Sunteți pe pagina 1din 27

Chapter 2

Solid Mechanics Fundamentals

Abstract In this chapter the mechanical fundamentals behind the numerical


applications presented in this work are developed. Firstly it is present a brief
exposition of the used continuum formulation, where the solid kinematics and
constitutive equations are shown. Following it is presented the used weak form and
the consequent generated discrete equation system. Next, the dynamic analysis
equations are presented and transient analysis basic concepts are introduced.

2.1 Continuum Formulation


The continuum mechanics is the foundation of the nonlinear numerical analysis.
Solids and structures subjected to loads or forces become stressed. The stresses
lead to strains, which can be interpreted as deformations or relative displacements.
Solid Mechanics and Structural Mechanics deals, for a given solid and boundary
condition (external forces and displacements constrains), with the relationship
between stress and strain and the relationship between strain and displacements
[13]. Solids can show different behaviours, depending on the solid material
stress-strain curve. In this work only linear elastic materials are considered. In
elastic materials the deformation in the solid caused by loading disappears fully
with the unloading, in contrast, plastic materials show a residual deformation
(which cannot be naturally recovered) that remains after the total unload process.
The material properties on the solid can also be anisotropic, i.e., the material
property varies with the direction [4]. On an anisotropic material the deformation
caused by a load applied in a given direction causes a different deformation if the
same load is applied in a distinct direction. Composite laminates are generally
constituted by layers of anisotropic material. There are many material constants to
be considered and defined in order to fully describe an anisotropic material, which
is the reason why so often the engineering problems reduce the analysis to an
isotropic material analysis. Isotropic materials are a special case of anisotropic
materials, where only two independent material properties need to be known, the
J. Belinha, Meshless Methods in Biomechanics, Lecture Notes in Computational
Vision and Biomechanics 16, DOI: 10.1007/978-3-319-06400-0_2,
 Springer International Publishing Switzerland 2014

15

16

2 Solid Mechanics Fundamentals

Young modulus and the Poisson ratio. In this chapter the rigid solid motion and
deformation are described, with emphasis on rotation, which plays an important
role in nonlinear continuum mechanics. Also the concepts of strain and stress in
nonlinear mechanics are introduced. The equilibrium and the constitutive equations are presented afterwards.

2.1.1 Kinematics
The general motion of a deformable body is represented in Fig. 2.1. The body, in
the initial position t 0, is considered to be an assemblage of material particles,
labelled by the coordinates X, with respect to the Cartesian basis e. The current
position of a particle is defined at time t by the coordinates x.
The motion can be mathematically described by a mapping function / between
initial and current particle positions,
x /X; t

2:1

It is considered the material description, the Lagragian description, since the


variation of the solid deformation is described with respect to the initial coordinates X, at time t.

2.1.1.1 Deformation Gradient


The deformation gradient F, is a key quantity in finite deformation analysis, since
it is involved in all equations relating quantities before deformation (initial configuration) with the correspondent quantities after the finite deformation (current
configuration). The deformation gradient tensor F can be defined as,
F

o/
r/
oX

2:2

Alternatively to Eq. (2.1) the motion can be expressed by,


x xX; t

2:3

which permits the deformation gradient to be written as,


F

ox
oX

2:4

2.1 Continuum Formulation

17

Fig. 2.1 General motion of a


deformable body

As so, for a three-dimensional deformation problem the deformation gradient


tensor of an initial material position X f X Y Z g in respect to a current
material position x f x y z g can be presented as,
2
F

ox
6 oX
oy
4 oX
oz
oX

ox
oY
oy
oY
oz
oY

ox
oZ
oy
oZ
oz
oZ

3
7
5

2:5

The determinant of the F is denoted by J and is called the Jacobian


determinant.
J detF

2:6

The Jacobian determinant can be used to relate the integral of a given functional
f in the current and in the initial configuration by,
Z
Z
f x; t dX
f /X; t; t J dX0 :
2:7
X

X0

2.1.1.2 Strain
Consider the change of the scalar product of the two elemental vectors from
dX 1 Q1  P and dX 2 Q2  P, initial configuration, to dx1 q1  p and
dx2 q2  p, current configuration, as a general measure of deformation. Where
Q1 and Q2 are two material particles in the neighbourhood of a material particle P
for the initial configuration and q1 and q2 and p the same respective material

18

2 Solid Mechanics Fundamentals

particles in the current configuration. Equation (2.4) permits the following relations dx1 F dX 1 and dx2 F dX2 , and the spatial scalar product dx1  dx2 can
be found in terms of the material vectors dX 1 and dX2 as,
dx1  dx2 dX1  FT FdX 2

