Vous êtes sur la page 1sur 43

Ruhr-University Bochum Faculty of Civil and Environmental Engineering Institute of Mechanics

An Introduction to Linear Continuum Mechanics


Klaus Hackl Mehdi Goodarzi

2010

ii

Contents
1 Fundamentals

1.1 Stress and equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1.1 Stress tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1.2 Balance equations: dierential form . . . . . . . . . . . . . . . . . . 1.2 Strain and compatibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2.1 Compatibility conditions . . . . . . . . . . . . . . . . . . . . . . . . . 2.1 Elasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Hooke's law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.1 Hooke's law for isotropic materials . . . . . . . . . . . . . . . . . . . 2.2.2 Alternative formulations of Hooke's law for isotropic materials . . . 2.2.3 Plane-strain and plane-stress problems . . . . . . . . . . . . . . . . . 2.3 Navier and Beltrami-Michell equations . . . . . . . . . . . . . . . . . . . . . 3.1 3.2 3.3 Scalar and vector potentials . . . . . . . . . . . . . . . . . . . . . . . . . . . Galerkin vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.1 Love's strain function . . . . . . . . . . . . . . . . . . . . . . . . . . Displacement functions of Papkovich-Neuber . . . . . . . . . . . . . . . . .

2 3 5 8 12
17

2 The Field Equations of Linear Continuum Mechanics

17 18 19 20 22 23

3 Displacement Functions

27

27 32 33 35

iii

iv

CONTENTS

Chapter 1

Fundamentals
The subject matter of mechanics is the study of motion, in how a physical object changes position with time and why. Here we shall conne ourselves within principles of classical mechanics. Studying how a body moves is called kinematics which is basically a geometrical investigation of motion dealing with fundamental concepts like space, time and frame of reference. In a second step Newton's laws of motion are introduced based on other fundamental concepts such as inertial frame of reference,mass and force to deal with the question of why things move, and this is called kinetics. The alternative strategy is formulation of mechanics based on conservation laws of energy, momentum and angular momentum which turns out to have broader range of validity than Newton's formulation. Scientic hypotheses try to explain physical observations based on a set of assumptions and more often than not, neglecting certain eects. One of the most universal aspects of observations is the scale, that is the range of values of quantities during an observation. For instance, having a model proposed by an observation over the ordinary length scales of laboratory, we can not draw the conclusion that the same model will hold over cosmic length scales, or on sub-atomic length scales unless it is somehow tested or proved. Considering length scales, mechanical model of a moving body can belong to one of the following in order of sophistication Particle is a model of a body of negligible dimensions so that it can be treated geometrically as a point, which means all the body mass is concentrated at its center of mass. For example, in study of planetary motion dimensions of planets are negligible compared to orbital distances therefore they are often treated as point masses. Kinematics of a particle deals with translation only. In this model, forces are free vectors which means they can move freely without their mechanical eect being altered. Model equations for particle motion consist of balance of energy and linear momentum. Rigid body is an object of nite dimensions and negligible change in shape during motion. That is where the distance between every two points on the body remains constant. A rigid body is assumed to have a continuous mass distribution although in reality matter is quantized at small scales. Kinematics of rigid body motion is expressed in terms of translation and rotation. Forces are sliding vectors viz they can slide on 1

CHAPTER 1.

FUNDAMENTALS

their line of action without their mechanical eects being altered. Model equations are balance of energy, linear momentum and angular momentum. Deformable body is a mechanical object of continuous mass distribution whose relative distance of points can change. This is ocially the starting point of continuum mechanics although a rigid body has a continuous mass distribution too. Continuum mechanics deals with the length scales large enough to neglect all the molecular effects and, at the same time, small enough to observe the shape changes in the body. Together with translations and rotations we have deformation in kinematics study of deformable bodies. As the point of action of forces is important here, forces are xed vectors. Balance equations of linear and angular momentum and energy are sucient to model deformation. In addition we will need the so called constitutive equations to have a closed model. In the following section we work out the concept of forces in continuum media.
not

1.1

Stress and equilibrium

In the view of causality, forces are the cause for change in momentum and also for deformation. According Newton's third law, forces exist only in couples, to wit the presence of any force is justiable only in the presence of its source. In this regard, we distinguish two types of forces Field forces which are applied from a distance through a eld such as electromagnetic or gravitational eld. Contact forces which are applied by direct contact of objects such as friction. Note that these are also manifestation of eld forces that act at inter-molecular distances, therefore observed as direct contact at large scales. Distribution of eld forces is usually expressed as force per unit volume or per unit mass, and distribution of contact forces is usually expressed as force per unit area. The exception is the so called concentrated force which is an attempt to model the situation when a nite amount of force is applied over a very small region of area or volume. In this specic case the force distribution will not be nite-valued anymore. Therefore concentrated forces are treated as individual force vectors rather than distributions, and presented by Dirac's delta function which will be introduced later on. Let be a body at t equilibrium with a nite and continuous distribution of solid mass m throughout its simply connected volume V , and be bounded by a smooth surface of area A. This is our typ ical picture of a body under consideration in continuum solid f (x) mechanics. You will see potato-like illustrations (Fig. 1.1)! Assume a distribution of eld forces presented by a volumetric force eld f (x) and contact force (also called traction ) distribution over denoted by t. Then equilibrium condi- Fig. 1.1: Solid body.

1.1.

STRESS AND EQUILIBRIUM

3
f dV + t dA = 0

(1.1) x t dA = 0 (1.2) x f dV + M =0 : This provides enough ground for analysis of the overall body motion which is a combination of translation and rotation, as if the body was rigid. However, deformation is a local phenomenon which depends on the state of forces at each body point.
F =0 :

tions for the body read

1.1.1 Stress tensor

In order to analyze the forces at each body point, we shall write the equilibrium equations for an innitesimal volume element at the body point. For an innitesimal element at
dX1 t3 dX3 t1 dX2 X1 e3 t2 X3 e1 f e2 n X2
(a) (b)

Fig. 1.2: a) Innitesimal volume element, and b) inintesimal tetrahedron. point x with volume dV the volumetric force is given by dF = f (x) dV . Also, there are tractions at element faces (Fig. 1.2.a), because an area element within a body of mass is an interface that experiences contact forces due to adhesion and/or cohesion. This traction force t at each point depends on the location x of area element dA, and on its direction denoted by unit normal vector n t = t(x, n) . (1.3) This is Cauchy's assumption which is substantial in continuum mechanics, although it might seem obvious. The functionality of t(x, n) is further specied by Cauchy's theorem through which the concept of stress tensor is introduced.
V

Cauchy's theorem

Assume that we want to write the equation of motion for an innitesimal tetrahedron (Fig. 1.2.b) whose edges coincide with the three Cartesian axes. The four triangular faces have unit normal vectors We identify face triangles of the tetrahedron by their normal vectors i.e. 1, 2, 3 and n. Notice that projection of the triangle n over coordinate planes are triangles 1, 2, 3. Therefore if
e1 , e2 , e3 , n .

4 the center point of triangle n has the position

CHAPTER 1.

