Download as pdf or txt
Download as pdf or txt
You are on page 1of 56

5 The Fundamental Equations of Continuum Mechanics

The Fundamental Equations


5
of Continuum Mechanics

5.1 Introduction

The fundamental equations of continuum mechanics are based on the conservation


principles of certain physical quantities. We consider five of these to establish the basic
equations that govern the Initial Boundary Value Problem (IBVP), namely:
ƒ The principle of conservation of mass;
ƒ The principle of conservation of linear momentum;
ƒ The principle of conservation of angular momentum;
ƒ The principle of conservation of energy;
ƒ The principle of irreversibility.
In this chapter we will address the fundamental principle of mechanics in the reference and
current configurations. At the end of the Chapter we will show that these principles are
insufficient to establish the IBVP set of partial differential equations, so, it is necessary to
add certain equations to fully resolve this problem. Then, we will introduce some concepts
and theorems to develop the concepts in this chapter.

5.2 Density
&
Density, denoted by f ( x , t ) , is a scalar function that measures the amount of a property
&
per unit volume around a material point ( x ) at a time t . One very important density
&
function is mass density, denoted by S ( x , t ) , which measures the amount of mass per unit
E.W.V. Chaves, Notes on Continuum Mechanics, Lecture Notes on Numerical 285
Methods in Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_7,
© International Center for Numerical Methods in Engineering (CIMNE), 2013
286 NOTES ON CONTINUUM MECHANICS

volume. Another density function we can quote is energy density, which measures stored
energy per unit volume. The term “specific” will be used to denote the amount of the
property per unit mass.

5.2.1 Mass Density


Any continuous medium is caused by a positive scalar quantity called mass. It is assumed
that the mass is continuously distributed throughout the continuum.
We will next review the concept of mass density introduced in Chapter 2. Let us consider a
sphere of infinitesimal radius centered at point P in the reference configuration, (see
Figure 5.1). The material contained in this sphere is denoted by 'm and the sphere volume
is represented by 'V0 . Then, the mass density S 0 , in the reference configuration, is
defined by the limit:
& & 'm dm
S ( x , t 0) S 0 ( X ) "im (5.1)
'V0 o0 'V0 dV0

(t 0) { t 0 t
'V0 o 0

'
V

o
0
 &  &
S0 ( X ) S ( x ,t )

x3

Reference configuration
x2
x1 Current configuration

Figure 5.1: Mass density.


Likewise, mass density in the current configuration is given by:
& 'm dm
S S ( x, t ) "im (5.2)
'V o 0 'V dV
& &
Then the functions S ( x , t ) and S 0 ( X ) are continuous density functions and are
interrelated to each other, (see Chapter 2), by:
& &
S0 (X ) JS ( x , t ) (5.3)

5.3 Flux

The properties conferred by density (e.g. mass, energy, entropy, etc.) are mobile and the rate
of change and direction of these quantities are assigned by the flux vector, usually denoted
& &
by q( x , t ) . With this information, we can define the amount of property that passes
through a differential area element da per unit time, (see Figure 5.2), as:
& &
q ˜ nˆ da q cos D da q n da (5.4)
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 287

& &
where n̂ is the unit normal vector, and D is the angle formed between q and qnˆ . Note
& &
that, only the normal vector qnˆ crosses the surface, since the tangential vector q sˆ remains
on the surface da . As an example of flux, we can mention the mass flux vector which is
represented by q S v . With regard to the SI unit we have: >q@
& & & kg &
2
where q represents
m s
the mass flux vector, and >q@
& J &
2
where q refers to the energy flux vector.
m s
& &
q qnˆ
D

Bt n̂ da

x3

x2
x1

Figure 5.2: Flux vector.

5.4 The Reynolds’ Transport Theorem


&
Let ) ( x , t ) be an Eulerian scalar field which describes a certain physical quantity per unit
&
volume. If ) ( x , t ) is continuous and differentiable, we can state that:

D & ª D & & D º


Dt V ³
) ( x , t )dV ³ «¬dV Dt )( x, t )  )( x, t ) Dt (dV )»¼
V

ª D
˜ v dV º»
& & &
³ «¬dV Dt )( x, t )  )( x, t )’
V
&
x
¼
(5.5)

ªD
˜ v º» dV
& & &
³ «¬ Dt )( x, t )  )( x, t )’
V
&
x
¼
whose equivalent in indicial notation is:
D & ªD & & wv k º
Dt V³) ( x , t )dV ³ «¬ Dt )( x, t )  )( x, t ) wx
V
» dV
k ¼
(5.6)

Then by using the material time derivative operator we can still state that:
D & ª w) ( x& , t ) &
w) ( x , t ) & wv º
Dt V³) ( x , t )dV ³ «¬«
V
w t
 vp
wx p
 ) ( x , t ) k » dV
wx k ¼»
& (5.7)
ª w) ( x , t ) w º
³ «
w t

wx
&

) ( x , t ) v p » dV
«
V ¬ p ¼»
This last equation is known as the Reynolds’ transport theorem and can be represented by the
following equations:
288 NOTES ON CONTINUUM MECHANICS

ªD
D
˜ v º» dV
& & & &
Dt V³) ( x , t )dV ³ «¬ Dt )( x, t )  )( x, t )’
V
&
x
¼
&
ª w) ( x , t ) & & º The Reynolds’ transport
³ « wt  ’ x& ˜ ) ( x , t ) v » dV (5.8)
V ¬ ¼ theorem
&
w) ( x , t ) & &
³ dV  ) ( x , t )v ˜ nˆ dS
³
V
w t S
t t

where Vt is the control volume, S t is the control surface, and n̂ is the outward unit
normal to the boundary S t of Bt . The first term on the right of equation is the local rate
of change of the property ) in the domain Vt , while the second term characterizes the
&
transport of )v , that leaves the domain Vt via the surface S t , (see Figure 5.3).

control volume &


(Gv )
St
&
Vt qn >(Gv& ) ˜ nˆ @ nˆ

&
& wG( x , t )
x
wt
control surface

Figure 5.3: Control volume.

Problem 5.1: Prove that Reynolds’ transport theorem is valid in the following equation:
D D
³
Dt V
) dV
Dt V
) JdV 0³ (5.9)
0

Solution:
D § D) DJ · § D) &· § D) &·
³) JdV 0 ³ ¨© J ) ¸dV 0 ³ ¨© J  J) ’ x& ˜ v ¸dV 0 ³ ¨© Dt  ) ’ x& ˜ v ¸ dV
Dt V V0
Dt Dt ¹ V0
Dt ¹ V ¹
0

5.4.1 Reynolds’ Transport Theorem for Volumes with


Discontinuities
Let us consider a material volume that is intersected by a discontinuous surface 6(t ) , a
&
singular surface which is moving over time at velocity Z , (see Figure 5.4). The surface 6(t )
divides the material volume into two parts viz. B  and B  . We can also define the
boundaries 6  and 6  situated ahead of and to the rear of the singular surface 6 as shown
in Figure 5.4.
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 289

S
B S

6(t )

B n̂
B


6(t )
&
Z

n̂
6
6

n̂ n̂

Figure 5.4: Material volume with discontinuity.

The mobile discontinuity 6(t ) is defined by the surface equation:


& &
f 6 ( x, t ) 0 x  6 (t ) (5.10)
Then, the unit normal vector n̂ on the surface 6(t ) is given by the equation:
& ’f 6
nˆ ( x , t ) (5.11)
’f 6
For material points belonging to the surface 6(t ) , the normal component of the velocity,
Z n , is defined as:
wf 6
Zn  wt (5.12)
’f 6
Then by combining the equation in (5.11) with (5.12) we obtain:
wf 6
& & ’f 6 & wf 6 & Df 6
Zn Z ˜ nˆ nˆ ˜ Z Ÿ  wt ˜Z Ÿ  ’f 6 ˜ Z 0 Ÿ 0 (5.13)
’f 6 ’f 6 wt Dt

Let A be a second-order tensor field and let us consider that A  and A  are the values of
A in the boundaries 6  and 6  , respectively. Then, we can define the jump of A as:
>>A @@ A  A (5.14)
Then by applying Gauss’ theorem (the divergence theorem) for the two domains B  and
B  we can obtain, respectively:
290 NOTES ON CONTINUUM MECHANICS

³ ’ ˜ A dV ³ A ˜ nˆ dS  ³ A ˜ nˆ  dS ³ ’ ˜ A dV ³ A ˜ nˆ dS  ³ A ˜ nˆ  dS
 

  
;   
(5.15)
V S 6 V S 6

Then, by summing up these two above equations we obtain:

³ ’ ˜ A dV ³ A ˜ nˆ dS  ³ A ˜ nˆ  dS  ³A ˜ nˆ  dS
 

     
(5.16)
V V S S 6 6

Additionally, if we bear in mind that nˆ  nˆ  nˆ and >>A @@ A   A  the above
equation becomes:

³ ’ ˜ A dV ³ A ˜ nˆ dS  ³ >>A @@˜ nˆ dS (5.17)


V  V  S  S  6

The Reynolds’ transport theorem can be modified for the case in which there is a singular
&
surface 6(t ) , which moves at the velocity Z , (see Figure 5.4). Then by applying the
equation in (5.8) to the two domains B  and B  , whose contours are S   6  and
S   6  , respectively, we obtain:
&
D & w) ( x , t ) & & & &
³
) ( x , t )dV ³ dV  ) ( x , t )v ˜ nˆ dS  )  ( x , t )Z ˜ nˆ  dS
³ ³
Dt V  V  w t S 
6 

& (5.18)
D & w) ( x , t ) & & & &
³
) ( x , t )dV ³ dV  ) ( x , t )v ˜ nˆ dS  )  ( x , t )Z ˜ nˆ  dS
³ ³
Dt V  V  w t S 
6 

Then by adding the two equations above, and once again considering that nˆ  nˆ  nˆ
and >>) @@ )   )  , we can conclude that:
&
w) ( x , t )
) ( x , t )v ˜ nˆ dS  >>) @@ Z ˜ nˆ dS
D & & & &
³
) ( x , t )dV
Dt V  V  ³ wt
dV  ³ ³ (5.19)
V  V  S  S  6

Additionally, by using the definition in (5.17) we can state that:


& & & & & &
³ )( x, t )v ˜ nˆ dS ³ ’ ˜ )( x, t )v dV  ³ >>)( x, t )v @@˜ nˆ dS
&
x (5.20)
S  S  V  V  6

Then by combining the above equation with that in (5.19) we obtain:


&
§ w) ( x , t )
D
˜ )( x, t )v ·¸dV  ³ >>)( x, t )v @@  >>)Z@@ ˜ nˆ dS (5.21)
& & & & & &
Dt V  V ³
) ( x , t )dV ³ ¨
© wt
 ’ x&
¹
V  V  6

which results in the Reynolds’ transport theorem for domains with discontinuities:
&
§ w) ( x , t )
D
˜ )( x, t ) … v ·¸dV  ³ >>) … v  Z @@ ˜ nˆ dS
& & & & &
Dt V 6³
) ( x , t )dV ³ ¨
V 6 ©
wt
 ’ x&
¹ 6

or (5.22)
&
D & § D) ( x , t ) & &· & &
³
) ( x , t )dV ³ ¨  ) ( x , t )’ x& ˜ v ¸ dV  >>) … v  Z @@ ˜ nˆ dS ³
Dt V 6 V 6 © Dt ¹ 6
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 291

5.5 Conservation Law

Conservation Law when applied to a particular physical quantity per unit volume, in a part
of the domain, states that no physical quantity (mass density, energy density, etc.) can be
created or destroyed, but merely moves from one place to another. The conservation law in
global form (weak) is established from Reynolds’ transport theorem:
&
ª w) ( x , t )
D
˜ )( x, t ) v º» dV
& & &
Dt V³) ( x , t )dV ³« wt
V ¬
 ’ x&
¼ (5.23)


z 0Ÿ source or sink

If the term on the left of the equation is nonzero this means that somewhere in the domain
there is a property source or sink, which can be represented locally by the variable Q .
Then, Q ! 0 indicates that there is a source, and Q  0 that there is a sink. For example, if
the property in question is mass density (mass per unit volume) in general Q 0 . However,
if there is a tumor (cells with abnormal growth) in a biological organism, we can establish a
law (at the macroscopic level) that indicates how the mass changes over time (source),
without regard to individual cells. Then, another example of a source we can cite is the
internal heat generated by a chemical reaction, such as in cement hydration. The effect of
the chemical reaction at the macroscopic level can be represented by a variable that
provides the amount of heat generated per unit volume and per unit time (the internal heat
source).
& &
Note that the term ()v ) shows the flux of the property ) . Then, if ()v ) represents the
& J
energy flux we have the following unit >)v @ , and if we are dealing with mass
m2s
& kg
transport we have >)v @ . As we have seen before, in general, the flux is represented
m2s
&
by q , with which we can establish the local form (strong) of conservation law and which is
denoted by the following continuity equation:
&
Q
w) ( x , t )
 ’ x&
& &
˜ q( x, t ) Continuity equation
>) @ (5.24)
wt s
where >) @ is the SI unit of the physical quantity per unit volume.