2:8

The right Cauchy-Green deformation tensor is defined by


C FT F

2:9

Which operates directly on the material vectors dX 1 and dX 2 . Alternatively the


initial scalar product dX 1  dX 2 can be obtained in terms of spatial vectors dx1 and
dx1 using the left Cauchy-Green deformation tensor b,
b F FT

2:10

The change in scalar product can now be found in terms of the material vectors
dX 1 and dX2 and the Lagrange or Green strain tensor E can be defined as,
1
dx1  dx2  dX 1  dX 2 dX1  E dX2
2

2:11

where the strain tensor E is expressed as


1
E C  I :
2

2:12

2.1.1.3 Polar Decomposition


The tensor F can be expressed as the product of the orthogonal rotation tensor R
by the symmetric stretch tensor U,
F RU

2:13

where
RT R I

and U UT

2:14

Such decomposition is called, polar decomposition, and Eq. (2.9) can be


expressed as,
C FT F U T RT RU U T IU UU

2:15

In order to actually obtain U from Eq. (2.15) it is first necessary to evaluate the
principal directions of C, represented by the eigenvectors set W i and the

2.1 Continuum Formulation

19

correspondent eigenvalues ki , with i f1; 2; 3g for the three-dimensional case. In


this manner C can be defined as,
C

3
X

k2i W i  W i

2:16

i1

since the eigenvectors W i are in fact orthogonal unit vectors, because C CT . As


so, with Eqs. (2.15) and (2.16) it is possible to write the material stretch tensor U
as,
U

3
X

ki W i  W i

2:17

i1

Once the stretch tensor U is known, the rotation tensor R can be obtained
without difficulty from Eq. (2.13).
R F U 1 :

2:18

2.1.1.4 Stress
In a large deformation analysis a body can experience a large rotation and/or a
large strain. The defined stress terms together with the obtained strain terms
enables to express the virtual work as an integral over the known body volume,
expressing in this manner the change in the body configuration. Both strain tensor
and stress tensor are referred to the same deformed state. The Cauchy stress tensor,
here defined as K, is a symmetric tensor and it represents the stresses of the current
configuration. For the three-dimensional case it can be defined as,
2

rxx
K 4 ryx
rzx

rxy
ryy
rzy

3
rxz
ryz 5
rzz

2:19

In this work it is used the Voigt notation, since the development of fourth order
tensors is less practical. In Voigt notation the tensors are expressed in column
vectors, so the stress tensor K is reduced to the stress vector r,
r f rxx

ryy

rzz

rxy

ryz

rzx gT

2:20

and the strain tensor E to the strain vector e,


e

exx

eyy

ezz

exy

eyz

ezx

T

The use of vectors is more practical in the programming process.

2:21

20

2 Solid Mechanics Fundamentals

2.1.1.5 Principal Stress


Another way of describing the Cauchy stress tensor, which completely defines the
stress state in an interest point, is through,
2

3 2
rxx
t^e1
K 4 t^e2 5 4 ryx
rzx
t^e3

rxy
ryy
rzy

3
rxz
ryz 5
rzz

2:22

where ^e1 , ^e2 and ^e3 are the versors of the coordinate system and t^ei is the stress
vector on a plane normal to ^e1 passing through the interest point, Fig. 2.2(a).
Following Cauchys stress theorem, if the stress vectors of three orthogonal planes,
with a common point, are known, then the stress vector on any other plane passing
through that point can be found through the coordinate transformation equations
[5]. Thus, the stress vector tn in a point belonging to an inclined plane,
Fig. 2.2(b), can be defined by,
2

tn n  rij n1

n2

rxx
n3   4 ryx
rzx

rxy
ryy
rzy

3
rxz
ryz 5
rzz

2:23

where n is the inclined plane normal vector. The relation in Eq. (2.23) leads to the
transformation rule of the stress tensor. The initial stress tensor rij , defined in the
xi coordinate system, can be transformed in a new stress tensor r0ij , defined in
another x0i coordinate system by the relation,
K0 A K AT

2:24

being A the rotation matrix. Developing Eq. (2.24),


2

rxx
6 0
4 ryx
0
rzx

rxy
0
ryy
0
rzy

0 3
2
rxz
a11
0 7
ryz 5 4 a21
0
a31
rzz

a12
a22
a32

32
rxx
a13
5
4
ryx
a23
rzx
a33

rxy
ryy
rzy

32
rxz
a11
ryz 54 a12
rzz
a13

a21
a22
a23

3
a31
a32 5
a33
2:25

The aij coefficients can be understood as the projection of the x0i coordinate
system versors in the xi coordinate system versors. Therefore, the angle between
the versors of each coordinate system can be defined as,
 
cij cos1 aij

2:26

Through Eq. (2.26) and Fig. 2.3 it is possible to comprehend better the physical
meaning of the aij coefficients and the respective angles.

