Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Section 2.

10

2.10 Convected Coordinates


In this section, the deformation and strain tensors described in 2.2-3 are now described using convected coordinates (see 1.16). Note that all the tensor relations expressed in & symbolic notation already discussed, such as U = C , FN i = i n i , F = lF , are independent of coordinate system, and hold also for convected coordinates.

2.10.1

Convected Coordinates

Introduce the curvilinear coordinates i . The material coordinates can then be written as
X = X( 1 , 2 , 3 ) so X = X i E i and (2.10.1)

dX = dX i E i = d i G i ,

(2.10.2)

where G i are the covariant base vectors in the reference configuration, with corresponding contravariant base vectors G i , Fig. 2.10.1, with

G i G j = ij
reference configuration current configuration

(2.10.3)

G2
X 2, x2

G1

g2

g1

E2 , e2
X ,x
3 3

E 1 , e1

X 1 , x1

Figure 2.10.1: Curvilinear Coordinates


The coordinate curves, curves of constant i , form a net in the undeformed configuration. One says that the curvilinear coordinates are convected or embedded, that is, the coordinate curves are attached to material particles and deform with the body, so that each material

Solid Mechanics Part III

279

Kelly

Section 2.10

particle has the same values of the coordinates i in both the reference and current configurations. In the current configuration, the spatial coordinates can be expressed in terms of a new, current, set of curvilinear coordinates

x = x (1 , 2 , 3 , t ) ,
with corresponding covariant base vectors g i and contravariant base vectors g i , with dx = dx i e i = d i g i ,

(2.10.4)

(2.10.5)

Example
Consider a motion whereby a cube of material, with sides of length L0 , is transformed into a cylinder of radius R and height H , Fig. 2.10.2.

R
H
L0

L0
Figure 2.10.2: a cube deformed into a cylinder
A plane view of one quarter of the cube and cylinder are shown in Fig. 2.10.3.

x2

P X

L0

x X1

p
x1

Figure 2.10.3: a cube deformed into a cylinder

Solid Mechanics Part III

280

Kelly

Section 2.10

The motion and inverse motion are given by

2R x = L0
1

(X ) (X ) + (X )
1 2 1 2

2 2

x = (X) ,

x2 = x3 =

2R L0

X 1X 2

(X ) + (X )
1 2

2 2

(basis: e i )

H 3 X L0

and
L0 2 2 x1 + x 2 2R L x2 2 X 2 = 0 1 x1 + x 2 2R x L0 3 x X3 = H X1 =

( ) ( )

X = 1 (x) ,

( ) ( )

(basis: E i )

Introducing a set of convected coordinates, Fig. 2.10.4, the material and spatial coordinates are
L X 1 = 0 1 2R L X 2 = 0 1 tan 2 2R X3 = and (these are simply cylindrical coordinates) L0 3 H

X = X(1 , 2 , 3 ) ,

x 1 = 1 cos 2
x = x (1 , 2 , 3 ) ,

x 2 = 1 sin 2 x 3 = 3

A typical material particle (denoted by p) is shown in Fig. 2.10.4. Note that the position vectors for p have the same i values, since they represent the same material particle.

Solid Mechanics Part III

281

Kelly

Section 2.10

X2

=
2

x2

2
X1

2
x1

1
1
1 = R
Figure 2.10.4: curvilinear coordinate curves

2.10.2

The Deformation Gradient

With convected curvilinear coordinates, the deformation gradient is


F = gi Gi ,

(2.10.6)

which is consistent with


dx = d j g j = d j g i G i G j = FdX

(2.10.7)

The deformation gradient F, the transpose F T and the inverses F 1 , F T , map the base vectors in one configuration onto the base vectors in the other configuration:
F = gi Gi F
1

FG i = g i F 1g i = G i F T G i = g i FTgi = Gi Deformation Gradient (2.10.8)

= Gi g

F T = g i G i FT = G i gi

Thus the tensors F and F 1 map the covariant base vectors into each other, whereas the tensors F T and F T map the contravariant base vectors into each other, as illustrated in Fig. 2.10.5.

