You are on page 1of 25

M.

Mirzaei, Elasticity -1-

Theory of Elasticity
Lecture Notes: 1
Majid Mirzaei, PhD
Associate Professor
Dept. of Mechanical Eng., TMU
mmirzaei@modares.ac.ir
http://www.modares.ac.ir/eng/mmirzaei/elasticity.htm

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 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 physical quantities are very often 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 correspond to
dyadics. Higher order tensors such as triadics, or tensors of order three, and tetradics, or
tensors of order four can also be defined.

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
M. Mirzaei, Elasticity -2-

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 three-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, since 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


Figure 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 C
Configuration

t=t0 R
u
C0
x2
R0
x1
X2
b

x3
X1
X3
Figure 1
M. Mirzaei, Elasticity -3-

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 = χ ( X, t ) (1-1)

Now we consider the two coordinate systems to be coincident, as shown in Figure 2. The
displacement of a material point is (see Fig 1):

u(t ) = x(t ) − X (1-2)

C0 u x
R0 X R
X2 x2 b C
t
X1 undeformed Deformed
x1 Configuration
Configuration
X3 x3
Figure 2

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

dS1 ds1

dS2 ds 2

undeformed Deformed
Configuration Configuration

Figure 3
M. Mirzaei, Elasticity -4-

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

ds1 ds2 (1-3)


=
dS1 dS2

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 (1-4)


(dS ) 2 = dX 12 + dX 22 + dX 32 = dX i dX i = dX ⋅ dX

u+du
dS ds
u

Deformed
undeformed
Configuration Configuration
Figure 4

Referring to Fig.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 (1-6)

and,

xi = xi ( X 1 , X 2 , X 3 )
∂xi ∂x ∂x
dxi = dX 1 + i dX 2 + i dX 3 (1-7)
∂X 1 ∂X 2 ∂X 3
dxi = xi , j dX j = F ⋅ dX
M. Mirzaei, Elasticity -5-

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 − δ ijδ ik ) dX j dX k

= ⎡( X i + ui ), j ( X i + ui ), k − δ jk ⎤ dX j dX k
⎣ ⎦ (1-9)
= ⎡⎣(δ ij + ui , j ) (δ ik + ui , k ) − δ jk ⎤⎦ dX j dX k

= ⎡⎣δ jk + u j , k + uk , j + ui , j ui ,k − δ jk ⎤⎦ dX j dX k
= ( u j ,k + uk , j + ui , j ui ,k ) dX j dX k

Writing the indices in a more proper form, we have,

(ds ) 2 − (dS ) 2 = ( ui , j + u j ,i + uk ,i uk , j ) dX i dX j (1-10)

= 2ε ijL dX i dX j

in which we distinguish the Lagrangian strain tensor ε ijL for characterization of the
deformation near a point.

1 (1-11)
ε ijL =
2
( ui, j + u j ,i + uk ,iuk , j )
Alternatively we may write,

X i = X i ( x1 , x2 , x3 )
∂X i ∂X ∂X
dX i = dx1 + i dx2 + i dx3 (1-12)
∂x1 ∂x2 ∂x3
dX i = X i , j dx j = F ⋅ dx

and proceed as follows,


M. Mirzaei, Elasticity -6-

(ds ) 2 − (dS ) 2 = dxi dxi − X i , j dx j X i ,k dxk = dx ⋅ dx − (F ⋅ dx) ⋅ (F ⋅ dx)


= (δ ijδ ik − X i , j X i ,k ) dx j dxk (1-13)
= ⎡δ jk − ( xi − ui ), j ( xi − ui ),k ⎤ dx j dxk
⎣ ⎦
= ⎡⎣δ jk − (δ ij − ui , j ) (δ ik − ui , k ) ⎤⎦ dx j dxk

= ⎡⎣δ jk − δ 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

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


= 2ε ijE dxi dx j (1-14)

This time, we define the Eulerian strain tensor ε ijL to characterize the deformation near a
point.