2.1 Continuum Formulation

21

Fig. 2.2 a Three-dimensional stress components. b Stress vector acting on a plane with normal
vector n

Fig. 2.3 Stress tensor transformation and respective angles

Let P be an interest point of a considered stressed body. There are at least three
planes, orthogonal with each other, crossing P where the corresponding stress
vector is normal to the plane. These planes are called principal planes and the

22

2 Solid Mechanics Fundamentals

normal vectors of each plane are called principal directions. The stress vectors are
parallel to the plane normal vectors and are called principal stresses.
The stress tensor is a physical quantity, independent of the coordinate system
chosen to represent it. Therefore, there are certain invariants associated with it
which are also independent of the coordinate system. Being a second order tensor,
the stress tensor has associated three independent invariant quantities. One set of
such invariants are the principal stresses of the stress tensor, which are just the
eigenvalues of the stress tensor. Their direction vectors are the principal directions
or eigenvectors. A stress vector parallel to the normal vector n is given by,
tn kn rn n

2:27

being k a constant of proportionality, and in this particular case the magnitude of


n
rn , the principal stress in n direction. Knowing ti rij nj and ni dij nj , where
dij is the Kronecker delta, the following development can be performed,
n

ti



kni ) rij nj kdij nj ) rij  kdij nj 0

2:28

which is a homogeneous system, three linear equations for three nj unknowns. To


obtain the nj non-zero solution, the matrix determinant must be equal to zero,



  rxx  k
rij  kdij   ryx

 rzx


rxz 
ryz  0
rzz  k 

rxy
ryy  k
rzy

2:29

which leads to the following cubic equation,




rij  kdij  k3 I1 k2  I2 k I3 0

2:30

being I1 , I2 and I3 the stress invariants,


I1 rkk
I2

1
r r
2 ii jj

 rij rji

 
I3 det rij

2:31


2:32
2:33

The three roots k1 r1 ; k2 r2 and k3 r3 of Eq. (2.30) are the eigenvalues or principal stresses, which are unique. Therefore the stress invariants have
always the same value regardless of the orientation of the chosen coordinate
system. For each eigenvalue k, exists a non-trivial solution n on Eq. (2.28). These
nj solutions, called eigenvectors, are the principal directions, which defines the
plane where the respective stress acts. Applying Eq. (2.25),

2.1 Continuum Formulation

r1
4 0
0

0
r2
0

3 2
0
n11
0 5 4 n21
r3
n31

23

32
rxx
n13
n23 54 ryx
rzx
n33

n12
n22
n32

32
rxz
n11
ryz 54 n12
rzz
n13

rxy
ryy
rzy

n21
n22
n23

3
n31
n32 5
n33
2:34

The principal stresses and principal directions characterize the stress in P and
are independent of the orientation of the coordinate system.

2.1.2 Constitutive Equations


The following relation between the stress rate and the strain rate is assumed,
dr c de

2:35

The material constitutive matrix is defined by c and if material nonlinear


relations exists between r and e, then c cep . With Eq. (2.35) the following
relation can be established,
de c1 dr

2:36

being s c1 and defined for the three-dimensional case as,


2

1
E
6 xxtxy
6  Exx
6 t
6  xz
6 Exx

s6
6 0
6
6 0
4
0

 Eyxyy
1
Eyy
t
 Eyzyy

0
0
0

 Etzxzz
t
 Ezyzz
1
Ezz

0
0
0

0
0
0
1
Gxy

0
0

0
0
0
0
1
Gyz

0
0
0
0
0

1
Gzx

3
7
7
7
7
7
7
7
7
7
5

2:37

The material constitutive matrix c is obtained by inverting the material compliance matrix s, which is here defined for an three-dimensional anisotropic
material. The elements on matrix s are obtained experimentally. Eii is the Young
modulus in direction i, tij is the Poisson ratio which characterizes the deformation
rate in direction j when a force is applied in direction i, Gij is the shear modulus
which characterizes the variation angle between directions i and j. Due to symmetry the following relation can be established,
Ei tji Ej tij

2:38

For the two-dimensional case the plane stress and plane strain [5] deformation
theory assumptions can be presumed. Considering the plane stress assumptions,

24

2 Solid Mechanics Fundamentals

Fig. 2.4 Projection of vector


n in the coordinate axis and in
the oxy plane

rzx rzy rzz 0, the material compliance matrix s is obtained directly from
the three-dimensional compliance matrix s,
2
s

