicted in Fig. 3.1. Under these conditions we may write for the stresses:

ij =

1 h/2 ij dx3 h h / 2

(3.18)

The bar sign over a quantity indicates that it has been averaged over the thickness. The stress tensor is:
11 12 ij ~ 22 0 0 0

(3.19)

Also, the equilibrium equations can be written as:


f = V, , V, = 0 ( 11 V ) + 12,2 = 0 ,1 ( 22 V ),2 + 12,1 = 0

(3.20)

which are exactly the same as those derived for plane strain condition, except for the mean value interpretation. Thus we may similarly write:

11 V = ,22 22 V = ,11 12 = 21 = ,12


and the related constitutive equations are: (3.21)

1 + = + 3 = 0
E E

(i) (3.22) (ii)

33 =
E

From Eq. (3.22i) we have:

M. Mirzaei, Elasticity

- 69 -

1 + 1 = 2 + =
E E E = E 1

(3.23)

(3.22ii) 33 =

which shows that in plane stress condition the strain in the x3 direction is a dependent quantity. We may also write stresses in terms of strains:

E 1 +

+ E

(3.24)

Substituting for we will have:

E 1 + 2G = 2G + + 2G

(3.25)

If we compare the above equations with those used in plane strain, we conclude that the entire system of equations in plane stress are analogous to those for plane strain, provided that we replace the barred quantities with the unbarred and define the followings:
G =G

2G + 2G

(3.26)

Hence, we can derive the fundamental biharmonic boundary value problem for the plane stress conditions as: 1 2 4 = 1 2 V

1 +

(3.27)

Assignment 7:
Find the stress field for a narrow rectangular beam under end loading.

M. Mirzaei, Elasticity

- 70 -

Airy Stress Functions in Cylindrical Coordinates


It is quite beneficial to solve the boundary value problems involving cylindrical regions using appropriate cylindrical coordinates.
X2

rr r rr
X1

Fig. 3.2
In case of zero or constant body forces, the state of stress is related to the Airy function using the following transformations:

rr = ri rj ij = i j ij r = ri j ij

(3.28)

which, by appropriate substitutions for the stress function , lead to the following expressions for the stresses:

rr = r

1 1 2 1 1 + 2 = ,r + 2 , 2 r r r r r 2 = 2 = ,rr r 1 1 1 = = 2 , ,r r r r r

(3.29)

The governing equations for this coordinate system will be:

1 1 2 = ,rr + ,r + 2 , r r 2 1 1 2 1 1 4 = 2 + + 2 + ,r + 2 , = 0 2 , rr r r r r r r

(3.30)

M. Mirzaei, Elasticity

- 71 -

As an example, we will obtain the stress field due to a line load of magnitude P per unit out-of-plane length, acting on the surface of a homogeneous isotropic half-space.

P
X1

r
X3

Fig. 3.3
Assume the Airy function in the form of:

r sin

(3.31)

The resulting state of stress can be written as:

rr =

2 P cos r

= r = 0

(3.32)

In order to find the displacement field we need to determine the strains by substituting the expression for stress into the constitutive relation and then integrating the strains to calculate the displacements. For the point force solution, one we can show that:
ur = 1

1 u = P sin log r + P sin 2

P cos log r

1 2 P sin 2

(3.33)

Axisymmetric Stress Distribution


When we have symmetrical stress distribution about an axis, in the absence of body forces, we may write:

M. Mirzaei, Elasticity

- 72 -

rr = r

1 1 = , r r r r 2 = 2 = ,rr r =0

(3.34)

and also:
1 2 = ,rr + ,r r 1 1 4 = r ( r,r ),r = 0 r r ,r ,r

(3.35)

The above expression can successively be integrated with respect to r to give the general form of the stress function for axisymmetric stress distribution in cylindrical coordinates as follows:

1 r ( r,r ),r = 0 ,r ,r r C 1 ( r,r ),r = 1 r ,r r

1 r ( r,r ),r = C1 ,r r 1 ( r,r ),r = C1 ln r + C2 r

( r,r ),r = C1r ln r + C2 r r2 r2 r2 r,r = C1 ln r + C2 + C3 = C1r 2 ln r + C2 r 2 + C3 4 2 2 C ,r = C1r ln r + C2 r + 3 r 2 2 r r r2 ln r + C2 + C3 ln r + C4 = C1 4 2 2 = C1r 2 ln r + C2 r 2 + C3 ln r + C4 Accordingly the stresses become: (3.36)

rr = C1 (1 + 2 ln r ) + 2C2 + C3 r

1 r2 1 = C1 ( 3 + 2 ln r ) + 2C2 C3 2 r =0

(3.37)

M. Mirzaei, Elasticity Finally, the displacements are:

- 73 -

1 1 + + C4 sin + C5 cos C1r (1 )( 2 ln r 1) 2 + 2C2 (1 ) r C3 E r 1 u = [ 4C1r + C4 cos C5 sin + C6 r ] E ur =


Axisymmetric Stress and Displacement fields

(3.38)

When we also have axisymmetric displacements in the absence of body forces, the displacement values do not depend on , so we have C1=0 and we may write:

rr = C2 + C3 r

1 r2 1 = C2 C3 2 r =0

(3.39)

Assignment 8:
Find the stress field for a cylindrical pressure vessel under internal and external pressures.

Curved Beam under constant bending moment


As an example for two-dimensional problems in cylindrical coordinates we consider the solution of a curved beam with rectangular cross section, under a constant bending moment M.

RO r

Ri

rr

Fig. 3.4

M. Mirzaei, Elasticity

- 74 -

For the upper and lower faces, we can write:

rr ( R0 ) = rr ( Ri ) = 0 r ( R0 ) = r ( Ri ) = 0
On the other hand, the equilibrium conditions dictate that:

(3.40)

dA = 0 r dA = M
A A

(i) (ii)

(3.41)

Notice that in this problem the stress field is axisymetric but the displacement field is not, so the stresses can be written as:

rr = C1 (1 + 2 ln r ) + 2C2 + C3 r

1 r2 1 = C1 ( 3 + 2 ln r ) + 2C2 C3 2 r =0

(3.42)

Next, we write the expression for rr at the boundaries: C1 (1 + 2 ln R0 ) + 2C2 + C3 1 =0 R02

1 C1 (1 + 2 ln Ri ) + 2C2 + C3 2 = 0 Ri

(3.43)

Note that we still need a third equation to be able to determine our three constants. In order to find it, we write equation (3.41i) in terms of the stress function as follows:

dA = B ,rr dr = B ,r R = 0
Ri
i

Ro

Ro

(3.44)

1 Since we have rr = ,r , the above expression is consistent with our first boundary r condition and indicates that ,r = 0 on the boundary. Also, the second equilibrium condition (3.41ii) can be written as:

M. Mirzaei, Elasticity

- 75 -

r dA = B r, rr dr = M
Ri Ro Ro Ri Ri

Ro

B r,rr dr = B r,r ( Ro ) ( Ri ) = M B

B ,r dr = M
Ri

Ro

(3.45)

Considering the general form of the stress function in cylindrical coordinates:

= C1r 2 ln r + C2 r 2 + C3 ln r + C4
we may write (3.45) as: C1 ( Ro 2 ln Ro Ri 2 ln Ri ) + C2 ( Ro 2 Ri 2 ) + C3 ln Ro M = Ri B

(3.46)

(3.47)

which gives the required third equation in terms of the three constants of the problem. Accordingly, we will have: 2M ( Ro 2 Ri 2 ) NB M 2 2 2 2 C2 = ( Ro Ri ) + 2 ( Ro ln Ro Ri ln Ri ) NB 4 M 2 2 Ro C3 = Ro Ri ln NB Ri C1 = where R N = ( Ro Ri ) 4 Ro Ri ln o Ri
2 2 2 2 2

(3.48)

Hence, the stresses can be found as:

rr = =

R r 4 M 2 2 Ro 1 + Ro 2 ln + Ri 2 ln i Ro Ri ln 2 NB Ri r Ri r R r 4 M 2 2 Ro 1 Ro 2 ln Ri 2 ln i Ro 2 + Ri 2 Ro Ri ln 2 NB Ri r Ro r

(3.49)

r = 0

M. Mirzaei, Elasticity

- 76 -

Sheet with a Circular Hole under Uniaxial Stress


As an example for two-dimensional problems in cylindrical coordinates we consider a thin sheet, with a circular hole, under uniaxial uniform tension as follows:

11 = T1 22 = 12 = 0

(3.50)

The above uniform stress is disturbed by the presence of the circular hole. Also note that it is more convenient to define the local stress distribution near the hole using cylindrical coordinates. Hence, we use the St.Venants principle and argue that the stress distribution near the hole remains unchanged if we construct a hypothetical circle, with a diameter equal to the sheet width, and apply equivalent forces on the boundary of this circle, as shown in Fig. 3.5a.

