Vous êtes sur la page 1sur 33

Brief introduction to continuum mechanics

Lecture notes to the course MEC-2410 Mechanics of materials - spring 2012

Reijo Kouhia March 7th , 2012


This is a very preliminary versio and contains many typos. I am very pleased to have a note of any error and any kind of idea on improving the text is wellcome. My e-mail address is reijo.kouhia@tut.fi.

Chapter 1 Introduction
1.1 The general structure of continuum mechanics

In principle the general structure of equations in continuum mechanics is threefold. First, there is a balance equation (or balance equations), stating the equilibrium or force balance of the system considerer. These equations relate e.g. the stress with external forces. Secondly, the stress is related to some kinematical quantity, like strain, by the constitutive equations. Thirdly, the strain is related to displacements by the kinematical equations. Balance equations are denoted as B = f , where B is the equilibrium operator, usually a system of differential operators. In the constitutive equations = C the operator C can be either an algebraic or differential operator. Finally the geometrical relation i.e. the kinematical equations are denoted as = Bu. These three equations form the system to be solved in continuum mechanics, and it is illustrated in gure 1.1. The equilibrium operator B is the adjoint operator of the kinematical operator B . Therefore there are only two independent quantities in the system. Example - axially loaded bar. The equilibrium equation in terms of the axial force N is dN = f, (1.1) dx where f is the distributed load, [force/length], in the direction of the bars axis. Thus, the equilibrium operator B is d (1.2) B = . dx The axial force is related to the strain via the elastic constitutive equation (containing the cross-section area as a geometric quantity) N = EA. 1 (1.3)

CHAPTER 1. INTRODUCTION

B = f
6 -

f B CBu = f

= C = Bu 

Figure 1.1: The general structure of equations in mechanics. In this case the constitutive operator C is purely algebraic constant C = EA. The kinematical relation is du = , (1.4) dx thus the kinematic operator d , (1.5) B= dx for which B is clearly the adjoint. The equilibrum equation expressed in terms of axial displacement is du d EA = f. (1.6) B CBu = dx dx Example - thin beam bending. The equilibrium equation in terms of the bending moment M is d2 M 2 = f, (1.7) dx where f is the distributed transverse load [force/length]. Thus, the equilibrium operator B is d2 B = 2 . (1.8) dx The bending moment is related to the curvature via the elastic constitutive equation (containing the inertiaa of the cross-section as a geometric quantity) M = EI. (1.9)

Again, the constitutive operator C is purely algebraic constant C = EI . The kinematical relation is d2 v (1.10) = 2, dx

1.1. THE GENERAL STRUCTURE OF CONTINUUM MECHANICS


thus the kinematic operator d2 , dx2

B=

(1.11)

for which B is clearly the adjoint, and also in this case B = B . The equilibrum equation expressed in terms of axial displacement is

B CBu =

d2 dx2

EI

d2 v dx2

= f.

(1.12)

Example - linear 3-D elasticity. The equilibrium, constitutive and kinematical equations are

div = b , = C , = sym grad u ,

(1.13) (1.14) (1.15)

where is the symmetric stress tensor, is the material density, b is the body force per unit mass, u is the displacement vector and C is the elasticity tensor. Thus the operators B , B and C are

B = div, B = sym grad C = C.

(1.16) (1.17) (1.18)

The formal adjoint of the B = div operator is the gradient operator.

CHAPTER 1. INTRODUCTION

1.2 Tensor analysis 1.3 Nomenclature


Strain and stress e , eij s , sij s1 , s2 , s3 oct ij oct v 1 , 2 , 3 , ij m oct 1 , 2 , 3 m oct = = = = = = = = = = = = = = = = = deviatoric strain tensor deviatoric stress tensor principal values of the deviatoric stress shear strain octahedral shear strain strain tensor octahedral strain volymetric strain princial strains normal stress stress tensor mean stress octahedral stress principal stresses shear stress mean shearing stress octahedral shear stress

Invariants I1 (A) = tr A = Aii 1 I2 (A) = 2 [tr(A2 ) (tr A)2 ] I3 = det A 1 J2 ( s ) = 2 tr s 2 J3 = det s , , = = = = = = = = = the rst invariant of tensor A second invariant third invariant second invariant of the deviatoric tensor s third invariant of a deviatoric tensor the Heigh-Westergaard stress coordinates hydrostatic length the length of the stress radius on the deviatoric plane the Lode angle on the deviatoric plane

1.4. ON THE REFERENCES


Material parameters E G Gf K k m , = = = = = = = = = Youngs modulus shear modulus fracture energy bulk modulus shear strength fc = mft Drucker-Prager murtoehdon parametreja Poissons ratio internal friction angle of the Mohr-Coulomb criterion

1.4

On the references

This lecture notes is mostly based on the following excellent books: 1. L.E. Malvern: Introduction to the Mechanics of a Continuous Medium. Beautifully written treatease on the topic. 2. G.A. Holzapfel: Nonlinear Solid Mechanics, A Continuum Approach for Engineers. A modern treatment of some basic material in Malverns book. Contains usefull material for understanding nonlinear nite element methods. 3. J. Lemaitre, J.-L. Chaboche: Mechanics of Solid Materials. 4. N.S. Ottosen, M. Ristinmaa: Mechanics of Constitutive Modelling.