Solid Mechanics Part III

282

Kelly

Section 2.10

contravariant basis

F T FT

G2
G2

g2

g2

G1
G1
covariant basis

g1

F
F 1

g1

Figure 2.10.5: the deformation gradient, its transpose and the inverses

Components of F

F has different components with respect to the different bases:


F = Fij G i G j = F ij G i G j = Fi j G i G j = F ji G i G j = f ij g i g j = f ij g i g j = f i j g i g j = f ji g i g j

F 1 = F 1 ij G i G j = F 1 G i G j = F 1 i G i G j = F 1 j G i G j
1 i ij

( ) = (f ) g
T i

gj

( ) = (f ) g
ij 1 ij
ij T ij

gj

( ) = (f ) g
j 1 j i
j i T j i i

gj
j

( ) = (f ) g
i 1 i j
T i j T i j

gj
j

F T = F T ij G i G j = F T G i G j = F T
ij

( ) = (f ) g
T

gj

( ) = (f ) g

gj
ij

( ) G G = (F ) G G = (f ) g g = (f ) g g
i i j j i

F T = F T ij G i G j = F T G i G j = F T
i ij

( ) = (f ) g

gj

( ) = (f ) g
T ij

j i

gj = f

T j i

G i G j = F T

i j

gi g j = f

T i j

Gi G j

gi g j

(2.10.9) The components of F with respect to the reference bases {G i }, G i are

{ }

Solid Mechanics Part III

283

Kelly

Section 2.10

Fij = G i FG j = G i g j =

X m x m i j

F ij = G i FG j = G jk G i g k Fi j = G i FG j = G jk G i g k F ij = G i FG j = G i g j = i x m X m j

(2.10.10)

and similarly for the components with respect to the current bases.
Components of the Base Vectors in different Bases

Now
g i = FG i = Fmj G m G j G i = F m G m G j G i j = Fmj G m i j = Fmi G m = F m G m i j j = Fim G m

(2.10.11)

showing that some of the components of the deformation gradient can be viewed also as components of the base vectors. Similarly,

G i = F 1g i = f
For the contravariant base vectors, one has
g i = F T G i = F T

( )
1

mi

gm = f

( )

1 m i

gm

(2.10.12)

( ) (G G )G = (F ) (G G )G = (F ) G = (F ) G = (F ) G = (F ) G
mj i m j T j m m j T mj T mi m i j T j m m i j m T i m m

(2.10.13)

and

G i = FTgi = f T

( )

mi

gm = f T

( )

i m

gm

(2.10.14)

2.10.3

Reduction to Material and Spatial Coordinates

Material Coordinates

Suppose that the material coordinates X i with Cartesian basis are used (rather than the convected coordinates with curvilinear basis G i ), Fig. 2.10.6. Then

Solid Mechanics Part III

284

Kelly

Section 2.10

i X i ,

X j X j Ej = E j = Ei i X i , i j X i j i i G = E = E =E X j X j Gi =

x j x j ej = ej i X i i j X i j gi = e = e x j x j gi =

(2.10.15)

and
F = g i G i = g i Ei = F
1

x j e j E i = Gradx i X X i i i = G i g = Ei g = E i e j = gradX j x

(2.10.16)

which are Eqns. 2.2.2, 2.2.4. Thus Gradx is the notation for F to be used when the material coordinates X i are used to describe the deformation.
reference configuration current configuration

E2

X2

E1

g2 g1

X3

X1
Figure 2.10.6: Material coordinates and deformed basis

Spatial Coordinates
Similarly, when the spatial coordinates x i are to be used as independent variables, then X j X j Ej = Ej i x i , i j x i j i G = E = E X j X j Gi = x j x j e j = i e j = ei i x i j x i j gi = e = j e = ei x j x gi =

i x i ,

(2.10.17)

and

Solid Mechanics Part III

285

Kelly

Section 2.10

F = g i G i = ei G i = F
1

x i e i E j = Gradx X j X j = G i g i = G i ei = E j e i = gradX i x

(2.10.18)

The descriptions are illustrated in Fig. 2.10.7. Note that the base vectors G i , g i are not the same in each of these cases (curvilinear, material and spatial).
X2

x2

G2

F = gi G

g2

G1
X1

g1
x1

F 1 = G i g i
X2
x2

E2 E1
X1

g2 g1

F=

x i e i E j = Grad x X j

x1

X2

x2

G2

F 1 =
X1

X E i e j = grad X j x
i

e2 e1
x1

G1

Figure 2.10.7: deformation described using different independent variables

Solid Mechanics Part III

286

Kelly

Section 2.10

2.10.4

Strain Tensors

The Cauchy-Green tensors


The right Cauchy-Green tensor C and the left Cauchy-Green tensor b are defined by Eqns. 2.2.10, 2.2.13,