T1
(a)
b

(b)

T1/2

X2

= +
(c)

(T1/2)(cos2)

T1

X1
Fig. 3.5

-(T1/2)(sin2)

M. Mirzaei, Elasticity

- 77 -

We start by a transformation of the Cartesian components of stress into the Polar components as:

rr = ri rj ij = 11 cos 2 + 2 12 sin cos + 22 sin 2 = i j ij = 11 sin 2 2 12 sin cos + 22 cos 2


2

r = ri j ij = ( 22 11 ) sin cos + 12 ( cos sin )


2

(3.51)

which, in combination with the stress state expressed by equation (3.50), give the stresses at the location of the hypothetical circle as follows:
T1 (1 + cos 2 ) 2 T ( b, ) = T1 sin 2 = 1 (1 cos 2 ) 2 T r ( b, ) = T1 sin cos = 1 sin 2 2

rr ( b, ) = T1 cos 2 =

(3.52)

As depicted in Fig. 3.5, we may separate the stresses along the outer boundary into two parts. For the first part (shown in Fig. 3.5b) we have the axisymmetric stress state:
(1) rr ( b, ) =

T1 2 (1) r ( b, ) = 0

(3.53)

Note that does not act on the boundary. The second part is defined for the circle depicted in Fig. 3.5c in the form of a -dependent stress state:

T1 cos 2 2 T r(2) ( b, ) = 1 sin 2 2


(2) rr ( b, ) =

(3.54)

The solution for the axisymmetric stress state (represented by Eqs. (3.53)) can be easily obtained from the solution of a circular disc under radial load (see Eq. (3.39)) as follows:

rr = C2 + C3 r

1 r2 1 = C2 C3 2 r =0

(3.55)

M. Mirzaei, Elasticity

- 78 -

We may find the constants using the following boundary conditions: T1 2 (1) r ( b, ) = 0
(1) rr ( b, ) = (1) rr ( a, ) = 0

(3.56)

r(1) ( a, ) = 0
Accordingly, the stresses can be written as;
(1) rr ( r , ) =

T1 b 2 T a 2b 2 1 2 2 b2 a 2 2 b a 2 T b2 T a 2b 2 (1) ( r , ) = 1 2 2 + 1 2 2 2 b a 2 b a

1 r2 1 r2

(3.57)

In order to find the solution to the second part, we start with the general form of the biharmonic equation. Due to the periodicity of the boundary conditions we consider a proper form for the stress function and proceed as follows: 2 1 1 2 1 1 4 = 2 + + 2 + ,r + 2 , = 0 2 , rr r r r r r r = F (r ) cos 2 2 1 1 2 1 4 cos 2 F,rr + F,r 2 F = 0 + 2 2+ 2 r r r r r r 2 9 9 F,rrrr + F,rrr 2 F,rr + 3 F,r = 0 r r r The above differential equation can be solved for F and will eventually result in the following expression for our stress function: (3.58)

= C1 + C2 r 2 + C3r 4 + C4

1 cos 2 r2

(3.59)

Equation (3.59) along with equations (3.29) can be used to calculate the stresses as follows:

M. Mirzaei, Elasticity

- 79 -

(2) rr = 4C1 (2)

r(2)

1 1 + 2C2 + 6C4 4 cos 2 2 r r 1 = 2C2 + 12C3 r 2 + 6C4 4 cos 2 r 1 1 = 2C1 2 + 2C2 + 6C3 r 2 6C4 4 sin 2 r r

(3.60)

The boundary conditions are:


T1 cos 2 2 T r(2) ( b, ) = 1 sin 2 2 (2) rr ( a, ) = 0
(2) rr ( b, ) =

(3.61)

r(2) ( a, ) = 0
which can be rewritten as: 4 a2 2 a2 4 b2 2 2 b 2 2 2 2 0 6a 2 0 6b 2 6 0 a4 C 1 6 0 4 a C2 T1 = 6 C3 2 4 b C4 T1 6 2 4 b

(3.62)

The above four equations can be solved for the four unknown constants. However, we may simplify Eqs.(3.62) for the case where the hole is small compared to the width of the plate. First, we multiply the forth equation by (a/b)2 and let a / b 0 , which gives C3=0. Next, we multiply the third equation by a2, let a / b 0 , and get C2=-T1/4. Hence, we may write the first and the second equations as: T 6 4 C1 + 1 4 C4 = 0 2 a 2 a 2 T 6 2 C1 1 4 C4 = 0 a 2 a from which we find:

(3.63)

M. Mirzaei, Elasticity
a2 a4 C1 = T1 and C4 = T1 2 4

- 80 (3.64)

Finally, we obtain the stresses using Eq. (3.60):


(2) rr = 2

a2 1 3 a4 T1 cos 2 2 2 2 r4 r 1 2 3 a4 T1 cos 2 2 r4 (3.65)

(2) = +

r(2) =

a2 1 3 a4 + sin 2 2 2 2 r4 r

In order to obtain the total solution, we reevaluate Eq. (3.57), by dividing the numerator and denominator of each term by b2, letting a / b 0 , and then adding the results to Eq. (3.65): T1 a 2 T1 a2 a4 rr ( r , ) = 1 2 + 1 4 2 + 3 4 cos 2 r r 2 r 2 T a2 T a4 ( r , ) = 1 1 + 2 1 1 + 3 4 cos 2 2 r 2 r

(3.66)

r ( r , ) =

T1 a2 a4 1 + 2 2 3 4 sin 2 2 r r

Now we may examine the stresses at particular points that are most important. For instance, the maximum hoop stress that occurs at r = a is:

( a, ) = T1 2T1 cos 2
Note that its largest value occurs at = a, = 3T1 2

(3.67)

, with the magnitude of: 2 2 (3.68)

which indicates that we have a stress concentration factor of 3 for the specified location.

Assignment 9:
Find the stress concentration factor for a small circular hole in a large sheet subjected to a remote tensile stress T1 and a compressive stress T2 .

M. Mirzaei, Elasticity

- 81 -

General Solution of Plane Problems using Complex Variable Functions


In this section we will use complex variable functions to obtain general solutions to the plane elasticity problems. Starting with the biharmonic equation we can show: 4 = 0 4 = 2 2 = (

),

4 4 4 +2 2 2 + 4 x14 x1 x2 x2 (3.69)

2 2 2 2 2 + 2 2 + 2 = 0 x1 x2 x1 x2 2 2 2 + 2 ( 22 + 11 ) = 0 x1 x2 2P 2P 2 + 2 = 2 P = P, = 0 x1 x2

The conclusion is that that P is harmonic. Now we assume that there is a complex variable function which has P as its real part:

f ( z ) = P + iQ
Since f ( z ) is analytic, its integral is also be analytic. Accordingly, we may write:

(3.70)

f ( z )dz = ( P + iQ ) dz =
( z ) = p + iq ( z ) =

lets say 4( z )

1 f ( z )dz 4 d ( z ) ,1 = z,1 = ( z ) dz 1 ( z ) = p,1 + iq,1 = f ( z ) 4 1 P Q p,1 + iq,1 = ( P + iQ ) p,1 = , q,1 = 4 4 4

(3.71)

Invoking the Cauchy-Riemann conditions we will have:

M. Mirzaei, Elasticity
P p,1 = q,2 = 4 (i) p = q = Q ,1 ,2 4 (i) 2 = P = 4 p,1 = 2 p,1 + 2q,2 2 2 p,1 2q,2 = 0 2 ( x1 p x2 q ) = 0

- 82 -

(3.72)

Note that ( x1 p x2 q ) is a solution of the Laplaces equation, so we may assume it to be the real part of a complex variable function:

( x1 p x2 q ) = p1
where we have:

= x1 p + x2 q + p1

(3.73)

( z ) = p1 + iq1

(3.74)

Moreover, = x1 p + x2 q + p1 may be assumed to be the real part of the following analytic function:

( x1 ix2 )( p + iq ) + p1 + iq1 or = x1 p + x2 q + p1 = e ( x1 ix2 )( p + iq ) + p1 + iq1

(3.75)

Finally, we may write the general form of the Muskhelishvili stress function in terms of two complex potentials and as follows:

= e [ z ( z ) + ( z ) ]

(3.76)

Plane Stress Formulation


Starting with the constitutive equations for plane stress we have:

1 + = +
E E which may be rewritten as:

(3.77)

M. Mirzaei, Elasticity Eu1,1 = 11 22 Eu2,2 = 22 11 G ( u1,2 + u2,1 ) = 12 (i) (ii) (iii)

- 83 -

(3.78)

Now, using Eqs. (3.72) we may write: Eu1,1 = (1 + ),11 + P Eu2,2 = (1 + ),22 + P P = 4 p,1 = 4q,2 4 2Gu1,1 = ,11 + 1 + p,1 2Gu = + 4 q 2,2 ,22 ,2 1 + The above equations can be integrated to give the displacements: 4 2Gu1 = ,1 + 1 + 2Gu = + 4 2 ,2 1 + p + f ( x2 ) q + f ( x1 )

(3.79)

(3.80)

If we substitute the above equations into Eq. (3.78iii), it can be shown that f(x2) and f(x1) are constants. Now we may substitute for the with our general complex variable stress function (Eq. (3.76)) but first we note: f ( z ) = + i f ( z ) + f ( z ) = 2 = 2e ( f ( z ) ) f ( z) = i so from (3.76) z ( z ) + ( z ) + z ( z ) + ( z ) = 2 If we substitute the above expression for the into Eq. (3.80), we will obtain the following general expression for the displacements in terms of the two complex potentials and : 2G (u1 + iu2 ) = 3 ( z ) z ( z ) ( z ) 1 + (3.82)

(3.81)

It should be noted that we can obtain the above equation for the plane strain condition by replacing with

M. Mirzaei, Elasticity

- 84 -

Having found the displacements, the general expressions for the stresses can be easily obtained as follows:

11 + 22 = 4e ( ( z ) )
22 11 2i 12 = 2 z ( z ) + ( z ) 22 11 + 2i 12 = 2 [ z ( z ) + ( z )] or (3.83)

The final conclusion is that the solution to a particular two-dimensional elasticity problem can be easily obtained if we can find the appropriate complex variable potentials for that problem. However, finding the required potentials is not usually that easy!

The Mode I Crack Problem


As an example for the solution of plane problems using complex variable potentials, we consider the Mode I crack problem. Here, we may assume the displacement field as u1 = u1(x1, x2), u2 = u2(x1, x2), and u3 = 0.

Fig. 3.6
We directly start from the general solution with the following form of Eq. (3.83):

11 + 22 = 4e ( ( z ) ) = 2 ( z ) + ( z )
22 11 2i 12 = 2 z ( z ) + ( z ) Due to symmetry with respect to the crack plane we choose a solution of the form: = Az +1 , = Bz +1

(3.84)

(3.85)

M. Mirzaei, Elasticity

- 85 -

where A, B, and are real constants. In order to have nonsingular displacements at the crack tip we must have > 1. The substitution of Eq. (3.85) into Eq. (3.84) yields:

22 i 12 = ( + 1)r { A [ 2 cos + cos( 2) ]


+ B cos i [ A sin( 2) + B sin ]}

(3.86)

The boundary conditions at the crack tip dictate that the stresses in Eq. (3.86) vanish for = . Consequently we have: A(2 + ) cos + B cos = 0 A sin + B sin = 0 for which a nontrivial solution exists if:
sin 2 = 0

(3.87)

(3.88)

which for > 1 has the following roots:

= , n / 2, n = 0,1, 2,...

1 2

(3.89)

The dominant contribution to the crack-tip stress and displacement fields occurs for = 1/2 for which A = 2B. An inverse square root singularity in the stress field exists at the crack tip. Substituting Eq. (3.87) with A = 2B and = 1/2 into Eqs. (3.86) and (3.82), we can show that:
11 1 sin( / 2) sin(3 / 2) KI cos( / 2) sin( / 2) cos( / 2) 12 = 1/ 2 (2 r ) 1 + sin( / 2) sin(3 / 2) 22

(3.90)

and also:
1 2 u1 K I r 2 cos( / 2) 1 + 2sin ( / 2) = 2 u2 2 2 sin( / 2) + 1 2 cos ( / 2)

(3.91)

The Mode I stress intensity factor K, which can be obtained using the global boundary conditions of the problem, is defined by: K I = lim {(2 r )1/ 2 22
r 0

=0

(3.92)

M. Mirzaei, Elasticity

- 86 -

General Solutions in Curvilinear coordinates


For many two-dimensional elasticity problems the solution procedure requires that we define our previously derived general solutions using curvilinear coordinates. In general, we may specify the position of a point in a plane as the intersection of two curves. This constitutes the curvilinear coordinates. The two curves may be represented by:
F1 ( x1 , x2 ) = F2 ( x1 , x2 ) =
X2

(3.93)

=constant

=constant
X1

Fig. 3.7
For instance, the usual polar coordinates can be considered as a form of curvilinear coordinates which specify the position of our point as an intersection of a radial line, at the angle from the initial line, with a circle of radius r. Then, we have:
F ( x , x ) = x2 + x2 = r 1 2 1 1 2 x2 F2 ( x1 , x2 ) = arctan = x1

(3.94)

We may also solve Eqs. (3.93) for x1 , x2 and obtain:


x1 = f1 ( , ) x2 = f 2 ( , )

(3.95)

By choosing different functions for f1 and f2 we can find various curvilinear coordinates. Let us assume:
x1 = f1 ( , ) = C cosh cos x2 = f 2 ( , ) = C sinh sin

(3.96)

M. Mirzaei, Elasticity

- 87 -

If we eliminate between the above two equations, we can show that:


2 x12 x2 + =1 C 2 cosh 2 C 2 sinh 2

(3.97)

which for different values of gives different ellipses with the same foci, i.e., a family of confocal ellipses. Alternatively, if we eliminate between the two equations, we will have:
x12 x2 2 2 2 =1 C 2 cos 2 C sin

(3.98)

which for different values of gives different hyprbolas with the same foci, or a family of confocal hyperbolas.
X2

X1

Fig. 3.8
We may also define the complex variable = + i in terms of our curvilinear coordinates by writing the following equalities using Eq. (3.96):
x1 + ix2 = C cosh( + i ) z = C cosh z = f ( )

(3.99)

Now we express the stresses in the curvilinear coordinates in terms of the stresses in the cartesian coordinates by a simple transformation caused by a rotation of degrees:

M. Mirzaei, Elasticity

- 88 -

cos 2 + 12 sin 2 2 2 + 22 11 22 cos 2 12 sin 2 = 11 2 2 = 12 cos 2 11 22 sin 2 2 +

11 + 22

11 22

(3.100)

which leads to:

+ = 11 + 22 + 2i = e 2i ( 22 11 + 2i 12 )

(3.101)

Finally, we use Eqs. (3.83) and (3.82) and write the general expressions for the stresses and displacements in terms of complex potentials in the curvilinear coordinates as:

+ = 4e ( ( z ) ) + 2i = 2e 2i [ z ( z ) + ( z )]
3 ( z ) z ( z ) ( z ) 2G (u iu ) = ei 1 + In order to find the factor e 2i , we write: f ( ) = df ( ) d
i

(3.102)

f ( ) = j cos + ij sin = jei f ( ) = je f ( ) e2i = f ( )

(3.103)

For example, for the elliptic coordinates we can show that: e2i = sinh sinh (3.104)

M. Mirzaei, Elasticity

- 89 -

Uniformly Stressed Sheet with an Elliptic Hole


Consider a large sheet, containing an elliptic hole with semiaxes, a and b, subjected to a far-field uniform stress 0 as depicted in Fig. 3.9. We may define our elliptic coordinates as: z = C cosh ; = + i ; ; dz = C sinh d

x1 = C cosh cos x2 = C sinh sin

C cosh 0 = a C sinh 0 = b

(3.105)

0
X2

b
2C

0
X1

Fig. 3.9
For this case, it is appropriate to use the general expressions for the stresses and displacements in terms of complex potentials in curvilinear coordinates (Eqs. (3.102)) with the following boundary conditions:
x1 , x2 ( ) 11 = 22 = 0 (i)