CHAPTER 1. INTRODUCTION

Chapter 2 Stress
2.1 Stress tensor and the theorem of Cauchy

Lets consider a body B in a 3-dimensional space occupying a volume domain , see gure 2.1. If the body B is divided into two parts by a surface S and the parts separated from each other. The force acting on a small surface S is denoted by f . A traction vector t is dened as f df t = lim = . (2.1) S 0 S dS The traction vector depends on the position x and also on the normal direction n of the surface, i.e. t = t (x , n ), (2.2) a relationship, which is called as the postulate of Cauchy.1 In the rectangular cartesian coordinate system, the traction vectors action in three perpendicular planes, parallel to the coordinate axes are denoted as t 1 , t 2 and t 3 , see gure 2.2. The components of the traction vectors are shown in the gure and expressed in terms of the unit vectors parallel to the coordinate axes e i the traction vectors are t 1 = 11 e 1 + 12 e 2 + 13 e 3 , t 2 = 21 e 1 + 22 e 2 + 23 e 3 , t 3 = 31 e 1 + 32 e 2 + 33 e 3 . (2.3) (2.4) (2.5)

To obtain the expression of the traction vector in terms of the components ij , let us consider a tetrahedra where the three faces are parallel to the coordinate planes and the remaining one is oriented in an arbitrary direction, see gure 2.3. In each of the faces, the average tranction is denoted as t i , where i = 1, 2, 3, n, and the area of the triangle A1 A2 A3 is denoted as S and S1 , S2 , S3 are the areas of triangles OA2 A3 , OA3 A1 and OA1 A2 , respectively. The body force acting on the tetrahedra is b V , where the volume element V = 1 hS , and h is the distance ON . 3
Sometimes traction vector t is also called as a stress vector. However, in this lecture notes this naming is not used since the stress has a tensor character.
1

CHAPTER 2. STRESS

Figure 2.1: A continuum body and the traction vector.

Figure 2.2: Traction vectors in three perpendicular directions.

2.1. STRESS TENSOR AND THE THEOREM OF CAUCHY

Figure 2.3: Cauchys tetrahedra. Figure from [3, page 75].

10 Equilibrium equation for the tetrahedra is

CHAPTER 2. STRESS

1 t n S + 3 b hS t 1 S1 t 2 S2 + t 3 S3 = 0,

(2.6)

which can be written as


1 S (t n + 3 b h n1 t 1 n2 t 2 + n3 t 3 ) = 0.

(2.7)

Now, letting h 0 we get t i t i and


3

tn =
i=1

n1 t i = ni t i = n1 (11 e 1 + 12 e 2 + 13 e 3 ) + n2 (21 e 1 + 22 e 2 + 23 e 3 ) + n3 (31 e 1 + 32 e 2 + 13 e 3 ), (2.8)

or

Notice the transpose in the stress tensor in the last expression. The stress tensori , expressed in rectangular cartesian coordinate system is 11 12 13 xx xy xz x xy xz = 21 22 23 = yx yy yz = yx y yz . (2.10) 31 32 33 zx zy zz zx zy z

n1 11 + n2 21 + n3 31 t n = n2 12 + n2 22 + n3 32 = n T = T n . n3 13 + n2 23 + n3 32

(2.9)

The form of the right hand side is know as von Krmn notation and the -symbol in it describes the normal component of the stress and the shear stresses, respectively. Such notation is common in engineering literature. The equation (2.9) is called the Cauchy stress theorem and it can be written as t (x , n ) = [ (x )]T n , (2.11)

expressing the dependent quantities explicitely. It says that the traction vector depends linearly on the normal vector n .