FUNDAMENTALS

(1.4) then the center points of the other three triangles are projections of x onto coordinate planes, given by x =x e +x e , x =x e +x e , x =x e +x e . (1.5) Furthermore, if the triangle n has the area dA then the triangles 1, 2, 3 (which are its projections) have the areas
x = x1 e 1 + x2 e 2 + x3 e 3 ,
c 1 2 2 3 3 c 2 1 1 3 3 c 3 1 1 2 2

dA1 = n1 dA dA2 = n2 dA dAi = ei n dA = ni dA : dA3 = n3 dA

(1.6)

Also note that the tetrahedron has the mass dV and acceleration a. Finally, the equation of motion is written as a dV = f dV + t(x, n) dA + t(x , e ) dA + t(x , e ) dA + t(x , e ) dA . (1.7) Substitution of dA 's according equation (1.6) gives a dV = f dV + t(x, n) dA + t(x , e ) n dA + t(x , e ) n dA + t(x , e ) n dA (1.8) and dividing both sides by dA dV dV a =f + t(x, n) + t(x , e ) n + t(x , e ) n + t(x , e ) n . (1.9) dA dA However the fraction dV /dA tends to zero as the volume shrinks. This is because dV is proportional to L and dA is proportional to L , where L is the characteristic length of the element, hence
c 1 1 1 c 2 2 2 c 3 3 3 i c 1 1 1 c 2 2 2 c 3 3 3 c 1 1 1 c 2 2 2 c 3 3 3 3 2

L0

Therefore equation (1.9) yields

dV L3 2 = L 0. dL L

c c t(x, n) = t(xc 1 , e1 ) n1 t(x2 , e2 ) n2 t(x3 , e3 ) n3 .

When the tetrahedron element shrinks, the center points of faces tend to each other
L0 :
c c xc 1 x2 x3 x

(1.10)

and equation (1.10) gives


t(x, n) = t(x, e1 ) n1 t(x, e2 ) n2 t(x, e3 ) n3 .

(1.11)

1.1.

STRESS AND EQUILIBRIUM

5 (1.12) (1.13) (1.14)

Now writing components gives


t1 (x, n) = t1 (x, e1 ) n1 t1 (x, e2 ) n2 t1 (x, e3 ) n3 t2 (x, n) = t2 (x, e1 ) n1 t2 (x, e2 ) n2 t2 (x, e3 ) n3 t3 (x, n) = t3 (x, e1 ) n1 t3 (x, e2 ) n2 t3 (x, e3 ) n3

and in matrix form

(1.15) The 3 3 matrix in the latter equation is a second order tensor which is only a function of x. It is called stress tensor and is usually denoted by . This equation can be written in the more compact form t(x, n) = (x) n , (1.16) which is called Cauchy's formula. Let's have a closer X look at the matrix form in equation (1.15). It is easy to show that t (x, e ) = t (x, e ). Then the com ponent at the i row and the j column is
3 33 i th j i j th 13 23 32

t1 (x, n) t1 (x, e1 ) t1 (x, e2 ) t1 (x, e3 ) n1 t2 (x, n) = t2 (x, e1 ) t2 (x, e2 ) t2 (x, e3 ) n2 . t3 (x, n) t3 (x, e1 ) t3 (x, e2 ) t3 (x, e3 ) n3

ij = ti (x, ej )
th th

31
21 12

which is the i component of the traction force applied to the surface normal to the j axis (Fig. 1.3). ThereX fore the second index j in the stress tensor shows X the surface to which the traction is applied, and the rst index i shows the direction of the traction compoFig. 1.3: Stress components. nent. The importance of equation (1.16) is that the functionality of traction on x and n is separated as the stress tensor is a function of x only at the right-hand-side. If you remember, when introducing tensors, we mentioned that second order tensors are linear mappings from vectors to vectors. Cauchy's formula is a physical example of this, where stress tensor maps directions denoted by n to tractions t exerted on an area element normal to n.
11 2 ij 1

22

1.1.2 Balance equations: dierential form

As mentioned before we need the local state of forces in order to analyze deformation. The integrals in equations (1.1) and (1.2) are writtenover the whole body , while they are equally true for any sub-domain of the body, say which includes the body point x f dV + t dA = 0 (1.17) x f dV + x t dA = 0 (1.18)

6 Let's focus at the rst equation. If we substitute Cauchy's formula into (1.17) it yields f dV + n dA = 0 . (1.19) is a closed surface, therefore the second integral according Gauss theorem Note that gives dV (1.20) n dA = Therefore equation (1.19) becomes f dV + dA = (f + ) dV = 0 . (1.21) only if the integrand equals zero This integral vanishes for any arbitrary sub-body +f =0 (1.22) which is the dierential equation for balance of forces. For the second balance equation (1.18), again we substitute Cauchy's formula to obtain x f dV + x ( n) dA = 0 (1.23) however it holds that x ( n) = (x ) n (1.24) therefore x f dV + (x ) n dA = 0 (1.25) and once again Gauss theorem gives (x ) n dA = (x ) dV (1.26) which turns equation (1.25) into x f dV + (x ) dV = (x f + (x ) ) dV = 0 . (1.27) There is an identity that can be easily veried (x ) = : + x ( ) (1.28) where is permutation tensor. Putting this into equation (1.27) gives x (f + ) + : dV = 0 (1.29)
CHAPTER 1. FUNDAMENTALS

1.1.

STRESS AND EQUILIBRIUM

7
: dV = 0

however the term in parentheses vanishes due to (1.22), then (1.30) (1.31) (1.32)
, therefore it must hold that which holds for every sub-body : = 0.

which can be true only if is a symmetric tensor


= T .

which is the local condition for balance of moments. Exercise 1. Show that : = 0 results in = . Exercise 2. Write equilibrium equations and Cauchy's formula in components. Exercise 3. Consider a body with some uid exposed to X hydrostatic pressure (Fig. 1.4).
T 3

with body force: f = ge . Solve equilibrium equations X X for p(x ) with boundary condition p(0) = p . Fig. 1.4: Body lled with uid. Example 1.1. Write equilibrium equations in cylindrical coordinates. Solution. The equilibrium equation is a vector equation or a set of three scalar equations given by (1.22) as
3 1 2 3 0

0 0 p(x3 ) 0 p(x3 ) 0 = p(x3 ) I = 0 0 p(x3 )

The divergence operator in cylindrical coordinates is given by


= 1 1 r rz (r rr ) + + er + r r r z 1 1 z (r r ) + + e + r r r z 1 1 z zz (r zr ) + + ez r r r z

+ f = 0.

and decomposition of body forces gives

f = fr er + f e + fz ez .

8 Substitution into equilibrium equation results in

CHAPTER 1.

FUNDAMENTALS

1 1 r rz (r rr ) + + + fr = 0 r r r z 1 1 z (r r ) + + + f = 0 r r r z 1 1 z zz (r zr ) + + + fz = 0 . r r r z

(1.33) (1.34) (1.35)

Using the transformation from Cartesian to cylindrical coordinates and the nabla operator in Cartesian coordinates, show that in cylindrical coordinates the nabla operator has the following form
Exercise 4.
1 er r + e + ez z . r

Calculate (r) for a tube loaded by torsion (Fig. 1.5). There are no body forces and the boundary condition (r ) = t is given. Note that
Exercise 5.
r r 1 1

r0 r1

= r (er e + e er ) .

Fig. 1.5: Tube under torsion.

1.2

Strain and compatibility

As pointed out before, the motion of a continuum medium is a combination of translation, rotation and deformation. Deformation is the change in shape. Since the shape of an object is characterized by the relative position of its points, in order to analyze deformation, we focus on the change of innitesimal line elements that connect the neighboring points of the body. Let us assume a typical image of a continuum body (Fig. 1.6) at initial/undeformed conguration whose points are denoted by the position vector X , and then in its moved/deformed conguration whose points are denoted by the position vector x. The basic assumption is that x is a one-to-one function of X . This guarantees that the material points are neither created, nor are they annihilated. So, the starting point is a description of motion by x = x(X , t) . (1.36) It is also customary to dene displacement eld as u = u(X , t) = x(X , t) X . (1.37)

1.2.