5.6 The Principle of Conservation of Mass. The


Mass Continuity Equation

The law of conservation of mass states that the total mass of a continuum does not change.
This implies that the total mass in the reference configuration is equal to the total mass in
the current configuration:
m ³S
V0
0 dV ³ S dV
V
>kg @ (5.25)
292 NOTES ON CONTINUUM MECHANICS

As a result of conservation of mass, the material time derivative of the total mass is zero,
D
i.e. m 0 , then:
Dt

D D & D & D & & D


Dt
m
Dt V ³
S ( x, t ) dV ³ Dt >S( x, t ) dV @ ³ dV Dt >S( x, t )@  S( x, t ) Dt >dV @
V V
0
(5.26)
ªD & & &º
³ « Dt >S ( x , t )@  S ( x , t )’ x ˜ v » dV 0
&

V ¬ ¼
or in indicial notation:
ªD & wv k º ª kg º
V
³ «¬ Dt >S( x, t )@  S wx » dV
k ¼
0 « s »
¬ ¼
(5.27)

If the above equation is valid for the entire domain, then it must also be satisfied locally:
DS ª kg º
 Sv k , k 0 « sm 3 » (5.28)
Dt ¬ ¼
which is the mass continuity equation in Eulerian description and is expressed in tensorial
notation as:
DS &
 S (’ x& ˜ v ) 0 The mass continuity equation (5.29)
Dt (Eulerian description)
Dx wx wx
Then by applying the material time derivative operator, i.e.  v k , the mass
Dt wt wx k
continuity equation (5.28) becomes:
DS wS wS wv wS w wS
 Sv k , k  vp S k  ( Sv k )  ( Sv k ) , k 0 (5.30)
Dt wt wx p wx k wt wx k wt

Hence, we have another way to express the mass continuity equation:


wS & The mass continuity equation
 ’ x& ˜ (S v ) 0 (5.31)
wt (Eulerian description)
We could have obtained the same equation in (5.31) by means of Reynolds’ transport
& &
theorem, i.e. in the equation (5.8) we substitute ) ( x , t ) for S ( x , t ) , which means:
& &
ª wS ( x , t ) wS ( x , t )
˜ S ( x, t ) v º» dV
D & & & & &
³
S ( x, t )dV ³ «¬  ’ x& 0Ÿ  ’ x& ˜ S ( x, t ) v 0 (5.32)
Dt V V
wt ¼ wt

We could also have obtained the mass continuity equation in (5.29) by means of the
principle of conservation of mass in a differential volume element dx1 dx 2 dx 3 , (see Figure
5.5), in which the following is satisfied:

Mass Inward mass Outward mass


accumulation = flux - flux

The rate of mass entering through face A is represented by the mass flux (Sv1 ) x1 dx 2 dx3 ,
ª w (Sv1 ) º
while the rate of mass that goes through face B is given by «Sv1  dx1 » dx 2 dx3 .
¬ wx1 ¼
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 293

Likewise, we can obtain the rate of change of mass in other faces. Moreover, by applying
the conservation of mass for the differential volume element we obtain:
­ª º
dx1 dx 2 dx3
wS
Sv1 dx 2 dx3  Sv 2 dx1 dx3  Sv3 dx1 dx 2  °®«Sv1  w(Sv1 ) dx1 » dx 2 dx3 
wt °̄¬ wx1 ¼
(5.33)
ª w ( Sv 2 ) º ª w ( Sv 3 ) º ½°
« Sv 2  dx 2 » dx1 dx 3  «Sv3  dx3 » dx1 dx 2 ¾
¬ wx 2 ¼ ¬ wx3 ¼ °¿

wS w (Sv1 ) w(Sv 2 ) w (Sv3 )


Then by simplifying the above equation we obtain    and
wt wx1 wx 2 wx3
by using the chain rule of derivative we find that:

wS wS wS wS § wv wv wv ·
 v1  v2  v3 S¨¨ 1  2  3 ¸
¸
wt wx1 wx 2 wx 3 w
© 1x wx 2 wx3 ¹
wS wS § wvi · DS & (5.34)
Ÿ  vi S ¨¨ ¸¸ Ÿ  S ’ x& ˜ v 0
wt wx i © w

xi ¹ Dt

& &
DS ’˜v Tr (’v )
Dt

face A ª w (Sv1 ) º
x2 face B «Sv1  dx1 » dx 2 dx 3
¬ wx1 ¼

Sv1 dx 2 dx 3 dx 2

dx3
x1 dx1
x1

x3
Figure 5.5: Conservation of mass in a differential volume element.

5.6.1 The Mass Continuity Equation in Lagrangian


Description
The mass continuity equation in (5.29) can also be expressed in Lagrangian description
(material). To do this, we can start from the conservation of mass which establishes:
& & &
³S
V0
0 ( X ) dV 0 ³ S( x, t ) dV ³ S
V
( x , t ) J dV
V0

& 0
(5.35)
f (X )

Since the above equation is valid for any volume it means that it will be valid locally too,
i.e.:
&
S0 (X ) J S (5.36)
Note that S 0 is independent of time and so is J S which results in the Lagrangian
description of the mass continuity equation:
294 NOTES ON CONTINUUM MECHANICS

D The mass continuity equation


(S J ) 0 (5.37)
Dt (Lagrangian description)
Problem 5.2: Show that
&
D & DPij ( x , t )
Dt V ³
S Pij ( x , t ) dV ³S
V
Dt
dV (5.38)
&
where Pij ( x , t ) is a continuum property per unit mass, which can be a scalar, a vector or
higher order tensor.
Solution: It was proven in equation (5.6) that:
D & ªD & & wv p º
Dt V ³
) ( x , t )dV ³ ««¬ Dt ) ( x, t )  ) ( x, t ) wx
V
»dV
p »
¼
Then by making ) S Pij , and by considering it in the above equation we obtain:
D ªD wv p º ª D DS wv º
³
Dt V
S Pij dV ³ «¬« Dt (SP
V
ij )  SPij » dV
wx p ¼» ³ «¬S Dt P
V
ij  Pij
Dt
 SPij k »dV
wx k ¼
ª D § DS wv · º
³
V
« S
¬ Dt
Pij  Pij ¨¨
Dt
S k
©
wx k
¸¸
¹
» dV
¼
0
mass continuity equation

Thus, we can conclude that:


D ª DPij º
Dt V ³
SPij dV ³ ««S Dt »¼
» dV
V ¬
Problem 5.3: Prove that the following relationship is valid:
& w & & &
Sa (S v )  ’ x& ˜ (S v … v ) (5.39)
wt
Solution: Based on the Reynolds’ transport theorem:
D w) &
) dV dV  ) (v ˜ nˆ ) dS
³ ³ ³
Dt V V
w t S
&
and if we consider that ) S v we obtain:
&
D & w (S v ) & &
S v dV ³ dV  S v … (v ˜ nˆ ) dS
³ ³
Dt V V
wt S
Then, the above equation in indicial notation becomes:
D w (S v i ) D w (S v i )
³ S vi dV ³
Dt V V
dV  ³ S v i (v k nˆ k ) dS Ÿ ³ S
wt S
v i dV
V
Dt


³
V
wt S
³
dV  (S v i v k )nˆ k dS
ai
Additionally, by applying the divergence theorem to the surface integral we obtain:
w (S v i ) ª w (S v i ) º
³ S ai dV ³
V V
dV  ³ (S v i v k ) ,k dV ³ «
wt V
 (S v i v k ) ,k » dV
V ¬ wt ¼
which in tensorial notation is:
& &
& ª w (S v ) & & º & w (S v ) & &
³ S a dV ³«  ’ x& ˜ (S v … v ) » dV Ÿ Sa  ’ x& ˜ (S v … v )
V V ¬ wt ¼ wt
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 295

5.6.2 Incompressibility
Compressibility is the ability to change the volume of a continuous medium. It is common
knowledge that gases are more compressible than liquids, but for practical purposes, the
liquid can be considered to be incompressible.
An incompressible medium is characterized by an isochoric motion, i.e. J 1 , hence the mass
density field (for all particle) is independent of time. In this case the mass continuity
equation in (5.29) boils down to:
DS & DS &
 S (’ x& ˜ v ) 0 Ÿ S ’ x& ˜ v Sv k ,k 0 (5.40)
Dt Dt
thus
& Mass continuity equation for a
’ x& ˜ v 0 (5.41)
incompressible medium
Thus, an incompressible medium can be characterized by:
D DS
>det( F )@ { J 0 ; { S 0 ; S S0 ; J 1 (5.42)
Dt Dt
or
wv1 wv 2 wv3
v k ,k 0 Ÿ Ÿ   Tr ( l ) Tr (D) 0 (5.43)
wx1 wx 2 wx3

where l denotes the spatial velocity gradient, and D is the rate-of-deformation tensor
which is equal to the symmetrical part of l , (see Chapter 2).

5.6.3 The Mass Continuity Equation for Volume with


Discontinuities
Now, let us consider a domain where there is a singular surface 6(t ) as established in
subsection 5.5.1, (see Figure 5.4). Based on the conservation of mass we have:
D
³
Dt V
SdV 0 (5.44)

and if we consider the Reynolds’ transport theorem with discontinuities, (see equation
(5.22)), in which ) S , we obtain:

D § DS &· & &


³
SdV  S’ x& ˜ v ¸dV  >>S v  Z @@ ˜ nˆ dS 0
³6 ¨© Dt ³ (5.45)
Dt V 6 V ¹ 6
& & &
where the mass density S ( x , t ) , and the velocity v ( x , t ) are continuous differentiable
& &
functions in V  6 , and >>S v  Z @@ is also a continuous differentiable function on 6 . The
global balance law is valid for any arbitrary parts of the volume and for the discontinuous
surface, hence it holds that:
DS &
 S’ x& ˜ v 0 in V  6 The mass continuity equation
Dt with discontinuities (5.46)
>>S v&  Z& @@ ˜ nˆ 0 on 6 (Eulerian description)
296 NOTES ON CONTINUUM MECHANICS

Problem 5.4: Let us consider the following velocity field:


xi
vi for t t 0
1 t
1) Find the mass density field;
2) Prove that this motion satisfies S x1 x 2 x 3 S 0 X 1 X 2 X 3 .
Solution: 1) By applying the mass continuity equation we obtain:
DS wv DS dS wv
S k 0 Ÿ { S k
Dt wx k Dt dt wx k
and by using the given velocity field, we find that:
wv i 1 wx i E ii 3
wx i 1  t wx i 1 t 1 t
Thus,
dS 3S dS 3dt
 Ÿ 
dt S 1 t
1 t
Then by integrating the both sides of the above equation we obtain:
dS 3dt
³ S ³  1  t Ÿ lnS 3 ln(1  t )  C
The constant of integration C is obtained by means of the above equation if we refer to
&
the initial condition t 0 , in which S ( x , t 0) S 0 , thus
ln S 0 3 ln(1  0)  C Ÿ C lnS 0
§ 1 · § S0 ·
ln S 3 ln(1  t )  ln S 0 ln¨¨ 3
¸  ln S 0
¸ ln¨¨ 3
¸
¸
© (1  t ) ¹ © (1  t ) ¹
Thus, we can conclude that:
S0
S
1  t 3
2) Then by using the velocity definition we obtain:
dx i xi dx i dt
vi Ÿ
dt 1 t xi 1 t
Additionally, by integrating the both sides of the above equation we obtain:
dx i dt
³ xi ³1 t Ÿ lnx i ln(1  t )  K i (5.47)
Then by applying the initial condition, i.e. at time t 0 Ÿ x i X i , we obtain:
ln(1  0)  K i
lnX i Ÿ K i lnX i
Additionally, by substituting the value of K i into the equation (5.47) we obtain:
lnx i ln(1  t )  ln X i Ÿ ln( x i ) ln> X i (1  t ) @
Hence we can conclude that x i X i (1  t ) , which gives us x1 X 1 (1  t ) , x 2 X 2 (1  t ) ,
S0
x3 X 3 (1  t ) , and if we consider that S , we obtain:
1  t 3
S t

1
t

1
t S0

1
Ÿ S x1 x 2 x 3 S0 X1 X 2 X 3
x1 x2 x3
X1 X2 X3
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 297

5.7 The Principle of Conservation of Linear


Momentum. The Equations of Motion

5.7.1 Linear Momentum


Let us consider the body, Bt , in motion which is subjected both to body forces (per unit
& & & &
mass), b( x , t ) , and to surface forces, t * ( x , t ) , acting on the surface S ı , (see Figure 5.6).
& &
Let v ( x , t ) be the Eulerian velocity field, then we can define the linear momentum of the
mass system Bt as:
& & &
L ³ v dm ³ S v dV Linear momentum
ª kg m º
« s » (5.48)
Bt V ¬ ¼

& &
n̂ t * ( x , nˆ , t )
Bt

dV & &
Sv ( x , t ) Sı

& &
x3 & Sb( x , t )
x

O
x2
x1
Figure 5.6: Continuum in motion.

5.7.2 The Principle of Conservation of Linear Momentum


The principle of conservation of linear momentum, based on Newton’s second law, states
that the rate of change of the linear momentum of an arbitrary part of a continuous
medium is equal to the resultant force (body and surface forces) acting on the part in
question, then:
&* & D & D
³t ³
dS  S b dV ³
S v dV ³ t dS  ³ S b dV ³ S v i dV
*
Dt V
i i
Dt V (5.49)
Sı V Sı V

The equation in (5.49) represents the global form of the principle of conservation of linear
momentum and by applying t *i V ij nˆ j we obtain:
D
³V

ˆ
ij n j dS ³
 Sb i dV
V
Dt V ³
Svi dV



Guass '  
(5.50)
Theorem

 ³ Uvi dV
³ Vij , j dV V
V

thus,
298 NOTES ON CONTINUUM MECHANICS

³ (V ij , j  Sb i  Svi )dV 0i ª kg m
« s2
º
N» (5.51)
V ¬ ¼
If the above equation is valid for the entire volume, it is also valid locally, i.e.:

V ij , j  Sb i  Svi 0i ª kg N Pa º
« s2 m2 (5.52)
¬ m3 m »¼
which are known as the equations of motion or Cauchy’s first equation of motion:
& & & The equations of motion
’ x& ˜ ı  Sb Sv Sa (5.53)
(Eulerian description)
Sometimes it is useful to express the equations of motion in the reference configuration.
To do this we can rewrite the equation in (5.49) in the undeformed configuration, i.e.:
&* & D & &
³t
S0
0 dS 0 ³
 S 0 b 0 dV0
V0
Dt V ³
S V J dV0 ³ S A J dV
V0
0
0

& & & (5.54)


D
³ P ˜N
ˆ dS  S b dV
0 0 0 ³
0 ³
S 0V dV0 ³ S 0 A dV0
S0 V0
Dt V V0
0

& & & & & &


where V { v ( X , t ) is the Lagrangian velocity, A { a ( X , t ) is the Lagrangian acceleration
&
field, P is the first Piola-Kirchhoff stress tensor, P Jı ˜ F T , and b 0 X , t is the body
&

forces vector per unit mass in the undeformed configuration. Note that P ˜ N ˆ ˜ P , since
ˆ zN
P is a non-symmetric tensor. Then by applying Gauss’ theorem (the divergence theorem)
to the surface integral we obtain:

³ ’
& & &
&
X
˜P  S 0 b0  S 0 A dV0 0
(5.55)
V0

Then, the local form of the equations of motion in material description (Lagrangian) can be
expressed as:
& &
’ X& ˜ P  S 0 b 0 S0 A The equations of motion
& & (5.56)
’ X& ˜ F ˜ S  S 0 b 0 S 0 A (Lagrangian description)

5.7.2.1 The Equilibrium Equations

In the exceptional cases when we have a static or quasi-static equilibrium, the acceleration
components are zero, thus we obtain the equilibrium equations as:
& & The equilibrium equations
’ x& ˜ ı  Sb 0 (5.57)
(Eulerian description)
Explicitly, the equations in (5.57), V ij , j  Sb i 0 i , can be expressed as:

­ wV11 wV12 wV13


°    Sb1 0
° wx1 wx 2 wx3
° wV 21 wV 22 wV 23
®    Sb 2 0 (5.58)
° wx1 wx 2 wx 3
° wV 31 wV 32 wV 33
° wx  wx  wx  Sb 3 0
¯ 1 2 3
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 299

Then by using both the engineering and Voigt notation, the equilibrium equations can be
expressed as follows:
V
ªw w w º ª« x º»
« x 0 0 0 »
«w wy wz » « V y » ª Sb º ª0 º
w w w «V z » 1
«0» Ÿ >L@T ^T`  ^M `
«0 0 0 » « »  ««Sb 2 »» « » ^0`
« wy wx wz » «W xy » (5.59)
« » «¬0»¼
« w w w » « W » ¬ Sb 3 ¼
«0 0 0 » « yz »
«¬ wz wy wx »¼ W

«¬ xz »¼
>L @T
Additionally, the equilibrium equations in the Lagrangian description are given by:
& &
’ X& ˜ P  S 0 b 0 0
& & The equilibrium equations (5.60)
’ X& ˜ ( F ˜ S )  S 0 b 0 0 (Lagrangian description)
Problem 5.5: Find the equilibrium equations in engineering notation by means of the
differential volume element equilibrium ( dxdydz ). For this purpose consider that the
Cauchy stress tensor field in the differential volume element varies as indicated in Figure
5.7.

z Rear face
Vx
W xy
wV z
Vz  dz
Rear face wz
wV yz
V yz  dz W xz
wz
wW xz
W xz  dz
wz wW yz dz
W yz  dy
W xy bz wy
Vy by wV y
wW xz Vy  dy
W xz  dx wy
wx bx wW xy
W xy  dy y
wy
W yz wW xy
W xy  dx dx
wV x wx
Vx  dx
wx

W xz

x W yz

Rear face
Vz

dy

Figure 5.7: The stress field in the differential volume element.