2.2 Coordinate transformation


If the stress tensor (or any other tensor) is known in a rectangular Cartesian coordinate system (x1 , x2 , x3 ) with unit base vectors e 1 , e 2 , e 3 and we would like to know its components in other recangular Cartesian coordinate system (x1 , x2 , x3 ) with unit base vectors

2.3. PRINCIPAL STRESSES AND -AXES

11

e 1 , e 2 , e 3 a coordinate transformation tensor is needed. Lets write the stress tensor in the xi -coordinate system as = 11 e 1 e 1 + 12 e 1 e 2 + 13 e 1 e 3 + 21 e 2 e 1 + 22 e 2 e 2 + 23 e 2 e 3 + 31 e 3 e 1 + 32 e 3 e 2 + 33 e 3 e 3 . (2.12) This kind of representation is called the dyadic form, and many times the base vector part e i e j is written either as e i e j , or in matrix notation e i e T j . It underlines the fact that a tensor contains not only the components but also the base in which it is expressed. Using the Einsteins summation convention it is briey written as = ij e i e j = ij e i e j , (2.13)

indicating the fact that the tensor is the same irrespectively in which coordinate system it is expressed. Now, taking a scalar product by parts with the vector e k from the left and with e p from the right, we obtain
ij e k e i e j e p = ij e k e i e j e p ki jp ki jp

(2.14)

thus
kp = ki pj ij

or in matrix notation [ ] = [ ][ ][ ]T ,

(2.15)

where the compnents of the transformation matrix are ij = e i e j . Notice that is the transformation from xi -system to xi -cordinate system.

2.3

Principal stresses and -axes


(ij ij )nj , (2.16)

The pricipal values of the stress tensor are obtained from the linear eigenvalue problem

where the vector ni denes the normal of the plane where the principal stress acts. The homogeneous system (2.16) has solution only if the coefcient matrix is singular, thus the determinat of it has to vanish, ad we obtain the characteristic equation 3 + I1 2 + I2 + I3 = 0. The coefcients Ii , i = 1, . . . , 3 are I1 = tr = ii = 11 + 22 + 33 , 2 1 [tr( 2 ) (tr )2 ] = 2 (ij ji I1 ), I2 = 1 2 I3 = det(ij ). (2.18) (2.19) (2.20) (2.17)

12

CHAPTER 2. STRESS

Solution of the characteristic equation gives the principal values of the stress tensor, i.e. principal stresses 1 , 2 ja 3 , which are often numbered as: 1 2 3 . The coefcients I1 , I2 and I3 are independent of the chosen coordinate system, thus they are called invariants.2 Notice, that the principal stresses are also independent of the chosen coordinate system. Invariants have a central role in the development of constitutive equations, as we will see in the subsequent chapters. If the cordinate axes are chosen to coincide to the principal directions n i (2.16), the stress tensor will be diagonal 1 0 0 = [ij ] = 0 2 0 . (2.21) 0 0 3 The invariants I1 , . . . , I3 have the following forms expressed in terms of the principal stresses I1 = 1 + 2 + 3 , I2 = 1 2 2 3 3 1 , I3 = 1 2 3 . (2.22) (2.23) (2.24)

2.4 Deviatoric stress tensor


The stress tensor can be additively decomposed into a deviatoric part, describing a pure shear state and an isoropic part describing hydrostatic pressure. ij = sij + m ij , (2.25)

1 1 I1 = 3 kk is the mean stress and sij the deviatoric stress tensor, for which where m = 3 the notation is also used in the literature. From the decomposition (2.25) it is observed that the trace of the deviatoric stress tensor will vanish

trs = 0. The principal valuse s of the deviatoric stress tensor s can be solved from |sij sij | = 0, giving the characteristic equation s3 + J1 s2 + J2 s + J3 = 0,
2

(2.26)

(2.27)

(2.28)

The invariants appearing in the characteristic equation are usually called as principal invariants. Notice, that in this note the second invariant is often dened as of opposite sign. However, we would like to dene the principal invariants of the tensor and its deviator in a similar way.

2.5. OCTAHEDRAL PLANE AND STRESSES

13

where J1 , . . . , J3 are the invariants of the deviatori stress tensor. They can be expressed as J1 = trs = sii = sx + sy + sz = 0, 1 1 1 [tr(s 2 ) (trs )2 ] = 2 tr(s 2 ) = 2 sij sji J2 = 2 2 2 2 2 2 1 = 6 [(x y ) + (y z ) + (z x )2 ] + xy + yz + zx (2.29) (2.30) (2.31) (2.32) (2.33)

=1 [(1 2 )2 + (2 3 )2 + (3 1 )2 ], 6 J3 = det s .

The deviatoric stress tensor is obtained from the stress tensor by substracting the isotropic part, thus the principal directions of the deviatoric stress tensor coincide to the principal directions of the stress tensor itself. Also the principal values of the deviatoric stress tensor are related to those of the stress tensor as s1 1 m s2 = 2 m . (2.34) s3 3 m The deviatoric invariants exressed in terms of the principal values are
2 ( s2 + s2 J2 = 1 2 + s3 ) , 2 1 3 1 3 J3 = 3 ( s1 + s3 2 + s3 ) = s1 s2 s3 .