C = FTF C 1 = F 1F T b = FF b
1 T 1

=F F

( )( ) = (G g )(g G ) = g G G (C ) G G = (g G )(G g ) = G g g b g g = (g G )(G g ) = G g g (b ) g g


= G i g i g j G j = g ij G i G j C ij G i G j
i j ij 1 ij i j i j i i j ij ij i j i j i j i j i j 1 i j i j ij ij

(2.10.19)

Thus the covariant components of the right Cauchy-Green tensor are the metric coefficients g ij , the covariant components of the identity tensor with respect to the convected bases in the current configuration, I g = g ij g i g j . It is possible to evaluate other components of C, e.g. C ij , and also its components with respect to the current basis through 2.10.14, but only the components C ij with respect to the reference basis will be used in the analysis. Similarly,

for b 1 , the components (b 1 )ij with respect to the current configuration will be used.

The Stretch
Now, analogous to 2.2.9, 2.2.12,

ds 2 = dx dx = dXCdX dS 2 = dX dX = dxb 1 dx
so that the stretches are, analogous to 2.2.17,

(2.10.20)

2 =
1

ds 2 dX dX = = dXCdX C 2 dX dX dS

dX i C ij dX dx b
i 1

dS 2 dx 1 dx = 2 = = dxb 1 dx b 2 dx dx ds

( ) dx
ij

(2.10.21)
j

The Green-Lagrange and Euler-Almansi Tensors


The Green-Lagrange strain tensor E and the Euler-Almansi strain tensor e are defined through 2.2.22, 2.2.24,

Solid Mechanics Part III

287

Kelly

Section 2.10

ds 2 dS 2 1 = dX (C I )dX dXEdX 2 2 2 2 ds dS 1 = dx I b 1 dx dxedx 2 2

(2.10.22)

The components of E and e can be evaluated through (writing G I , the identity tensor expressed in terms of the base vectors in the reference configuration, and g I , the identity tensor expressed in terms of the base vectors in the current configuration)

E=

1 (C G ) = 1 g ij G i G j Gij G i G j = 1 (g ij Gij )G i G j Eij G i G j 2 2 2 1 1 1 e = g b 1 = g ij g i g j Gij g i g j = (g ij Gij )g i g j eij g i g j 2 2 2 (2.10.23)

Note that the components of E and e with respect to their bases are equal, Eij = eij (although this is not true regarding their other components, e.g. E ij e ij ).

2.10.5

Intermediate Configurations

Stretch and Rotation Tensors

The polar decompositions F = RU = vR have been described in 2.2.5. The decompositions are illustrated in Fig. 2.10.8. In the material decomposition, the material is first stretched by U and then rotated by R. Let the base vectors in the associated intermediate configuration be {g i } . Similarly, in the spatial decomposition, the material is first rotated by R and then stretched by v. Let the base vectors in the associated intermediate configuration in this case be {G i }. Then, analogous to Eqn. 2.10.8, {Problem 1}
U = gi Gi U 1 = G i g i U T = g i G i UT = Gi gi
UG i = g i U 1g i = G i U T G i = g i UTgi = Gi (2.10.24)

v = gi Gi v 1 = G i g i v T = g i G
i

vG i = g i v 1g = G
i i T i

vT = Gi gi

v G =g vTgi = Gi

(2.10.25)

Solid Mechanics Part III

288

Kelly

Section 2.10

Note that U and v symmetric, U = U T , v = v T , so


U = gi Gi = Gi gi U 1 = G i g i = g i G i UG i = g i , Ug i = G i U 1g i = G i , U 1G i = g i

(2.10.26)

v = gi G i = Gi gi v 1 = G i g i = g i G i

vG i = g i , vg i = G i v 1g i = G i , v 1G i = g i

(2.10.27)

Similarly, for the rotation tensor, with R orthogonal, R 1 = R T , R = Gi Gi = Gi Gi RT = Gi Gi = Gi Gi R = gi gi = gi gi RT = gi gi = gi gi RG i = G i , RG i = G i R TG i = G i , R TG i = G i Rg i = g i , Rg i = g i R Tgi = gi , R Tgi = gi (2.10.28)

(2.10.29)

The above relations can be checked using Eqns. 2.10.8 and F = RU , F = vR , v 1 = RF 1 , etc.