= 0 = = 0
From Eqs. (3.102) we have:

(ii)

(3.106)

M. Mirzaei, Elasticity

- 90 (i) (ii)

+ = 4e ( ( z ) ) + 2i = 2e 2i [ z ( z ) + ( z )]

(3.107)

which, in combination with our first boundary condition, indicates that we must have:

+ = 4e ( ( z ) ) = 2 0 e ( ( z ) ) = 0 2 z ( z ) + ( z ) = 0

at infinity

(3.108)

In order to find the appropriate complex potentials, we note that they should conform to our elliptic boundary and be periodic in . Hence, we assume the following potentials and check if our boundary conditions are satisfied: ( z ) = AC sinh 2 ( z ) = BC ( z ) = AC cosh d cosh =A = A coth dz sinh

coth 1 e( z ) = A = ( z ) =
BC cosh ( z ) = B sinh sinh 3 A 1 ( z ) = C sinh 3 z ( z ) + ( z ) = 0

0
2

(3.109)

the far field boundary conditions are satisfied. In order to 2 find B, we proceed with the local boundary conditions. From Eq. (3.107) we have: (i) (ii) i = 2e ( ( z ) ) e2i [ z ( z ) + ( z ) ] i = ( z ) + ( z ) e 2i [ z ( z ) + ( z )] where e 2i can be found from Eq. (3.104) as: e2i = sinh sinh (3.111) (3.110)

It is clear that with A =

Accordingly, Eq. (3.110) becomes:

M. Mirzaei, Elasticity cosh cosh + i = A sinh sinh

- 91 sinh cosh cosh +B + A 3 sinh 3 sinh sinh

1 A sinh sinh ( + ) + cosh + B cosh = 2 sinh sinh which may be written at the boundary of the ellipse as:

(3.112)

= 0 + = 2 0 = 2 0 1 = ( A cosh 20 + B ) cosh 2 sinh sinh

(3.113)

Invoking the second boundary condition in Eq. (3.106ii) we have: 1 B = A cosh 2 0 = 0 cosh 2 0 2 1 ( z ) = 0C sinh 2 ( z ) = 1 0 cosh 2 0 C 2 2

(3.114)

Although all the boundary conditions are now satisfied, we must make sure that there would be no discontinuity in the resulting displacement field. Hence, we use Eq. (3.102) to obtain the Cartesian components of displacement as: 2G (u + iv) = 3 BC AC sinh AC cosh coth 1 + sinh (3.115)

which shows that displacements are periodic in and there would be no discontinuity in displacements. The most interesting component of the stress is in fact the largest value of at the hole and can be obtained from Eq. (3.107) as follows:

+ = 4e ( ( z ) ) = 2 0e coth
2 0 sinh 2 sinh 2 i sinh 2 + = cosh 2 cos 2 cosh 2 cos 2 2 0 sinh 2 0 ( ) = = 0 cosh 2 0 cos 2 coth = ( ) =0, =
max

(3.116)

2 0 sinh 2 0 cosh 2 0 1

M. Mirzaei, Elasticity Using Eq. (3.105) we may show that: C 2 = a 2 b2 2ab sinh 2 0 = 2 C a 2 + b2 cosh 2 0 = C2

- 92 -

(3.117)

which, in combination with Eq. (3.116), results in:

( ) ( )

max

= 0,
min

= 2 0

a b b a

= ,

3
2 2

= 2 0

(3.118)

Eqs. (3.118) show that as b 0 (as the ellipse becomes a crack) a stress singularity develops at the crack tip. For a=b, the ellipse becomes a circle with a stress concentration factor of two, which is in agreement with the results we already obtained for the sheet with a circular hole.

Assignment 10:
Find the maximum stresses for an elliptic hole in a large sheet subjected to remote tensile stress T1 . The major axis of the elliptic hole makes an angle with the x1 direction.

M. Mirzaei, Elasticity

- 93 -

Chapter 4 Linear Thermoelasticity


Temperature change can often be a source of significant stresses in elastic bodies. In general, we should define a reference temperature at which we consider the body to be free of thermal stresses. As the temperature is raised or lowered with respect to the reference temperature the body expands or contracts and thermal stresses may arise. It should be note that for a body which is not constrained, a uniform temperature change with respect to the reference temperature will not cause thermal stresses. However, this is not the case for a constrained body, where the thermal stresses are developed even if the temperature changes are uniform throughout the body. In either case, the existence of temperature gradients within an elastic body will result in the development of thermal stresses which are generally tensile at locations of lower temperatures and compressive at locations of higher temperatures. Although later we will examine some special cases where the thermal stresses can be absent also for bodies under temperature gradients. Usually, the solution of the thermoelasticity problem begins with finding the time and position dependent temperature field. We may write the general form of the heat conduction equation as: S = a,ii + Cv t ; a= k Cv (4.1)

in which is the temperature field , S is the strength of the heat sources in the field , is the density, Cv is the specific heat, and k is the thermal conductivity. For a stationary field with no heat sources we have: ,ii = 0 Since we are dealing with linear thermoelasticity, the strain tensor is defined by: (4.2)

ij =

1 ( ui, j + u j ,i ) 2

(4.3)

and also the equilibrium equations are:

ij , j + fi = 0

(4.4)

However, we may split the total strain tensor into the mechanical and thermal parts as:

M. Mirzaei, Elasticity

- 94 -

ij = ij m + ij T
For the mechanical stresses we have:

(4.5)

ij m =

1 + ij kk ij E E E G= 2 (1 + ) ij m = 1 ij kk ij 2G 1 +

(4.6)

In order to obtain an expression for the thermal strains we may write:


dST = (1 + ) dS0

(4.7)

Here dS0 is the initial length of a line element, dST is its length as a result of a temperature change , and is the coefficient of thermal expansion. Accordingly, the thermal strains are:

ij T = ij

(4.8)

We can combine Eqs. (4.6) and (4.8) to obtain the total strains, resulting in the so-called Duhamel-Neumann Constitutive equations:

ij = ij m + ij T =

1 ij kk ij + ij 2G 1 +

(4.9)

In order to find the stresses in terms of the strains, we first contract on i and j indices to obtain:

ll =

1 1 2 ll + 3 2G 1 + 2G (1 + ) ll = ( ll 3 ) 1 2

(4.10)

Substituting back into Eq. (4.9) we have:

M. Mirzaei, Elasticity

- 95 -

ij =

1 2G ij 1 2 ( ll 3 ) ij + ij 2G

1 +1 = ij ll ij + ij 2G 1 2 1 2
Hence, the stresses are:

(4.11)

ij = 2G ij +

1 2

ll ij

2G (1 + ) ij 1 2

(4.12)

Based on our new set of constitutive equations and implementing the procedure we followed before, we may obtain a generalized form for the Navier displacement equations as follows: ui , jj + 2 (1 + ) 1 1 u j ,ij + f i = G 1 2 (1 2 ) ,i (4.13)

We may also derive a generalized form for the Beltrami-Michell equations:

(1 + ) ij ,kk + kk ,ij = (1 + ) fi , j + f j ,i +

f k , k ij 1 1 + 2G (1 + ) ,ij + , kk ij 1

(4.14)

A Thin Circular Disc Problem


As a first example in solving thermoelasticity problems, we will find the thermal stresses in a thin circular disc. We assume that the temperature field is symmetrical about the center of the disc and does not vary over the thickness. For this case the equilibrium equations simplify to: d rr rr + =0 dr r (4.15)

We now repeat the procedure implemented above for determination of the thermoelastic constitutive equations for our specific problem. Starting with the strains we write: 1 ( rr ) E 1 = ( rr ) E which can be written in terms of the stresses as:

rr =

(4.16)

M. Mirzaei, Elasticity

- 96 -

rr =

E rr + (1 + ) 1 2 E = + rr (1 + ) 1 2

(4.17)

Substituting the above expressions into our equilibrium equation (Eq. (4.15)) we have: r d d ( rr + ) + (1 )( rr ) = (1 + ) r dr dr (4.18)

which can be rewritten in terms of displacements to give:


d 2ur 1 dur ur d + 2 = (1 + ) 2 dr r dr r dr d 1 d (rur ) d = (1 + ) dr dr r dr

(4.19)

This is in fact nothing but the simplified form of the generalized Navier equations for the current problem. The above differential equation can be solved for the radial displacement to give: ur = (1 + ) C 1 r a rdr + C1r + r2 r (4.20)

Accordingly, the stresses become:

rr = E

1 C1 (1 + ) C2 (1 ) r 2 1 r E 1 = E 2 rdr E + C 1 + ) C2 (1 ) 2 2 1( a r 1 r 1 r2

rdr +

E 1 2

(4.21)

The constants can be found from the boundary conditions. For a solid disc, a can be taken as zero and we also have: lim 1 r rdr = 0 r 0 r 0 (4.22)

Note that Eq. (4.20) implies that we must have C2 = 0 to ensure that the displacements at the center of the disc have finite values. Moreover, at the outer edge of the disc we have rr = 0 , so we may write:

M. Mirzaei, Elasticity C1 = (1 )

- 97 (4.23)

b
2

rdr

Finally, the stresses can be shown to be:

rr = E

rdr 1 b 1 r = E + 2 rdr + 2 rdr 0 0 b r

1 2 b

rdr

1 r2

(4.24)

Assignment 11:
Find the thermal stresses for a hollow sphere for which the inner and outer surfaces are kept at constant temperatures Ti and To respectively.