STRAIN AND COMPATIBILITY

9
dx

dX

Fig. 1.6: Undeformed and deformed congurations of a continuum body. Excluding dynamic cases from our discussion, the time dependence is dropped. Moreover, we assume that the displacement eld is innitesimal u 1. (1.38) Now imagine two neighboring material points in the undeformed conguration connected by the dierential line element dX , and in the deformation conguration connected by dx. We would like to study the relation dX dx. According denition of displacement eld we have dx = dX + u(X + dX ) u(X ) . (1.39) Then substitution of Taylor's expansion as u(X + dX ) = u(X ) + u dX + (1.40) will result in dx = dX + u dX = x dX . (1.41) Here u is the displacement gradient and x the deformation gradient which are related clearly by x = I + u . (1.42) The displacement gradient is a second order tensor and can be decomposed into symmetric and antisymmetric parts designated by 1 = (u + u) (1.43) 2 1 = (u u) (1.44) 2 the symmetric part is called strain tensor and the antisymmetric part is the rotation tensor. As their names suggest the symmetric part reects deformation and the antisymmetric part carries information about rotation only. It is immediately clear that = , = . (1.45)
T T

(1.46) Every antisymmetric tensor has three independent elements only, and therefore can be specied by a vector. So the rotation tensor can be expressed by the so called rotation vector 1 1 = : . (1.47) w = (w ) = 2 2 Exercise 6. Show that = w . By expanding the equations (1.47) and (1.44), one can show that 1 w = u. (1.48) 2
1, 1.
i jik jk ij ikj k

10 To stay within linear continuum theory the fundamental assumption is


CHAPTER 1.

FUNDAMENTALS

Physical meaning of the decomposition

Putting the decomposition u = + into (1.41) we obtain


dx = dX + u dX = dX + dX + dX .
r s r

We show that dx = dX is the rotational part of deformation and dx deformational part of deformation. Let us focus at dx rst [ dX ] = dX = w dX (1.50) which gives dv dx = w dX . (1.51) This, according assumption (1.46), means that dx is obtained by a small rotation of dX around w. Note that dx is normal dv to dX , and since w 1 then the ratio dx / dX which is equal to the angle of rotation (Fig. 1.7) is also small. In short form notation one can state w w 1 Fig. 1.7: Rotational part of
i ij j ikj k j r r r r r

(1.49) = dX is the

1.

Now let us have a look at dx . Deformation is expressed in terms of change in lengths and angles. The length of the line element in the deformed conguration is obtained by
s

dxr / dX = w

displacement.

(1.52) where higher order terms, denoted by , are neglected due to assumption (1.46). Also we know that
= dX dX + 2dX (w dX ) + 2dX dX + dX (w dX ) = 0

dx dx = (dX + w dX + dX ) (dX + w dX + dX )

1.2.

STRAIN AND COMPATIBILITY

11 (1.53)

because dX is orthogonal to (w dX ). Therefore and then


dx2 = dX 2 + 2dX dX dx dX dX = 1+2 dX dX 2
1/2

(1.54) where Taylor's expansion is used because 1. The change of length is nally obtained by 1 dX dX . (1.55) l = dx dX = dX Remember that this result is based on the assumption (1.46) of small strain and rotation. To study the changes in angle, assume two adjacent innitesimal vectors normal to each other in the undeformed conguration denoted by dX and dX , and their deformed counterparts denoted by dx and dx respectively. Inner multiplication of the two vectors will give us information on the angle between them dx dx = (dX + w dX + dX ) (dX + w dX + dX ) (1.56)
1+ dX dX , dX 2
1 2 1 2 1 2 1 1 1 2 2 2

where the higher order terms are neglected. The rst term vanishes due to the assumption of orthogonality dX dX = 0, and the sum of second and third terms is also zero because
1 2

= dX 1 dX 2 + dX 1 (w dX 2 ) + (w dX 1 ) dX 2 + 2dX 1 dX 2 +

Therefore On the other hand

dX 1 (w dX 2 ) = (w dX 1 ) dX 2 . dx1 dx2 2dX 1 dX 2 .

(1.57)

(1.58) where /2 is the angle in the undeformed conguration. Finally comparing (1.58) and (1.57) yields dX dX = 2 . (1.59) dX dX Again, note that this formula is obtained under the assumption (1.46). As can be seen from equations (1.55) and (1.59), deformation depends on the strain tensor only. Therefore our objective of interpreting the additive decomposition of displacement gradient into symmetric (strain) and antisymmetric (rotation) parts is fullled. Exercise 7. For the given displacement gradient
= dX1 dX2 sin() dX1 dX2
1 2 1 2

dx1 dx2 = dx1 dx2 cos dX1 dX2 cos = dX1 dX2 sin(/2 )

u = 103

1 2 0 3 0 0 0 1 1

12 1. calculate , , w.
1 a = 0 mm 1

CHAPTER 1.

FUNDAMENTALS

2. For the two orthogonal vectors and calculate the change of their lengths and the angle between them. Remark 1. Strain is dened as the symmetric part of displacement gradient. Sometimes this is indicated by the short notation = u (1.60) called symmetric gradient of displacement.
s

1 b = 1 mm 1

1.2.1 Compatibility conditions

If the equation (1.43) is expanded as


11 = 22 33 u1 , X1 u2 = , X2 u3 = , X3

12 = 21 = 23 = 32 31 = 13

1 2 1 = 2 1 = 2

we see that the six independent components of the strain tensor are obtained from the three components of the displacement vector, assuming that u is dierentiable. Now suppose that the strain components are given and the displacement components are to be determined from these equations. In this case since we have six equations and three unknowns there is not necessarily a unique acceptable solution for displacement eld. In other words the set of equations (1.61) poses a restriction on the strain eld, i.e. we can not choose strain components arbitrarily and they are somehow connected. Denition 2. A compatible strain eld is the one for which a single-valued and continuous displacement eld can be found by solving the equation = u. Our objective then will be to nd the condition(s) under which a given strain eld is compatible. One would naturally think of three additional equations to connect strain components to each other. Applying the following identities from tensor calculus u = 0 and u = 0 . (1.62) applying the curl operator from the left-hand-side to equation (1.43) we obtain 1 1 1   = (u + u) = (   u + u) = u (1.63) 2 2 2
s

u1 u2 + X2 X1 u2 u3 + X3 X2 u3 u1 + X1 X3

, , .

(1.61)

1.2.

STRAIN AND COMPATIBILITY

13 (1.64)

and then from the right-hand-side


= 1 1   u =  (u  )  2 2

which nally leaves us with = 0, (1.65) the so called compatibility condition. This condition is obtained from the denition of strain, therefore it is a necessary condition for compatibility, viz if the displacement eld is acceptable then the above condition should be fullled. In a next step we show that this condition is also sucient, i.e. if the condition is fullled then the strain eld is compatible, that is to say, a physically meaningful displacement eld exists such that = u. Since the expression is to vanish for a compatible strain eld, one would dene its value as a measure of incompatibility. Therefore, the incompatibility of the strain is dened as Inc() , (1.66) which makes sense as the compatibility condition can be restated as Inc() = 0 . (1.67) Exercise 8. Prove that the alternative form of compatibility condition
s

(1.68) Now we intend to show that the condition (1.65) is sucient for the strain eld to be compatible in a simply connected domain. For this we assume that the strain eld is given in a simply connected body , and it fullls (1.65), and then it has to be proven that a continuous single-valued displacement u can be obtained from (1.43) if the displacement is given at one point over the body. We denote the point where the displacement is given by X and the given value by u . From (1.41) we have X du = dX + dX (1.69) whose integration gives X dX + dX (1.70) u(X ) = u(X ) + X X Fig. 1.8: Integration path. where the integral is taken over any continuous path C connecting X and X . Using index notation the last term can be calculated as = = X dX dX I dX X (1.71)
+ tr() = + .
0 0 X X 0 0 X0 X0 0 X X X X0 X0 X X0 X0

can also be expressed as

tr(I I ) = 0

k X j Xj dX k = ij X j Xj ij

X X0

X0

k ij X j Xj dX k

14 here integration by parts is used. Then we have


j Xj ij X
X X0

CHAPTER 1.