300 NOTES ON CONTINUUM MECHANICS

Solution:
To obtain the equilibrium equations we apply the force equilibrium condition in the
volume element. First, we evaluate the equilibrium force according to the x -direction:

¦F x 0

§ wV x · § wW xy ·
Sb x dxdydz  ¨ V x  dx ¸ dydz  V x dydz  ¨¨ W xy  dy ¸¸dxdz
© wx ¹ © wy ¹
§ wW ·
 W xy dxdz  ¨ W xz  xz dz ¸ dxdy  W xz dxdy 0
© wz ¹
Then by simplifying the above equation we obtain:
wV x wW xy wW
Sb x dxdydz  dxdydz  dxdydz  xz dxdydz 0
wx wy wz
wV wW xy wW xz
Sb x  x   0
wx wy wz

The equilibrium force according to the y -direction, ¦ Fy 0 , can be expressed as follows

§ wV y · § wW yz ·
Sb y dxdydz  ¨¨ V 22  dy ¸¸dxdz  V y dxdz  ¨¨ W yz  dz ¸¸dxdy
© wy ¹ © wz ¹
§ wW xy ·
 W yz dxdy  ¨¨ W xy  dx ¸¸ dydz  W xy dydz 0
© wx ¹
Then by simplifying the above equation we obtain:
wW xy wV y wW yz
Sb y    0
wx wy wx z

Finally, the equilibrium according to the z -direction, ¦ Fz 0 , is given by:

§ wV z · § wW ·
Sb z dxdydz  ¨ V z  dz ¸ dxdy  V z dxdy  ¨ W xz  xz dx ¸ dzdy
© wz ¹ © wx ¹
§ w W yz ·
 W xz dzdy  ¨¨ W yz  dy ¸¸ dxdz  W yz dxdz 0
© w y ¹
Additionally, by simplifying the above equation we obtain:
wW xz wW yz wV z
Sb z    0
wx wy wz
Then, the equilibrium equations in engineering notation become:
­ wV x wW xy wW xz
°    Sb x 0
° wx wy wz
°° wW xy wV y wW yz
®    Sb y 0
° wx wy wx z
° wW wW wV z
° xz  yz   Sb z 0
¯° wx wy wz
Problem 5.6: Let ı be the Cauchy stress tensor field, which is represented by its
components in the Cartesian basis as:
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 301

V11 x12 ; V 22 x 22 ; V 33 x12  x 22


V12 V 21 2 x1 x 2 ; V 23 V 32 V 31 V13 0
Considering that the body is in equilibrium, find the body forces acting on the continuum.
& &
Solution: By applying the equilibrium equations, ’ x& ˜ ı  Sb 0 , we obtain:
­ wV 11 wV12 wV13
°    Sb 1 0
° wx1 wx 2 wx 3
­2 x1  2 x1  Sb 1 0
°° wV 21 wV 22 wV 23 °
V ij , j  Sb i 0i Ÿ ®    Sb 2 0 Ÿ ® 2 x 2  2 x 2  Sb 2 0
° wx1 wx 2 wx 3 °Sb
° wV 31 wV 32 wV 33 ¯ 3 0
°    Sb 3 0
¯° wx1 wx 2 wx 3
Thus, to satisfy the equilibrium equations the following condition must be met:
4 x1  Sb 1 Ÿ Sb 1 4 x1
4 x 2  S b 2 Ÿ S b 2 4 x 2
Ÿ Sb 3 0
&
Sb 4( x1 eˆ 1  x 2 eˆ 2 )
Problem 5.7: The equations of motion of a body are given, in Lagrangian description, by:
­ x1 X 1  BtX 3
°
® x 2 X 2  BtX 3
°x
¯ 3 X 3  Bt ( X 1  X 2 )
where B is a constant scalar. Find the mass density in the current configuration (S ) in
terms of the mass density of the reference configuration (S 0 ) , i.e. S S (S 0 ) .
Solution:
We can apply the equation S 0 JS , where J is the Jacobian determinant and is given by:
wx1 wx1 wx1
wX 1 wX 2 wX 3
1 0 Bt
wx i wx 2 wx 2 wx 2
J F 0 1 Bt 1  2(Bt ) 2
wX j wX 1 wX 2 wX 3
wx 3 wx 3 wx 3  Bt  Bt 1
wX 1 wX 2 wX 3
S0 S0
Thus, we obtain S
J 1  2(Bt ) 2

5.7.3 The Equations of Motion with Discontinuities


Let us consider again a domain with a singular surface 6(t ) such as that discussed in
subsection 5.5.1, (see Figure 5.4). Then, the principle of conservation of linear momentum
becomes:
D & &
³
S v dV
Dt V  6 S 6
³ ı ˜ nˆ dS  ³ S b dV
V 6
(5.61)

Then by applying the divergence theorem with discontinuities, (see Eq. (5.17)), we obtain:

³ ’ ˜ ı  S b dS  ³ >>ı @@ ˜ nˆ
D & &
Dt V 6 ³
S v dV
V 6
&
x
6
dV (5.62)
302 NOTES ON CONTINUUM MECHANICS

&
Additionally, by using Reynolds’ transport theorem, (see equation (5.22)), with ) Sv , we
obtain:
&
D & § D ( Sv ) & &· & & &
³
Sv dV ³ ¨  Sv ’ x& ˜ v ¸ dV  >>Sv … v  Z @@ ˜ nˆ dS
³ (5.63)
Dt V 6 V 6 ©
Dt ¹ 6

Then by combining the above equation with the equation in (5.62) and by considering that
& &
D ( Sv ) & D ( S ) D (v )
v S we obtain:
Dt Dt Dt
& &· &
& § D (S ) & · § D (v ) & & &
³ v¨  S’ x& ˜ v ¸  ¨ S  ’ x& ˜ ı  S b ¸dV  >>Sv … v  Z  ı @@ ˜ nˆ dS
³ 0 (5.64)
V 6 © Dt ¹ © Dt ¹ 6

D (S ) &
Bearing in mind the mass continuity equation,  S’ x& ˜ v 0 , the equation in (5.64)
Dt
becomes:
& &· &
§ D (v ) & & &
³6 ¨© S  ’ x& ˜ ı  S b ¸dV  >>Sv … v  Z  ı @@ ˜ nˆ dS
³ 0 (5.65)
V
Dt ¹ 6

Then, the local form can be expressed as:


& & The equations of motion with
’ x& ˜ ı  Sb Sa in V  6
& discontinuities (5.66)
>>Sv& … v&  Z&  ı @@ ˜ nˆ 0 on 6 (Eulerian description)
For a static or quasi-static problem the equations in (5.66) become:
& &
’ x& ˜ ı  Sb 0 in V  6 The equations of motion with
&
­°>>ı @@ ˜ nˆ 0 discontinuities (static problem) (5.67)
® on 6
°̄Ÿ ı  ˜ nˆ ı  ˜ nˆ (Eulerian description)

5.8 The Principle of Conservation of Angular


Momentum. Symmetry of the Cauchy Stress
Tensor

5.8.1 Angular Momentum


Once again let us consider Figure 5.6, and we can define the angular momentum of a mass
system with respect to the origin by:
& & &
HO ³ ( x š Sv ) dV
V
Angular momentum (5.68)
H O i (t ) ³ (.
V
ijk x j S v k ) dV

&
The SI unit of H O is H O >& @ kg m 2
s ¬
&
, and ª« H O º»
¼
kg m 2
s2
Nm J.
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 303

5.8.2 The Principle of Conservation of Angular Momentum


The principle of conservation of angular momentum states that the rate of change of
angular momentum with respect to a point is equal to the resultant moment (with respect
to this point) produced by all forces acting on the body under consideration.
Then by obtaining the resultant momentum with respect to the origin, (see Figure 5.6), and
by applying the principle of angular momentum, we obtain:
& &* & & D & &
³ (x š t

³
)dS  ( x š Sb)dV
V
Dt V³( x š S v )dV >Nm@ (5.69)

NOTE: The equation in (5.69) is valid for those continuous media in which the forces
between particles are equal, opposite and collinear, and without any distributed moments.Ŷ
The equation in (5.69) can be rewritten in indicial notation as:
D D
³ (.

ijk ³
x j t *k )dS  (. ijk x j Sb k )dV
V
Dt V ³
(. ijk x j S v k )dV ³ S Dt (.
V
ijk x j v k )dV

³ S (.
V
x v  . ijk x j v k )dV
ijk
,j k
(5.70)
v

j

0i

Then by substituting t *k V kl nˆ l into the first integral of (5.70), and by applying the Gauss’
theorem, we obtain:

³.
V
ijk ³
( x j V kl ) ,l dV  (. ijk x j Sb k )dV
V
³ (.
V
ijk x j Sa k )dV (5.71)

³ (.
V
ijk x j ,l V kl  . ijk x j V kl ,l  . ijk x j Sb k )dV
, ³ (.
V
ijk x j Sa k )dV
(5.72)
E jl

ª º
³ «¬ .
V
ijk V kj  . ijk x j (V kl ,l  Sbk  Sa k ) » dV

¼
0i Ÿ ³
V
. ijk V kj dV 0i
(5.73)
0 k
Equations of motion

. ijk V kj 0i
 (5.74)
V jk V kj

Thus obtaining Cauchy’s second law of motion, also known as the Boltzmann postulate, the
symmetry of the Cauchy stress tensor is:

ı ıT Cauchy’s second law of motion (5.75)

Then bearing in mind the relationship ı J 1P ˜ F T , the Boltzmann postulate in the
reference configuration becomes:
ı ıT ½
°
§1 ¾T Ÿ P˜FT F ˜ PT (5.76)
1 T ·
P˜FT ¨ P ˜ F ¸ °
J © J ¹ ¿
and considering that P F ˜ S , where S is the second Piola-Kirchhoff stress tensor, we
obtain:
304 NOTES ON CONTINUUM MECHANICS

F ˜S ˜ F T F ˜ (F ˜ S)T ; F ˜S ˜ F T F ˜ ST ˜ F T (5.77)

Thus

S ST (5.78)
Problem 5.8: Find the linear and angular momentum for a solid subjected to rigid body
motion.
& x3c
F(n ) &
F( 2 )
Rigid body Bt
G
x2c
&
& v
x3 x
x1c

O
& G - mass center
x2 F(1)
x1

Solution: According to Problem 2.16 in Chapter 2, we obtained the velocity for rigid body
motion as:
& & & & &
v c  Ȧ š ( x  c )
&
where Ȧ is the axial vector (angular velocity) associated with the antisymmetric tensor W
(the spin tensor).
Linear momentum:
³ S c  Ȧ š ( x  c) dV ³ S c dV  ³ S Ȧ š x dV  ³ S Ȧ š c dV
& & & & & & & & & & &
L ³ S v dV
V V V V V
& & & & &
c S dV  Ȧ š S x dV  Ȧ š c S dV
³ ³ ³
V V V
& &
By definition ³
V
S x dV mx is the first moment of inertia, where m is the total mass, and
&
x k is the vector position of the center of mass G . The first moment of inertia is equal to
& & &
zero if the Cartesian system originates at the center of mass, so, S x c dV mx c 0 . ³
V
&
L >
& & & &
m c  Ȧ š ( x  c )
&
@ (Linear momentum for rigid body motion)
mv
& & & & &
where v c  Ȧ š ( x  c ) is the velocity of the center of mass.
Angular momentum:

³ ( x š Sv ) dV ³ >x š S c  Ȧ š ( x  c) @
& & & & & & & &
HO dV
V V
Thus
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 305

& & & & & & & & &


HO ³ S x š c dV  ³ S x š (Ȧ š x ) dV  ³ S x š (Ȧ š c) dV
V V V

ª º & ª º (5.79)
& & & & & & &
«¬V
³ »¼ V «¬V
³
« S x dV » š c  S x š (Ȧ š x ) dV  « S x dV » š (Ȧ š c )
»¼
³
Next, we discuss the second integral of the previous equation.
& & &
It was proven in Chapter 1 that given three vectors a , b , c , the relationship
& & & & & & & & & & &
a š (b š c ) (a ˜ c )b  (a ˜ b)c holds, thus when a c it holds that
& & & & & & & & & & & & & & & & & &
a š (b š a) (a ˜ a)b  (a ˜ b)a , so, ³ S x š (Ȧ š x) dV ³ S >( x ˜ x)Ȧ  ( x ˜ Ȧ) x @ dV , with
V V
which we obtain:
³ S >x k x k Zi  x p Z p xi @ dV
V
³ S >x
V
k x k Z p E pi  x p Z p x i dV @ ³ S >x
V
k @
x k E pi  x p x i Z p dV

³ S >x
V
k x k E pi  x p x i dV Z p @ I O ip Z p

or in tensorial notation:
& & & ª & & & & º & &
³ S x š (Ȧ š x ) dV «¬«³ S >( x ˜ x ) 1  ( x … x )@ dV »¼» ˜ Ȧ
V V
IO ˜ Ȧ

& & & &


where I O ³ S >( x ˜ x ) 1  ( x … x )@ dV is the inertia tensor with respect to the origin O . As
V
we can observe, I O is a second-order pseudo-tensor, since it depends on the reference
system, and the components I O ij ³ S >x
V
k @
x k E ij  x i x j dV can be expressed explicitly as:

³ S >( x x  x 2 x 2  x 3 x 3 )E 11  x1 x1 @ dV ³ S >x @
2
I O 11 1 1 2  x 32 dV
V V

³ S >x @ ³ S >x @
2
I O 22 1  x 32 dV ; I O 33 2
1  x 22 dV
V V

I O 12
V
³ S >( x x 1 1  x 2 x 2  x 3 x 3 )E 12  x1 x 2 @ dV  S >x1 x 2 @ dV
³
V
 I O 12

I O 13  S >x1 x 3 @ dV
³  I O 13 ; I O 23  S >x 2 x 3 @ dV
³  I O 23
V V
where I O 11 , I O 22 , I O 33 , are moments of inertia of the body relative to the reference point O ,
and I O 12 , I O 13 , I O 23 , are the products of inertia of the body relative to the reference point
O.
Returning to the equation in (5.79) we can state that:
&ª & º & & & & ª & º & &
«¬V »¼
³V «¬V »¼
³
« S x dV » š c  S x š (Ȧ š x ) dV  « S x dV » š (Ȧ š c )
HO ³
& & & & & & & & & & &
m x š c  I O ˜ Ȧ  m x š (Ȧ š c ) m x š c  (Ȧ š c )  I O ˜ Ȧ
& & &
> @
Then by adding and subtracting the term m x š Ȧ š x in the above equation we obtain:
& & & & &
& & & &
& & & & &
H O m x š c  Ȧ š c  I O ˜ Ȧ m x š c  Ȧ š ( x  c )  m x š (Ȧ š x )  I O ˜ Ȧ
&
> @
& & & &
>
& & & & & & & &
@
& & &
m x š v  m (x ˜ x) 1  ( x … x) ˜ Ȧ  IO ˜ Ȧ m x š v  m ( x … x)  (x ˜ x) 1  IO ˜ Ȧ
& &
^> @ `
&
m x š v  I ˜Ȧ
& & &
m x š v  HG
306 NOTES ON CONTINUUM MECHANICS