(2.35) (2.36)

2.5

Octahedral plane and stresses

Octahedral plane is a plane whose normal makes equal angles with each of the princial axes of stress. In the principal stress space the normal to the octahedral plane takes the form 1 n = [n1 , n2 , n3 ]T = [1, 1, 1]T (2.37) 3 The normal stress on the octahedral plane is thus
2 2 1 oct = ij ni nj = 1 n2 1 + 2 n2 + 3 n3 = 3 (1 + 2 + 3 ) = m

(2.38)

and for the shear stress on the octahedral plane, the following equation is obtained
2 2 oct = ti ti oct = ij ik nj nk (ij ni nj )2 .

(2.39)

Expressed in terms of principal stresses, the octahedral shear stress is


2 2 2 2 1 oct =3 (1 + 2 + 3 ) 1 (1 + 2 + 3 )2 9 1 =9 [(1 2 )2 + (2 3 )2 + (3 1 )2 ].

(2.40) (2.41)

By using the second invariant of the deviatoric stress tensor, the octahedral shear stress has the form oct = 2 J. (2.42) 3 2

14

CHAPTER 2. STRESS

2.6 Principal shear stresses


It is easy to see with the help of Mohrs circles that the maximun shear stress is one-half of the largest difference between any two of the principal stresses and occurs in a plane whose unit normal makes an angle of 45 with each of the corresponding principal axes. The quantities
1 |2 3 |, 1 = 2 1 2 = 2 |1 3 |,

3 = 1 | 2 | 2 1

(2.43)

are called as principal shear stresses and max = max(1 , 2 , 3 ) or max = 1 | 3 |, 2 1 if the convention 1 2 3 is used. (2.45) (2.44)

2.7 Geometrical illustration of stress state and invariants


The six-dimensional stress space is difcult to elucidate, therefore the principal stress space is more convenient for illustration purposes. Lets consider a three-dimensional euclidean space where the coordinate axes are formed from the principal stresses 1 , 2 and 3 , see gure 2.4. Considering the stress point P (1 , 2 , 3 ), this the vector OP can be thought to represent the stress. The hydrostatic axis is dened through relations 1 = 2 = 3 , and it makes equal angle to each of the principal stress axes, and thus the unit vector parallel to the hydrostatic axis is 1 (2.46) n = [1, 1, 1]T . 3 Since the deviatoric stress tensor vanishes along the hydrostatic axis, the plane perpendicular to it is called the deviatoric plane. The special deviatoric plane going through the origin, i.e. 1 + 2 + 3 = 0, (2.47) is called the -plane, A stress state on the -plane is a pure shear stress state. The vector OP can be divided into a component parallel to the hydrostatic axis ON and a component lying on the deviatoric plane NP , which are thus perpendicular to each other. The length of the hydrostatic part ON is 1 = |ON | = OP n = I1 = 3m = 3oct , 3 (2.48)

2.7. GEOMETRICAL ILLUSTRATION OF STRESS STATE AND INVARIANTS 15


1 P (1 , 2 , 3 ) hydrostatic axis n O N 3 2

Figure 2.4: Princial stress space. and its component representation has the form 1 m 1 ON = m = 3 I1 1 . 1 m

(2.49)

Respectively, the component NP on the devatoric plane is 1 m s1 NP = 2 m = s2 . 3 m s3 The length NP squared is


2 2 2 2 2 = |NP |2 = s2 1 + s2 + s3 = 2J2 = 3oct = 5m .

(2.50)

(2.51)

The invariants I1 and J2 have thus clear geometrical and physical interpetation. The cubic deviatoric invariant J3 is ralated to the angle , dened on the deviatoric plane as an angle between the the projected 1 -axis and the vector NP , see gure 2.5. The vector e 1 is a unit vector in the direction of the projected 1 -axis, and has the form 2 1 (2.52) e 1 = 1 . 6 1 The angle can then be determined by using the dot product of vectors NP and e 1 as NP e 1 = cos , (2.53)

16 1 P e1 N 2 3

CHAPTER 2. STRESS

Figure 2.5: Deviatoric plane. The projections of the principal stress axes are shown with dashed line. which gives 3 s1 21 2 3 1 (2s1 s2 s3 ) = = . cos = 2 3 J2 2 3 J2 2 3 J2 From the trigonometric identty it is obtained cos 3 = 4 cos3 3 cos and 2 J3 3 3 J3 cos 3 = = 3 . 3 / 2 2 J2 oct (2.55) (2.54)

(2.56)

A stress space described by the coordinates , ja is called the Heigh-Westergaard stress space.