FUNDAMENTALS

= (X ) (X X ) (X 0 ) (X 0 X ) = (X 0 ) (X 0 X )

and using the result in exercise 6


k k ij X j Xj = X j Xj = w X X X X = 1 1 u X u X = 2 2 X X =
ilj wl

(1.72)

which nally gives (1.73) the so called Michell-Cesaro formula. Now we have to show that the integral term in the above formula is path-independent i.e. u(X ) is a single-valued function. However, we remember that path-independence of a line integral is fullled if and only if the integral vanishes over every closed path, viz we should prove that + X = 0. X dX (1.74) Using Stokes' theorem and the compatibility condition (1.65) we have + X = X (1.75) X X dX dAn + X = 0 . (1.76) = dAn ( ) X Proof of the second step is left as exercise 10. x Exercise 9. Let the tensor A be given as
X

u(X ) = u(X 0 ) + (X 0 ) (X X 0 ) +

+ X X dX

X0

For which values of a and b holds A = 0? Calculate


X3

A = ax2 1 e1 e1 + bx1 x2 e2 e2 .

X2

X3

u(X 3 ) =
0

A ds

for the two paths indicated in Fig. 1.9. Exercise 10. Prove the following identity

Fig. 1.9:
tion.

X1

x1

Paths of integra-

+ ( ( x x) = x x) .

(1.77)

1.2.

STRAIN AND COMPATIBILITY

Let us have a look at the classical solution to the cantilever beam problem, considering its compatibility. The solution to the cantilever beam problem shown in the y gure is of the form = (x) y e e . (1.78) x x where (x) is the curvature at each cross section. The beam is Fig. 1.10: Cantilever beam. built-in at its left end x = 0, where u(0) = 0 and (0) = 0. Check if is compatible. If yes, nd the displacement eld components u (x, y) and u (x, y). Solution. We start with the compatibility condition, applying the curl operator from the right-hand-side to the strain = ((x) y e e ) ( e + e + e ) = (x) e e (1.79) and then from the left = ( e + e + e ) ((x) e e ) = 0 . (1.80) Now with the compatibility condition being fullled, we want to calculate the displacement eld from Michell-Cesaro formula. Applying the boundary conditions at x = 0 we obtain ( + u(x) = dx x x) (1.81) in which
Example 1.2.
xx x x 0 x y x x x x y y z z x z x x y y z z x z x 0

15

( x x) = (( x) ex ez ) (( x x) ex + ( y y ) ey + ( z z ) ez ) = ( x) ( x x) ex ey ( y y ) ex ex .

Substitution into (1.81) yields


x

(1.82) (1.83)

u(x) =
0 x

[ex dx + ey dy ] ( x) y ex ex ( x x) ex ey + ( y y ) ex ex dx ( x) [y ex ( x x) ey ] .
0 x

Therefore and

ux (x, y ) = y
0 x

( x) dx

(1.84) (1.85) (1.86)

uy (x, y ) =

( x) ( x x) dx .
0

Then it is also clear that

uy = (x)

and

ux = y uy .

16
Exercise 11.

CHAPTER 1.

FUNDAMENTALS

For the given strain tensor and boundary conditions


= x2 2 e1 e1 + x1 x2 (e1 e2 + e2 e1 ) u(0) = 0 , (0) = 0 (e1 e2 + e2 e1 )

calculate the displacement eld u(x).

Chapter 2

The Field Equations of Linear Continuum Mechanics


Until now we have developed the basic kinematical and dynamical notions concerning material bodies. As mentioned before, balance laws of linear momentum, angular momentum and energy are not sucient to model mechanical behavior of deformable bodies, and we additionally need the so called constitutive laws. Constitutive laws try to express material deformation versus loading, based on specic thermo-mechanical assumptions. Depending on the type of constitutive formulation one can classify materials to e.g. elastic, plastic, viscoelastic etc. In this chapter we briey address constitutive laws and governing equations of linear elasticity. Derivation of these laws and the respective thermodynamics is postponed to the second part of these lectures in nonlinear continuum mechanics where hyper-elasticity is studied.
2.1 Elasticity

So far we have established two separate sets of equations, namely equilibrium equations (1.22) and kinematic equations of strain-displacement (1.61), and now we want to study stress-strain relations. As will be seen, this completes the set of equations, i.e. there will be as many independent equations as unknowns. We shall conne our study within the so called elastic range, that is the range of strain for which the material behavior remains elastic. A body of material is called elastic if at each body point the strain is a one-to-one function of stress at that point regardless of the history of loading. In formal notation (x) (x) . (2.1) This ensures that stress and strain are both single-valued functions of each other which in turn means that the material follows the same stress-strain curve during loading and unloading, and shows no hysteresis eect. This is more rigorously presented within the framework of hyper-elasticity later on. 17

18

CHAPTER 2.

THE FIELD EQUATIONS OF LINEAR CONTINUUM MECHANICS

2.2

Hooke's law

Experimental data for most hard solids shows a linear relation between load and strain within a specic interval during loading and unloading. In the most general representation this linearity is of the form of generalized Hooke's law as (x) = C (x) : (x) (2.2) which says that each of stress components at any point of a body is a linear function of the strain components at that point. Here the so called stiness tensor C is introduced which is a fourth-order tensor. This linearity could be also represented as (x) = S (x) : (x) (2.3) where S is called compliance tensor which is a fourth-order tensor satisfying the relation S : C = C : S = I. (2.4) and I is the fourth-order identity tensor I= ee e e . (2.5) The functionality of position x is emphasized in equation (2.2), however we usually write = C : . (2.6) Specially for a homogeneous body when C does not depend on x, we have
ik jl i j k l

To further clarify the above equations we can write them in index notation as
ij = Cijkl kl .

(x) = C : (x) .

(2.7) The stiness tensor, being of fourth-order, has 3 3 3 3 = 81 components. However, these components are not independent. Looking into equation (2.2), since stress and strain tensors are both symmetriceach having six independent componentsthe stiness tensor can only have 6 6 = 36 independent components. Because the most general linear relation between the six stress components and the six strain components can take the form of
K K K K K K 11 12 13 14 15 16 11 11 K21 22 22 33 33 = K31 23 223 31 231 212 12 K61 K66

...

..

(2.8)

which is called contracted notation. Here the original stress and strain tensors are recast in a vectorial form. The factor 2 in front of shear strains has technical reasons beyond our scope, however it can be omitted in principle.

A homogeneous body is the one whose physical properties are the same for all of its points.

2.2.

HOOKE'S LAW

19
Cijkl = Cjikl = Cijlk

Remark .