1
Exx
6  txy
4 Exx

 Eyxyy
1
Eyy

3
0
0 7
5

2:39

1
Gxy

For the plane strain deformation theory it is considered ezx ezy ezz 0 and
the material compliance matrix s is defined as,
2
s

tzx txz
1
Exx  Exx
6  txy  tzy txz
4 Exx
Exx

t t

 Eyxyy  zxEyyyz
tzy tyz
1
Eyy  Eyy
0

3
0
0 7
5

2:40

1
Gxy

In the case of an anisotropic material, it is possible to rotate the material


constitutive matrix c and orientate the material directions with a vector. Consider a
known vector n in the Euclidean space R3 , Fig. 2.4, and the respective projections
on the coordinate axis and in the oxy plane. As it is known,
h cos

1

n n
 oxy ox
noxy   knox k

!
and

x cos

1

n n
 oxy
noxy   knk

!
2:41

With the obtained angle information it is now possible to rotate the material
matrix using the rotational transformation matrix and therefore align the material
ox axis with the known vector n. The rotational transformation matrix that permits
an anticlockwise rotation along the ox axis of a known angle b can be defined as,

2.1 Continuum Formulation

T ox

1
60
6
60
6
60
6
40
0

25

0
cos2 b
sin2 b
0
sin b  cos b
0

0
0
sin2 b
0
cos2 b
0
0
cos b
 sin b  cos b
0
0
 sin b

0
 sin 2b
sin 2b
0
cos2 b  sin2 b
0

3
0
0 7
7
0 7
7
sin b 7
7
0 5
cos b
2:42

Along the oy axis,


2

T oy

cos2 b
6
0
6
2
6
sin
b
6
6
0
6
4
0
 sin b  cos b

0
1
0
0
0
0

sin2 b
0
0
0
0
0
cos2 b
0
0
0
cos b  sin b
0
sin b cos b
sin b  cos b
0
0

3
sin 2b
7
0
7
 sin 2b 7
7
7
0
7
5
0
cos2 b  sin2 b
2:43

Along the oz axis,


2

cos2 b
6 sin2 b
6
6
0
T oz 6
6 sin b  cos b
6
4
0
0

sin2 b
cos2 b
0
 sin b  cos b
0
0

0
0
1
0
0
0

 sin 2b
sin 2b
0
cos2 b  sin2 b
0
0

0
0
0
0
cos b
sin b

3
0
0 7
7
0 7
7
0 7
7
 sin b 5
cos b
2:44

The material matrix after rotation can be defined as,


h
 i
T
ccurrent T oz Th T oy x cinitial T oy x T oz h :

2:45

2.2 Weak Form


The strong form system equations are the partial differential system equations
governing the studied physic phenomenon. In contrast, the weak form requires a
weaker consistency on the adopted approximation (or interpolation) functions. The
ideal would be obtaining the exact solution from strong form system equations,
however this is usually an extremely difficult task in complex practical engineering
problems. Formulations based on weak forms are able to produce stable algebraic
system equations and to give a discretized system of equations which leads to more

26

2 Solid Mechanics Fundamentals

accurate results. These are the reasons why so many prefer the weak form to obtain
the approximated solution.
In this work the discrete equation system is obtained using the Galerkin weak
form, which is a variational method [6]. For meshless methods used in this book
the discrete system of equations is obtained similarly with the FEM, with some
differences inherent to the meshless approach. The discrete equations for the static
and the dynamic approach are developed and shown for the basic three-dimensional deformation theory.

2.2.1 Weak Form of Galerkin


Consider the solid with a domain X bounded by C, Fig. 2.5. The continuous solid
surface on which the external forces t are applied is denoted as Ct (natural
boundary) and the surface where the displacements are constrained is denoted as
Cu (essential boundary).
The Galerkin weak form is a variational principle based on the energy principle.
Of all possible displacement configurations satisfying the compatibility conditions,
the essential boundary conditions (kinematical and displacement) and the initial
and final time conditions, the real solution correspondent configuration is the one
which minimizes the Lagrangian functional L,
L T  U Wf

2:46

being T the kinetic energy, U is the strain energy and Wf is the work produced by
the external forces. The kinetic energy is defined by,
T

1
2

qu_ T u_ dX

2:47

where the solid volume is defined by X and u_ is the displacement first derivative
with respect to time, i.e., the velocity. The solid mass density is defined by q. The
strain energy, for elastic materials, is defined as,
Z
1
eT r dX
2:48
U
2

being e the strain vector and r the stress vector. The work produced by the external
forces can be expressed as,
Wf