Chapter 3 Kinematical relations


3.1 Motion of a continuum body

Motion of a continuum body B embedded in a three-dimensional Euclidean space and occupying a domain will be studied. Consider a point P which has an initial position X at time t = 0. At time t > 0 the body occcupies another conguration and the motion of the particle P is described by mapping x = (X , t), or in index notation xi = i (Xk , t). (3.1)

The motion is assumed to be invertibe and sufciently many time differentiable. The displacement vector is dened as u = x X. (3.2)

3.2

Deformation gradient

The most important measure of deformation in non-linear continuum mechanics is the deformation gradient, which will be introduced next. Consider a material curve at the initial conguration, a position of a point on this curve is given as X = ( ), where denotes a parametrization, see gure 3.1. Notice that the material curve does not depend on time. During the motion, the material curve deforms into curve x = (, t) = (( ), t). (3.3)

The tangent vectors of the material and deformed curves are denoted as dX and dx , respectively, and they are dened as dX = ( )d, dx = (, t)d = ( )d = F dX , X 17 (3.4) (3.5)

18

CHAPTER 3. KINEMATICAL RELATIONS

Figure 3.1: Deformation of a material curve, gure from [1, page 70]. since on the deformed curve x = (, t) = (( ), t). The quantity F is called the deformation gradient and it describes the motion in the neighbourhod of a point. It is dened as i F = , or in indicial notation Fij = . (3.6) X Xj The deformation gradient reduces into identity tensor I if there is no motion, or the motion is a rigid translataion. However, rigid rotation will give a deformation gradient not equal to the identity.

3.3 Denition of strain tensors


Lets investigate the change of length of a line element. Denoting the length of a line element in the deformed conguration as ds and as dS in the initial conguration, thus 1 1 1 [(ds)2 (dS )2 ] = (dx dx dX dX ) = (F dX F dX dX dX ) 2 2 2 1 T (3.7) = dX (F F I )dX = dX E dX , 2 where the tensor 1 E = (F T F I ) 2

(3.8)

is called the Green-Lagrange strain tensor. Lets express the Green-Lagrange strain in terms of displacement vector u . It is rst observed that the deformation gradient takes the form F = u =I + , X X (3.9)

3.3. DEFINITION OF STRAIN TENSORS

19

where the tensor u / X is called the displacement gradient. Thus the Green-Lagrange strain tensor takes the form E= = or in index notation Eij = 1 2 I+ u X
T

I+
T

u X u X

I
T

1 u + 2 X 1 2

u X

u X

(3.10)

ui uj uk uk + + Xj Xi Xi Xj

(3.11)

If the elements of the displacement gradient are small in comparison to unity, i.e. ui 1, Xj (3.12)

then the quadratic terms can be neglected and the innitesimal strain tensor can be dened as the symmetric part of the displacement gradient ij = 1 2 ui uj + Xj Xi Eij . (3.13)

Lets dene a stretch vector in the direction of a unit vector n 0 as = F n 0 , (3.14)

and the length of the stretch vector = || is called the stretch ratio or simply the stretch. The square of the stretch ratio is 2 = = n 0 F T F n 0 = n 0 C n 0 , (3.15)

where the tensor C = F T F is called the right Cauchy-Green strain tensor. The atribute right comes from the fact that the deformation gradient operates on the right hand side. The right Cauchy-Green strain tensor is symmetric and positive denite tensor, i.e. C = C T and n C n > 0, n = 0. For values 0 < < 1 a line element is compressed and elongated for values > 1. The deformation gradient can also be decomposed multiplicatively as F = R U = V R, (3.16)

where R is an othogonal tensor (R T R = R RT = I ) describing the rotation of a material element and U and V are symmetric positive denite tensors describing the

20

CHAPTER 3. KINEMATICAL RELATIONS

deformation. The decomposition (3.16) is also called the polar decomposition. The tensor U is called as the right stretch tensor and V the left stretch tensor. The square of the stretch can be expressed as 2 == = n 0 U T RT R U n 0 = n 0 U T U n 0 = n 0 U 2 n 0 . Other strain measures can be dened as E (m) = 1 (U m I ). m (3.18) (3.17)

For m = 2 we obtain the Green-Lagrange strain tensor which have already been discussed. With m = 0 we obtain the Hencky or logarithmic strain tensor E (0) = ln U . (3.19)

which is called the Biot strain tensor. If the defrmation is rotation free, i.e. R = I , the Biot strain tensor coincides the small strain tensor . It is much used in dimensionally reduced continuum models, like beams, plates and shells.