3 Note that the symmetry of stress and strain tensors is fullled only if

which is called minor symmtery of the stiness tensor. We will see later on that the existence of strain energy potential requires the fourth order tensor C to satisfy C =C (2.10) the so called major symmetry, which in turn requires the symmetry of the stiness matrix K in equation (2.8). Therefore the matrix K has only 21 independent components. The above arguments which prove the reduction of number of independent stiness components from 81 to 21 can be summarized as follows Summary 4. Since the stiness tensor has to fulll major and minor symmetries C =C =C =C (2.11) it has only 21 independent components for a linear elastic material.
ijkl klij ijkl klij jikl ijlk

(2.9)

2.2.1 Hooke's law for isotropic materials

As yet, we have simplied the general Hooke's law based on symmetry properties (2.11). Now we consider material symmetries based on which we can further reduce the number of required stiness components. Solids have crystal structures at the atomic scale. These crystal structures are classied based on their symmetry properties. A detailed study of how these symmetries inuence the stiness tensor goes beyond our agenda (see [4]), but regardless of crystallography a typical solid material is composed of microscopic grains with random distributions and orientations throughout the body. Although each grain has a specic crystal structure with its respective symmetries, the overall behavior of the material at the macro scale seems isotropic due to this randomness. Denition 5. An isotropic material is the one that has identical thermo-mechanical properties in all directions at each material point. This requires that the elastic constants be invariant under an arbitrary rotation of coordinate system C =C (2.12) and consequently we come up with two substantial results.
ijkl ijkl

Theorem 6. For an isotropic material, the principal axes of the strain tensor coincide with the principal axes of the stress tensor.

Components of stiness tensor are also called elastic constants.

20

CHAPTER 2.

THE FIELD EQUATIONS OF LINEAR CONTINUUM MECHANICS

Theorem 7. For an isotropic material, the stiness tensor in principal coordinates follows the relationships
C1111 = C2222 = C3333 = C C1122 = C2233 = C3311 = C2211 = C1133 = C3322 =

(2.13) (2.14)

and the rest of components are zero.

So, there are only two independent stiness components for an isotropic material in principal coordinates. But this holds in any other coordinates as well because the rotation transformation from principal coordinates to any other coordinates leaves us with a stiness tensor whose components are expressed in terms of the two components above. For an arbitrary coordinate system one can show the following constitutive equation = tr() I + 2 (2.15) where and are Lam parameters. Where is the same as in (2.14), and is related to the constant C in (2.13) by 2 = C . The above equation in index notation is given by = + 2 . (2.16) The inverse form of equation (2.15) is given by 1 = tr( ) I . (2.17) 2 2 (3 + 2) In view of Hooke's law as given by (2.2) and (2.3) stiness and compliance tensors for an isotropic material are given in terms of Lam parameters as 1 C = I I + 2I and S = I I + I, (2.18) 2 (3 + 2) 2 where I and I are the second-order and fourth-order identity tensors respectively.
ij kk ij ij

2.2.2 Alternative formulations of Hooke's law for isotropic materials

We proceed with reformulations of Hooke's law which are suitable for dierent types of problems. The strain tensor can be additively decomposed into the so called volumetric and deviatoric parts 1 1 (2.19) = vol() + dev () : vol() = tr() I , dev () = tr() I . 3 3 Kinematically, the volumetric part reects dilatation which is the relative change of volume dv dV = tr() (2.20) dV
Volumetric-deviatoric decomposition

21 and the deviatoric part reects distortion. Interesting is that if we decompose the stress in the same fashion into deviatoric and volumetric parts as 1 1 = vol( ) + dev ( ) : vol( ) = tr( ) I , dev ( ) = tr( ) I (2.21) 3 3 then the Hooke's law is decomposed such that volumetric stress depends on volumetric strain only, and deviatoric stress depends on deviatoric strain only by = (3 + 2) , = 2 (2.22) The so called bulk modulus or compression modulus of elasticity is introduce as 2 K =+ (2.23) 3 which based on (2.22) yields vol( ) = 3K vol() (2.24) Remark 8. The deviatoric parts of stress and strain tensors are traceless i.e. tr(dev()) = 0. Maybe the most important aspect of the above decomposition is that deviatoric and volumetric parts are orthogonal to each other in the sense that vol( ) : dev ( ) = 0 , vol() : dev () = 0 . (2.25) Then it follows that : = vol( ) : vol() + dev ( ) : dev () (2.26) which means the elastic energy is decomposed into the volumetric elastic energy and the deviatoric elastic energy.
2.2. HOOKE'S LAW

vol

vol

dev

dev

Young's modulus, shear modulus and Poisson's ratio

Hooke's law for isotropic solids is often expressed in terms of a set of material constants which are more favorable in applied mechanics because they can be directly measured in experiments. Those constants are Young's modulus (2 + 2) E= , (2.27) +
Poisson's ratio
=

and usually renaming one of Lam parameters to shear modulus Then Hooke's law is reformulated as
= G = . E E tr( ) I + (1 + ) (1 2 ) 1+ = 1+ tr( ) I . E E

, 2 ( + )

(2.28) (2.29) (2.30) (2.31)

and its inverse as

22

CHAPTER 2.

THE FIELD EQUATIONS OF LINEAR CONTINUUM MECHANICS

Exercise 12. During the tension test of a prismatic bar with square cross section having the dimensions l = 500mm and a = 30mm, the changes in the dimensions of l = 8mm and d = 0.2mm are obtained after application of a force F = 2000N (Fig. 2.1). Calculate the
l + l a F F a a

Fig. 2.1: Prismatic bar subjected to tension. material constants: , , E, .


2.2.3 Plane-strain and plane-stress problems

Whenever a general theory is developed there are special cases that draw special attention, sometimes due to their broad range of application and other times for the impact they have in development of models. Two well know examples of such cases in elasticity are plane-strain and plane-stress problems. As the name suggests, plane-strain is a type of deformation in a material body when given a specic plane, all the o-plane components of strain are zero or negligibly small. Assume that the plane of strain is normal to X axis in Cartesian coordinates. Then we may express the plane-strain condition by =0 for i = 3 or j = 3 . (2.32) Then Hooke's law for an isotropic material becomes = ( + ) + 2 (2.33) = ( + ) + 2 (2.34) = 2 (2.35) = ( + ) (2.36) = =0 (2.37)
3 ij 11 11 11 22 22 11 22 22 12 33 13 12 11 22 23

Plane-strain

Plane-stress

Now the meaning of plane-stress should be immediately clear by analogy. Plane stress is a condition where for a given plane all the o-plane stress components are zero. If the plane of stress is normal to X axis in Cartesian coordinates then we state the plane-stress condition as =0 for i = 3 or j = 3 . (2.38)
3 ij

2.3.

NAVIER AND BELTRAMI-MICHELL EQUATIONS

23 (2.39) (2.40) (2.41) (2.42) (2.43) (2.44)

With introduction of a new constant the Hooke's law for isotropic materials becomes
12 = 2 12 . = 2 + 2 (11 + 22 ) + 2 11 11 = (11 + 22 ) + 2 22 22 =

Also note that as a result of condition (2.38) we have


33 = 13

Exercise 13.

Prove the following relationships


= 3K E , 6K = 3K 2 , 3

(11 + 22 ) + 2 = 23 = 0 .

E=

(1 + ) (1 2 ) .