>& & & &


@
where I I O  m ( x … x )  ( x ˜ x ) 1 is the inertia pseudo-tensor, which is related to the
reference system at the center of mass. By means of this equation we can calculate the
inertia tensor in any reference system if we know the inertia tensor at the center of mass:
I O ij Iij  m>x i x j  ( x12  x 22  x 32 )E ij @ . Explicitly, these components can be expressed as:
I O 11 I11  m( x 22  x32 ) ; I O 12 I12  m( x1 x 2 )
I O 22 I 22  m( x12  x32 ) ; I O 23 I 23  m( x 2 x 3 )
I O 33 I33  m( x12  x 22 ) ; I O 13 I13  m( x1 x3 )
Note that, the above equations represent the parallel axis theorem (Steiner’s theorem) from
Classical Mechanics.
Problem 5.9: Obtain the principle of conservation of linear momentum and angular
momentum for a solid subjected to rigid body motion.
Solution: We can start from the definition of the principle of conservation of linear
momentum which states that:
& D & &
¦F ³
Dt V
S v dV L
& &
Then we use the equation of linear momentum obtained in Problem 5.8, L m v , to
obtain:
& D & & & &
¦F Dt V ³
S v dV L mv ma

Then we have:
& &
¦F ma
Now let us consider the principle of conservation of angular momentum which states:
& D & & D & &
¦M O
Dt V ³
( x š S v )dV
Dt
HO { HO

By which we obtain:
& & & &
¦M O HO or ¦M G HG
&
where the equation of angular momentum H O was obtained in Problem 5.8. The set of
& & & &
equations ¦F ma and ¦M G H G inform us that the following systems are
equivalent:
&
& & HG
F(n ) F( 2 )

G G
= &
ma

&
F(1) G - center of mass
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 307

5.9 The Principle of Conservation of Energy. The


Energy Equation

The principle of conservation of energy states:


“The rate of change of the kinetic energy plus the rate of change of the internal
energy is equal to the sum of the rate of change of the work done by the system plus (5.80)
the rate of change of any other energy supplied to, or removed from, the system”.
The energy supplied to, or removed from, the system per unit time can be any of three
kinds: thermal; chemical; or electromagnetic energy. In this publication we only consider
thermal energy as the energy added to the system. In such circumstances, the principle of
conservation of energy is known as the first law of thermodynamics. Mathematically, the
principle of conservation of energy, for continuum thermodynamics, is given by:
DK DU DW DQ ªJ º
  «s W» (5.81)
Dt Dt Dt Dt ¬ ¼
where K is the kinetic energy, U is the internal energy, W is the work done by the
system, and Q is the energy added to the system.
Next, we will introduce the types of energy involved in the energy equation.

5.9.1 Kinetic Energy


The kinetic energy of the system represented in Figure 5.6 is given by:
1 & & 1
K (t ) S (v ˜ v )dV
³ ³
S (vi vi )dV Kinetic energy >J @ (5.82)
2V 2V

kg m m Nm
The SI unit of the energy is the joule: >K@ ³m3
dV ³m 3
dV Nm J.
V
s s V

Then, the rate of change of the kinetic energy becomes:


D D ª1 º 1 D 1
Dt
K(t ) { K
Dt ³
« S (vi vi )dV »
«¬ 2 V »¼
³
S (vi v i )dV
2 V Dt 2V ³
S (vi vi  vi vi )dV (5.83)

Thus
D
Dt
K(t ) { K ³ S v v
V
i i dV (5.84)

5.9.2 External and Internal Mechanical Power


Let us consider the equations of motion V ij , j  Sb i Svi , and if we substituting those into
the rate of kinetic energy given in (5.84) we obtain:
K ³ v (V
V
i ij , j  Sb i )dV (5.85)

Then the term vi V ij , j can be substituted by:


308 NOTES ON CONTINUUM MECHANICS

(v i V ij ) , j v i , j V ij  vi V ij , j Ÿ v i V ij , j (vi V ij ) , j  vi , j V ij
, (5.86)
l ij

&
where ’ x& v { l is the spatial velocity gradient, which can be broken down into a
symmetric and an antisymmetric part, i.e. l D  W , (see Chapter 2), where D is the rate-
of-deformation tensor and W is the spin tensor. The components of these tensors can be
expressed in terms of Eulerian velocity as:
1 §¨ wvi wv j ·¸ 1 §¨ wv i wv j ·¸
l ij vi, j    ª m º
2 ¨ wx j wxi ¸ 2 ¨ wx j wxi ¸ «m s» (5.87)
©
¹ ©
¹ ¬ ¼
Dij Wij

Returning to the equation in (5.85), and considering the relationships in (5.86) and (5.87),
the rate of change of the kinetic energy becomes:
K
V
³ > v V
i ij , j  v i , j V ij  Sb i vi dV @ ³ >(v V
V
i ij ) , j @
 V ij (D ij  Wij )  Sb i vi dV
(5.88)

V
³ Sb v dV  ³ v V
i i
V
i ij , j ³
dV  V ij D ij dV
V

where we have taken into account that the double scalar product of a symmetric and
antisymmetric tensor is equal to zero, i.e. V ij Wij 0 or ı : W 0 . Then by applying the
divergence theorem to the second integral of the right side of the equation in (5.88), we
find that:

³ v V ³v V ˆ
³ v t dS
*
i ij , j dV i ij n j dS i i (5.89)
V Sı Sı

By combining the above relationship with the equation in (5.88), we can still express the
rate of change of the kinetic energy as:
K ³ Sb v dV  ³ v t ³
*
i i i i dS  V ij D ij dV
D
V

Sı V

Ÿ K Pext (t )  Pint (t ) (5.90)
Pext (t ) Pint ( t ) Dt
External Mechanical Power Internal Mechanical Power

or
D
K  Pint (t ) Pext (t ) (5.91)
Dt
where we have introduced the external mechanical power Pext (t ) , which is the rate of change
DW
of the work done by the external forces , as:
Dt
& & & &
Pext (t ) ³ t * ˜ v dS  S b ˜ v dV³
Sı V ªJ º
The external mechanical power « W» (5.92)
Pext (t ) ³

t *i vi dS ³
 S b i vi dV
V
¬s ¼

and the internal mechanical power, also known as the stress power, which is the rate of change of
the work done by the internal forces:

³V D ³ ı : D dV ³ Tr ı ˜ D dV ³ Tr ı ˜ D dV
T
Pint ij ij dV The stress power (5.93)
V V V V
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 309

NOTE: The SI unit of power is the watt, W { J / s , i.e. one joule ( J ) per second ( s ),
m N m J J
which is equal to >Pint @ ³ Pa m s dV ³ m 2
dV ³m 3
dV W.Ŷ
V V
ms V
s s

We can also define the stress power per unit volume, denoted by w int (t ) , as:

w int (t ) ı : D Tr (ı ˜ D) Stress power per unit volume (5.94)

Then by starting from the stress power in the current configuration we can also express the
stress power as a function of the other stress tensors, i.e.:

³ ı : D dV ³ ,
V
Jı : D dV
IJ V0
³ IJ : D dV 0
V0
0 (5.95)

Bearing in mind that W ij Pik F jk (Kirchhoff stress tensor components), P ˜ F T F ˜ PT


(Cauchy’s second law of motion in the reference configuration), and D ij F pi1 E pl Flj1 , we
obtain:

³ IJ : D dV ³P ³P
1  1
0 ik F jk F pi E pl Flj dV0 ik F jk D ij dV0
V0 V0 V0

³P ³P ( l ij  Wij ) dV0
1 
ik F pi E pl F 1 F jk dV0 ik F jk
lj

V0 V0
Glk

³P ³F ³P ik l ij F jk
1  1 
ik F pi E pk dV0 pi Pik E pk dV0 dV0 (5.96)
V0 V0 V0

³S
V0

pk E pk dV0
V0
³P  dV
ik Fik 0
V0
³ P : F dV 0

1 1 S
³ S : E dV
V0
0
2V³S : C dV0 ³ J P : F dV ³ S
V V 0
P : F dV
0

which proves that the rate of change of the deformation gradient and the first Piola-
Kirchhoff stress tensor are conjugate quantities ( P : F ). Other conjugate quantities are: the
second Piola-Kirchhoff stress tensor and the rate of change of the Green-Lagrange strain
tensor ( S : E ); the Kirchhoff stress tensor and the rate-of-deformation tensor ( IJ : D ).
Furthermore, we can show that T : U is already a conjugate pair. To prove this, let us
consider the relationship P R ˜ T , where T U ˜ S is the Biot stress tensor, and R is the
orthogonal tensor from the polar decomposition, and U UT is the right stretch tensor.
Then if we refer to the right polar decomposition, i.e. F R ˜ U Ÿ F R ˜ U  R ˜ U , we
obtain:
P : F R ˜ T : R ˜ U  R ˜ U Pij Fij R ip T pj R ik U kj  R ik U kj
R ˜ T : R ˜ U  R ˜ T : R ˜ U R ip T pj R ik U kj  R ip T pj R ik U kj
T ˜ UT : R T ˜ R  T : U ( T U )(R R )  T U
pj kj ip ik kj kj (5.97)
U ˜ S ˜ UT : R T ˜ R  T : U (U pq S qj U kj )(R ip R ik )  Tkj U kj

T :U T Ukj kj
310 NOTES ON CONTINUUM MECHANICS

Note that U ˜ S ˜ U : R T ˜ R 0 , since the tensor (U ˜ S ˜ UT ) T U ˜ S ˜ UT U ˜ S ˜ U is


symmetrical and R T
˜ R is an antisymmetric tensor. Thus, the equation in (5.97)
becomes:
P : F  T :H
T :U  (5.98)
where H U  1 is the Biot strain tensor, (see Chapter 2) and if we know that U is
symmetrical, it is also possible to express the above relationship as:
P : F  T sym : U
( T sym  T skew ) : U  (5.99)
Then, if we take into account all the equations obtained before, we can summarize the
stress power per unit volume by:
S 1 The stress power per unit current
w int ı :D P : F P : F (5.100)
S0 J volume

1  T :H
 The stress power per unit reference
w int IJ : D S : E P : F S : C T :U (5.101)
2 volume

5.9.3 The Balance of Mechanical Energy


If we compare the equation given in (5.91) with the energy equation (5.81), i.e.:
D DK DU DW DQ
K  Pint (t ) Pext (t ) œ  
(5.102)
Dt Dt Dt Dt Dt
,
0
we can observe that the equation in (5.91) is an exceptional case of the energy equation
where only mechanical energy is considered. In this case the principle of conservation of
energy is known as the balance of mechanical energy which is otherwise known as the theorem of
power extended:

D 1 2
³ Sv dV  V ij D ij dV ³ ³ Sb v dV  ³ v t dS
*
i i i i
Dt V 2 V V Sı

D 1 2 & & &* & Balance of


Sb ˜ v dV  ˜ v dS (5.103)
³
Dt V 2
Sv dV  ı : DdV ³ ³ ³t mechanical energy

V
V




Rate of change Stress power External mechanical power
of the Kinetic energy

OBS.: In rigid body motion D 0 is satisfied, so, the stress power (internal
mechanical power) is zero Pint (t ) 0 , then it holds that K Pext (t ) .

If K is discarded, which characterizes a static or quasi-static regime, it holds that


Pint (t ) Pext (t ) .
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 311

Problem 5.10: Find the kinetic energy related to rigid body motion in terms of the inertia
tensor, (see Problem 5.8 and Problem 5.9).
& & & & &
Solution: The rigid body motion velocity can be expressed as v c  Ȧ š ( x  c ) . Then, the
kinetic energy becomes:
K (t )
1
1 ­ & &
& &
S ® c  Ȧ š ( x&  c& ) ˜ c&  Ȧ
S (v ˜ v )dV
2V ¯
2V ³ &
³ & & ½
>
š ( x  c) ¾ dV
¿
@> @
& & & &
Using the following vector sum x x  x c , where x is the mass center vector position,
&
and xc is the particle vector position with respect to the system that has its origin in the
center of mass, the energy equation becomes:
K (t )
1
2V
& &
^> & & & & & & &
@>
&
S c  Ȧ š (( x  x c)  c) ˜ c  Ȧ š (( x  x c)  c ) dV
³ @`
1
2V ³
& &
^> & & & & & &
& &
@ >
& &
S c  Ȧ š ( x  c )  (Ȧ š x c) ˜ c  Ȧ š ( x  c )  (Ȧ š x c) dV @`
& & & & &
Note that v c  Ȧ š ( x  c ) is the center of mass velocity, thus:

³ ^> @> @`
1 & & & & & &
K (t ) S v  (Ȧ š x c) ˜ v  (Ȧ š x c) dV
2V
or:
1 & & 1 & & & 1 & & & 1 & & & &
K (t ) ³ Sv ˜ v dV  Sv ˜ (Ȧ š x c) dV 
³ S (Ȧ š x c) ˜ v dV   S (Ȧ š x c) ˜ (Ȧ š x c) dV
³ ³
2V 2V 2V 2V

Then by simplifying the above equation we obtain:


1 & & & & & 1 & & & &
K (t ) Sv ˜ v dV  Sv ˜ (Ȧ š x c) dV 
³ ³ S (Ȧ š x c) ˜ (Ȧ š x c) dV ³
2V V
2 V
Next, we discuss separately the terms of the previous equation:
1 & & 1 & 2 1
1) Sv ˜ v dV
³ v ³ S dV mv 2
2V 2 V
2
& & & & ª& & º & & &
2) ³ Sv ˜ (Ȧ š x c) dV
v ˜ «Ȧ š Sx c dV » v ˜ (Ȧ š m , ³
x c) 0
&
V ¬« V ¼» 0
&
Note that, the system xc is located at the center of mass ( G ), hence the center of mass
&
vector position related to the system xc is zero.
& & & &
3) S (Ȧ š x c) ˜ (Ȧ š x c) dV
³
V
& & & &
³ S >(Ȧ š x c) ˜ (Ȧ š x c)@ dV ³ S .
V V
ijk Z j x kc . ipq Z p x qc dV
V
³ S (E jp E kq  E jq E kp )Z j x kc Z p x qc dV

³S Z
V
j (E jp E kq x kc x qc  E jq E kp x kc x qc )Z p dV

³S Z
V
j (E jp x kc x kc  x cp x cj )Z p dV

§ ·
Z j ¨ S (E jp x kc x kc  x cj x cp ) dV ¸Z p
³
¨ ¸
©V ¹
Z j I jp Z p
or in tensorial notation as:
312 NOTES ON CONTINUUM MECHANICS