Z
X

uT b dX

Z
Ct

uT t dC

2:49

2.2 Weak Form

27

Fig. 2.5 Continuous solid


subject to volume forces and
external forces

where u represents the displacement, b the body forces and Ct the traction
boundary where the external forces t are applied. By substitution the Lagrangian
functional L can be rewritten as,
L

1
2

qu_ u_

1
dX 
2

e r dX

u b dX

uT t dC

2:50

Ct

and then minimized,


Zt2
d

2
61
42

t1

qu_ T u_ dX 

1
2

eT r dX

uT b dX

3
7
uT t dC5 dt 0

2:51

Ct

Moving the variation operator d inside the integrals,


Zt2
t1

2
61
42



1
d qu_ T u_ dX 


d eT r dX

duT b dX

3
7
duT t dC5 dt 0

Ct

2:52
Since all operations are linear, changing the order of operation does not affect
the result. In the first term of Eq. (2.52) the time integral can be moved inside the
spatial integral,
2
3

Zt2
Z
Z Zt2




1
1
4 d qu_ T u_ dt5 dX
d qu_ T u_ dX dt
2:53
2 X
2
t1

t1

Using the chain rule of variation and then the scalar property, the integral can
be rewritten as,

28

2 Solid Mechanics Fundamentals

Zt2


d qu_ u_ dt q
T

t1

Zt2

du_ u_ u_ du_ dt 2q

t1

Zt2


du_ T u_ dt

2:54

t1

And knowing that u_ T u_ is a scalar and u_ = ou=ot,


Zt2


du_ T u_ dt

t1


Zt2
oduT ou
dt
ot ot

2:55

t1

Then integrating by parts, with respect to time,


Zt2
t1




t2
Zt2
2

oduT ou
o
u
ou

duT 2 dt duT
dt
ot
ot 
ot ot
t1

2:56

t1

Notice that u satisfies, by imposition, the conditions at the initial time, t1 , and
final time, t2 , leading to a null du at t1 and t2 . Therefore the last term in Eq. (2.56)
vanishes. Considering the last development and switching the integration order
again, Eq. (2.53) becomes,
Zt2
Z
t1



1
d qu_ T u_ dX
2 X


Zt2
Z
 T 
dt
q
du u
dt 
t1

2:57


being
u = o2 u ot2 the acceleration. The second term on Eq. (2.52) can also be
developed. The integrand function in the second integral term can be written as
follows,
 
d eT r deT r + eT dr

2:58

as the two terms in Eq. (2.58) are in fact scalars, the transpose does not affect the
result, as so,

T
eT dr eT dr drT e

2:59

Using the constitutive equation r ce and the symmetric property of the


material matrix, cT c, it is possible to write,
drT e dceT e deT cT e deT ce deT r

2:60

2.2 Weak Form

29

Therefore Eq. (2.58) becomes,


 
d eT r 2deT r

2:61

simplifying the second term in Eq. (2.52),


Zt2
t1

2
41
2

2
3
Zt2 Z
d eT r dX5dt 4 deT rdX5dt

t1

2:62

Equation (2.52) now becomes,


Zt2

2
6
4q

t1


dX
duT u

Z
X

deT r dX

Z
X

duT b dX

3
7
duT t dC 5dt 0 2:63

Ct

To satisfy Eq. (2.63) for all possible choices of the integrand of the time
integration has to be null, leading to the following expression,
Z
Z
Z
Z
 T 
dX 
q
du u
deT r dX duT b dX
duT t dC 0
2:64
X

Ct

This last equation is known as the Galerkin weak form, which can also be
viewed as the principle of virtual work. The principle of virtual work states that if
a solid body is in equilibrium, the virtual work produced by the body inner stresses
and the body applied external forces should vanish when the body experiments a
virtual displacement. Considering the stress-strain relation, r c e, and the straindisplacement relation, e L u, Eq. (2.64) can be rearranged in the following
expression,
Z
Z
Z
Z


T
T
T
dX 0 2:65
dL u cL udX  du b dX 
du t dC q duT u
X

Ct

which is the generic Galerkin weak form written in terms of displacement, very
useful in solid mechanical problems. In static problems the fourth term of
Eq. (2.65) disappears.

2.3 Discrete System of Equations


The discrete equations for meshless methods are obtained from the principle of
virtual work by using the meshless shape functions as trial and test functions. The
domain X is discretized in a nodal distribution, and each node possesses an

30

2 Solid Mechanics Fundamentals

influence-domain, which imposes the nodal connectivity between the neighbouring nodes. The meshless trial function uxI is given by,
uxI

n
X

ui xI ui

2:66

i1

being ui xI the meshless approximation or interpolation function and ui are the


nodal displacements of the n nodes belonging to the influence-domain of interest
node xI . Considering the NNRPIM, it is known that the NNRPIM interpolation
function satisfies the condition,
ui xj dij

2:67

where dij is the Kronecker delta, dij 1 if i j and dij 0 if i 6 j. Following


Eq. (2.66), the test functions (or virtual displacements) are defined as,
duxI

n
X

ui xI dui

2:68

i1

where dui are the nodal values for the test function.