1 (1-15)
ε ijE =
2
( ui, j + u j ,i − uk ,iuk , j )
We need to make a number of assumptions to simplify the equations of linear elasticity.
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 linear measure of deformation in linear elasticity, and define the
infinitesimal strain tensor:

1
ε ijL = ε ijE = ε ij =
2
( ui, j + u j ,i ) (1 − 16)

In order to check the relationship between the above definition of strain with the
conventional definition, we consider a one-dimensional case,

(ds ) 2 − (dS ) 2 = 2ε ijL dX i dX j = 2ε 22 dX i dX i = 2ε 22 (dS ) 2


(ds ) 2 − (dS ) 2 (ds + dS )(ds − dS ) ds − dS (1-17)
ε 22 = = =
2(dS ) 2 dS (ds + dS ) dS
M. Mirzaei, Elasticity -7-

The linear strain tensor in Cylindrical and Spherical


Coordinate Systems

Cylindrical

⎛ ∂ur 1 ∂ur ∂uθ uθ ∂ur ∂u z ⎞


⎜ ε rr = ∂r 2ε rθ =
r ∂θ
+
∂r

r
2ε rz =
∂z
+
∂r ⎟
⎜ ⎟ (1-18)
⎜ 1 ∂uθ ur ∂u 1 ∂u z ⎟

εθθ = + 2ε θ z = θ +

r ∂θ r ∂z r ∂θ
⎜ ⎟
⎜ ∂u ⎟
⎜ ε zz = z ⎟
⎝ ∂z ⎠

Spherical
⎛ ∂ur 1 ∂ur ∂uθ uθ 1 ∂ur ∂uϕ uϕ ⎞
⎜ ε rr = 2ε rθ = + − 2ε rϕ = + − ⎟
⎜ ∂r r ∂θ ∂r r r sin θ ∂ϕ ∂r r ⎟ (1-19)
⎜ 1 ∂uθ ur 1 ∂uθ 1 ∂uϕ uθ cot ϕ ⎟
⎜ εθθ = + 2εθϕ = + − ⎟
⎜ r ∂θ r r sin θ ∂ϕ r ∂θ r ⎟
⎜ 1 ∂uϕ ur uθ cot θ ⎟
⎜⎜ ε ϕϕ = + + ⎟⎟
⎝ r sin θ ∂ϕ r r ⎠

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.

ε ll ,mm = ul ,lmm
ε mm ,ll = um ,mll
(1-20)
1
ε lm ,lm = (ul ,mlm + um ,llm )
2
1
= (ul ,lmm + um ,mll )
2
⇒ 2ε lm,lm = ε ll ,mm + ε mm ,ll
M. Mirzaei, Elasticity -8-

In general, we have

ε ij ,kl + ε kl ,ij = ε ik , jl + ε jl ,ik (1-21)

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

ε11,22 + ε 22,11 = 2ε12,12


ε 22,33 + ε 33,22 = 2ε 23,23
ε 33,11 + ε11,33 = 2ε 31,31
ε12,13 + ε13,12 − ε 23,11 = ε11,23 (1-22)
ε 23,21 + ε 21,23 − ε 31,22 = ε 22,31
ε 31,32 + ε 32,31 − ε12,33 = ε 33,12

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

∇ 2φ = ∇ ⋅∇φ = φ,ii
∇ ⋅ (φ v ) = ∇φ ⋅ v + φ∇ ⋅ v
∇ 2 (φψ ) = φ∇ 2ψ + 2∇φ ⋅∇ψ + ψ∇ 2φ
∇3 (φψ ) = ∇∇ 2 (φψ ) = ?
∇ ⋅ (φ∇ψ ) = ?
M. Mirzaei, Elasticity -9-

Kinetics of Deformable Solids


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

C0
R0 R
X2 b C
t
X1 undeformed Deformed
Configuration
Configuration
X3
Figure 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:

dP (1-23)
t = lim
dA→0 dA

in which dA is an element of area on a surface subjected to a force dP.

dP

dA
Figure 6

The body force vector denotes the external force acting on the interior of a solid per unit
volume and can be defined as:

dP (1-24)
b = lim
dV →0 dV
M. Mirzaei, Elasticity - 10 -

in which dV denotes an infinitesimal volume element subjected to a force dP.

dP

dV
Figure 7

Internal forces induced by external loading


The solid body shown in figure 8 is being 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 t
Pn
R R
t t
b C
e2 -n
e1

e3

Figure 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:

dPn (1-25)
Tn = lim
dA→0 dA
M. Mirzaei, Elasticity - 11 -

in which dA is an element of area in the interior of the solid, with normal n.

dPn
n

dA
Figure 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. 10.

e2
T2

σ22

σ21
σ23
σ12
σ32
T3
σ11
σ13
σ33 σ31
e3 e1

T1
Figure 10

Here, we may write:

T1 = σ 11e1 + σ 12e 2 + σ 13e3


T2 = σ 21e1 + σ 22e 2 + σ 23e3 or Ti = σ ij e j
(1-26)
T3 = σ 31e1 + σ 32e 2 + σ 33e3
M. Mirzaei, Elasticity - 12 -

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. 11.