Instantaneous Point Source in an Infinite Body


In this example we consider a quantity of heat Q deposited at time t = 0 at Xi = 0 inside an infinite linear elastic body. The heat will gradually spread into the body, and every point will experience an increase and later decrease in temperature with time. Eventually, the body will return to its initial temperature after a long period of time. In order to obtain the resulting thermal stresses, we must first find the time and position dependent temperature field. The problem can be treated as spherically symmetric, so we may start with the proper form of the equation of heat conduction in spherical coordinates: 1 = a 2 r2 t r r r Defining the Laplace transform as:
( s ) = (t )e st dt 0

(4.25)

(4.26)

we have: = s (4.27)

t The Laplace transform of Eq. (4.25) results in:

M. Mirzaei, Elasticity a d 2 d r 2 r dr dr

- 98 (4.28)

s =

The solution to the above differential equation can be shown to be:


A = e r s r a B + e r s r a

(4.29)

Note that we should set B = 0 to ensure that remains bounded as r . In order to find the constant A, we write:
Q = 4 r 2 Cv dr
0

(4.30)

We also define Q using the Heaviside step function as:

Q = QH (t )

(4.31)

Next, we take the Laplace transform of both sides of Eq. (4.30) and proceed as follows:
1 Q = 4 r 2 Cv dr s 0

= 4 Cv A e
0

s a r

rdr

(4.32)

= 4 Cv A

a s

Accordingly, we find the constant A as: A= Q 4 Cv a (4.33)

Finally, we find the time and position dependent temperature field for our problem as follows:
= =
Q e 4 Cv ar s r a

(4.34)
e
r2 4 at

Q
3/ 2 ( 4 at ) Cv

M. Mirzaei, Elasticity

- 99 -

The next step in solving our thermoelasticity problem is the determination of the displacement field. We start with the generalized form of the Navier equations (Eqs. (4.13)): ui , jj + 2 (1 + ) 1 1 u j ,ij + f i = G 1 2 (1 2 ) ,i (4.35)

The solution to the above equation consists of the general solution uih plus the particular integral uip . Now, we may assume that the particular solution can be derived from the thermoelastic potential as follows:
uip = ,i

(4.36)

If we substitute the above expression into the Generalized Navier equations, we will have: 2 (1 + ) 1 , jj = 0 , jj + 1 2 1 2 ,i (1 + ) , jj = 2 = (1 )

(4.37)

Assuming a non-stationary source-free temperature distribution, a solution for the above equation can be written as: =

(1 + ) a t dt + + t 0 1 (1 ) 0

(4.38)

where 2 1 = 0 , and 0 represents the initial thermoelastic potential. We now turn our attention to the homogeneous form of the Navier equations. For the current problem, since there are no body forces or surface tractions, we may assume: uih = 0 (4.39)

Substituting the obtained temperature field into the general solution for the thermoelastic potential, we can write: =
t e (1 + ) Q a 3/ 2 0 t 3/ 2 (1 )( 4 a ) Cv

r2 4 at

dt + 0 + t 1

(4.40)

In order to ease the formulations, we may further consider;

M. Mirzaei, Elasticity

- 100 -

r r2 V2 = V= 4at 2 at 2 r r3 t 3/ 2 = 3/ 2 3 t= 2 8a V 4aV 2 r dt = dV 2aV 3 which upon substitution into Eq. (4.40), results in:

(4.41)

(1 + ) Q 2 2 rat eV dV 0 t1 = (1 ) 4 r Cv (1 + ) Q 2 0 eV dV + 2 eV dV = r (1 ) 4 r Cv 2 at 0 r (1 + ) Q 2 2 at eV dV + 2 1 = (1 ) 4 r Cv 0 2 (1 + ) Q 1 erf r = (1 ) 4 r Cv 2 at
2 2 2 2

(4.42)

The error function is: r erf 2 at 2 =

r 2 at 0

e V dV

(4.43)

Now we will determine 1 and 0 values of the thermoelastic potential. After a long period of time the displacements and stresses return to zero, so we may write: 1 = 0 (1 + ) Q 0 = t = = (1 ) 4 r Cv (1 + ) Q erf r = (1 ) 4 r Cv 2 at

(4.44)

Having found the thermoelastic potential, the displacement, strain, and stress fields can be determined as follows:

M. Mirzaei, Elasticity
ui = uih + uip = ,i

- 101 -

ij = ,ij ij = 2G ,ij +

1 2

,kk ij

(1 + )
1 2

(4.45)
ij

Two-Dimensional Thermoelastic Problems


Plane strain
Here we start by reviewing some of the plane strain formulations:
u1 = u1 ( x1 , x2 ) u = u ( x ) u2 = u2 ( x1 , x2 ) u3 = 0, constant 11 1 ij = 2 ( ui , j + u j ,i ) , + f = 0 f = V, 11 V = ,22 22 V = ,11 12 = 21 = ,12

, = 1, 2

12 0 22 0
0

(4.46)

The two-dimensional form of the Duhamel-Neumann constitutive relations is:

= 2G +

1 2

(1 + ) 1 2

(i) (ii) (iii) (4.47)

3 = 2G 3 = 0 33 = 2G
(1 + ) 1 2 1 2

In order to find the appropriate expression for 33 we need to find . We start by contracting on and indices in Eq. (4.47i), as follows:

M. Mirzaei, Elasticity

- 102 -

= 2G +

1 2 = + 2 (1 + ) 2G

2 2 (1 + ) 1 2 1 2

(4.48)

Substituting for in Eq. (4.47iii) results in:

33 = 2G

(1 + ) 2G 33 = E

(4.49)

which indicates that 33 may be determined from a knowledge of and . Furthermore, we have:

1 + (1 + ) 2G

(4.50)

which is the two-dimensional form of the Duhamel-Neumann relation. Now, using the above expression, and implementing the same approach we followed in Chapter 3, we can derive the biharmonic equation for linear thermoelasticity as follows: 1 2 2 1 + 2 4 = V 2G 1 1 If body forces are negligible or constant we may write: 1 + 4 = 2G 1 2 (4.52) (4.51)

The solution to the above biharmonic equation is:

= h + p

(4.53)

where h can be found from the elasticity part of the problem without considering the thermal effects. In order to find the particular integral, we use the definition of the thermoelastic potential, Eqs. (4.36), in two-dimensional form:
up = ,

(4.54)

which satisfies the two-dimensional form of Eq. (4.37):

M. Mirzaei, Elasticity

- 103 (4.55)

, = 2 =

(1 + ) (1 )

If we differentiate the above equation twice we will have:

4 =

(1 + ) 2 (1 )

(4.56)

Substituting into our biharmonic equation (Eq. (4.52)) we can write:


4 p = 2G4 p = 2G

(4.57)

Naturally, the isothermal deformation satisfies the homogeneous form of Eq. (4.52), and as a result the complete solution for the stress function is:

= h + p = h 2G

(4.58)

Plane Stress Formulation


We start by reviewing some of the plane stress formulations:

ij = ij

1 h/2 ij dx3 h h / 2 11 12 0 22 0 0

(4.59)

The bar sign over a quantity indicates that it has been averaged over the thickness. The equilibrium equations can be written as: f = V, , V, = 0 (4.60)

which are exactly the same as those derived for plane strain condition, except for the mean value interpretation, so we may similarly write:

M. Mirzaei, Elasticity

- 104 -

11 V = ,22 22 V = ,11 12 = 21 = ,12


The constitutive equations are: 1 1 + + 2G 1 3 = 3 = 0 2G 1 33 = 1 + + 2G (4.61)

(4.62)

which can be inverted in a standard manner to give:

2G (1 ) + (1 + ) 1

(4.63)

Now, similar to our plane strain formulation, upon substitution for the strains into the only nonzero compatibility equation we find that: 4 = E2 The solution to the above biharmonic equation is: (4.64)

= h + p

(4.65)

where h can be found from the corresponding isothermal elasticity problem. In order to find the particular integral, we use the plane stress form of the thermoelastic potential expression to write: 2 = , = (1 + ) 4 = (1 + ) 2 using (4.49) 4 = p = 2G E 4 (1 + ) (4.66)

M. Mirzaei, Elasticity

- 105 -

Solution Strategies for Two-Dimensional Problems


As mentioned before, the solution of the thermoelasticity problems begins with finding the time and position dependent temperature field. For the two-dimensional problems, the next step is the solution of either of the following expressions for the thermoelastic potential: 2 = , =

(1 + ) (1 )

; 2 = , = (1 + )

(4.67)

which using p = 2G will give the particular solution. The solution to the isothermal elasticity problem can be found using the methods introduced earlier, or it can be constructed as a finite or infinite series of properly chosen biharmonic functions. In Cartesian coordinates we may use: e x2 cos x1 ; x1e x2 cos x1 ; x2 e x2 cos x1 (4.68)

where is an arbitrary constant. All other combinations obtained by interchanging x1 and x2 and/or replacing cos x1 by sin x1 can also be used. In Polar coordinates, a complete set of biharmonic functions which are periodic in are:
r 2 ; log r ; r 2 log r r log r cos ; r n cos n ; r 2 n cos n ; n = 0,1, 2,K

(4.69)

and the corresponding expressions with cos n replaced by sin n can also be used.

Assignment 12:
Using the stress function approach, find the thermal stresses for a long circular hollow cylinder for which the inner and outer surfaces are kept at constant temperatures Ti and To respectively. The cylinder is also under internal pressure Pi.

Two-Dimensional Simply-Connected Regions with Steady Heat Flow


Here, we will show that there are special cases where the thermal stresses can be absent even if the body experiences temperature gradients. Let us consider a steady heat flow which is confined to the x1x2 plane, for example, consider a long cylinder with no variation of temperature in the axial direction. The temperature field will satisfy the 2D form of the Laplaces equation:

M. Mirzaei, Elasticity

- 106 -

, = 0 and the constitutive equations are:

(4.70)

11 (1 + ) =

1 2 22 11 E 1

1 2 22 (1 + ) = 11 22 E 1 Now, in order to maintain 11 = 22 = 0 , we should have:

(4.71)

11 = (1 + ) 22 = (1 + )

(4.72)

The above strains should satisfy the compatibility equations, which in this case simplify to:

11,22 + 22,11 = 0

(4.73)

which is identically satisfied if we substitute Eqs. (4.72) into Eq. (4.70). Hence we may conclude that the thermal stresses can be absent for two-dimensional simply-connected regions with steady heat flow (even for bodies under temperature gradients). As we will see in the next example the situation is different for non-simply-connected regions.

Insulated Circular Hole inside an Infinite Plate subjected to Uniform Heat Flow
In this example we assume a uniform heat flow in the direction of the negative x2 axis in an infinite plate. For this case we may write: = qx2 (4.74)

where q is the constant temperature gradient. As depicted in Fig. 4.1, the uniform heat flow is disturbed by the presence of an insulated circular hole.

M. Mirzaei, Elasticity

- 107 -

X2

X1

Fig. 4.1 In order to obtain the thermal stresses, we should first find the temperature field. The two-dimensional form of Eq. (4.2) for plane stress is:
, = 0

(4.75)

which can be written in cylindrical coordinates as: 2 1 1 2 + + =0 r 2 r r r 2 2 (4.76)

We may solve the above partial differential equation using the separation of variables method as follows: = R(r ) X ( ) Accordingly we may write: dR X = r dr 2 d 2 R = 2 X dr r 2 dX =R d 2 d2X ; =R d 2 2 ; (4.77)

(4.78)

Substituting the above expressions into Eq. (4.76), we have:

M. Mirzaei, Elasticity d 2R 1 dR R d2X =0 X+ X+ 2 dr 2 r dr r d 2 r 2 d 2 R r dR 1 d 2 X + + =0 R dr 2 R dr X d 2

- 108 -

(4.79)

which upon splitting into two parts results in: r2 R 1 X d 2 R r dR + = k2 2 dr R dr 2 d X = k 2 d 2 (i) (4.80) (ii)

Equation (4.80i) can be rearranged as the Cauchy-Eulers equation: r2 d 2R dR +r k 2R = 0 2 dr dr (4.81)

which can be solved using the following considerations: r = e z ; z = ln r dR dR dz 1 dR = = dr dz dr r dz d 2R 1 dR 1 d 2 R 2 = 2 + dr r dz r 2 dz 2 With proper substitutions into Eq. (4.81), the solution can be obtained as follows: d 2R = k 2R 2 dz R = A1e kz + A2 e kz R = A1r k + A2 r k Now we turn our attention to Eq. (4.80ii), which can be solved as the Eulers equation:
d2X + k2X = 0 d 2 X = B1eik + B2 eik X = C1 sin k + C2 cos k

(4.82)

(4.83)

(4.84)

Finally, the temperature field can be written as:

M. Mirzaei, Elasticity

- 109 (4.85)

= RX = ( A1r k + A2 r k ) ( C1 sin k + C2 cos k )

for which the four constants can be determined from the boundary conditions. Also note that we have:
= k ( A1r k 1 A2 r k 1 ) ( C1 sin k + C2 cos k ) r

(4.86)

The first temperature boundary condition is:

= 0 ; = 0 r
C2 = 0 Thus, Eqs. (4.85) and (4.86) can be rewritten as:
= ( A1C1r k + A2C1r k ) sin k = ( D1r k + D2 r k ) sin k = ( D1r k 1 D2 r k 1 ) k sin k r

(4.87)

(4.88)

The second temperature boundary condition is: = 0 r D2 = D1a k 1+ k +1 = D1a 2 k r=a ; This time, Eq. (4.88) can be rewritten as: a 2k = D1r 1 + 2 k sin k r a 2k = D1kr k 1 1 2 k sin k r r
k

(4.89)

(4.90)

The third temperature boundary condition is:

= =q r x2 =q

a 2k D1kr k 1 1 2 k r

(4.91)

The above expression must remain valid as r so we may conclude that:

M. Mirzaei, Elasticity

- 110 -

k 1 = 0 k = 1 D1 = q

(4.92)

Finally, the temperature distribution can be found as: a2 = q r sin + sin r a2 = q x2 + sin r

(4.93)

The above expression shows that the uniform temperature field qx2 remains unchanged at a2 sin term. As r discussed before, the uniform temperature field has no influence on the thermal stresses and may be disregarded in the process of determination of the thermoelastic potential from its governing equation as follows: the far field and the effect of the circular hole is represented by the q 2 = (1 + ) a2 2 1 1 2 + + 2 = (1 + ) q sin r r 2 r r r 2 2 (1 + ) qa r log r sin = 2 (4.94)

Having found the thermoelastic potential, we may easily calculate p from p = 2G . In order to have some clue to the proper form of h , we may proceed as follows:

= h + p = h 2G
h p rr = rr + rr

1 1 2 h = rr 2G + 2 2 r r r 2G (1 + ) qa 2 1 h = rr sin 2 r E qa 2 = sin 2r
h rr

(4.95)

Now we may apply our first overall stress boundary condition:

M. Mirzaei, Elasticity

- 111 -

rr = 0 at r = a E qa h rr = sin r =a
2 1 1 2 h + 2 2 r r r
r =a

E qa = sin 2

(4.96)

which suggests the following form for h :

h =

E qa f (r ) sin ; f (a) = 1 2

(4.97)

Similarly, for the shear stresses we have:

r = rh + rp
1 = rh 2G r r E qa 2 = rh + cos 2r (4.98)

Applying our second overall stress boundary condition, we will have:

r = 0 at r = a E qa rh = cos r =a
2 1 h r r
r =a

E qa = cos 2

(4.99)

which approves the previously suggested form for h . However, according to our third overall boundary condition all the stresses should vanish as r . Accordingly, we may modify the expression (4.97) as:

h =

E qa 1 f sin ; f (a ) = 1 2 r

(4.100)