2.3

Navier and Beltrami-Michell equations

Now we shall be able to write the governing equations of isotropic elasticity as Hooke's law completes the set of equations. Let us restate all those equations together 1 1 = (u + u) = ( u + u ) (2.45) 2 2 +f =0 +f =0 (2.46) = tr() I + 2 = + 2 . (2.47) In the rst step we would like to have an equation in terms of displacement eld only. For this we have to substitute from (2.45) into (2.47) which yields
ij i j j i i ij ij j kk ij ij

and then plug in this result into (2.46)

1 1 ij = (k uk + k uk ) ij + 2 (i uj + j ui ) 2 2 = k uk ij + (i uj + j ui )

(2.48) (2.49)

i ( k uk ij + (i uj + j ui )) + fj = i k uk ij + i i uj + i j ui + fj = j k uk + i i uj + j k uk + fj = ( + ) j k uk + i i uj + fj = 0 .

(2.50) which is the so called Navier equation. Since u is a vector the equation is basically a system of three partial dierential equations.
( + ) u + u + f = 0 ,

This equation can be written in operator notation as

24

CHAPTER 2.

THE FIELD EQUATIONS OF LINEAR CONTINUUM MECHANICS

Exercise 14.

Prove the alternative form of the Navier equation as


( + 2) u u + f = 0 .

Another step would be to set up a system of dierential equations in terms of stress tensor. Stress tensor has six independent components. The equation (2.46) provides three partial dierential equations which is not sucient. On the other hand equations (2.45) and (2.47) do not provide more equations in terms of stress. Therefore we need additional equations. It turns out the additional equation is the alternative form of compatibility condition (1.68) that we repeat here + tr() = + . (2.51) From equation (2.22) we have 1 2 tr() = tr( ) , (2.52) E and the inverse form of Hooke's law (2.31) 1+ = tr( ) I . (2.53) E E Now we put (2.53) and (2.52) into (2.51) to obtain (2.54) For the second term on the left-hand-side based on equilibrium equation (2.46) we have tr( ) = = = f . (2.55) For the rst term on the right-hand-side again based on (2.46) we have + = (f + f ) , (2.56) because = = f . Considering the second term on the right-hand-side ( (tr( ) I )) = = = (tr( ) ) . (2.57) The same can be easily shown for (tr() I ) therefore (tr( ) I ) = (tr( ) I ) = tr( ) . (2.58) Substituting all these results into (2.54) yields 1 + tr( ) + f I + f + f = 0 , ( = 1) (2.59) 1+ 1 the so called Beltrami-Michell formulation. Note that Beltrami-Michell equation is in fact compatibility equation in terms of stress which together with equilibrium equation + f = 0, establishes a complete system of partial dierential equations for the stress eld.
m m kk ij i k mm kj i j mm ij

1 2 1+ tr( ) I + tr( ) = E E E 1+ ( + ) ( (tr( ) I ) + (tr( ) I ) ) . E E

2.3.

NAVIER AND BELTRAMI-MICHELL EQUATIONS

25

Exercise 15. Exercise 16.

Assume f = const., using Beltrami-Michell equation show that


= 0 and = 0 .

Assuming that
= (r) and f = f (r) ,

derive Beltrami-Michell equation in cylindrical coordinates.

26

CHAPTER 2.

THE FIELD EQUATIONS OF LINEAR CONTINUUM MECHANICS

Chapter 3

Displacement Functions
An elastostatics problem is basically a boundary value problem, with boundary conditions of the type Dirichlet, Neumann or mixed. In chapter 2 we derived the governing equations of linear isotropic elasticity in terms of displacement eld (Navier equation) and in terms of stress eld (Beltrami-Michell and equilibrium equations). In this chapter we present some classical solutions to the displacement formulation in a coherent way. All the methods are meant to the solution of the original boundary value problem. We shall often assume that the body forces are not present which is not a serious restriction as in practical situations, nding the particular solution belonging to body forces is not complicated . Throughout our derivations we frequently employ tensorial and vectorial identities. Our point of departure is the Navier equation with zero body force ( + ) u + u = 0 , (3.1) with prescribed boundary conditions u(x) = u , x (3.2) (x) n = t , x . (3.3) The idea is to substitute u with a combination of scalar or vector functions, called displacement functions, or their derivatives so that the governing equation in terms of these functions takes the form of harmonic or biharmonic equations which are basically simpler to solve than the Navier equation. However, the boundary conditions expressed in terms of displacement functions become more complicated.
simplify
u

3.1

Scalar and vector potentials

The rst approach we are going to present is based on the so called Helmholtz theorem as follows.
We remember that the solution to an inhomogeneous dierential equation is the sum of the general solution to the homogenized equation and the particular solution to the inhomogeneous equation.

27

28

CHAPTER 3.

DISPLACEMENT FUNCTIONS

We know that a curl-free vector eld has the form of v = , and a divergence-free vector eld has the form of w = . Therefore if the displacement eld u fullls the assumptions of Helmholtz theorem we can write u = + , (3.4) where is a scalar and is a vector eld. Now if u is substituted from (3.4) into (3.1) it gives
( + ) u + u = ( + ) [ + ] + [ + ]
( ( = ( + ) [ + ( ( ( ( )] + + (

Theorem 9. Every nite vector eld which is uniform, continuous and vanishing at innity may be decomposed into the sum of a curl-free (irrotational) and a divergence-free (solenoidal) eld.

which nally gives ( + 2 ) + = 0 . (3.6) Any set of and that satises this equation will also satisfy (3.1) when u = + , and for any solution u to (3.1) there exits at least one set of and that satises (3.6), which is not unique because u is expressed in terms of rst derivatives of and . However this is not important because we are interested in nding the solution in terms of u. One particular solution of (3.6) is the solution to = const. and = const. (3.7) and as long as this solution is capable of satisfying the boundary conditions our objective is fullled. The two above equations are called Laplace's equations or harmonic equations. Remark 10. The reader may wonder why we reduced the original problem to the particular case of harmonic functions. The answer is: simplication. However it might be the case that the problem is over-simplied, but we do not know until we try the harmonic solution and see if the boundary conditions are fullled. The good news is that harmonic solutions are rich enough to reect the boundary conditions in many engineering problems we face. Example 3.1. Wave propagation. The dynamic case of wave propagation can be studied with Helmholtz decomposition. Instead of equilibrium equation we have the balance of momentum , + f = u (3.8) is acceleration. Consequently Navier equation becomes where is density and u . ( + ) u + u + f = u (3.9) After substitution of Helmholtz decomposition and derivations we get the particular case of , = , ( + 2) = (3.10)

= ( + ) + + = 0 ,

(3.5)

3.1.

SCALAR AND VECTOR POTENTIALS

29

(3.11) with k being a unit vector called wave-number vector or propagation vector and c and c are compression wave speed and shear wave speed. In order to obtain the wave speed we have (k x c t) = = = k k = = (3.12) = =c which by substitution into the wave equation for yields = c , ( 2) (3.13) therefore + 2 . (3.14) c = In the same way we can show that the shear wave speed is 2 . (3.15) c =
(k x cp t) , = (k x cs t) , =
p s i i i i j j p t t 2 p 2 p p s

which are called wave equations. The solution will be of the form

z p(r) h

Fig. 3.1: Axially symmetric slab.


Example 3.2.