2.3.1 Weak Form of Galerkin


The meshless formulation can be established in terms of a weak form of the
differential equation under consideration, Eq. (2.64). In the solid mechanics context this implies the use of the virtual work equation.
L

Z
X

r de dX 

b  du dX 

t  du dC q


dX 0
duT u

2:69

The virtual deformation de is defined by,


de Bdu

2:70

where B is the deformation matrix. Thus, the virtual work of the first term in
Eq. (2.69), using Eq. (2.70), can be expressed as,
L1

duT B r dX

2:71

The strain vector can be divided in two parts, the linear part and the nonlinear
part,

2.3 Discrete System of Equations

31

e e0 + eNL

2:72



1
1
e |{z}
Lh Ah L A h
2
2
|{z}
e0

2:73

which can also be presented as,

eNL

Matrix L is defined as,


2

eT1

6
6 0
6 13
6
6 0
6
6 13
L6 T
6 e2
6
6
6 0
6 13
4 T
e3

13
eT2

13
eT1

eT3
0

13

13 7

2
0 7
1
7
13 7
6
0
7
eT3 7 6
0
7 6
76
0 7 6
60
13 7
40
7
eT2 7
0
7
5
T
e1

0
0
0
1
0
0

0
0
0
0
0
1

0
0
0
1
0
0

0
1
0
0
0
0

0
0
0
0
1
0

0
0
0
0
0
1

0
0
0
0
1
0

3
0
07
7
17
7
07
7
05
0

2:74

Being ei the coordinate i director column vector,


I e1

e2

e3 

2:75

The column vector h is defined by,


h Gu

2:76

The geometric matrix G is defined by,


3

2 ou

ou
oz

6
GT 4 0

ou
oy

ou
ox

ou
oy

ou
oz

ou
ox

ou
oy

7
05

ox

2:77

ou
oz

which produces the following column vector h,


3
hx
h 4 hy 5
hz

2 ou 3

on

being

6 ov 7
hn 4 on
5

2:78

ow
on

The current configuration displacement is considered in matrix A, which corresponds in Eq. (2.73) to the actualized component.

32

2 Solid Mechanics Fundamentals

hTx

6
6 0
6 1x3
6
6 0
6
6 1x3
A6 T
6 hy
6
6
6 0
6 1x3
4
hTz

1x3
hTy

1x3
hTx

hTz
0

1x3

7 2 ou
0 7
7 6 ox
1x3 7
60
hTz 7
7 6
0
7 6
76
ou
6
0 7 6 oy
1x3 7
7 6
40
hTy 7
7
ou
5
oz
T
hx
1x3

ov
ox

ow
ox

0
0

0
ow
oy

0
0

0
0

0
0

ou
oy

ov
oy

ov
ox
ov
oz

ow
ox
ow
oz

ov
oz

ou
ox
ou
oz

ou
oz

ov
oy

ow
oy

ow
oz

ou
oy
ou
ox

ov
oy
ov
ox

ov
oz

3
0
07
7
ow 7
7
oz 7
07
7
ow 7
oy 5

2:79

ow
ox

The deformation matrix B, dependent of u, can be defined as,


B B0 BNL u

2:80

since it varies with the deformation of the solid. The linear part of the deformation
matrix is represented by B0 and the nonlinear contribution by BNL . For the threedimensional case,
2 ou

6
BT0 4 0

ou
oy

ou
oz

ox

ou
oy
ou
ox

0
ou
oz
ou
oy

ou 3
oz

7
05

2:81

ou
ox

and
BNL A G

2:82

The nonlinear deformation is actualized through matrix A, which contains the


displacement current configuration.