Here, we consult the list of the biharmonic functions in polar coordinates and finalize our expression as:

h =

E qa 1 Ar sin 2

(4.101)

Accordingly, we can write:

M. Mirzaei, Elasticity

- 112 -

h rr

E qa sin r =a 2 1 1 2 h AE qa sin r =a = + 2 2 a3 r r r =
3

(i) a3 (i),(ii) A = 2 (ii)

E qa a = sin 2 r
h rr

(4.102)

rr =

E qa a a 3 sin 2 r r3

The other stress components can be calculated using h as follows:

h =
r

E qa 4 sin 4r E qa a a 3 = cos 2 r r3 E qa a a 3 + sin 2 r r3

(4.103)

Zero Thermal Stresses in a Three-Dimensional Body


As mentioned before, there are some special cases where the thermal stresses can be absent, even if the body experiences temperature gradients. Previously, we examined the steady heat flow in two-dimensional simply-connected regions. We may also arrive at zero thermal stresses for an unrestrained body when the temperature distribution is a linear function of rectangular space coordinates. For the stresses all to be zero, the six stress compatibility equations result in:

,11 = ,22 = ,33 = ,12 = ,23 = ,31 = 0

(4.104)

Hence, the thermal strain distribution must be a linear function of the rectangular space coordinates like:

= a0 + a1 x1 + a2 x2 + a3 x3

(4.105)

M. Mirzaei, Elasticity

- 113 -

Note that the coefficients may be time dependent. Accordingly, if the thermal properties of the material are constant, the zero stress condition dictates a linear temperature distribution.

Assignment 13:
Consider an unrestrained circular disc for which we have = Ar . Based on our previous discussions concerning the special cases for zero thermal stress, can we conclude that the disc is stress free? Why?

M. Mirzaei, Elasticity

- 114 -

Chapter 5 Variational Principles and Applications


Leonhard Euler: Since the fabric of the universe is most perfect, and is the work of a most wise creator, nothing whatsoever takes place in the universe in which some relation of maximum and minimum does not appear. Wherefore, there is absolutely no doubt that every effect in universe can be explained as satisfactorily from final causes, by the aid of the method of Maxima and Minima, as it can from the effective causes themselves. Ref: Timoshenko, S.,
History of Strength of Materials (McGraw-Hill Book Company, Inc., New York, 1953)

Energy Principles
As clearly stated by Leonhard Euler in the 18th century, the field equations of the theory of elasticity may also be developed from the energy considerations. In particular, these methods provide powerful tools for obtaining approximate solutions and implementation of numerical solution techniques.

Principle of Virtual Work


The virtual work can be defined as the work done on a deformable body, by all the forces acting on it, as the body is given a small hypothetical displacement which is consistent with the constraints present. The virtual displacements are represented by the symbol , known as the variation operator. In general, the loadings consist of body forces and surface tractions. The later are prescribed over a part of the boundary designated by S . Over the remaining boundary (designated by Su ) the displacement field u is prescribed. However, it must be ensured that u = 0 on Su to avoid violating the constraints.
T

Su Su
f

Fig. 5.1 Accordingly, the virtual work is defined by:

M. Mirzaei, Elasticity

- 115 -

Wvirt = T. u + f . u
S V

(5.1)

= Ti ui dS + f i ui dV
S V

Now we may invoke the stress boundary relations and implement the divergence theorem to obtain:

T u dS + f u dV =
i i i i S V S

ij

n j ui dS + fi ui dV
V

= ( ij ui ) + f i ui dV ,j V
=0 = ( ij , j + f i ) ui + ij ( ui , j ) dV V

(5.2)

The first grouping of terms within the last integral can be set to zero to ensure the equilibrium. The product of the symmetric stress tensor with the skew-symmetric part of ui , j is also zero. Since the symmetric part of ui , j is nothing but a variation in the strain tensor, we may write the following expression, known as the principle of virtual work, PVW:

T u dS + f u dV =
i i i i ij S V V

ij

dV

(5.3)

It is interesting to note that the above equation is independent of any constitutive law and can be applied to all continuous materials within the limitations of small deformations.

Illustrative Example I
Consider a planar truss as depicted in Fig. 5.2. It is assumed that all the members have the same cross-sectional area and Youngs modulus. If we neglect the gravitational effects, the PVW for this problem can be written as: P u = ij ij dV
V

(5.4)

M. Mirzaei, Elasticity

- 116 -

8m (1) 30o A 30o

(2) 10m

X1
(3) 10m

X2
1000N
Fig. 5.2

For this problem, we may only consider virtual displacements for the point A. The changes in the length of each rod due to the imposed virtual displacements u are:

L1 = u1 cos 60 ; X L2 = u1 cos 30 ; X L3 = u1 cos 30 ;


X1
1 1

L1 = u2 sin 60 X L2 = u2 sin 30 X L3 = u2 sin 30


X2
2 2

(5.5)

In the above, X Li represent the components of the virtual elongation (or contraction) of the members in the X directions. Note that, u have been chosen positive in the positive direction of the axes. Accordingly, we can calculate the components of the virtual strains as: cos 60 ; 8 cos 30 X1 2 = u1 ; 10 cos 30 X1 3 = u1 ; 10
X1

1 = u1

sin 60 8 sin 30 X2 2 = u2 10 sin 30 X2 3 = u2 10


X2

1 = u2

(5.6)

where, X i represent the components of the virtual strain of the members in the X directions. Now, we will calculate the real stress in each rod, using the real axial strain caused by the real displacement of the point A.

M. Mirzaei, Elasticity

- 117 -

cos 60 sin 60 cos 60 sin 60 + u2 1 = Eu1 + Eu2 8 8 8 8 cos 30 sin 30 cos 30 sin 30 2 = u1 + u2 2 = Eu1 + Eu2 10 10 10 10 cos 30 sin 30 cos 30 sin 30 3 = u1 u2 3 = Eu1 Eu2 10 10 10 10

1 = u1

(5.7)

Accordingly, we can apply the PVW in the horizontal direction as follows:


cos 60 cos 60 sin 60 cos 60 P u1 = 8 AE u1 1 u1 + u2 u1 8 8 8 8 cos 30 cos 30 sin 30 cos 30 +10 AE u1 u1 + u2 u1 10 10 10 10 cos 30 cos 30 sin 30 cos 30 +10 AE u1 u1 u2 u1 10 10 10 10

(5.8)

which may be simplified to give: P 1 = 0.181u1 + 0.054u2 AE and also vertically as:
cos 60 sin 60 sin 60 sin 60 P2 u2 = 8 AE u1 u2 + u2 u2 8 8 8 8 cos 30 sin 30 sin 30 sin 30 +10 AE u1 u2 + u2 u2 10 10 10 10 cos 30 sin 30 sin 30 sin 30 +10 AE u1 u2 u2 u2 10 10 10 10

(5.9)

(5.10)

which upon simplifying gives the second required equation: P2 = 0.054u1 + 0.143u2 AE Now, we may set: P = (0 , 1000) , and solve Eqs. (5.9) and (5.11) to give: u1 = 2342 AE ; u2 = 7846 AE (5.12) (5.11)

M. Mirzaei, Elasticity

- 118 -

Principle of Complementary Virtual Work


In order to establish another useful variational principle, we may vary the stress field and applied forces while keeping the displacement field fixed.
Wvirt = u. TdS + u. fdV

(5.13)

= ui Ti dS + ui fi dV
S V

Using an approach similar to what used for PVW, we may write the Principle of Complementary Virtual Work, PCVW, as:

u T dS + u f dV =
i i i i ij S V V

ij

dV

(5.14)

Illustrative Example II
Consider a simple pin-connected structure as depicted in Fig. 5.3. The aim is to determine the vertical movement of the point C. We assume linear elastic behavior, identical cross section, and same modulus of elasticity for all the members.
1000N D 500 1433 10m 10m 612 E 60o 1000 -866 -1412 1000N C F 1433 10m -2027 0.5F -0.707F 0.5F 0.707F F

Fig. 5.3

M. Mirzaei, Elasticity

- 119 -

Since we require the vertical displacement of the point C, we apply a virtual vertical load at this point. Next, we will obtain the virtual stresses due to this virtual load.