{G }
i

{G i }
{g i }

{g i }
R

Figure 2.10.8: the material and spatial polar decompositions Various relations between the base vectors can be derived, for example,
G i g j = (RG i ) (Rg j ) = G i R T Rg j = G i g j Gi g j = Gi g j = Gi g j = L L L = Gi g j = Gi g j = Gi g j (2.10.30)

Solid Mechanics Part III

289

Kelly

Section 2.10

Deformation Gradient Relationship between Bases

The various base vectors are related above through the stretch and rotation tensors. The intermediate bases are related directly through the deformation gradient. For example, from 2.10.26a, 2.10.28b,
g i = UG i = UR T G i = F T G i

(2.10.31)

In the same way,


gi = FTG i g i = F 1G i G i = F T g i G i = Fg i

(2.10.32)

Tensor Components
The stretch and rotation tensors can be decomposed along any of the bases. For U the most natural bases would be {G i } and {G i }, for example,
U = U ij G i G j , U ij = G i UG j = G i g j U = U ij G i G j , U ij = G i UG j = G im G j g m U = U ij G i G j , U ij = G i UG j = G i g j U = U i j G i G j , U i j = G i UG j = g i G j with U ij = U ji , U ij = U ji , U ij = U ji , U i j = U ij . One also has (2.10.33)

v = vij G i G j , vij = G i vG j = G i g j v = v ij G i G j , v ij = G i vG j = G im G j g m v = vij G i G j , vij = G i vG j = G i g j v = vi j G i G j , vi j = G i vG j = g i G j with similar symmetry. Also, (2.10.34)

Solid Mechanics Part III

290

Kelly

Section 2.10

U 1 = U 1 ij g i g j , U 1 U 1 U 1 and v 1 = v 1 ij g i g j , v 1 v 1 v 1
1 ij i i i 1 ij i i i

( = (U = (U = (U

1 i j 1 j i

) )g )g )g

gj, g j, g j,

(U (U (U (U

1 ij 1 i j 1 j i

) ) ) )

ij

= g i U 1g j = G i g j = g i U 1g j = g im G m g j = g i U 1g j = g i G j = g i U 1g j = G i g j (2.10.35)

( = (v = (v = (v

1 i j 1 j i

) )g )g )g

g j, gj, gj,

(v (v (v (v

1 ij 1 i j 1 j i

) ) ) )

ij

= g i v 1g j = G i g j = g i v 1g j = g mj G m g i = g i v 1g j = g i G j = g i v 1g j = G i g j (2.10.36)

with similar symmetry. Note that, comparing 2.10.33a, 2.10.34a, 2.10.35a, 2.10.36a and using 2.10.30, U = U ij G i G j v = vij G i G j U v
1

( ) g g = (v ) g g = U
1
i ij

U ij = U 1

( )

ij

= vij = v 1

( )

ij

(2.10.37)

ij

Now note that rotations preserve vectors lengths and, in particular, preserve the metric, i.e.,

Gij = G i G j g ij = g i g j

= =

Gij = G i G j g ij = g i g j

(2.10.38)

Thus, again using 2.10.30, and 2.10.33-2.10.36, the contravariant components of the above tensors are also equal, U ij = (U 1 ) = v ij = (v 1 ) .
ij ij

As mentioned, the tensors can be decomposed along other bases, for example,

v = v ij g i g j , v ij = g i vg j = G i g j

(2.10.39)

2.10.6

Eigenvectors and Eigenvalues

Analogous to 2.2.5, the eigenvalues of C are determined from the eigenvalue problem

Solid Mechanics Part III

291

Kelly

Section 2.10

det (C C I ) = 0 leading to the characteristic equation 1.11.5


2 3 I C C + II C C IIIC = 0 C

(2.10.40)

(2.10.41)

with principal scalar invariants 1.11.6-7


I C = trC = Aii = C1 + C 2 + C3 II C =
1 2

[(trC)

tr(C 2 ) =

] (C C
1 2 i i

j j

C ij C i j = C1C 2 + C 2 C3 + C3 C1

(2.10.42)

IIIC = det C = ijk C1i C 2j C 3k = C1C 2 C3


The eigenvectors are the principal material directions N i , with

(C i I )N i The spectral decomposition is then

=0

(2.10.43)

C = i2 N i N i
i =1

(2.10.44)

where Ci = i2 and the i are the stretches. The remaining spectral decompositions in 2.2.37 hold also. Note also that the rotation tensor in terms of principal directions is (see 2.2.35)
R = ni Ni = ni Ni

(2.10.45)

where n i are the spatial principal directions.