e2
dAn n

dA1 Tn dAn
e1
-e1

T(-e1) dA1
e3

Figure 11

⎛1 ⎞
Tn dAn − Ti dAi + f ⎜ hdAn ⎟ = 0, Σn
⎝3 ⎠
h → 0 ⇒ Tn dAn − Ti dAi = 0
dAi = dAn Cos(n, ei ) (1-27)
Tn dAn − Ti dAnn ⋅ ei = 0
Tn = Ti ei , Ti = σ ij e j , n ⋅ ei = ni
Ti ei − σ ij e j ni = 0 , Ti ei = σ ij e j ni = σ ji ei n j

Accordingly, we will arrive at the following which are called: the stress boundary
equations.

Ti = σ ji n j (1-28)

Principle Stresses

For practical purposes, it would be convenient to break down the traction vector into
normal and shear components as,

σ nn = Tn ⋅ n = Ti ei ⋅ n = Ti ni ∑n (1-29)
= σ ji n j ni
M. Mirzaei, Elasticity - 13 -

and,

1 (1-30)
σ ns = (TT
i i − σ nn )
2 2
∑n

In order to find the extremum values of the normal components of the stress tensor, which
are in fact the principle stresses, we may write:

σ nn = σ ji n j ni ∑n (1-31)
= σ ji n j ni − λ (ni ni − 1)

where λ is the lagrangian multiplier. Next, we write,

∂σ nn
= σ ji n j − λ ni = 0 ∑n
∂ni
(1-32)
= σ ji n j − λδ ij n j = 0
= (σ ij − λδ ij )n j = 0

which implies that,

σ 11 − λ σ 12 σ 13
σ 21 σ 22 − λ σ 23 = 0
(1-33)
σ 31 σ 31 σ 33 − λ

The characteristic polynomial is,

λ 3 − I I λ 2 + I II λ − I III = 0 (1-34)

The roots of the above characteristic polynomial are the eigenvalues of our problem, or
the principle stresses.

⎧ I I = σ ii

⎪ 1 (1-35)
⎨ I II = (σ iiσ jj − σ ijσ ji )
⎪ 2
⎪ I III = σ ij

Finally, for each eigenvalue, we may find the eigenvectors which are the principle stress
directions.
M. Mirzaei, Elasticity - 14 -

λ1 → (σ ij − λδ ij )n j = 0 ⇒ n11 , n12 , n31


λ2 → (σ ij − λδ ij )n j = 0 ⇒ n12 , n22 , n32 (1-36)
λ3 → (σ ij − λδ 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.

⎛5 0 0 ⎞
σ ij = ⎜ ⋅ −6 −12 ⎟⎟

⎜⋅ ⋅ 1 ⎟⎠

Conservation of Linear Momentum


In order to find 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 = ∫ ρudV


S V V
(1-37)

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

∫σ
S
ji n j dA + ∫ f i dV = ∫ ρ ui dV
V V
(1-38)

Now we use the Divergence theorem of Gauss and write:

∫ G ⋅ ndA = ∫ ∇ ⋅ GdV
S V
(1-39)
∫σ ji , j dV + ∫ f i dV ≡ ∫ (σ ji , j + f i ) dV = ∫ ρ ui dV
V V V V

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

σ ji , j + fi = ρ ui (1 − 40)
M. Mirzaei, Elasticity - 15 -

The static form of the above equations, usually known as the equilibrium equations,
reads:

σ ji , j + f i = 0 (1-41)

The Equations of Motion in Cylindrical and Spherical Coordinate Systems can be


written as,

∂σ rr 1 ∂σ rθ ∂σ rz σ rr − σ θθ
+ + + + f r = ρ ur
∂r r ∂θ ∂z r
∂σ rθ 1 ∂σ θθ ∂σ θ z 2σ rθ (1-42)
+ + + + fθ = ρ uθ
∂r r ∂θ ∂z r
∂σ rz 1 ∂σ θ z ∂σ zz σ rz
+ + + + f z = ρ uz
∂r r ∂θ ∂z r

and,

∂σ rr 1 ∂σ rθ 1 ∂σ rϕ 1
+ + + (2σ rr − σ θθ − σ ϕϕ − σ rθ cot ϕ ) + f r = ρ ur
∂r r ∂θ r sin θ ∂ϕ r
(1-43)
∂σ rθ 1 ∂σ θθ 1 ∂σ θϕ 1
+ + + [3σ rθ + (σ θθ − σ ϕϕ ) cot θ ] + fθ = ρ uθ
∂r r ∂θ r sin θ ∂ϕ r
∂σ rϕ 1 ∂σ θϕ 1 ∂σ ϕϕ 1
+ + + (3σ rϕ + 2σ ϕθ cot θ ) + fϕ = ρ uϕ
∂r r ∂θ r sin θ ∂ϕ r