The logarithmic strain1 has a special position in non-linear continum mechanics, especially in formulating constitutive equations, since it can be additively decomposed into volymetric and isochoric parts similarly as the small strain tensor . For m = 1 we obtain E (1) = U I , (3.20)

3.4 Geometric intepretation of the strain components


Lets investigate the extension = 1 of a line element, for instance in a direction n 0 = (1, 0, 0)T , thus (1) = C11 ,
1 E11 = 2 (C11 1)

C11 = 1 + 2E11 1 + 2E11

1 + 2E11 1 (3.21)

Secondly, lets compute the angle change of two unit vectors N 1 and N 2 . In the deformed conguration they are n 1 = F N 1 and n 1 = F N 2 and the angle between them can be determined from cos 12 =
1

n 1 n2 N 1 C N 2 = . |n 1 ||n 2 | N 1 C N 1 N 2 C N 2

(3.22)

The logarithmic strain is sometimes called also as the true strain. Such naming is not used in this text, all properly dened strain measures are applicable, since the denition of strain is a geometrical construction. Naturally, the choice of strain measure dictates the choise of the stress. However, deeper discussion on this topic is beyong the present lecture notes.

3.5. DEFINITION OF INFINITESIMAL STRAIN


If we choose the directions N 1 and N 2 as 1 0 , N1 = 0 cos 12 =

21

then

0 N2 = 1 , 0 2E12 . (1 + 2E11 )(1 + 2E22 )

(3.23)

C12 C12 = = (1) (2) C11 C22

(3.24)

Using the trigonometric identity sin( 1 12 ) = cos 12 2 and if E11 , E22 1 then
1 2

(3.25) (3.26)

Thus, the component E12 is approximately one half of the angle change of the two direction vectors.

12 2E12 .

3.5

Denition of innitesimal strain

Lets investigate the motion of two neghbouring points, which are denoted as P Q in the undeformed conguration. After deformation these points occupy the positions marked by p and q . Displacement of the point Q relative to P is dened as du = u Q u P . Length of the vector P Q is denoted as dS , thus dui ui dxj = , dS xj dS (3.28) (3.27)

where the Jacobian matrix J = u / x can be divided additively into a symmetric and an antisymmetric part as J = + , (3.29) where the symmetric part is the innitesimal strain tensor 11 12 13 xx xy xz x 1 = 21 22 23 = yx yy yz = 2 yx 1 zx zy zz 31 32 33 2 zx ij = 1 (ui,j + uj,i ). 2
1 2 xy 1 2 xz 1 2 yz

y 1 2 zy

and the antisymmetric part is the innitesimal rotation tensor. Written in displacement components the strain tensor has the expression (3.31)

(3.30)

It should be emphasised that the rotation matrix near the point P describes the rigid body rotation ony if the elements ij are small.

22 3.5.1 Principal strains

CHAPTER 3. KINEMATICAL RELATIONS

The principal strains are obtained from the linear eigenvalue problem (ij ij )nj = 0, (3.32)

where the vector ni denes the normal direction of the principal strain plane. Thus, the characteristic polynomial has the form 3 + I1 2 + I2 + I3 = 0, where the strain invariants Ii , i = 1, . . . , 3 are I1 = tr = ii = 11 + 22 + 33 , 2 ), I2 = 1 [tr(2 ) (tr)2 ] = 1 ( I 1 2 2 ij ji I3 = det(ij ). (3.34) (3.35) (3.36) (3.33)

The invariants I1 , . . . , I3 expressed in terms of the principal strains 1 , . . . , 3 have the forms I1 = 1 + 2 + 3 , I2 = 1 2 2 3 3 1 , I3 = 1 2 3 . 3.5.2 Deviatoric strain As in the case of the stress tensor, the innitesimal strain tensor can be additively decomposed into a deviatoric part and an isotropic part as ij = eij + 1 , 3 kk ij (3.41) (3.38) (3.39) (3.40)

If the coorinate axes are chosen to coincide with the axes of principal strains, the strain tensor will be symmetric 1 0 0 = [ij ] = 0 2 0 . (3.37) 0 0 3

where the deviatoric strain tensor is denoted as e . In the literature also the notation is used. By denition the deviatoric strain tensor is traceless tre = 0. The eigenvalues of the deviatoric strain ei can be solved from the equation |eij eij | = 0, (3.43) (3.42)

3.5. DEFINITION OF INFINITESIMAL STRAIN


and the characteristitic equation is e3 + J1 e2 + J2 e + J3 = 0, where the invariants J1 , . . . , J3 have expressions J1 = tre = eii = ex + ey + ez = 0, 1 1 [tr(e 2 ) (tre )2 ] = 1 tr(e 2 ) = 2 eij eji J2 = 2 2 2 2 =1 [(1 2 ) + (2 3 ) + (3 1 )2 ], 6 J3 = det e = tr(e 3 ) = e1 e2 e3 .