2.3.2 Stiffness Matrix


The tangential stiffness matrix K T can be determined considering the variation of
the virtual work of Eq. (2.71), in order to the generalized displacements du,
2
3
Z
T
dL1 d 4 B rdX5
X

2:83

2.3 Discrete System of Equations

33

which can be developed as,


Z
Z
T
T
dL1 dB rdX B drdX K T du
X

2:84

Using Eqs. (2.35) and (2.70) the following relation is obtained,


dr cBdu

2:85

In the deformation matrix B only the nonlinear part dBNL is dependent of u,


Eq. (2.80), thus dB dBNL and therefore,
Z
Z
T
T
2:86
dL1 dBNL r dX B c B dX
X

Where the stiffness matrix can be presented as,


K T K r K 0 K NL

2:87

Being,
Z

Kr

dBTNL r dX

2:88

BT0 c B0 dX

2:89

K0

Z
X

K NL


BT0 c BNL BTNL c BNL BTNL c B0 dX

2:90

The initial stress matrix or geometric matrix K r is defined as,


K r du

GT dAT r dX

2:91

The variation of matrix A in order to u can be defined as,


2

dhx

6
6 0
dAT 6 31

4
0
31

31

31

dhy

dhy

dhx

dhz

dhy

31

31

dhz

31

31

dhz

7
0 7
31 7
5
dhx

2:92

34

2 Solid Mechanics Fundamentals

As so, the term dAT r can be represented as,


2

rxx

33
6
6
I
dAT r 6 syx 33
4
szx I
33

sxy

sxz

syz

szy I

rzz

33

ryy

33
33

33 7

I 7

dh
33 7
5

Z dh ZG du

2:93

33

and therefore Eq. (2.88) can be presented as,


Kr

GT Z G dX

2:94

Therefore, the initial stress matrix K r takes into consideration the actualized
stress field.

2.3.3 Mass Matrix


The virtual work of the last term in Eq. (2.69) can be expressed and developed as,
2
dL4 d 4q

dX5 M d
duT u
u

2:95

where the mass matrix M can be defined as,


M

HT q H dX

2:96

being H the interpolation function matrix for the interest point i defined as,
H i ui I

2:97

Where ui is the interpolation function for interest node i and I is the identity
matrix defined in Eq. (2.75). The density diagonal matrix can be defined as,
q qI
being q the solid material density.

2:98

2.3 Discrete System of Equations

35

2.3.4 Force Vector


The virtual work of the middle terms in Eq. (2.69) can be expressed and developed
as,
2
dL2 d4

3
b  du dX5 f b

2:99

and
2
dL3 d4

3
t  du dC5 f t

2:100

being the total force vector f defined as,


fb ft f

2:101

Thus, the total force vector f can be developed in a matrix form,


Z
Z
f HT b dX HT t dC:

2:102

2.3.5 Essential Boundary Conditions Imposition


If the shape functions of the meshless method possess the Kronecker delta property, then the boundary conditions can be imposed directly as in the FEM. The
continuum analysis involves two types of boundary conditions, the essential
boundary conditions (displacement related) and the natural boundary conditions
(force related). Neglecting dumping effects and assuming that the matricial form of
the equilibrium equations resulting from virtual work expression, Eq. (2.69), can
be presented as,
Ku M
uf

2:103

Such equation can be rewritten as,

K cc
K dc

K cd
K dd

uc
M cc

ud
M dc

M cd
M dd



uc
f
c
ud
fd

2:104

where uc are the unknown displacements and ud the known, or prescribed, displacements. The vectors f c and f d correspond respectively to the known applied

36

2 Solid Mechanics Fundamentals

Fig. 2.6 Essential boundary condition nonaligned with the global axis

loads (external and body forces) and to the unknown reactions due the imposed
displacement constrains. With the Eq. (2.104) it is assumed that the displacement
components considered are axial aligned with the prescribed displacements. If this
is not the case it is required the identification of all prescribed displacement
orientations and transform locally the discrete equilibrium equations to correspond
to the global axis. Thus,
u Tu

2:105

where u is the vector of nodal point degrees of freedom in the required directions.
The transformation matrix T is defined by Eq. (2.106) and Fig. 2.6, which is a
typical representation of the constrained displacements in 2D and 3D analysis.

T 2D

u
x
uy

vx
vy


and

T 3D

ux
4 uy
uz

vx
vy
vz

3
wx
wy 5
wz

2:106

Using Eqs. (2.105) and (2.106) it is possible to write,


f
KuMu

2:107

M T TM T

2:108

K T TK T

2:109

f T Tf

2:110

where,

Notice that the matrix multiplications in Eqs. (2.108), (2.109) and (2.110)
involve changes only in those columns and rows of M, K and f that are actually

2.3 Discrete System of Equations

37

affected by the prescribed displacement. In practice, the transformation can be


effectively carried out on the local level just prior to adding the local matrices to
the global assembled matrices.