& & & & & ª & & & & º &
³ S >(Ȧ š x c) ˜ (Ȧ š x c)@ dV
V
Ȧ ˜ « S >( x c ˜ x c) 1  ( x c … x c) @ dV » ˜ Ȧ
³
«¬V »¼
& &
Ȧ˜ I ˜Ȧ
where I is the inertia pseudo-tensor related to the system located at the center of mass,
(see Problem 5.8).
Then if we bear in mind all the above considerations, the kinetic energy equation for rigid
body motion becomes:
1 & & 1 & & & 1 & & & &
K (t ) ³ Sv ˜ v dV  2Sv ˜ (Ȧ š x c) dV 
³ S (Ȧ š x c) ˜ (Ȧ š x c) dV³
2V 2 2V
V

0

1 1& &
K (t ) mv 2  Ȧ ˜ I ˜ Ȧ
2 2
Additionally, if we take into account that:
ª º
³ >
« S x 2c  x 3c dV @  S >x1c x 2c @ dV
³  S >x1c x 3c @ dV »
³
2 2

«V » ª I11  I12  I13 º


> @
V V
«  S >x c x c @ dV « »
Iij « ³ 1 2 ³ S x1c  x3c dV
2 2
 S >x 2c x 3c @ dV »»
³ «  I12 I 22  I 23 »
« V » « I  I 23 I 33 »¼
³ > @
V V
«  S >x1c x 3c @ dV  S >x 2c x 3c @ dV ¬ 13

¬« V
³ ³
V V
S x1c 2  x 2c 2 dV »
¼»
we obtain an explicit equation for the kinetic energy as:
1 1
K (t ) mv 2  Z k I kj Z j
2 2
ª I11  I12  I13 º ª Z1 º
1 1 « »
mv 2  >Z1 Z2 Z 3 @ « I12 I 22  I 23 » ««Z 2 »»
2 2 « I
¬ 13  I 23 I 33 »¼ «¬Z 3 »¼
1
2
1
>
mv 2  I11Z12  I 22 Z 22  I 33 Z32  2 I12 Z1 Z 2  2 I13 Z1 Z3  2 I 23 Z 2 Z3
2
@

K (t )
1
2
1
>
mv 2  I11Z12  I 22 Z 22  I 33 Z32  2 I12 Z1 Z 2  2 I13 Z1Z 3  2 I 23 Z 2 Z 3
2
@

5.9.4 The Internal Energy


If we take a handful of atoms (the material point) and we evaluate the average of all forms
of energy present in it we obtain what is known as the internal energy. Continuum
thermodynamics usually presents the rate of change of the internal energy as:
DU D ªJ º
Dt Dt V³Su dV
V
³ Su dV «s»
¬ ¼
(5.104)

J
where u is the specific internal energy, i.e. energy per unit mass, >u @ . For example, for an
kg
p
ideal gas the specific internal energy is given by u c v T c p T  , where T is the
, S
I
temperature, c v is the specific heat capacity at a constant volume, I is the specific entropy,
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 313

c p is the specific heat capacity at a constant pressure, p is the thermodynamic pressure,


and S is the mass density. We can give another example with the mechanical problem,
which was discussed in the previous subsection, where the rate of change of the internal
DU
energy is given by
Dt ³ ı : D dV .
V

5.9.5 Thermal Power


We define thermal power as the rate of increase of total heat in the continuum, which is
DQ
denoted by . The contribution of thermal power considered here is caused by:
Dt
ƒ The Cauchy heat flux (non convective, i.e. without mass transport);
ƒ The heat sources.
1) The Cauchy heat flux
Let us assume that there is a temperature gradient in the continuum, so there is scientific
evidence of energy transfer (heat) from the hottest to the colder region. Then, we can
represent this transferred energy per unit area per unit time by the thermal flux vector
& &
q( x , t ) , which is also known as the Cauchy heat flux or true heat flux. Now, let us consider the
domain B bounded by the surface S , (see Figure 5.8). The amount of energy which is
transferred through the surface dS per unit time, (see Figure 5.8), is represented by
& &
q( x , t ) ˜ nˆ dS , where n̂ is the outward unit normal to the area element dS . Meanwhile, the
tangential component remains on the surface. Thus, the rate of increase of total heat, due
to thermal flux, in the continuum is given by:
& &
³  q( x, t ) ˜ nˆ dS ªJ
«s
º
W» (5.105)
S ¬ ¼
2) The heat sources
If in a continuum there is a nuclear or chemical reaction which results in the release of
heat, we can represent this by means of the heat sources, (see Figure 5.8).
We represent the rate of increase of total heat in the continuum cause by the heat source
as:

³ S r dV ªJ
«s
º
W» (5.106)
V ¬ ¼
&
where r ( x , t ) is the radiant heat constant (also called the heat source) per unit mass per unit
& J &
time, a scalar function, and the SI unit is >r ( x , t )@ , and S ( x , t ) is the mass density.
s kg
Then by considering the heat flux (incoming) and the heat source, we can define the thermal
power (the rate of thermal work) as:
DQ & ªJ º
Dt ³ S r dV  ³ q ˜ nˆ dS
V S
The thermal power «s
¬

¼
(5.107)
314 NOTES ON CONTINUUM MECHANICS

& & &


q n ( x , t ) (q ˜ nˆ ) nˆ q n nˆ
Current configuration
& &
q( x , t )
Bt n̂
x3

dS
dV
&
Sr ( x , t )
>q& @ J
x2 m2s
J
x1 >r @
kg s

Figure 5.8: Heat flux and heat source.

5.9.6 The First Law of Thermodynamics. The Energy


Equation
Once we know what forms of energy are involved in a system we can provide the energy
equation by starting from that in (5.81):
DK DU DW DQ
  (5.108)
Dt Dt Dt Dt
The mechanical power and the thermal power are not exact differentials ( Dx
Dt
), but there is
experimental evidence showing that the sum of mechanical and thermal power is already an
exact differential, (Mase(1977)).
Considering only the mechanical and thermal energy, the principle of conservation energy
becomes what is known as the first law of thermodynamics, which postulates the
interchangeability of mechanical and thermal energy. Then, the equation in (5.108)
becomes:
Vijnˆ j
-
D vv
Dt V ³
S i i dV  S udV
2 V
³ ³

³
t *i vi dS  Svi b i dV  S rdV  q i nˆ i dS
V
³
V
³
Sq
(5.109)

Then by using divergence theorem to transform the surface integral into the volume
integral we obtain:

³ V v dV  ³ Sv b dV  ³ S rdV  ³ q
D vv
Dt V ³
S i i dV  SudV
2 ³ ij i , j i i i ,i dV

³ > V v  Sv b  S r  q @ dV
V V V V V

³ Sv v dV  ³ SudV
V
i i
V V
ij i , j i i i ,i (5.110)

³ SudV ³ V
V V
ij , j v i  V ij v i , j  Sb i vi  S r  q i ,i  Svi vi dV
Additionally, by rearranging the above equation we obtain:
ª º
³ SudV ³ «¬ v V  S
b  S v
V V
i ij , j   V i i ij v i , j  Sr  q i ,i » dV
¼ (5.111)
0i
the equations of motion
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 315

Then if we bear in mind that vi , j D ij  Wij , the above equation becomes:

³ Su dV ³ >V D
V V
ij ij
 Wij  Sr  q i ,i dV Ÿ @ ³ Su dV ³ V D
V V
ij ij
 Sr  q i ,i dV (5.112)

The local form of the above equation is known as the energy equation:

Su V ij D ij  Sr  q i ,i ª J W º
« m3 s (5.113)
¬ m 3 »¼
which is expressed in tensorial notation as:
&
S u ı : D  ’ x& ˜ q  Sr The energy equation (current configuration) (5.114)

NOTE: For a purely mechanical problem in which there is no internal heat production
& &
( r 0 ) nor heat flux q 0 , the energy equation becomes:
1 ª J º
u ı :D « » (5.115)
S ¬ s kg ¼
ª1 º m3 N m Nm J
where the SI unit can easily be verified >u @ « ı : D» Ŷ
¬S ¼ kg m 2 m s s kg s kg

5.9.6.1 The Energy Equation in Lagrangian Description

The energy equation (5.114) can also be established in Lagrangian description (material
description). From the equation in (5.112), the integral related to the integral energy can be
written in the reference configuration as:
& & & &
³ S( x, t )u( x, t )dV ³ JSu dV ³ S
V V0
0
V0
0 ( X )u
 ( X , t )dV0 (5.116)

The integral associated with stress power can be established in the reference and current
configuration, (see equations (5.100) and (5.101)), as shown bellow:

1 S
³ ı : DdV ³ ,
V
Jı : D dV
IJ
V0
³ IJ : DdV ³ S : E dV ³ 2 S : CdV ³ P : FdV ³ S
0
V0
0
V0
0
V0
0
V0
0
V 0
P : F dV

(5.117)
Similarly for the integral related with the heat source, i.e.:
& & & &
³ S( x, t ) r ( x, t ) dV ³ JS r dV ³ S
V V0 V0
0 0 (X ) r ( X , t ) dV0
(5.118)






current configuration reference configuration

Finally, we can address the integral related to the heat flux. The amount of heat that passes
through the area element da in the current configuration must in theory be the same as,
that which passes through the area element dA in the reference configuration, (see Figure
5.9). Then the following relationship must be met:
& & & &
q 0 ˜ dA q ˜ da (5.119)
&
where q 0 is the heat flux in the reference configuration. Then if we use Nanson’s formula
& &
da JF T ˜ dA , obtained in Chapter 2, the equation in (5.119) becomes:
316 NOTES ON CONTINUUM MECHANICS

& & & & & & & &


q 0 ˜ dA J q ˜ F  T ˜ dA Ÿ q 0 J q ˜ F T Ÿ q J 1 q 0 ˜ F T (5.120)

&
& F q
q0
&
dA
& &
& da
da JF T ˜ dA
& &
q J 1 q 0 ˜ F T

Reference configuration
Current configuration

Figure 5.9: Heat flux.


&
Thus, the integral ³ ’ x& ˜ qdV can be written in the reference configuration as:
V

& wq i w §1 ·
³’ &
x ˜ qdV ³q i ,i dV ³ J wx dV0 ³ J wx ¨ q 0 k Fik ¸ dV0
V V V0 i V0 i ©J ¹
(5.121)
wq § 1 · w §1 ·
V0
wx i © J ¹
³
J 0 k ¨ Fik ¸  J q 0 k
wx i
¨ Fik ¸ dV0
© J ¹
&

It was proven in Chapter 2 that ’ x& ˜ J 1 F 0 , thus, the above equation becomes:
& wq 0 k § 1 · wq 0 k § wx i · wq 0 k
³’ & ˜ qdV ³J ¨ Fik ¸ dV ³ ¨ ¸¸ dV0 ³ dV0
wxi ¨© wX k
x
V V0
wx i © J ¹ V0 ¹ V0
wX k
(5.122)
&
³ ’ X& ˜ q 0 dV0
V0

Bearing in mind the equations in (5.116), (5.117), (5.118) and (5.122), the energy equation
in the reference configuration can be established as:

³ S : E  ’ ˜ q 0  S 0 r dV0
&
³S
V0
0 u dV0
V0
&
X (5.123)

Additionally, the local form of the above equation is:


& & & The energy equation
S 0 u ( X , t ) S : E  ’ X& ˜ q 0  S 0 r ( X , t ) (5.124)
(reference configuration)

5.9.7 The Energy Equation with Discontinuity


In this subsection we obtain the energy equation for a domain with a singular surface 6(t )
as discussed in subsection 5.5.1, (see Figure 5.4). In this case the energy equation becomes:
DK DU DW DQ
 
Dt Dt Dt Dt
(5.125)
D ª1 & & º & & & &
« ³ S (v ˜ v )dV  SudV » Sb ˜ v dV  v ˜ ı ˜ ndS  S rdV  q ˜ nˆ dS
³ ³ ³ ³ ³
Dt ¬« 2 V 6 V 6 ¼» V 6 S  S  V 6 S  S 
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 317

For the terms on the left of the equation in (5.125) we can apply Reynolds’ transport
& &
theorem, (see the equation in (5.22)), to ) S (v ˜ v )  Su , thus:

D ª1 & & º
« S (v ˜ v )  Su dV »
³
Dt «¬ 2 V 6 »¼
§ D S 2 (v ˜ v )  Su
1 & &

 S 12 (v ˜ v )  Su ’ x& ˜ v ¸¸dV  >> S 12 (v ˜ v )  Su … v  Z @@ ˜ nˆ dS
& & & & & &
³
V 6 ©
¨
¨ Dt ¹ 6
³
(5.126)
Then by mathematically manipulating the terms of the volume integral we can see that:
D S 12 (v ˜ v )  Su D 1 (v ˜ v )
& & & &
& & DS
 S 12 (v ˜ v )  Su ’ x& ˜ v
& & & 1
2
(v ˜ v ) S 2 
Dt Dt Dt
(5.127)
DS Du
 S 12 (v ˜ v )  Su ’ x& ˜ v
& & &
u S
Dt Dt
Moreover, by reorganizing the above equation, we find that:
D S 12 (v ˜ v )  Su D 12 (v ˜ v )
& & & &
 S (v ˜ v )  Su ’ x& ˜ v
& & &
1
2
S 
Dt Dt
DS 1 & & Du 1 & &
 (v ˜ v )  u  S  2 (v ˜ v )  u S’ x& ˜ v &
(5.128)
Dt 2 Dt
D 1 (v ˜ v )
& &
12 (v& ˜ v& )  u §¨ DS  S’ x& ˜ v& ·¸  S Du  S 2
© Dt

¹
Dt Dt
0

Thus,
D S 12 (v ˜ v )  Su
& &
Du
 S 12 (v ˜ v )  Su ’ x& ˜ v
& & & & &
S  Sv ˜ v (5.129)
Dt Dt
Then if we return to the equation in (5.126), and if we refer to (5.129) we can conclude
that:
D ª1 & & º
« S (v ˜ v )  Su dV »
³
Dt ¬« 2 V 6 ¼»
(5.130)
 v ˜ v ¸dV  >> S 12 (v ˜ v )  Su … v  Z @@˜ nˆ dS
§ Du & & · & & & &
³
V 6

© Dt ¹ 6
³
For the surface integrals on the right side of the equation in (5.125) we can apply Gauss’
theorem to a volume with discontinuity, (see equation (5.17)):

³ ’ ˜ v ˜ ı  q dV  ³ >>v ˜ ı  q@@˜ nˆ dS
& & &
³ v ˜ ı  q ˜ nˆ dS
& & &
&
x (5.131)
S  S  V  V  6

Then if we know that vi V ij , j


& & &
V ij , j vi  V ij v i , j Ÿ ’ x& ˜ v ˜ ı v ˜ ’ x& ˜ ı  ı : ’ x& v , and if
we have observed the spatial velocity gradient has been broken down into a symmetric and
&
an antisymmetric part, we obtain ı : ’ x& v ı : D  ı : W ı : D , so, we can conclude that:

˜ ı  ı : ’ x& v  ’ x& ˜ q dV  ³ >>v ˜ ı  q@@˜ nˆ dS


& & &
³ v ˜ ı  q ˜ nˆ dS ³ v ˜ ’
& & & &
&
x
    6
S S V V
(5.132)
³
&
v& ˜ ’ x& ˜ ı  ı : D  ’ x& ˜ q dV  ³ >>v& ˜ ı  q& @@˜ nˆ dS
V  V  6
318 NOTES ON CONTINUUM MECHANICS