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 × ρ u)dV


S V V
(1-44)
∫ε
S
ijk X iT j dA + ∫ ε ijk X i f j dV = 0
V

which, using (1-28) and the Divergence theorem of Gauss, can be written as,
M. Mirzaei, Elasticity - 16 -

∫ε
S
ijk X i (σ lj nl )dA + ∫ ε ijk X i f j dV = 0
V

∫ε
V
ijk ( X iσ lj ),l dV + ∫ ε ijk X i f j dV = 0
V
(1-45)
∫ ε ijk [( X iσ lj ),l + X i f j ]dV = 0
V

∫ε
V
ijk [ X i ,lσ lj + X iσ lj ,l + X i f j ]dV = 0

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

⎡ ⎛ =0
⎞ ⎤
∫ ε ijk ⎢ X i,lσ lj + X i ⎜ σ lj ,l + f j ⎟⎥ dV = 0
V ⎣ ⎝ ⎠⎦
(1-46)
∫ ε ijkδ ilσ lj dV = 0
V

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

ε ijkσ ij = 0 (1-47)
⇒ σ ij = σ ji

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 ψ, which is a quadratic function of
the strain components.

∂ψ (1-48)
σ ij =
∂eij

Accordingly, we may write the most general form of the Hooke’s Law as:

σ ij = Eijkl ekl (1-49)


M. Mirzaei, Elasticity - 17 -

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.

⎧σ 11 ⎫ ⎡ E1111 E1122 E1133 2 E1112 2 E1123 2 E1131 ⎤ ⎧ ε11 ⎫


⎪σ ⎪ ⎢ E2222 E2233 2 E2212 2 E2223 2 E2231 ⎥⎥ ⎪⎪ε 22 ⎪⎪
⎪ 22 ⎪ ⎢
⎪⎪σ 33 ⎪⎪ ⎢ E3333 2 E3312 2 E3323 2 E3331 ⎥ ⎪⎪ε 33 ⎪⎪
⎨ ⎬=⎢ ⎥⎨ ⎬
⎪σ 12 ⎪ ⎢ 2 E1212 2 E1223 2 E1231 ⎥ ⎪ε12 ⎪ (1-50)
⎪σ 23 ⎪ ⎢ 2 E2323 2 E2331 ⎥ ⎪ε 23 ⎪
⎪ ⎪ ⎢ ⎥⎪ ⎪
⎪⎩σ 31 ⎪⎭ ⎣⎢ 2 E3131 ⎦⎥ ⎩⎪ε 31 ⎭⎪

which can be simplified to:

⎧σ 11 ⎫ ⎡C11 C12 C13 C14 C15 C16 ⎤ ⎧ ε11 ⎫


⎪σ ⎪ ⎢ C22 C23 C24 C25 C26 ⎥⎥ ⎪⎪ε 22 ⎪⎪
⎪ 22 ⎪ ⎢
⎪⎪σ 33 ⎪⎪ ⎢ C33 C34 C35 C36 ⎥ ⎪⎪ε 33 ⎪⎪
⎨ ⎬=⎢ ⎥⎨ ⎬
⎪σ 12 ⎪ ⎢ C44 C45 C46 ⎥ ⎪ε12 ⎪ (1-51)
⎪σ 23 ⎪ ⎢ C55 C56 ⎥ ⎪ε 23 ⎪
⎪ ⎪ ⎢ ⎥⎪ ⎪
⎪⎩σ 31 ⎪⎭ ⎣⎢ C66 ⎦⎥ ⎩⎪ε 31 ⎭⎪

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 the following Figure.

x3 x3′

x1′

x2 x2′

x1

Figure 12: single plane symmetry


M. Mirzaei, Elasticity - 18 -

This corresponds to a coordinate transformation with the direction cosines shown in table
1.

axes x1 x2 x3
x1′ -1
x2′ 1
x3′ 1

Table 1

Using the transformation σ ij′ = α ikα jlσ kl , in which α ik and α jl are the direction cosines,
we find:

σ 11′ = σ 11 σ 22
′ = σ 22 σ 33′ = σ 33 (1-52)
σ 12′ = −σ 12 σ 23
′ = σ 23 σ 31′ = −σ 31

Similarly, we may find

ε11′ = ε11 ε 22
′ = ε 22 ε 33′ = ε 33 (1-53)
ε12′ = −ε12 ε 23
′ = ε 23 ε 31′ = −ε 31