2.3.6 Dynamic Equations


The equilibrium equations governing the linear dynamic response can be represented as in Eq. (2.103). The fundamental mathematical method used to solve
Eq. (2.103) is the separation of variables. In order to change the equilibrium
equations to the modal generalized displacements [7] it is proposed the following
transformation:
ut U xt

2:111

where U is a m  m square matrix containing m spatial vectors independent of the


time variable t, xt is a time dependent vector and m 2N for the 2D case and
m 3N for the 3D case, being N the total number of nodes in the problem domain.
_ U xt
_
t U xt. The comFrom Eq. (2.111) also follows that ut
and u
ponents of u(t) are called generalized displacements. For which the solution can be
presented in the form,
ut / sinx t  t0

2:112

being / the vector of order m, t the time variable, the constant initial time is
defined by t0 and x is the vibration frequency vector. Substituting Eqs. (2.112) into
(2.103) the generalized eigenproblem is obtained, from which / and x must be
determined,
K / x2 M /

2:113

Equation (2.113) yields the m eigensolutions,


2

K /1 x21 M /1
6 K /2 x2 M /2
2
6
4:
K /m x2m M /m

2:114

The vector / i is called the ith mode shape vector and xi is the corresponding
frequency of vibration. Defining a matrix U whose columns are the eigenvectors / i ,
U /1

/2

. . . /m 

2:115

38

2 Solid Mechanics Fundamentals

and a diagonal matrix X which stores the eigenvalues xi ,


2

x21
6 0
6
X6 .
4 ..

0
x22
..
.



..
.

   x2m

0
0
..
.

3
7
7
7
5

2:116

the m solutions can be written as,


KU M U X

2:117

It is required that the space functions satisfy the following stiffness and mass
orthogonality conditions,
UT K U X

2:118

UT M U I

2:119

and

After substituting Eq. (2.111) and its time derivatives into Eq. (2.103) and premultiplying it by UT , the equilibrium equation that corresponds to the modal
generalized displacement is obtained,
xt X xt UT Ft

2:120

The initial conditions on xt are obtained using Eq. (2.111) and considering
the the M-orthonormality of UT at time t 0,

x0 UT Mu0
x_ 0 UT M u_ 0

2:121

Equation (2.120) can be represented as m individual equations of the form,

xi t x2i xi t fi t
fi t / Ti Ft

2:122

with the initial conditions,

/ Ti Mu0
xt0
i
t0
x_ i / Ti M u_ 0

2:123

For the complete response, the solution to all m equations in Eq. (2.122) must
be calculated and then the modal point displacements are obtained by superposition of the response in each mode.

2.3 Discrete System of Equations

39

ut

m
X

/ i xi t

2:124

i1

Therefore the response analysis requires, first, the solution of the eigenvalues
and eigenvectors of the problem, Eq. (2.113), then the solution of the decoupled
equilibrium equations in Eq. (2.122) and, finally, the superposition of the response
in each eigenvector as expressed in Eq. (2.124).

2.3.7 Forced Vibrations


In this book when forced vibrations are imposed only three different timedependent loading conditions are considered, f t f  gt. A time constant
loadload case A,
gA t 1

2:125

A transient loadload case B,




gB t 1 if
gB t 0 if

t  ti
t [ ti

2:126

And a harmonic loadload case C,


gC t sinc t

2:127

The solution of each equation in Eq. (2.123) can be calculated using the
Duhamel integral,
1
xi t
xi

Zt

fi s sinxi t  sds ai sinxi t bi cosxi t

2:128

where ai and bi are determined from the initial conditions: Eq. (2.123) and
fi t / Ti f t. For load case A and load case B the obtained solution is defined as,
xi t

fi t
x_ t0
1  cosxi t i sinxi t xt0
cosxi t
i
2
xi
xi

2:129

For load case C the obtained solution is,


xi t



fi t
c
sin

c
t


sin

x
t

i
xi
x2i  c2

2:130

40

2 Solid Mechanics Fundamentals

References
1. Fung YC (1965) Foundations of solid mechanics. Englwood Cliffs, Prentice-Hall, New Jersey,
USA
2. Malvern LE (1969) Introduction of the Mechanics of a Continuous Medium. Englwood Cliffs,
Prentice-Hall, New Jersey, USA
3. Hodge PG (1970) Continuum mechanics. Mc Graw-Hill, New York
4. Lekhnitskii SG (1968) Anisotropic Plates. Gordon and Breach Science Publishers, New YorkLondon-Paris
5. Timoshenko S, Goodier JN (1970) Theory of Elasticity. 3rd ed. Singapore, McGraw Hill
6. Reddy JN (1986) Applied functional analysis and variational methods in engineering.
McGraw-Hill, Singapore
7. Bathe KJ (1996) Finite element procedures. Prentice-Hall, Englewood Cliffs

http://www.springer.com/978-3-319-06399-7

S-ar putea să vă placă și