Then by substituting the equations (5.130) and (5.132) into the energy expression in (5.125)
we obtain:

 v ˜ v ¸dV  >> S 12 (v ˜ v )  Su … v  Z @@ ˜ nˆ dS
§ Du & & · & & & &
³ S¨
© Dt ¹
³
V 6 6
& (5.133)
v& ˜ ’ x& ˜ ı  ı : D  ’ x& ˜ q dV  >>v& ˜ ı  q@@˜ nˆ dS  Sb ˜ v&dV  S rdV
& &
³
V 6
³
6
³
V
³
V

or

³ S¨
§ Du & & ·
© Dt ¹
&
& & &
 v ˜ v ¸   v ˜ ’ x& ˜ ı  ı : D  ’ x& ˜ q  Sb ˜ v  S r dV
V 6
(5.134)
 >> S (v ˜ v )  Su … v  Z  v ˜ ı  q@@˜ nˆ dS
& & & & & &
³
6
1
2
0

³
§ Du ·

© Dt ¹
&
¸   ı : D  ’ x& ˜ q  S r  v ˜
& &



&
&
Sv  ’ x& ˜ ı  Sb dV


V 6
0 (5.135)
 >> S 12 (v ˜ v )  Su … v  Z  v ˜ ı  q@@ ˜ nˆ dS
& & & & & &
³
6
0

with which we can conclude that:

˜ q  S rdV  ³ >> S 12 (v ˜ v )  Su … v  Z  v ˜ ı  q@@˜ nˆ dS


& & & & & & &
V 6
³ Su  ı : D  ’ &
x
6
0 (5.136)

which thereby results in the energy equation for volumes with discontinuity:
&
Su ı : D  ’ x& ˜ q  S r in V The energy equation with
(5.137)
>> S 12 (v& ˜ v&)  Su … v&  Z&  v& ˜ ı  q& @@˜ nˆ 0 on 6 discontinuity

5.10 The Principle of Irreversibility. Entropy


Inequality

5.10.1 The Second Law of Thermodynamics


Before applying the second law of thermodynamics, we define entropy which is a state
function. In thermodynamics, entropy is the physical quantity that measures the energy that
can not be used to produce work. In a broader sense, entropy is interpreted as the
measurement of system disorder. The entropy unit is J / K , joules per Kelvin and a process
characterized by constant entropy is called the isentropic process.
The second law of thermodynamics imposes restrictions on the possible direction of the
thermodynamics process. For example, the first law of thermodynamics does not establish
the direction of the heat flux.
The second law of thermodynamics states that “the rate of change of the total entropy H is never
&
less than the sum of the entropy flow s that enters through the surface of the continuum plus the entropy
created inside the continuum B ”.
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 319

The total entropy of the system ( H ) is given by:


& & ªJ º
H (t )
V
³ S I ( x, t ) dV ³ S V0
0 I ( X , t ) dV0 «K »
¬ ¼
(5.138)

& J
where I ( x , t ) is the specific entropy (per unit mass), >I @ .
kgK
The entropy supplied to the system ( B ) is given by:
& & ª J º
B ³ Sb ( x, t )dV ³ S
V V0
0 b ( X , t ) dV0 « sK »
¬ ¼
(5.139)

J
where b is the source of local entropy per unit mass per unit time >b @ .
kg s K
Then the entropy flux that enters the system through the material surface is defined by:
&
 ³ s ˜ n̂ dS ª J º
« sK » (5.140)
S s& ¬ ¼
Thus, we can set the entropy inequality as:
D & & &
*(t ) S I ( x , t ) dV t Sb ( x , t )dV  s ˜ nˆ dS
³ ³ ³
Dt V V S& s
(5.141)
& & &
³ S I ( x , t ) dV t Sb ( x , t )dV  ³ ³ s ˜ nˆ dS
V V S s&

Then by applying the divergence theorem to the surface integral, we obtain:

The second law of thermodynamics


*(t ) ³ S I dV t ³ SbdV  ³ ’ &
x ˜ s& dV (5.142)
V V V
(Entropy inequality)

NOTE: The global form of the entropy inequality in (5.142) implies that: if entropy occurs
then the process is irreversible, that is, we can not return to the original system without
adding work to the system. And, the equality of (5.142) represents a reversible process. Ŷ
The local form of the equation in (5.142) is given by:
& &
SI ( x, t ) t Sb  ’ x& ˜ s (5.143)
and if we consider that:
&
& q & (1) r
s s ; b  b (1) (5.144)
T T
& &
where T ( x , t ) t 0 is the absolute temperature, >T @ K , and by assuming that s (1) and b (1)
are equal to zero, the entropy inequality in (5.143) becomes:
&
r §q· r 1 & 1 &
SI t S  ’ x& ˜ ¨¨ ¸¸ S  ’ x& ˜ q  2 q ˜ ’ x& T (5.145)
T ©T ¹ T T T
Thus,
320 NOTES ON CONTINUUM MECHANICS

& &
& r ( x, t ) §q·
SI ( x , t )  S  ’ x& ˜ ¨¨ ¸¸ t 0
T ©T ¹ Entropy inequality
& (5.146)
& r ( x, t ) 1 & & 1 & (current configuration)
SI ( x , t )  S  ’ x ˜ q  2 q ˜ ’ x& T t 0
T T T
We can also express the entropy inequality given in (5.143) in the reference configuration
as:
& & & &
S 0 I ( X , t ) t S 0b ( X , t )  ’ X& ˜ S ( X , t ) (5.147)
&
where S is the entropy flux vector in Lagrangian description. For thermal processes, the
entropy flux vector and entropy source can be established, respectively, as:
& &
& & q0 & & r( X , t)
S ( X , t)  S1 ; b ( X , t)  b1 (5.148)
T T
&
Then if we take into account the equation in (5.148) where S 1 and b1 are equal to zero,
the equation in (5.147) becomes:
& &
& r( X , t) §q ·
S 0 I ( X , t ) t S 0  ’ X& ˜ ¨¨ 0 ¸¸
T © T ¹ Entropy inequality
& (5.149)
& r( X , t) 1 & & 1 & (reference configuration)
S 0 I ( X , t ) t S 0  ’ X ˜ q 0  2 q 0 ˜ ’ X& T
T T T
& &
Then if we refer to Eq. (5.120), where we obtained q 0 J q ˜ F T , or in indicial notation
& &
q0 i J q k Fik1 , it is true that:

& & wT & wT wx p & wT & wT & wT


q 0 ˜ ’ X& T q 0 i J q k Fik1 J q k Fik1 F pi J q k E pk J qk
wX wx p wX i wx p wx p wx k

i

Material Spatial

Then, we can prove the following relationship is valid:


& & & & & &
q 0 ( X , t ) ˜ ’ X& T ( X , t ) J q( x , t ) ˜ ’ x& T ( x , t ) (5.150)

5.10.2 The Clausius-Duhem Inequality


If we combine the entropy inequality in (5.146) with the energy equation given in (5.114),
& &
S u ı : D  ’ x& ˜ q  Sr Ÿ S u  ı : D Sr  ’ x& ˜ q , we obtain:
r 1 & & 1 & 1 & 1 &
SI  S  ’ x ˜ q  2 q ˜ ’ x& T SI  S r  ’ x& ˜ q  2 q ˜ ’ x& T t 0
T T T T T
(5.151)
1 1 &
Ÿ SI  S u  ı : D  2 q ˜ ’ x& T t 0
T T
In this scenario, the entropy inequality is called the Clausius-Duhem inequality, and is given by:
& 1 1 1 & The Clausius-Duhem inequality
S I ( x , t )  ı : D  S u  2 q ˜ ’ x& T t 0 (5.152)
T T T (current configuration)
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 321

We can also express the Clausius-Duhem inequality in the reference configuration. From
& &
the equation in (5.124) we obtained S 0 r ( X , t )  ’ X& ˜ q 0 S 0 u  S : E and by substituting
this into the entropy inequality given in (5.149) we obtain:
&
S 0 I ( X , t ) t
1
T
& &
T
1 &
S 0 r ( X , t )  ’ X& ˜ q 0  2 q 0 ˜ ’ X& T
& (5.153)
1
1 &
S 0 I ( X , t ) t S 0 u  S : E  2 q 0 ˜ ’ X& T
T T
or:
& 1 1 1 &
S 0 I ( X , t )  S : E  S 0 u  2 q 0 ˜ ’ X& T t 0
T T T
The Clausius-Duhem inequality
or (5.154)
& 1 1 1 & (reference configuration)
S 0 I ( X , t )  P : F  S 0 u  2 q 0 ˜ ’ X& T t 0
T T T

5.10.3 The Clausius-Planck Inequality


&
Note that the inequality q ˜ ’ x& T d 0 is always valid, since the orientation of the heat flux
&
vector ( q ) is always opposite to the temperature gradient ( ’ x& T ), (see Figure 5.10). Then,
we can formulate the heat conduction inequality:
&
q ˜ ’ x& T t 0 (current configuration)
Heat conduction inequality & (5.155)
q 0 ˜ ’ X& T t 0 (reference configuration)

If we now incorporate the restrictions in (5.155) into the Clausius-Duhem inequality


(5.152) and in (5.154) we will have a less restrictive inequality known as the Clausius-Planck
inequality:
& 1 1 &
Dint SI ( x, t )  ı : D  S u ( x, t ) t 0 (current configuration)
Clausius-Planck T T
(5.156)
inequality & &
Dint S 0 I ( X , t )  P : F  S 0 u ( X , t ) t 0 (reference
1 1
T T configuration)
where Dint is the internal energy dissipation, which requires positiveness at any time,
D int t 0 .

5.10.4 The Alternative Form to Express the Clausius-Duhem


Inequality
An alternative form of entropy inequality is that expressed in terms of the Helmholtz free
energy, Z , which is a thermodynamic potential per unit mass and is given in Eulerian description
by:
ªJ º
Z u  TI The Helmholtz free energy « » (5.157)
¬ kg ¼
322 NOTES ON CONTINUUM MECHANICS

NOTE: A thermodynamic potential indicates the amount of energy available in the


system. In this chapter we only work with the potentials u u ( E , I ) and Z ( E , T ) . We can
also use another potential, e.g. Gibbs free energy ( G(S, T ) ), or Enthalpy ( H(S, I ) ), (see Chapter
10). The choice to adopt one or the other depends on the independent variables under
consideration, ( E -“volume”, I -entropy, S -“pressure”, T -temperature). For more details
about these potentials see the chapter on Thermoelasticity. Ŷ

&
q & & &
­°qnˆ (q ˜ nˆ )nˆ qnˆ nˆ
®
°̄’ x& T ’ x& T nˆ
’ x& T & &
q ˜ ’ x& T ’ x& T q ˜ nˆ

T1
&
qnˆ T1 ! T2 ! T3
T2

T3 & & & &


ܳ q ˜ ’ x& T (qnˆ q sˆ ) ˜ ’ x& T qnˆ ˜’ x& T

Figure 5.10: Temperature gradient and heat flux vector.

If we calculate the rate of change of the Helmholtz free energy, we obtain:


TSI Su  SIT  SZ
Z u  IT  TI Ÿ TI 
u  IT  Z Ÿ
>
Su  S IT  Z @ (5.158)

Then if we consider that T ! 0 (absolute temperature) and the entropy inequality given in
(5.145), we obtain:
r 1 & & 1 & & 1&
SI t S  ’ x ˜ q  2 q ˜ ’ x& T Ÿ STI t S r  ’ x& ˜ q  q ˜ ’ x& T (5.159)
T T T T
Afterwards by combining the above inequality with the equation in (5.158) we obtain:

> @ & 1&


Su  S IT  Z t S r  ’ x& ˜ q  q ˜ ’ x& T
T
(5.160)
&
Then by also considering the energy equation in (5.114), i.e. S u ı : D  ’ x& ˜ q  Sr , we
obtain:
&
>
ı : D  ’ x& ˜ q  S r  S IT  Z x@
 t S r  ’ & ˜ q&  1 q& ˜ ’ & T
T
x
(5.161)
>
Ÿ ı : D  S IT  Z
T
@
 t 1 q& ˜ ’ & T
x

by which we obtain the Clausius-Duhem inequality (current configuration) in terms of the


Helmholtz free energy:

>
ı : D  S I7  Z
T
@
  1 q& ˜ ’ & T t 0
x
Clausius-Duhem inequality
(current configuration)
(5.162)
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 323

The Clausius-Duhem inequality in the reference configuration, (see Eq. (5.154)),


1& 1&
S 0TI  S : E  S 0 u  q 0 ˜ ’ X& T t 0 Ÿ S : E  S 0 >u  TI @  q 0 ˜ ’ X& T t 0 (5.163)
T T
can also be written in terms of the Helmholtz free energy Z . To do this let us consider the
& & &
Helmholtz free energy in Lagrangian description Z u ( X , t )  T ( X , t )I ( X , t ) . Additionally,
the rate of change is given by Z u  TI  TI Ÿ u  TI Z  TI with that the Clausius-
Duhem inequality in the reference configuration becomes:

> T
0 X
@
  TI  1 q& ˜ ’ & T t 0
S : E  S 0 Z Clausius-Duhem inequality
(5.164)
P : F  S 0 Z > T
0 X
@
  TI  1 q& ˜ ’ & T t 0 (reference configuration)

&
The Helmholtz free energy per unit reference volume is denoted by : S 0 ( X )Z , and it holds
that : S 0 Z . Proof of this can be shown by:
D: D (S 0 Z ) D(S 0 ) D (Z )
: { Z  S0 S 0 Z .
Dt Dt Dt

Dt
0

5.10.5 The Alternative Form of the Clausius-Planck


Inequality
The Clausius-Planck inequality can also be expressed in terms of Helmholtz free energy.
&
Then, if we consider the heat conduction inequality,  q ˜ ’ x& T t 0 , the equation in (5.162)
becomes:

>
Dint ı : D  S IT  Z t 0 @ (5.165)
which in the reference configuration is given by:

>
Dint S : E  S 0 IT  Z t 0 @ (5.166)

5.10.6 Reversible Process


A thermodynamic process is said to be reversible if there is no dissipation of energy, i.e.
*(t ) 0 , (see equation (5.142)). A reversible process is characterized by:
ƒ The work done by the forces between two points being independent of the path;
ƒ The work done in a closed cycle being zero.
If we take into account that the dissipation of energy is equal to zero in a reversible process
we obtain:

Dint ı : D  SZ 0 Ÿ SZ ı : D (5.167)


Then the equation in (5.167) in the reference configuration becomes:
Dint S : E  S 0 Z 0 1
1 Ÿ S 0 Z S : C S : E (5.168)

S : C  S 0 Z 0 2
2
324 NOTES ON CONTINUUM MECHANICS

5.10.7 Entropy Inequality for a Domain with Discontinuity


If we applying the entropy inequality in (5.141) for a volume with discontinuity we obtain:
D & & &
*(t ) ³
S I ( x , t ) dV t Sb ( x , t )dV  s ˜ nˆ dS
³ ³ (5.169)
Dt V  V  V  V  S  S 

For the surface integral on the right side of the inequality in (5.169) we can apply the
divergence theorem with discontinuity given in Eq. (5.17), the result of which is:
& & &