CD = AC AB

0.707 F A 0.707 F = A 0.5 F = BD = A

(5.15)

Now we obtain the real strains in the relevant members. Note that we only consider those members which are affected by the virtual load.

CD = AC AB

612 AE 2027 = AE 1433 = BD = AE

(5.16)

Accordingly, the principle of complementary virtual work will be:

Fv =

612 0.707 F 2027 0.707 F 14.14 A + 14.14 A A A AE AE 1433 0.5 F +2 10 A AE A

(5.17)

which can be solved for the displacement to give: v= 40720 AE (5.18)

Principle of Total Potential Energy


The general form of the constitutive expressions for an elastic continuum can be written as:

ij =

ij

(5.19)

M. Mirzaei, Elasticity

- 120 -

in which is the density and is the strain energy density. If we substitute the above expression for stresses in Eq. (5.3), we will have:

T u dS + f u dV =
i i i i S V V S V V

ij

ij dV
(5.20)

Ti ui dS + fi ui dV = dV = dV = U
V

where U is the total strain energy stored in the body. The left side of Eq. (5.20) may be defined as a variation in the potential energy, V , so we may write: V = U (U + V ) = 0 or = 0 in which is the total potential energy of the body. (5.21)

Strain Energy
For elastic bodies the strain energy can be defined as: U =

ij

ij d ij dV

(5.22)

If we substitute the constitutive equation for the linear elastic response of isotropic materials into Eq. (5.22), we have: U = kk jj + ij ij dV V 2 for which the one-dimensional form is: U = (5.23)

EededV =

1 1 2 V Ee dV = 2 V edV 2

(5.24)

Assignment 14:
Solve the illustrative example I, using the principle of total potential energy.

M. Mirzaei, Elasticity

- 121 -

Principle of Total Complementary Energy


The principle of total complementary energy can be derived from the principle of complementary virtual work by considering a specific complementary strain energy density as a function of stresses such that:

ij =

ij

(5.25)

Accordingly, we may write:

ui Ti dS + ui fi dV =
S V V V

ij dV ij
V

= dV = dV = U

(5.26)

where U is the complementary strain energy stored in the body. The left side of Eq. (5.26) may be defined as a variation in the potential energy, V , so we may write: V = U (U + V ) = 0 or = 0 in which is the total complementary energy of the body. (5.27)

Complementary Strain Energy


For elastic bodies the complementary strain energy can be defined as: U =

ij

ij d ij dV

(5.28)

If we substitute the constitutive equation for the linear elastic response of isotropic materials into Eq. (5.28), we will have: 1 U = kk jj + ij ij dV V G 2E for which the one-dimensional form is: (5.29)

M. Mirzaei, Elasticity U =

- 122 1 2 1 V E dv = 2 V edV 2 (5.30)

d dV =

It is clear that only for linear elasticity we have U = U .

The First Castigliano Theorem


In this section we will derive the first Castigliano theorem using the principle of total potential energy. Consider an elastic body maintained in a state of static or dynamic equilibrium as depicted in Fig. 5.4. The body rests on a system of rigid supports while being under the action of externally applied loads.
QN

Q2

Fig. 5.4 Hence, we may write:


= U ( i ) Q j j ; V = Q j j
j =1 j =1 N N

(5.31)

Implementing the principle of total potential energy, we have:

=
i =1 N

i = 0 i

U i Q i i = 0 i =1 U Q i = i ; i = 1,K , N

(5.32)

M. Mirzaei, Elasticity

- 123 -

The above equation shows that if the strain energy can be computed as a function of displacements, the related force or torque for a particular displacement component can be determined by differentiation.

Illustrative Example III


Consider the problem shown in Fig. 5.5. We intend to determine the force required to cause the joint to descend a given distance .

Fig. 5.5

The deformed length of each member is:


L + cos

(5.33)

Hence, we may write:


e= cos L
e

e2 U = 2 dedV = 2 ALE V 0 2 EA cos 2 2 = L

(5.34)

Now, using the first castiglianos theorem, we have: P= dU 2 AE = cos 2 d L (5.35)

M. Mirzaei, Elasticity

- 124 -

The Second Castigliano Theorem


Here, we utilize the principle of total complementary energy.
= U ( Q i ) Q j j ; V = Q j j
j =1 j =1 N N

(5.36)

Accordingly, we may write:

=
i =1 N

Qi = 0 i Q

U i Q i = 0 i i =1 Q U i = ; i = 1,K , N Q i

(5.37)

The above equation shows that if the complementary strain energy can be computed as a function of forces, the related displacement for a particular force component can be determined by differentiation.

An Introduction to Energy Functionals


In general, we may present a partial differential equation by:

Lu = f

(5.38)

where L is an operator acting on the dependent variable u , and f is a function of the independent variable and is called the driving or forcing function. We may also define the inner product of two functions, g and h , over the domain of the problem, V, as:

( g, h ) = V g hdV

or

( g , h ) = V ghdV

(5.39)

If the operator L is self-adjoint or symmetric, we will have:

( Lu, v ) = ( u , Lv )

(5.40)

In our discussions we will also require that L be positive definite for all functions u satisfying homogenous boundary conditions:

( Lu, u ) > 0

(5.41)

M. Mirzaei, Elasticity

- 125 -

The general expression for homogenous boundary condition is:

ui +

ui =0 n

(5.42)

where and are constants and n is the normal to the boundary. Now we may define a function I (u ) , called a quadratic functional, such that: I (u ) = 1 ( Lu, u ) (u, f ) 2 (5.43)

It can be shown that if the equation Lu = f has a solution, and if L is a self-adjoint positive definite operator, then the function u that minimizes I (u ) is the solution to Lu = f . The minimization procedure can also be used to find approximate solutions.

Variational Principles Applied to Beams


We start with the definition of the strain energy for a linear elastic beam:
U =U =

ij

ij d ij dV

1 l h/2 = b 1111dx3 dx1 2 0 h / 2 l h/2 1 2 = bE 11dx3 dx1 0 h / 2 2 l h/2 2 1 = bE ( x3,11 ) dx3 dx1 0 h / 2 2 2 EI l = 0 (,11 ) dx1 2 The potential V is defined by:
V = f i ui dV Ti ui dS = q dx1
V S l

(5.44)

(5.45)

Finally, the total potential energy can be written as:


l EI 2 = U + V = (,11 ) q dx1 0 2

(5.46)

which has the same form as Eq. (5.43).

M. Mirzaei, Elasticity

- 126 -

The Ritz Technique


Based on the above definitions it may be shown, either by the application of variational principles to the Navier displacement equations of equilibrium or by comparing Eq.(5.46) with Eq.(5.43), that for homogenous boundary conditions the quadratic functional is: I (u ) = (5.47)

Hence, in order to find approximate solutions to our beam problem, we may apply the Ritz technique to the total potential energy. The procedure starts with the approximate displacement field defined as:
ui( n ) = ai( j )i( j )
j =1 n

(5.48)

The ai( j ) are called the Ritz coefficients and i( j ) are known as coordinate functions. The coefficients ai( j ) are determined so as to minimize the total potential energy ( n ) , which implies that: ( n ) = 0 , j = 1,L , n ai( j ) The above differentiation results in 3n equations for the unknown coefficients. (5.49)

Illustrative Example IV
Let us find approximate expressions for the deflection of the beam shown in Fig. 5.6.

P L/2 X1

X3
Fig. 5.6

M. Mirzaei, Elasticity

- 127 -

The exact solution to this problem is:


3 PL3 x1 x1 L = 3 4 ; x1 < 48 EI L L 2

(5.50)

We start by finding the total potential energy as follows:


L EI 2 = U + V = (,11 ) dx1 P x = L 1 0 2 2

(5.51)

We may choose our coordinate function as:

(1) = a (1) sin

x1
L

(5.52)

Accordingly the total potential energy can be written as: (1) = EI (1) 2 L ( a ) L 2 Pa(1) 2
4

(5.53)

Now we may find the coefficients through the minimization of (1) and determine the displacement function as follows:
(1) EIL (1) a P=0 = a (1) 2 L 2 PL3 a = 4 EI x 2 PL3 (1) = 4 sin 1 L EI
(1) 4

(5.54)

Finally, we compare the obtained result with the exact solution at two different points: L 4 (exact) 0.01432 0.01452 L 2 0.02083 0.02053

EI PL3

EI (1) (Ritz) PL3 which shows a very good agreement.

Potrebbero piacerti anche