23

(3.44)

(3.45) (3.46) (3.47) (3.48)

For small strains the rst invariant I1 = x + y + z v describes the relative volume change. The octahedral strains are dened similarly as for the stress
1 oct = 1 I =3 v , 3 1 2 8 oct =3 J2 .

(3.49) (3.50)

For the rst sight the equation (3.50) might look strange as copared to the expression of the octahedral stress, but we have to remember that xy = 2xy , etc.

24

CHAPTER 3. KINEMATICAL RELATIONS

Chapter 4 Constitutive models


4.1 Introduction

Constitutive equations describe the response of a material to applied loads. In continuum mechanics, distinction between uids and solids can be characterized in this stage. It is important to notice that the balance equations and the kinematical relations described in the previous sections are equally valid both fr uids and solids. In this lecture notes only phenomenological models will be introduced, which roughly means that mathematical expressions are tted to experimental data. Phenomenological models are not capable to relate the actual physical mechanisms of deformation to the underlying mcroscopic physical structure of the material. The constitutive equations should obey the thermodynamic principles, (i) the conservation of energy and (ii) the dissipation inequality, i.e. the nonnegativity of the entropy rate. Excellent texts for materials modelling are [2, 4].

4.2

Elastic constitutive models

Elasticity means that the response of a material is independent of the load history. The most general form of elasticity is called as Cauchy-elasticity and it essentially means that there exists one-to-one relation between stress and strain ij = fij (kl ), or ij = gij (kl ). (4.1)

The tensor valued tensor functions fij and gij are called as response functions. For nonlinear Cauchy-elastic models, the loading-unloading process may yield hysteresis, which is incompatible with the notion of elasticity, where the response should be reversible. However, in this lecture notes the Couchy-elasticity is not treated. Another form of elasticity, where the constitutive equations are expressed in rate-form ij = fij (kl , mn ) 25 (4.2)

26

CHAPTER 4. CONSTITUTIVE MODELS

is called hypo-elastic. If the material is incrementally linear, it can be written in the form ij = Cijkl(mn ) kl . (4.3)

The most rigorous form of elasticity is called as hyper-elasticity, and the constitutive equations of a hyper-elastic model can be derived from a potential, i.e. the strain energy function W = W (ij ) as W ij = . (4.4) ij 4.2.1 Isotropic elasticity A material which behaviour is independent of the direction in which the response is measured is called isotropic. Therefore also the strain energy density should be an isotropic tensor valued scalar function W = W () = W ( ) = W ( T ) = W (I1 , I2 , I3 ), (4.5)

where I1 , I2 and I3 are the principal invariants of the strain tensor and is the transformation tensor from the x -coordinate system to the x -system, i.e. x = x . Alternatively the strain energy density function W can be written as W = W ( I1 , J 2 , J 3 ) , or 2 , I 3 ), W = W ( I1 , I (4.6)

2 , I 3 the generic where J2 and J3 are the invariants of the deviatoric strain tensor and I invariants dened as 3 = 1 tr(3 ). 2 = 1 tr(2 ), I (4.7) I 2 2 Equations (4.5) and (4.6) are spacial forms of representation theorems, for which an alternative form can be written as: the most general form of an isotropic elastic material model can be written as = a0 I + a1 + a2 2 , (4.8) where the coefcients a0 , a1 and a2 can be non-linear functions of the strain invariants. Alternatively (4.8) can be formulated as = b0 I + b1 + b2 2 , (4.9)

where b0 , b1 and b2 can be non-linear functions of stress invariants. From (4.8) and (4.9) it can be easily seen that the principal directions of the strainand stress tensors coincide.

4.2. ELASTIC CONSTITUTIVE MODELS


4.2.2 Transversally isotropic elasticity

27

A material is called transversally isotropic if the behaviour of it is isotropic in a plane and different in the direction of the normal of that isotropy plane. The strain energy density function can now be written as W = W (, M ) = W ( T , M T ), (4.10)

where M = mm T is called the structural tensor and the unit vector m denes the normal of the isotropy plane. The representation theorem of a transversally isotropic solid says that the strain energy density function can depend on ve invariants W = W ( I1 , I2 , I3 , I4 , I5 ) , where the invarinats Ii are I1 = tr , I2 = 1 tr(2 ), 2 I3 = 1 tr(3 ), 3 I4 = tr(M ), I5 = tr(2 M ). (4.12) (4.11)

The invariants I4 and I5 can also be written as I4 = tr(M ) = m T m , The constitutive equation is thus = W W W 2 W W W = I+ + + M+ (M + M ). I1 I2 I3 I4 I5 W I1 W I2 W I3 W I4 W I1 (4.14) I5 = tr(2 M ) = m T 2 m . (4.13)

If we restrict to linear elasticity, the cefcients W/Ii has to satisfy = 1 I1 + I4 , = 2 , = 0, = 3 I 1 + 4 I 4 , = 5 , (4.15) (4.16) (4.17) (4.18) (4.19)

since all the terms in (4.14) have to be linear in . Due to the identity 2W 2W = , Ii Ij Ij Ii (4.20)

28 we have now

CHAPTER 4. CONSTITUTIVE MODELS

I4

W I1

I1

W I4

thus

= 3 .