³ s ˜ nˆ dS
 
³ ’ ˜ s dV  ³ >>s @@˜ nˆ dS
 6
(5.170)
S S V V

For the volume integral on the left side of the inequality in (5.169) we can apply the
Reynolds’ transport theorem given in (5.22) in which ) S I , then:

D ª º § D S I &· & &


³
« S I dV » ³
¨  S I ’ x& ˜ v ¸dV  >>S I … v  Z @@ ˜ nˆ dS
³ (5.171)
Dt «¬V  6 »¼ V 6 ©
Dt ¹ 6

Then by substituting the equations in (5.171) and (5.170) into (5.169), we obtain:
§ D S I &· & & &
³¨  S I ’ ˜ v ¸ dV  >>S I … v  Z @@ ˜ nˆ dS t
³ Sb ( x , t )dV  ³
V 6 ©
Dt ¹ V  V 
6 (5.172)
& &
 ’ x& ˜ s dV  >>s @@ ˜ nˆ dS³ ³
V 6 6

Additionally by regrouping the integrands we obtain:


§ DI DS & & · & & &
³ ¨S I  S I ’ x& ˜ v  ’ x& ˜ s  Sb ¸dV  >>S I … v  Z  s @@ ˜ nˆ dS t 0
³
V 6 © Dt Dt ¹ 6

(5.173)
DS & § DS &·
Note that the equation I  S I ’ x& ˜ v I¨  S’ x& ˜ v ¸ 0 is valid due to the mass
Dt © Dt ¹
continuity equation. Then, the equation in (5.173) becomes:
§ DI & · & & &
³ ¨S  ’ x& ˜ s  Sb ¸dV  >>S I … v  Z  s @@ ˜ nˆ dS t 0
³ (5.174)
V 6 © Dt ¹ 6

The local form of the above equation is expressed as:


DI &
S t ’ x& ˜ s  Sb in V  6
Dt Entropy inequality with
(5.175)
>>S I … v&  Z&  s& @@ ˜ nˆ t 0 on 6 discontinuity

Problem 5.11: 1) Consider a continuum motion in which the stress power is equal to zero.
&
Also, consider that the heat flux is given by q K (T ) ˜ ’ x& T , which is known as Fourier’s
law of thermal conduction, where K (T ) is a second-order tensor called the thermal conductivity
wu (T )
tensor (the thermal property of the material), and c , where c is the specific heat
wT
capacity at a constant deformation (the thermal property of the material) and is expressed in
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 325

J
units of joule per kelvin, i.e. >c@
. Taking into account all previous considerations, find
K
the energy equation for this process. Then also provide the unit of K (T ) in the
International System of Units (SI).
2) Consider the stress power is equal to zero, and that there is a continuous medium with
no internal heat source. Also consider that there is a heterogeneous material where
&
K K ( x ) is an arbitrary second-order tensor (not necessarily symmetrical). a) Show that
the thermal conductivity tensor is semi-definite positive, b) Check in which scenario the
&
skew part of K ( x ) does not affect the outcome of the heat conduction problem. c) Taking
into account that the material is isotropic, in what format is K ?
Solution: For this problem we know that the stress power is equal to zero, ı : D 0 . It then
follows that, the energy equation becomes:
wu wT & &
S u S ı
, : D  ’ x& ˜ q  Sr ’ x& ˜ q  Sr
wT wt 0
wT & wT
Ÿ Sc ’ x& ˜ q  Sr Ÿ Sc ’ x& ˜ > K (T ) ˜ ’ x& T @  Sr
wt wt
or
wT
’ x& ˜ >K (T ) ˜ ’ x& T @  Sr Sc
wt
The above equation is called the heat flux equation which is applied to the thermal
conduction problem.
wT º
Then if we take into account the following units: >q@
& J W ª & K
, ’ xT { & » , we
m2s m 2 «¬ wx ¼ m
can ensure that the units are consistent if the following is met:
>q& @ >K @ ˜ >’ x& T @
ª J W º ª J W ºª K º
« m2s m 2 »¼ « s m K m K »« m »
¬ ¬ ¼¬ ¼
ª J W º
thus, we can draw the conclusion that >K @ «s m K m K » .
¬ ¼
NOTE: As we will see later, when the stress power is equal to zero, we can decouple the
thermal and mechanical problem. That is, we can study these problems separately. Ŷ
2) a) We start from the heat conductivity inequality:
& &
 q ˜ ’ x& T (K ( x ) ˜ ’ x& T ) ˜ ’ x& T t 0  q i T,i ( K ij T, j )T,i t 0
& or
’ x& T ˜ K ( x ) ˜ ’ x& T t 0 T,i K ij T, j t 0
& &
Remember that the arbitrary tensor A is semi-definite positive if it holds that x ˜ A ˜ x t 0
& & &
for all x z 0 thereby demonstrating that K ( x ) is a semi-definite positive tensor. Then, as a
&
result the eigenvalues of K ( x ) are all real values greater than or equal to zero, i.e. K 1 t 0 ,
&
K 2 t 0 , K 3 t 0 . Also remember that since K ( x ) is not symmetric, the principal space of
&
K ( x ) does not define an orthonormal basis. Moreover, it is noteworthy that: the
&
antisymmetric part of K ( x ) does not affect the heat conduction inequality since:
&
’ x& T ˜ K ( x ) ˜ ’ x& T >
’ x& T ˜ K sym  K skew @ ˜’ xT
& t0
’ x& T ˜ K sym
˜ ’ x& T  ’ x& T ˜ K skew
˜ ’ x& T t 0
’ x& T ˜ K sym ˜ ’ x& T  K skew : (’ x& T … ’ x& T ) t 0
326 NOTES ON CONTINUUM MECHANICS

Notice that K skew : (’ x& T … ’ x& T ) 0 , since the double scalar product between an
antisymmetric tensor ( K skew ) and a symmetric one (’ x& T … ’ x& T ) is equal to zero, then:
&
0 d ’ x& T ˜ K ( x ) ˜ ’ x& T
’ x& T ˜ K sym ˜ ’ x& T t 0
&
That is, the above inequality is always true whether K ( x ) is symmetric or not.
b) For the proposed problem the only remaining governing equation is the energy
Du & &
equation: S { Su ı : D  ’ x& ˜ q  Sr ’ x& ˜ q , where u is the specific internal
Dt
energy, ı : D is the stress power, and Sr is the internal heat source per unit volume. Then:
Su q i ,i (K ij T, j ) ,i K ij ,i T, j  K ij T, ji (’ x& ˜ K T ) ˜ (’ x& T )  K : ’ x& (’ x& T )
> @
(’ x& ˜ K T ) ˜ (’ x& T )  K sym  K skew : ’ x& (’ x& T )
(’ x& ˜ K ) ˜ (’ x& T )  K
T sym
: ’ x& (’ x& T )  K skew : ’ x& (’ x& T )
(’ x& ˜ K T ) ˜ (’ x& T )  K sym : ’ x& (’ x& T )
where we have considered the symmetry of >’ x& (’ x& T )@ij T, ji . If the material is
T,ij
&
homogeneous the implication is that the K field does not depend on ( x ) , so K ij ,i 0 j . In
this scenario the heat equation reduces to:
Su K sym : ’ x& (’ x& T )
Therefore, when the material is homogeneous, the antisymmetric part of K does not affect
the outcome.
c) The feature of isotropic materials is that their properties (at one material point) do not
change if the coordinate system is changed. It follows then that K must be an isotropic
tensor. An isotropic second-order tensor has the format of a spherical tensor, (see Chapter
1), then the tensor K must be of the type: K K1 , where K is a scalar.

5.11 Fundamental Equations of Continuum


Mechanics

Then, we sum up the fundamental equations of continuum mechanics in the current


configuration as:
Fundamental Equations of the Continuum Mechanics
(Current configuration)
Mass Continuity Equation DS &
 S (’ x& ˜ v ) 0 (1 equation) (5.176)
(Principle of conservation of mass) Dt
Equation of Motion & &
(Principle of conservation of linear ’ x& ˜ ı  Sb Sv (3 equations) (5.177)
momentum)
Symmetry of the Cauchy Stress Tensor
(Principle of conservation of angular ı ıT (6 unknowns) (5.178)
momentum)
Energy Equation &
S u ı : D  ’ x& ˜ q  Sr (1 equation) (5.179)
(Principle of conservation of energy)
Entropy Inequality & 1 1 1 &
SI ( x , t )  ı : D  S u  2 q ˜ ’ x& T t 0 (5.180)
(Principle of irreversibility) T T T
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 327

The entropy inequality is not one more problem equation. Rather, it is used to establish
restrictions on the problem variables. The symmetry of the Cauchy stress tensor reduces
the number of ı -unknown from 9 to 6.
The mass continuity equation, the equations of motion and the energy equation give us in
&
total 5 equations. The unknowns are: the three components of velocity v , temperature T ,
mass density S , six components of the Cauchy stress tensor ı , the specific internal energy
&
u , three components of the heat flux vector q , and the entropy I , with a total of 16
unknowns.
To achieve the well-posedness of the problem eleven equations must be added. We must
add equations that connect the stress, heat and energy with other fields. These equations
are called constitutive equations, which is the subject of the next chapter.

5.11.1 Particular Cases


5.11.1.1 Rigid Body Motion

When we are dealing with rigid body motion without the effect of temperature, the only
principles needed to establish the set of equations are: the principle of conservation of the
linear momentum and the principle of conservation of angular momentum. Then, the
& & & &
governing equations are characterized by ¦ F m a and ¦ M G H G , (see Problem 5.9).
The problem can then be solved by introducing the appropriate initial and boundary
conditions.

5.11.1.2 Flux Problems

For problems which only involve the transport of a physical quantity (mass, energy, or
otherwise) the only principle necessary to establish the governing equations is the
conservation law of the physical quantity, or in its strong form: the physical quantity
continuity equation, (see equation (5.24)):
& &
Q
w) ( x , t ) & w) ( x, t )
 ’ x& ˜ ()v ) {  ’ x&
& &
˜ q( x, t ) >) @ (5.181)
wt wt s
The case in Problem 5.11 was related to energy transport (not mass transport). Said energy
transport exists due to the agitation of atoms, in which the degree of agitation at the
macroscopic level is characterized by the temperature. If a particle starts to increase the
degree of agitation, then neighboring particles also start behaving in a similar fashion. In
this way the energy in solids is transported without any mass transport. This energy
& &
transport at the macroscopic level is represented by means of the flux, q )v .
When we are working in the field of continuum mechanics we do not go down to the
atomic level and measure the average velocity (vibration) of a handful of atoms to establish
the flux. What we do is: we go to the laboratory with the material with which we want to
establish the heat flux (energy flow), we vary the temperature and we verify
macroscopically that the flux can be characterized by the following phenomenological law
q k’T (in a one-dimensional case), where k is a thermal property of the material. This
procedure was performed by Fourier, thereby establishing Fourier’s law of heat conduction.
Fourier also verified in the laboratory that heat flux is opposite to the temperature gradient,
a fact already proven by the second law of thermodynamics. The law q k’T is a
phenomenological law or constitutive equation of heat flux, and connects two thermal
328 NOTES ON CONTINUUM MECHANICS

variables. It is also interesting to observe that the fundamental equations of continuum


mechanics (5.176)-(5.180) do not have such a relationship.
In Problem 5.11, the physical quantity in question is given by ) ScT . According to the
kg J
SI units we have: >T @ K , >S @ , >c @ with which we can verify the following SI
m3 kg K
units:
kg J J
>) @ >ScT @ K (unit of energy per unit volume - energy density)
m 3 kg K m3

>q& @ { >)v& @ J m J
(unit of energy flux)
m3 s m2s
There are several engineering problems which are characterized by the continuity equation,
some of which are: heat conduction problems (energy flux); filtration problems in porous
media (mass transport); diffusion problems (e.g. transport of contaminant in an aqueous
medium); and the Saint-Venant torsion problem (stress flux).

5.12 Flux Problems

5.12.1 Heat Transfer


Heat flow is a form of energy transfer in a continuous medium which occurs in three ways,
namely via: conduction; convection; radiation.

5.12.1.1 Thermal Conduction

Thermal Conduction: Transfer of energy in the form of heat, which is caused by the collision
and vibration of molecules and atoms (no mass transport).
&
Temperature: The temperature ( T ( x , t ) ! 0 ) is not a form of energy. Rather it is a
measurement of how hot a particle is. In experiments it has been proven that the hot
particles tend to give heat to cooler particles. The SI unit of absolute temperature is the
kelvin, >T @ K . Absolute zero T 0 K | 273,15º C is a theoretical temperature when even
atoms and electrons cease to move.
When a continuous medium undergoes a non-uniform temperature variation, heat is
transferred from a higher to a lower temperature region. When this phenomenon occurs
without mass transport, this is known as a heat conduction problem. The phenomenological law
(constitutive equation of heat flux) that governs heat conduction behavior can be defined by
means of Fourier’s law of heat conduction, which states that the heat flux is proportional to the
temperature gradient:
& wT Fourier’s law of heat ª J º
q K ˜ & K ˜ ’ x& T «m2s » (5.182)
wx conduction ¬ ¼
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 329

where q is the heat flux per unit area per unit time, and its SI unit >q@
& & J W
; ’ x& T is
m2s m2
K
the temperature gradient whose SI unit is >’ x& T @ , and K is the thermal conductivity
m
W
tensor, whose SI unit is >K @ , (see Problem 5.11).
mK
NOTE: Fourier’s law of heat conduction is not universal as there are complex materials in
which heat flow is governed by more complex laws. ɶ
The negative sign in Fourier’s law is there because the heat flux vector is always opposite to
the temperature gradient. The temperature gradient vector ( ’ x& T ) points from the coldest
&
to the warmest region, while the heat flux vector ( q ) points from the warmest to the
coldest region (physical fact), (see Figure 5.10).

conduction
conduction

ܳ
Figure 5.11: Heat conduction.
The thermal conductivity tensor contains the thermal properties of the material, which are
obtained in the laboratory, and depends on porosity, mass density, composition, etc.
Explicitly, the components of K are:
ªK 11 K 12 K 13 º ª1 0 0º
(K ) ij «K K 23 »» isotropic material
   o K ««0 1 0»»
« 21 K 22 (K ) ij (5.183)
«¬K 31 K 32 K 33 »¼ «¬0 0 1 »¼

For isotropic materials, i.e. those that have the same property in any direction, the thermal
conductivity tensor is represented by a spherical tensor, (see Problem 5.11).
&
If we are dealing with homogenous material, K is not dependent on x . For isotropic
&
materials, the components of the heat flux vector ( q K ˜ ’ x& T ) are obtained as follows:

ª wT º ª wT º
« » « »
« wx1 » « wx1 »
­ q1 ½ ª1 0 0º « » « »
& ° °
 « » « wT » K «
wT »
(q) i q
® 2¾ K «0 1 0» « wx » (5.184)
°q ° « wx 2 »
¯ 3¿ «¬0 0 1»¼ « 2 » « »
« wT » « wT »
« » « »
¬ wx3 ¼ ¬ wx3 ¼
330 NOTES ON CONTINUUM MECHANICS

&
and the normal component q n q ˜ nˆ , (see Figure 5.2), is evaluated as:
wT ˆ wT ˆ wT ˆ
qn q i nˆ i q1nˆ 1  q 2 nˆ 2  q 3nˆ 3 K n1  K n2  K n3 (5.185)
wx1 wx 2 wx3

5.12.1.2 Thermal Convection Transfer

Heat transfer by convection occurs in a fluid environment where there are moving particles
between regions with different temperatures, (see Figure 5.12). In other words it shows the
transfer of energy (heat) due to the movement of fluid particles. This phenomenon is
governed by Newton’s Law of Cooling, which is:
q B T  Text Newton’s law of cooling (5.186)

where q is the thermal energy; B is the heat transfer coefficient per unit area; T is the
temperature of the body’s surface, and Text is the temperature of the surrounding
environment.

warm air

radiator cold air

Figure 5.12: Thermal convection.