Hence, we may write:

⎧σ 11′ ⎫ ⎧ σ 11 ⎫ ⎡ ⎤ ⎧ ε11′ ⎫
⎪σ ′ ⎪ ⎪ σ ⎪ ⎢ ⎥ ⎪ε ′ ⎪
⎪ 22 ⎪ ⎪ 22 ⎪ ⎢ ⎥ ⎪ 22 ⎪
⎪⎪σ 33
′ ⎪⎪ ⎪⎪ σ 33 ⎪⎪ ⎢ ⎥ ⎪⎪ε 33
′ ⎪⎪
⎨ ⎬=⎨ ⎬= ⎢ ⎥⎨ ⎬
σ ′
⎪ ⎪ ⎪
12 − σ 12 ⎪ ⎢ C ⎥ ⎪ε12′ ⎪
⎪σ 23
′ ⎪ ⎪ σ 23 ⎪ ⎢ ⎥ ⎪ε 23
′ ⎪
⎪ ⎪ ⎪ ⎪ ⎢ ⎥⎪ ⎪
⎪⎩σ 31
′ ⎪⎭ ⎩⎪−σ 31 ⎭⎪ ⎣ ⎦ ⎩⎪ε 31
′ ⎭⎪
⎡ ⎤ ⎧ ε11 ⎫ (1-54)
⎢ ⎥⎪ ε ⎪
⎢ ⎥ ⎪ 22 ⎪
⎢ C ⎥ ⎪⎪ ε 33 ⎪⎪
=⎢ ⎥⎨ ⎬
⎢ ⎥ ⎪−ε12 ⎪
⎢ ⎥ ⎪ ε 23 ⎪
⎢ ⎥⎪ ⎪
⎣ ⎦ ⎩⎪−ε 31 ⎭⎪

which results in:


M. Mirzaei, Elasticity - 19 -

⎧σ 11 ⎫ ⎡C11 C12 C13 −C14 C15 −C16 ⎤ ⎧ ε11 ⎫


⎪σ ⎪ ⎢ C22 C23 −C24 C25 −C26 ⎥⎥ ⎪⎪ε 22 ⎪⎪
⎪ 22 ⎪ ⎢
⎪⎪σ 33 ⎪⎪ ⎢ C33 −C34 C35 −C36 ⎥ ⎪⎪ε 33 ⎪⎪
⎨ ⎬=⎢ ⎥⎨ ⎬
⎪σ 12 ⎪ ⎢ C44 −C45 C46 ⎥ ⎪ε12 ⎪ (1-55)
⎪σ 23 ⎪ ⎢ C55 −C56 ⎥ ⎪ε 23 ⎪
⎪ ⎪ ⎢ ⎥⎪ ⎪
⎩⎪σ 31 ⎭⎪ ⎣⎢ C66 ⎦⎥ ⎩⎪ε 31 ⎭⎪

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 in two
dimensions we need 4 elastic constants to describe the behavior of an orthotropic
material).

axes x1 x2 x3
x1′ -1
x2′ -1
x3′ 1

We may have 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 axes x1 x2 x3
x1′ 1 0 0 x1′ 0 1 0
x2′ 0 0 1 x2′ 1 0 0
x3′ 0 1 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 cosθ sinθ
x3′ 0 - sinθ cosθ
M. Mirzaei, Elasticity - 20 -

⎧σ 11 ⎫ ⎡ 2 μ + λ λ λ 0 0 0 ⎤ ⎧ ε11 ⎫
⎪σ ⎪ ⎢ λ 2μ + λ λ 0 0 0 ⎥⎥ ⎪⎪ε 22 ⎪⎪
⎪ 22 ⎪ ⎢
⎪⎪σ 33 ⎪⎪ ⎢ λ λ 2μ + λ 0 0 0 ⎥ ⎪⎪ε 33 ⎪⎪
⎨ ⎬=⎢ ⎥⎨ ⎬
⎪σ 12 ⎪ ⎢ 2μ 0 0 ⎥ ⎪ε12 ⎪ (1-56)
⎪σ 23 ⎪ ⎢ 2μ 0 ⎥ ⎪ε 23 ⎪
⎪ ⎪ ⎢ ⎥⎪ ⎪
⎩⎪σ 31 ⎭⎪ ⎣ 2 μ ⎦ ⎩⎪ε 31 ⎭⎪

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

σ ij = 2 με ij + λδ ijε kk (1-57)

We may invert the above equation to express the strains in terms of stresses:

1 λ (1-58)
ε ij = σ ij − δ ijσ kk
2μ 2 μ (2μ + 3λ )

The Lamé 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 υ, as
well. Table 2 shows the relationships between different elastic constants.