(4.21)

Transversally isotropic linear solid has thus ve material coefcients, and the constitutive equation can be written as = (1 tr + 3 tr(M ))I + 2 +(3 tr + 4 tr(M ))M + 5 (M + M ). (4.22) 4.2.3 Orthotropic material A material is called orthotropic if it has three perpendicular symmetry planes. Lets denote the unit vectors normal to the symetry planes as m 1 , m 2 and m 3 . Due to the orthogonality m i m j = ij . The structural tensors associated with these direction vectors are M i = mm T , and they satisfy M1 + M2 + M3 = I, (4.23) due to the orthogonality. Thus, only two structural tensors are necessary to describe the behaviour of an orthotropic material W = W (, M 1 , M2 ) = W ( T , M 1 T , M 2 T ). (4.24)

The representation theorem of an orthotropic solid says that the strain energy density function can depend on seven invariants W = W (tr ,
1 2

tr(2 ),

1 3

tr(3 ), tr(M 1 ), tr(M 2 ), tr(2 M 1 ), tr(2 M 2 )). (4.25)

It can be written in a form, where all the structural tensors M i are symmetricaly present. Notice that M 1 + M 2 M 3 = (M 1 + M 2 + M 3 ) = , M 1 + M 2 + M 3 = (M 1 + M 2 + M 3 ) = , thus summing by parts gives
1 1 1 = 2 (M 1 + M 1 ) + 2 ( M 2 + M 2 ) + 2 (M 3 + M 3 ),

(4.26) (4.27)

(4.28) (4.29)

and tr = tr(M 1 ) + tr(M 2 ) + tr(M 3 ). In a similar way it can be deduced tr (2 ) = tr(2 M 1 ) + tr(2 M 2 ) + tr(2 M 3 ). (4.30)

In other words, the invariants tr , tr(M 1 ) and tr(M 2 ) can be replaced by the invariants I1 = tr(M 1 ), I2 = tr(M 2 ) and I3 = tr(M 3 ). In a similar way the invariants

4.2. ELASTIC CONSTITUTIVE MODELS

29

tr(2 ), tr(2 M 1 ) and tr(2 M 2 ) can be replaced by the invariants I4 = tr(2 M 1 ), I5 = tr(3 ), the tr(2 M 2 ) and I6 = tr(2 M 3 ). If we now denote the cubic invariant I7 = 1 3 strain energy density function for an orthotropic material can be written as a function of these seven invariants as W = W ( I1 , . . . , I7 ) , (4.31) and the constitutive equation has the form W = =
7

i=1

W Ii Ii

W W W W = M1 + M2 + M3 + ( M 1 + M 1 ) I1 I2 I3 I4 W W 2 W ( M 2 + M 2 ) + ( M 3 + M 3 ) + . + I5 I6 I7

(4.32)

If we now restrict to a linear model, the coefcients W/Ii has to satisfy the following conditions W I1 W I2 W I3 W I4 W I5 W I6 W I7 I2 I3 I3 W I1 W I1 W I2 = 1 I1 + 1 I2 + 2 I3 , = 2 I1 + 3 I2 + 3 I3 , = 4 I 1 + 5 I 2 + 6 I 3 , = 7 , = 8 , = 9 , = 0. (4.33) (4.34) (4.35) (4.36) (4.37) (4.38) (4.39)

Due to the identity of the second derivatives (4.20), we have I1 = I1 = I2 = W I2 W I3 W I3 1 = 2 , 2 = 4 , 3 = 5 . (4.40) (4.41) (4.42)

30

CHAPTER 4. CONSTITUTIVE MODELS

Bibliography
[1] G.A. Holzapfel. Nonlinear Solid Mechanics, A Continuum Approach for Engineers. John Wiley & Sons, 2000. [2] J. Lemaitre and J.-L. Chaboche. Mechanics of Solid Materials. Cambridge University Press, 1990. [3] L.E. Malvern. Introduction to the Mechanics of a Continuous Medium. Prentice Hall, Englewood Cliffs, New Jersey, 1969. [4] N.S. Ottosen and M. Ristinmaa. The Mechanics of Constitutive Modeling. Elsevier, 2005.

31

Vous aimerez peut-être aussi