A two-dimensionally innite slab is loaded at the top with a given axially symmetric distribution p(r) in cylindrical coordinates, and constrained with a frictionless surface at the bottom. We want to nd the solution in terms of Helmholtz functions. Solution. To understand the problem we have to express boundary conditions. At the top we have a vertical force distribution therefore z=h : = p(r) , = 0 (3.16) and at the bottom there is no vertical movement and tangential force (no friction!) z=0 : u = 0 , = 0. (3.17)
zz rz z rz

30 Axial symmetry means that u, and are independent of . Back to equations (3.7) we assume the special case =0 (3.18) = 0 , (3.19) and we have to see if the solution can fulll all the boundary conditions. If it does not then we will assume a more sophisticated case. Now we need to express the boundary conditions in terms of . Starting from the above assumption 1 = (u + u) = , (3.20) 2 therefore based on Hooke's law (2.15)  = tr() I + 2 =  I + 2 = 2 . (3.21) Then in cylindrical coordinates u = , u = (3.22) 1 (3.23) = 2 , = 2 r = 2 , = 2 , (3.24) and the boundary conditions (3.16) and (3.17) in terms of are obtained as z = h : 2 = p(r) , 2 = 0 (3.25) z = 0 : = 0 , 2 = 0 (3.26) The Laplace's equation (3.19) in cylindrical coordinates reads 1 = + + = 0 . (3.27) r For separation of variables we put (r, z ) = f (r) g (z ) , (3.28) which results in 1 f g + f g + fg = 0 (3.29) r after factorization f + 1/r f g = = k , (3.30) f g where k reects the possibility of a positive or negative constant. Since the rst expression in the above equation is a function of r only and the second is a function of z only, therefore the only possible way for them to be equal is to be constant.
CHAPTER 3. DISPLACEMENT FUNCTIONS

rr

2 r

rz

zz

2 z

r z

2 z

r z

r z

2 r

2 z

Separation of variables is a standard method to reduce PDE's to ODE's.

31 The negative choice k leads to contradiction with boundary values. Therefore we take the positive constant +k which yields two ordinary dierential equations 1 f + f k f = 0 , g + k g = 0. (3.31) r The rst equation is the so called Bessel's equation of zeroth order which has the general solution f (r) = A J (kr) + B H (kr) , (3.32) where J is Bessel function of zeroth order and H is Hankel function of zeroth order. The second equation has the well know real solution g (z ) = C sin(kz ) + D cos(kz ) . (3.33) Since H (x) for x 0 (3.34) in order to have a nite solution in the neighborhood of r = 0 it must hold that B = 0, then the general solution becomes = J (kr) (C sin(kz ) + D cos(kz )) . (3.35) Applying the boundary conditions gives u (0) = (0) = J (kr) k C = 0 C=0 (3.36) (0) = 2 (0) = 2 k J (kr) k C = 0 (3.37)
3.1. SCALAR AND VECTOR POTENTIALS

rz

r z

(3.38) As you can see there are innitely many possible values for k and therefore there are innitely many solutions of the form (r, z ) = J (k r) cos(k z ) , (3.39) which are called normal modes. The general solution is the linear combination of all possible solutions as = a . (3.40) This situation is a bit tricky because there are innitely many constants a 's to determine but there is only one remaining boundary condition to be fullled as
n 0 n n n n n=0 n 2 zz (h) = 2z = 2 n

rz (h) = sk J0 (kr) (D k sin(kh)) = 0 sin(kh) = 0 kn =

n , n = 0, 1, . . . h

kn a2 n J0 (kn r ) cos(kn h) = 2
n 2 (1)n an kn J0 (kn r) = p(r) .

(3.41)

32 The key idea is the orthogonality property of Bessel functions as J (k r) J (k r) r dr = 0 for i = j . (3.42) Therefore if we multiply both sides of (3.1) with J (k r) r and then integrate over [0, ] we obtain p(r) J (k r) r dr (3.43) (J (k r)) r dr = 2(1) a k because all the terms except the i term vanish due to (3.42), and a is obtained as p(r) J (k r) r dr 1 a = (3.44) 2(1) k (J (k r)) r dr which nally yields p(r) J (k r) r dr 1 J (k r) cos(k z ) . (3.45) = 2(1) k (J (k r)) r dr
CHAPTER 3. DISPLACEMENT FUNCTIONS

2 i i

th

i 2 i

0 0

n=0

i 2 n

0 0

potentials

Exercise 17.

Using Helmholtz theorem calculate u, and for the given scalar and vector
2 = k x2 1 x2 ;

A circular plate with radius R and thickness h is constrained at the top and bottom by frictionless surfaces (Fig. 3.2). A load distribution p(z) is applied on its vertical boundary. Using Helmholtz theorem derive solution for axially symmetric problem.
Exercise 18.
z R p (z ) h r

2 x2 1 x2 . 0 = l 0

Fig. 3.2: Circular plate under radial loading. Hint: If the ansatz = J (kr) [C sin(kz ) + D cos(kz )] is used, the solution in the form = J (k r) cos k z will be obtained. Then Fourier expansion in z should be used.
0 n 0 n n

3.2

Galerkin vector

Helmholtz decomposition gives displacement eld in terms of rst derivatives of a scalar and a vector potential. It is also possible to express u in terms of second derivatives of a vector eld. Let us introduce the Galerkin vector V which is related to u by u = k V V (3.46) where k is a constant.

3.2.

GALERKIN VECTOR

33
u

Theorem 11. For any vector eld satises the relation (3.46).

and constant k there exists a vector eld V that

Therefore the relation (3.46) provides a completely general solution if V is determined somehow. If the Galerkin vector is plugged into Navier equation (2.50) it gives ( + ) k V ( + ) V + k V V + f = 0 (3.47) and then [( + ) k ( 2)] V + k V + f = 0 , (3.48) as we are allowed to chose the constant k, the best choice is the one that simplies the above equation the most, which turns out to be + 2 k= = 2 (1 ) , (3.49) + and this gives 2 (1 ) V + f = 0 , (3.50) which is a vectorial biharmonic equation. Biharmonic equations are well studied in analysis and nding the answer to the above equation is basically simpler than the original Navier equation. Then we obviously need the relation (3.46) to transform the answer in terms of V to u. Remark 12. Galerkin vector is related to Helmholtz potentials by = V , = 2 (1 ) V . (3.51) There is a special case of Galerkin vector which can be used in axially symmetric problems. Assuming that the only non-vanishing component of Galerkin vector is the X component we have V =V =0, V =L (3.52) where L is Love's strain function. Then (3.50) simplies to 2 (1 ) L + f = 0 . (3.53) Note that f = f = 0 which is the symmetry requirement. The relation (3.46) in terms of Love's function reads u = L , u = L , u = 2 (1 ) L L . (3.54) Example 3.3. In the case of spherical symmetry L is a function of r only (3.55) L = L(r) : r= xx = x
3 1 2 3 3 1 2 1 1 3 2 2 3 3 2 3 i i

3.2.1 Love's strain function

34 and Laplacian in spherical coordinates is given by

CHAPTER 3.

DISPLACEMENT FUNCTIONS

Therefore in the absence of body forces equation (3.53) becomes


L =

2 2 r + r . r

(3.56) (3.57)

The general solution to this equation has the form


i

2 2 r + r r

2 4 3 2 4 r + r L = r L + r L = 0. r r

(3.58) with C 's being integration constants to be determined by application of boundary conditions.
L = C1 r1 + C2 + C3 r + C4 r2 ,

The classical problem of Kelvin deals with a concentrated force in an innite medium. This is an important case whose solution is used to solve more complicated problems. Let us assume the concentrated force at the origin of coordinate sysf e tem and directed towards X axis as f = f (x) e (3.59) where (x) is Dirac's delta function in 3-dimensional space. Due to the axial symmetry we use the Love's formulation of Fig. 3.3: Kelvin's problem. Navier equation with body force 2 (1 ) L = f (x) . (3.60) Integrating both sides within a sphere of radius R surrounding the origin yields 2 (1 ) L dV = f (x) dV = f , (3.61) where we have used the fundamental property of Dirac's delta (x) dV = 1. On the other hand we know that L = L, and using Gauss theorem L dV = L dV = n L dS = L dS . (3.62) The general solution of the form (3.58)
3 0 3 0 3 0 0 0 rR rR r R r r R r R r=R r =R

Kelvin's problem

is simplied to

L = C1 r 1 + C2 + C3 r + C4 r 2 L = C1 r1 + C2

(3.63)

3.3.