λ, μ E ,ν μ ,ν E, μ k ,ν
νE 2 μν μ ( E − 2μ ) 3kν
λ λ
(1 +ν )(1 − 2ν ) (1 − 2ν ) 3μ − E (1 +ν )
E 3k (1 − 2ν )
μ μ μ μ
2 (1 +ν ) 2 (1 +ν )
2 E 2ν (1 + ν ) μE
k λ+ μ k
3 3 (1 − 2ν ) 3 (1 − 2ν ) 3 ( 3μ − E )
μ ( 3λ + 2μ )
E E 2 μ (1 + ν ) E 3k (1 − 2ν )
(λ + μ )
λ E
ν ν ν −1 ν
2 (λ + μ ) 2μ

Table 2
M. Mirzaei, Elasticity - 21 -

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:

1 (I)
ε ij =
2
( ui, j + u j ,i )
σ ji , j + f i = 0 (II)

σ ij = 2με ij + λδ ijε kk or
1 λ
ε ij = σ ij − δ ijσ kk (III)
2μ 2μ (2μ + 3λ )

along with the compatibility and stress boundary equations,

ε ij ,kl + ε kl ,ij = ε ik , jl + ε jl ,ik (IV)

Ti = σ ji n j (V)

Classical Displacement Formulation

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

σ ij = λδ ij uk ,k + μ (ui , j + u j ,i ) (1-59)

Next, we substitute the above expression for the stresses in Eqns(II) to give:

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


λuk ,kj + μ (ui , ji + u j ,ii ) + f j = 0 (1-60)

Finally, using i for the dummy index in all terms we may write:

μu j ,ii + (λ + μ )ui ,ij + f j = 0 (1-61)


M. Mirzaei, Elasticity - 22 -

or alternatively,

(λ + 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 ∂ω z ∂ωθ ⎞
(λ + 2G ) − 2G ⎜ − ⎟ + fr = 0
∂r ⎝ r ∂θ ∂z ⎠
∂I ⎛ ∂ω ∂ω ⎞
(λ + 2G ) e − 2G ⎜ r − z ⎟ + fθ = 0
r ∂θ ⎝ ∂z ∂r ⎠
∂I e ⎛ ∂ (rωθ ) ∂ωr ⎞
(λ + 2G ) − 2G ⎜ − ⎟ + fz = 0
∂z ⎝ ∂r ∂θ ⎠
where (1-63)
1 ∂ (rur ) 1 ∂uθ ∂u z
I e = eii = ∇ ⋅ u = + +
r ∂r r ∂θ ∂z
1 ∂u z ∂uθ ∂u ∂u
2ωr = − ; 2ωθ = r − z ;
r ∂θ ∂z ∂z ∂r
1 ⎛ ∂ (ruθ ) ∂ur ⎞
2ω z = ⎜ − ⎟
r ⎝ ∂r ∂θ ⎠
M. Mirzaei, Elasticity - 23 -

Spherical:

∂I e 2G ⎡ ∂ (ωϕ sin θ ) ∂ωθ ⎤


(λ + 2G ) − − + fr = 0
∂r r sin θ ⎢⎣ ∂θ ∂ϕ ⎥⎦
∂I 2G ⎡ 1 ∂ (ωr ) ∂ (rωϕ ) ⎤
(λ + 2G ) e − − + fθ = 0
r ∂θ r ⎢⎣ sin θ ∂ϕ ∂r ⎥⎦
(λ + 2G ) ∂I e 2G ⎡ ∂ (rωθ ) ∂ (rωr ) ⎤
− − + fϕ = 0
r sin θ ∂ϕ r ⎢⎣ ∂r ∂θ ⎥⎦
where (1-64)
1 ∂ ( r 2 ur ) 1 ∂ (uθ sin θ ) 1 ∂uϕ
I e = eii = ∇ ⋅ u = + +
r 2 ∂r r sin θ ∂θ r sin θ ∂ϕ
1 ⎡ ∂ (uϕ sin θ ) ∂uθ ⎤ 1 ∂ur 1 ∂ (ruϕ )
2ωr = ⎢ − ⎥ ; 2ωθ = − ;
r sin θ ⎣ ∂θ ∂ϕ ⎦ r sin θ ∂ϕ r ∂r
1 ⎛ ∂ (ruθ ) ∂ur ⎞
2ωϕ = ⎜ − ⎟
r ⎝ ∂r ∂θ ⎠