2.10.7

Displacement and Displacement Gradients

Consider the displacement u of a material particle. This can be written in terms of covariant components U i and u i : u = x X U i G i = ui g i . The covariant derivative of u can be expressed as (2.10.46)

u = U m i G m = um i g m i
Solid Mechanics Part III 292

(2.10.47)

Kelly

Section 2.10

The single line refers to covariant differentiation with respect to the undeformed basis, i.e. the Christoffel symbols to use are functions of the Gij . The double line refers to covariant differentiation with respect to the deformed basis, i.e. the Christoffel symbols to use are functions of the g ij . Alternatively, the covariant derivative can be expressed as

u x X = = gi Gi i i i
and so

(2.10.48)

g i = G i + U m i G m = im + U m G m = Fim G m G i = g i u m i g m = im u m
The last equalities following from 2.10.11-12.

)g = ( f ) g
m 1 m i

(2.10.49)

The components of the Green-Lagrange and Euler-Almansi strain tensors 2.10.23 can be written in terms of displacements using relations 2.10.49 {Problem 2}:

E ij =

1 (g ij Gij ) = 2 1 eij = (g ij Gij ) = 2

1 Ui j + U j + U n iU n i j 2 1 ui j + u j u n i u n i j 2

( (

)
j

(2.10.50)

In terms of spatial coordinates, i = X i , G i = E i , g i = (x j / X i )e j , U i

= U i / X j , the

components of the Euler-Lagrange strain tensor are


E ij =
m n 1 (g ij Gij ) = 1 x i x j mn ij = 1 U ij + U ij + U ki U kj 2 X 2 2 X X X X X

(2.10.51)

which is 2.2.46.

2.10.8

The Deformation of Area and Volume Elements

Differential Volume Element

Consider a differential volume element formed by the elements d i G i in the undeformed configuration, Eqn. 1.16.36:

Solid Mechanics Part III

293

Kelly

Section 2.10

dV = G d1 d 2 d 3

(2.10.52)

where, Eqn. 1.16.13,


G = det Gij , Gij = G i G j

[ ]

(2.10.53)

The same volume element in the deformed configuration is determined by the elements d i g i :
dv = g d1 d 2 d 3

(2.10.54)

where

g = det g ij , g ij = g i g j
From 1.16.22 et seq., 2.10.11,
g = g1 g 2 g 3 = F1i F2j F3k G i G j G k = F1i F2j F3k ijk G = G det F

[ ]

(2.8.55)

(2.10.56)

where ijk is the Cartesian permutation symbol, and so the Jacobian determinant is (see 2.2.53)

J=

g dv = = det F dV G

(2.10.57)

and det F is the determinant of the matrix with components F ij .


Differential Area Element

Consider a differential surface (parallelogram) element in the undeformed configuration, bounded by two vector elements dX (1) and dX ( 2) , and with unit normal N . Then the vector normal to the surface element and with magnitude equal to the area of the surface is, using 1.16.23, given by
(G NdS = dX (1) dX ( 2) = d (1) i G i d ( 2 ) j G j = eijk ) d (1)i d ( 2 ) j G k

(2.10.58)

Solid Mechanics Part III

294

Kelly

Section 2.10
(G where eijk ) is the permutation symbol associated with the basis G i , i.e.

(G eijk ) = ijk G i G j G k = ijk G .

(2.10.59)

Using G k = F T g k , one has NdS = ijk G d (1)i d ( 2) j F T g k


Similarly, the surface vector in the deformed configuration with unit normal n is
(g nds = dx (1) dx ( 2 ) = d (1)i g i d ( 2 ) j g j = eijk ) d (1)i d ( 2 ) j g k (g where eijk ) is the permutation symbol associated with the basis g i , i.e.

(2.10.60)

(2.10.61)

(g eijk ) = ijk g i g j g k = ijk g .

(2.10.62)

Comparing the two expressions for the areas in the undeformed and deformed configurations, 2.10.60-61, one finds that nds =

g T F NdS = (det F )F T NdS G

(2.10.63)

which is Nansons relation, Eqn. 2.2.59.

2.10.9
1. 2.

Problems

Derive the relations 2.10.24. Use relations 2.10.49, with g ij = g i g j and Gij = G i G j , to derive 2.10.50 1 (g ij Gij ) = 1 U i j + U j i + U n i U n 2 2 1 1 eij = (g ij Gij ) = u i j + u j u n i u n i j 2 2

E ij =

( (

Solid Mechanics Part III

295

Kelly

You might also like