DISPLACEMENT FUNCTIONS OF PAPKOVICH-NEUBER

35

because otherwise

(3.64) which is not acceptable as the solution has to stay nite. Then plugging (3.63) into the surface integral in (3.62) gives 1 L dS = 4R 2C = 8C (3.65) R then putting back into (3.61) results in 16 (1 ) C = f (3.66) and nally the answer is f L= r. (3.67) 16 (1 ) Exercise 19. For the Love strain function in (3.67) calculate u and .
x : L
r 2 3 r=R 2 3 3 0 0

3.3

Displacement functions of Papkovich-Neuber

So far we have managed to replace the Navier equation by the fourth-order biharmonic equation (3.50) and the third-order equation (3.6). Note that the harmonic solutions to (3.6) are special cases that satisfy the original third-order equation, not the general solution to it. Now it would be nice if we could have an equivalent formulation to Navier equation which is of the second-order i.e. one order lower with respect to (3.6) and two orders lower with respect to (3.50). Starting from Galerkin vector (3.46), if the vector eld A and the scalar eld b are introduced such that 1 A = V , b = V (3.68) 2 then we can write displacement eld in terms of A and b u = 4 (1 ) A b . (3.69) Therefore 1 A = V = 0 , see (3.50) (3.70) 2 b = V = V = 2 A . (3.71) Exercise 20. For the Papkovich-Neuber solution u = 4 (1 ) A b show that (A X ) = 2 A . (3.72)

36 Using the result in the recent exercise the general form of b is given by
CHAPTER 3.

DISPLACEMENT FUNCTIONS

(3.74) A = 0 , a = 0 . (3.75) This is called Papkovich-Neuber solution which is equivalent to Galerkin's solution, and since Galerkin's solution is equivalent to Navier equation, then the above formulation is completely general and equivalent to our original formulation of Navier. Example 3.4. Problem of Leon. Find the displacement eld in a uni-axially loaded innite medium with a spherical cavity (Fig. 3.4). Solution. Laplace's equation in spherical coordinates (independent of due to axial symmetry of our problem), reads f 2 f cos f 1 f f = + + + =0 (3.76) r r r r sin r R and we have to solve (3.75) A = 0 , a = 0 . (3.77) Because of axial symmetry around the axis of loading the only nonzero component of A is A , which means A = A = 0 and we are left with A = 0 , a = 0 . (3.78) Fig. 3.4: Leon's problem. By separation of variables we seek solutions of the form f (r, ) = R(r) () (3.79) that ends up in 1 d dR 1 d d r = sin = l (l + 1) , l {0, 1, 2, . . .} . (3.80) R dr dr sin d d The special form of the constant l (l + 1), for l being a non-negative integer, guarantees a convergent solution of Legendre equation that appears here in terms of = cos d d 2 + l (l + 1) = 0 . (3.81) 1 d d
u = 4 (1 ) A [A x + a]

Note that a can be any harmonic function in general. Then this all comes down to

b=Ax+a,

a = 0 .

(3.73)

To derive (3.80) into (3.81) we need to know from the chain rule that
d d = sin d d d 1 d = d sin d

3.3.

DISPLACEMENT FUNCTIONS OF PAPKOVICH-NEUBER

37

The general solution is given by (3.82) where P (x) is Legendre polynomial of degree l. Then the general solution to (3.78) is obtained 1 P (cos ) (3.83) a= cr +d r
f= Arl + B 1 rl+1 Pl (cos ) ,
l l l l l+1 l l=0

Az =
l=0

Cl r l + D l

1 rl+1

Pl (cos ) .

(3.84)

(3.85) r = R : = = 0. (3.86) In order to homogenize the boundary condition the following re-parameterization is introduced = , = e e . (3.87) Then according transformation rules for second-order tensors = cos , = sin , = sin cos . (3.88) the boundary conditions are And in terms of 0 c = 0, C = 0 r : (3.89) = . (3.90) r=R : = , Considering the number of boundary conditions to be fullled we can start from a limited number of terms in (3.83) and (3.84), say l = 0, 1, 2 to see if the boundary conditions are fullled. If we need more terms we can add them later on, so let us assume 1 1 1 a=d +d cos + d 3 cos 1 (3.91) r r r 1 1 1 cos + D 3 cos 1 . (3.92) A =D +D r r r (3.93) To express the boundary conditions in terms of a and A we use the following result whose proof is left as an exercise u = 4 (1 ) A cos (A r cos + a) (3.94) 1 u = 4 (1 ) A sin (A r cos + a) . (3.95) r
r : zz =
rr

The boundary conditions are given as

z z

rr

rr

rr

1 2

2 3

1 2

2 3

38 We also need the following formulas


rr = r ur ,

CHAPTER 3.

DISPLACEMENT FUNCTIONS

(3.96) (3.97) (3.98) It is left to the reader as an important exercise to apply the boundary conditions and obtain the following results 1 6 5 d = R , D = 0, (3.99a) 4 7 5 5 1 d = 0, D = R , (3.99b) 4 7 5 1 1 R , D = 0. (3.99c) d = 4 7 5
1 1 = u + ur r r 1 1 1 r = ur + r u u 2 r r 1 1 = ur + cot u . r r
0 3 0 1 1 3 2 5 2

Exercise 21.

Using Papkovich-Neuber solution


u = 4 (1 ) A b , b=AX +a

for the ansatz


a=
n=0

cn rn + dn

1 rn+1

pn (cos )

A1 = A2 = 0 A3 =
n=0 n

Cn r n + D n

1 rn+1

pn (cos )

with p (cos ) being Legendre polynomial, show that Using Papkovch-Neuber solution for a sphere with radius R and surface displacement u (R, ) = 0 and u (R, ) = u sin nd (r, ). Hint: use ansatz of the previous assignment but only up to n = 1 for a and A .
Exercise 22.
r 0 3

ur = 4 (1 ) A3 cos r (A3 r cos + a) 1 u = 4 (1 ) A3 sin (A3 r cos + a) . r

Bibliography
[1] A.I. Borisenko and I.E. Tarapov. Vector and Tensor Analysis with Applications. Dover Pub. Inc, 1968. [2] P.C. Chou and N.J. Pagano. Elasticity: Tensor, Dyadic, and Engineering Approaches. Dover Pub. Inc, 1992. [3] G.A. Holzapfel. Nonlinear Solid Mechanics: A Continuum Approach for Engineering. Wiley, 2000. [4] A.E. Love. Treatise on the Mathematical Theory of Elasticity. Dover Pub. Inc, 4 edition, 1927. [5] R.W. Ogden. Non-Linear Elastic Deformations. Dover Pub. Inc, 1997. [6] M.R. Spiegel. Mathematical Handbook of Formulas and Tables. McGraw-Hill (Schaum's outline), 1979. [7] T.C.T. Ting. Anisotropic Elasticity: Theory and Applications. Oxford University Press, 1996.
th

39

Vous aimerez peut-être aussi