Classical Force Formulation


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

ν (1-65)
σ ij ,kl + σ kl ,ij − σ ik , jl − σ jl ,ik =
1 +ν
(δ σ
ij tt , kl − δ klσ tt ,ij − δ ikσ tt , jl − δ jlσ tt ,ik )

Only six of the above equations are independent:

ν (1-66)
σ ij ,kk + σ kk ,ij − σ ik , jk − σ jk ,ik =
1 +ν
(δ σ ij tt , kk − δ kkσ tt ,ij − δ ikσ tt , jk − δ jkσ tt ,ik )

which are nine equations with free indices i and j. Now we substitute the equilibrium
equations (II) in the above and simplify to give:

1 ⎛ ν ⎞ (1-67)
∇ 2σ ij + σ mm ,ij = − ⎜ fi , j + f j ,i + δ ij f m,m ⎟
1 +ν ⎝ 1 −ν ⎠
M. Mirzaei, Elasticity - 24 -

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

1 ∂ ⎛ ∂σ rr ⎞ 1 ∂ 2σ rr ∂ 2σ rr 1 ∂ 2 It ν ⎡ 1 ∂ ( rf r ) 1 ∂ ( fθ ) ∂ ( f z ) ⎤ ∂ ( fr )
⎜r ⎟+ 2 + + =− ⎢ + + ⎥−2
r ∂r ⎝ ∂r ⎠ r ∂θ 2
∂z 2
1 + ν ∂r 2
1 −ν ⎣ r ∂r r ∂θ ∂z ⎦ ∂r
1 ∂ ⎛ ∂σ θθ ⎞ 1 ∂ σ θθ ∂ σ θθ
2 2
1 ∂ It
2
ν ⎡ 1 ∂ ( rf r ) 1 ∂ ( fθ ) ∂ ( f z ) ⎤ ∂ ( fθ )
⎜ r ⎟ + 2 + + =− ⎢ + + ⎥−2
r ∂r ⎝ ∂r ⎠ r ∂θ 2
∂z 2
1 + ν ∂θ 2
1 −ν ⎣ r ∂r r ∂θ ∂z ⎦ ∂θ
1 ∂ ⎛ ∂σ zz ⎞ 1 ∂ σ zz ∂ σ zz
2 2
1 ∂ 2 It ν ⎡ 1 ∂ ( rf r ) 1 ∂ ( fθ ) ∂ ( f z ) ⎤ ∂ ( fz )
⎜r +
⎟ 2 + + =− ⎢ + + ⎥−2
r ∂r ⎝ ∂r ⎠ r ∂θ ∂z 1 + ν ∂z 1 −ν ⎣ r ∂r r ∂θ ∂z ⎦ ∂z
2 2 2

1 ∂ ⎛ ∂σ θ z ⎞ 1 ∂ σθ z ∂ σθ z
2 2
1 ∂ 2 It ⎡ ∂ ( fθ ) ∂ ( f z ) ⎤
⎜r +
⎟ 2 + + = −⎢ + ⎥
r ∂r ⎝ ∂r ⎠ r ∂θ ∂z 2 1 +ν ∂θ∂z ⎣ ∂z ∂θ ⎦
2

1 ∂ ⎛ ∂σ rz ⎞ 1 ∂ σ rz ∂ σ rz
2 2
1 ∂ 2 It ⎡ ∂ ( f z ) ∂ ( fr ) ⎤ (1-68)
⎜r +
⎟ 2 + + = −⎢ + ⎥
r ∂r ⎝ ∂r ⎠ r ∂θ ∂z 1 +ν ∂r ∂z ⎣ ∂r ∂z ⎦
2 2

1 ∂ ⎛ ∂σ rθ ⎞ 1 ∂ 2σ rθ ∂ 2σ rθ 1 ∂ 2 It ⎡ ∂ ( f r ) ∂ ( fθ ) ⎤
⎜ r ⎟ + 2 + + = −⎢ + ⎥
r ∂r ⎝ ∂r ⎠ r ∂θ 2
∂z 2
1 + ν ∂r ∂θ ⎣ ∂θ ∂r ⎦

It = σ rr + σθθ + σ zz
M. Mirzaei, Elasticity - 25 -

Beltrami-Michell Equations in Spherical Coordinates


1 ∂ ⎛ 2 ∂σ rr ⎞ 1 ∂ ⎛ ∂σ rr ⎞ 1 ∂ 2σ rr 1 ∂ 2 It
⎜r +
⎟ 2 ⎜ sin θ ⎟ + + =
r 2 ∂r ⎝ ∂r ⎠ r sin θ ∂θ ⎝ ∂θ ⎠ r 2 sin 2 θ ∂ϕ 2 1 +ν ∂r 2
ν ⎡ 1 ∂ ( r fr ) 1 ∂ ( fϕ ) ⎤ ∂ ( fr )
2
1 ∂
− ⎢ 2 + ( sin θ fθ ) + ⎥−2
1 −ν ⎢ r ∂r r sin θ ∂θ r sin θ ∂ϕ ⎥ ∂r
⎣ ⎦
1 ∂ ⎛ 2 ∂σ θθ ⎞ 1 ∂ ⎛ ∂σ θθ ⎞ 1 ∂ σ θθ
2
1 ∂ 2 It
⎜r ⎟+ 2 ⎜ sin θ ⎟+ 2 2 + =
r ∂r ⎝
2
∂r ⎠ r sin θ ∂θ ⎝ ∂θ ⎠ r sin θ ∂ϕ 2
1 + ν ∂θ 2
ν ⎡ 1 ∂ ( r fr ) 1 ∂ ( fϕ ) ⎤ ∂ ( fθ )
2
1 ∂
− ⎢ 2 + ( sin θ fθ ) + ⎥−2
1 −ν ⎢ r ∂r r sin θ ∂θ r sin θ ∂ϕ ⎥ ∂θ
⎣ ⎦
1 ∂ ⎛ 2 ∂σ ϕϕ ⎞ 1 ∂ ⎛ ∂σ ϕϕ ⎞ 1 ∂ 2σ ϕϕ 1 ∂ 2 It
⎜ r ⎟ + ⎜ sin θ ⎟ + + =
r 2 ∂r ⎝ ∂r ⎠ r 2 sin θ ∂θ ⎝ ∂θ ⎠ r 2 sin 2 θ ∂ϕ 2 1 +ν ∂ϕ 2

ν ⎡ 1 ∂ ( r fr ) 1 ∂ ( fϕ ) ⎤ ∂ ( fϕ )
2
1 ∂
− ⎢ 2 + ( sin θ fθ ) + ⎥−2 (1-69)
1 −ν ⎢ r ∂r r sin θ ∂θ r sin θ ∂ϕ ⎥ ∂ϕ
⎣ ⎦
1 ∂ ⎛ 2 ∂σ θϕ ⎞ 1 ∂ ⎛ ∂σ θϕ ⎞ 1 ∂ 2σ θϕ 1 ∂ 2 It ⎛ ∂ ( fθ ) ∂ ( fϕ ) ⎞
⎜r ⎟+ 2 ⎜ sin θ ⎟+ 2 2 + = −⎜
⎜ ∂ϕ
+ ⎟
r ∂r ⎝
2
∂r ⎠ r sin θ ∂θ ⎝ ∂θ ⎠ r sin θ ∂ϕ 2
1 +ν ∂θ∂ϕ ∂θ ⎟
⎝ ⎠
1 ∂ ⎛ 2 ∂σ rϕ ⎞ 1 ∂ ⎛ ∂σ rϕ ⎞ 1 ∂ σ rϕ
2
1 ∂ 2 It ⎛ ∂ ( f r ) ∂ ( fϕ ) ⎞
⎜r ⎟+ 2 ⎜ sin θ ⎟+ 2 2 + = −⎜
⎜ ∂ϕ
+ ⎟
r ∂r ⎝
2
∂r ⎠ r sin θ ∂θ ⎝ ∂θ ⎠ r sin θ ∂ϕ 2
1 +ν ∂r ∂ϕ ⎝ ∂r ⎠⎟

1 ∂ ⎛ 2 ∂σ rθ ⎞ 1 ∂ ⎛ ∂σ rθ ⎞ 1 ∂ 2σ rθ 1 ∂ 2 It ⎛ ∂ ( f r ) ∂ ( fθ ) ⎞
⎜ r +
⎟ 2 ⎜ sin θ +
⎟ 2 2 + = −⎜ + ⎟
r ∂r ⎝
2
∂r ⎠ r sin θ ∂θ ⎝ ∂θ ⎠ r sin θ ∂ϕ 2
1 +ν ∂r∂θ ⎝ ∂θ ∂r ⎠

It = σ rr + σθθ + σ ϕϕ

These notes have been prepared as a student aid and should not be considered as a book. No
originality is claimed for these notes other than selection, organization, and presentation of the
material.

The following references have been used for preparation of the lecture notes and are recommended
for further study in this course.

References:

• 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"

You might also like