If we consider a room in which there is a radiator, the air particles in contact with the hot
surface of the radiator increases their temperature and their mass density decreases, so that
the hot ascending particles, displace the cooler particles moving downwards, (see Figure
5.12) and because of this movement, the heat will be transferred to the whole room.

5.12.1.3 Thermal Radiation

Thermal radiation is the process by which thermal energy is transferred between two
surfaces, obeying the laws for electromagnetic radiation (photon transport). To give an
example we can mention how heat is transferred from the Sun to the Earth. The
phenomenological law governing this phenomenon is the Stefan-Boltzmann law.

5.12.1.4 The Heat Flux Equation

Next, we can obtain the partial differential equation that governs the heat transfer problem,
by means of an energy balance, i.e.:

Heat that Heat Heat that Change of


+ generated _ = internal
enters into leaves the
the system internally system energy
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 331

Let us consider a differential volume element, (see Figure 5.13), in which there are inflows
and outflows of heat. In addition let us consider energy generated internally represented by
J
Q Sr (per unit volume per unit time), whose SI unit is >Q @ . The scalar function r
m3s
describes the heat generated which could be caused by a phenomenon such as a chemical,
J
or nuclear reaction, and whose SI unit is >r @ . As the temperature of the body
kg s
increases, part of the thermal energy is stored in the body. For a differential volume
element ( dx1 dx 2 dx3 ) this stored energy is governed by the expression:
wT
Sc v dx1 dx 2 dx3 (5.187)
wt
where S is the mass density; and the material property c v is the specific heat capacity at a
J
constant volume whose SI unit is >c v @ .
kg K

wq z wq y
qz  dz qy  dy
wz wy

dy
wq x
qx qx  dx
wx
Q
dz
z y
dx
qy
x qz

Figure 5.13: Source and heat flux in a differential volume element.

In the following demonstration, let us consider the following change of nomenclature:


coordinates: x1 { x , x 2 { y , x 3 { z ; heat flux components; q1 { q x , q 2 { q y , q 3 { q z .
Notice that we are employing the engineering notation.
Then by applying the energy conservation law throughout the differential element, we
obtain:
ª § wq ·
q x dy dz  q y dx dz  q z dx dy  Q dx dy dz  « ¨ q x  x dx ¸dydz
¬ © wx ¹
§ wq y · § wq · º wT
 ¨¨ q y  dy ¸¸ dxdz  ¨ q z  z dz ¸ dxdy » c v Sdxdydz (5.188)
© wy ¹ © wz ¹ ¼ wt
ª wq wq y wq z º wT
ŸQ« x   » Sc v
¬ wx wy w z ¼ wt
which results in the heat equation:
& wT wT
Q  ’ x& ˜ q Sc v Ÿ Q  ’ x& ˜ K ˜ ’ x& T Scv The heat flux equation (5.189)
wt wt
332 NOTES ON CONTINUUM MECHANICS

wT
where we have considered Fourier’s law of heat conduction q i K ij . Notice that the
wx j
above equation was obtained in Problem 5.11 and it should be pointed out that we have
one equation in (5.189) and one unknown (temperature). The solution of equation (5.189)
is unique if we are given the appropriate boundary and initial conditions. The governing
equation in (5.189) together with the boundary and initial conditions are called the Initial
Boundary Value Problem (IBVP) of thermal conduction.
Then by considering an isotropic homogeneous material, the heat equation in (5.189)
becomes:
w § wT · w § wT · w § wT · wT
Q K ¨ ¸  K ¨¨ ¸¸  K ¨ ¸ S cv (5.190)
wx © wx ¹ wy © wy ¹ wz © wz ¹ wt
Q w 2 T w 2 T w 2T 1 wT
Ÿ    (5.191)
K wx 2 wy 2 wz 2 L wt
where L is known as the thermal diffusivity:
K ª m2 º
L « » (5.192)
S cv ¬ s ¼
Particular Cases
&
ƒ A steady state temperature field, i.e. T T ( x ) :
wT
0 (5.193)
wt
The equation in (5.191) becomes:

Q w 2 T w 2 T w 2T Q
   0Ÿ  ’ 2x& T 0 The Poisson’s equation (5.194)
K wx 2 wy 2 wz 2 K

From a mathematical point of view, the above equation is known as the Poisson’s
equation.
ƒ A steady state problem, and without internal heat generation:
wT
0 ; Q 0 (5.195)
wt
In this scenario the equation in (5.191) becomes Laplace’s equation:

w 2 T w 2T w 2 T
  0 Ÿ ’ 2x& T 0 Laplace’s equation (5.196)
wx 2 wy 2 wz 2
&
ƒ Transient problem, T T ( x , t ) (time dependent), but in the absence of internal heat
generation, Q 0 , the equation in (5.191) becomes Fourier’s equation:

w 2 T w 2 T w 2T 1 wT 1 wT
  Ÿ ’ 2x& T Fourier’s equation (5.197)
wx 2 wy 2 wz 2 L wt L wt

Initial and Boundary Conditions


1. Prescribed value of the temperature:
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 333

T ( x, y , z , t ) T * to t !0 on S1 (5.198)
Mathematically this condition is known as Dirichlet boundary condition.
2. Flux boundary condition:
wT ˆ wT ˆ wT ˆ
Q K nx  K ny  K nz 0 to t !0 on S 2 (5.199)
wx wy wz
Mathematically this condition is known as a Neumann boundary condition.
A combination of the boundary conditions of Dirichlet and Neumann is known as the
Robin boundary condition, i.e.:
wT ˆ wT ˆ wT ˆ
Q K nx  K ny  K n z  B (T  Text ) 0 to t!0 on S 3 (5.200)
wx wy wz

where nˆ x , nˆ y , nˆ z are the components of the outward unit normal vector on the surface.
Initial conditions T ( x, y, z , t 0) T0 .

&
q* y ­nˆ x cos D
t z n̂ °
S3 ®nˆ y cos E
°ˆ
J E
¯n z cos J
B D
x
x3 Text
dV
T*
Sr
O S1
x2

x1

Figure 5.14: Heat flux problem and boundary conditions.


334 NOTES ON CONTINUUM MECHANICS

5.13 Fluid Flow in Porous Media (filtration)

Let us consider a reservoir as shown in Figure 5.15. To obtain the partial differential
equation that governs the fluid flow in porous media we will make the same approach as
that made to the heat flux problem, but in this case we will consider the two-dimensional
case and steady state regime.
y

dam (impermeable)

h1
material point - P
h2
wq y
qy  dy
wy
Flux
P
wq x
L qx P qx  dx
dy wx
dx
soil (permeable)
qy
x
rock (impermeable)
Figure 5.15: Fluid flow throughout the porous medium.

The partial differential equation governing the fluid flow in porous media for a steady state
case can be obtained by means of the differential element equilibrium, (see Figure 5.15), i.e.:
ª § wq x · º ª § wq y · º wq wq y
«q x  ¨ q x  dx ¸ » dy  «q y  ¨¨ q y  dy ¸¸ » dx 0 Ÿ  x dxdy  dxdy 0 (5.201)
¬ © wx ¹¼ ¬« © wy ¹ »¼ wx wy

The phenomenological law of mass flux in porous media is governed by Darcy’s law,
&
q K ˜ ’ x& G , where G is the total potential (water level), and K is the permeability tensor
&
which depends on the material. In the two-dimensional case, the components of q are
given by:
wG wG
qx K ; qy K (5.202)
wx wy
where we have considered an isotropic material (one which has the same permeability in all
directions). Then by substituting the components of the flux vector in Equation (5.201), we
obtain:
w § wG · w § wG ·
¨K ¸  ¨K ¸ 0 (5.203)
wx © wx ¹ wy ¨© wy ¸¹
§ w 2 G w 2G ·
K ¨¨ 2  2 ¸¸ 0 Ÿ ’ 2 G 0 Laplace’s equation (5.204)
© wx wy ¹
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 335

The boundary conditions are:


wG wG
ƒ There is no flow at x f and x f : 0 ; 0
wx x f wx x f

wG
ƒ There is no flow at the border (soil-rock interface): 0
wy y 0

wG b b
ƒ There is no flow at the soil-dam interface: ( x, L ) 0 ;  dxd
wy 2 2
Additionally, the total potential is prescribed at the water-soil interface:
G ( x, L ) x
b h1 ; G ( x, L ) x!
b h2 (5.205)
2 2

5.14 The Convection-Diffusion Equation

Diffusion: An irreversible physical process is one where particles which are in a high
concentration region tend to move to a region of low concentration. In general
this process is governed by Fick’s law of diffusion:
& ª mol º
J D ˜ ’ x& c Fick’s law of diffusion «m2s » (5.206)
¬ ¼
&
where D ! 0 is the diffusion tensor (or diffusivity tensor), and c( x , t ) is the solute
mol
concentration whose SI unit is >c @ . This concentration is defined as follows:
m3
solute mass
c (5.207)
fluid mass
or we can express this concentration as:
1 solute mass Ss
c (5.208)
Sf volume Sf
where S f , S s are the mass densities of the fluid and solute, respectively.
In general, when we have a process where there is mass transport (a fluid+solute) two
mechanisms take place, namely: convection and diffusion. In this case the matter (solute) is
defined by the concentration c and we must consider the matter to be diffused throughout
the aqueous medium. Here, we can assume that the amount of the matter is too little to
affect the fluid velocity field.
& &
Let us consider the solute flux q c v (a convective term) to which we add the diffusive
term to obtain the total flux:
& & wc
q c,v D˜ &
Convective w

x (5.209)
term Diffusive
term

Then, to obtain the partial differential equation for the convection-diffusion problem we
consider the one-dimensional case, (see Figure 5.16).
336 NOTES ON CONTINUUM MECHANICS

wq x
qx qx  dx
wx

x x  dx

Figure 5.16: Mass transport (solute).

Here we can put the conservation law down to:

solute that solute solute that Change of


+ generated _ = the solute
enters into leaves the
the system internally system internally

Then, mathematically, the above expression becomes:


§ wq · wc wq x wc
q x dy  Qdxdy  ¨ q x  x dx ¸ dy dxdy Ÿ Q (5.210)
© wx ¹ wt wx wt
Additionally, by substituting the flux given in (5.209) into the above equation, we obtain:
§ wc ·
w¨ c v x  D ¸
© wx ¹ wc w c v x w § wc · wc (5.211)
Q Ÿ Q  ¨D ¸
wx wt wx wx © wx ¹ wt
Therefore we can summarize the convection-diffusion equation in three dimensions as:
& wc Convection-diffusion
Q  ’ x& ˜ (v c)  ’ x& ˜ (D ˜ ’ x& c) (5.212)
wt equation

wc &
where is the local variation of the concentration with respect to time, v ˜ ’ x& c is the
wt
convection term caused by fluid motion, and ’ x& ˜ (D ˜ ’ x& c) is the diffusion.
Next, we assume that at a material point there are two types of material that are
represented by a physical quantity per unit volume in such a way that c c f  c s , and the
& & &
following holds v v f  v s , (see Figure 5.17).

&
vf
cf
material point - P

V
cs
&
P v &
vs

Figure 5.17: Heterogeneous medium.


5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 337

Then, from the continuity equation for this physical quantity we obtain:

Q
w)
wt
 ’ x&
&
˜ )v Ÿ Q
w (c f  c s ) w
wt
> & &
 & (c f  c s )(v f  v s )
wx
@ (5.213)

thus

Q
w (c f  c s ) w
wt
> &
 & (c f  c s )(v f  v s )
wx
&
@
ŸQ
w (c f  c s ) w f & f
wt
> & &
 & c v  c f v s  csv f  csv s
wx
&
@
(5.214)
ŸQ
wc f wc s
wt

wt
&
> & & &
 ’ x& ˜ c f v f  c f v s  c s v f  c s v s @
ŸQ «
ª wc f
w t
& º wc s
 ’ x& ˜ (c f v f )» 
w t
& &
> &
 ’ x& ˜ (c s v f )  ’ x& ˜ c f v s  c s v s @
¬ ¼

wc f &
If we assume that there is no ( f )-material source, then  ’ x& ˜ (c f v f ) 0 holds,
wt
which is the continuity equation of the quantity c f with which the equation in (5.214)
becomes:

Q
wc s
wt
&
> & &
 ’ x& ˜ (c s v f )  ’ x& ˜ c f v s  c s v s @
wc s & & &
ŸQ  ’ x& ˜ (c s v f )  ’ x& ˜ (c s v s )  ’ x& ˜ (c f v s ) (5.215)
wt
wc s & & & &
ŸQ  ’ x& ˜ (c s v f )  ’ x& ˜ (c s v s )  ’ x& c f ˜ v s  c f ’ x& ˜ v s
wt
&
If the physical quantity c f does not change with x , then the gradient of c f becomes
&
’ x& c f 0 . In addition if we consider the medium ( s ) to be incompressible we obtain
&
’ x& ˜ v s 0 . These simplifications indicate that the material ( s ) does not affect the velocity
field of the material ( f ). So, if the amount of the material ( s ) is significant, this approach
is no longer valid. Then, with these approximations we obtain:
wc s & & wc s & &
Q  ’ x& ˜ (c s v f )  ’ x& ˜ (c s v s )  ’ x& ˜ (c s v f )  ’ x& ˜ q ( D ) (5.216)
wt wt
& &
Notice that the term (c s v s ) { q ( D ) represents the flux caused by the ( s )-material
& &
concentration, the diffusive term. The term (c s v f ) { q (C ) is related to mass transport, the
&
convective term. Then, if q (D ) is defined by Fick’s law we refer back to the equation in
(5.212).
338 NOTES ON CONTINUUM MECHANICS

5.14.1 The Generalization of the Flux Problem


Flux problems can be found in many branches of physics or engineering. These problems
are only governed by the continuity equation (sometimes called the transport equation):
wG
Q  ’ x& ˜ D ˜ ’ x& G S c (5.217)
wt

where G ( x1 , x 2 , x3 , t ) is the scalar variable to be solved. Depending on the problem the


variables take the following meanings:

Q  ’ x& ˜ D ˜ ’ x& G 0

Scalar field Flux vector


Phenomenological
Equation D Q &
G q law

Heat Heat flux


Temperature Thermal Fourier’s law
generated vector
Heat flux conductivity &
T tensor Q & q K ˜ ’ x& T
q

piezometric head Darcy’s law


Fluid flow in the (or hydraulic head) permeability water Volume flux
&
porous media tensor source vector q K ˜ ’ x& G
h

Constitutive Ion flux


ion concentration Fick’s law
matrix for vector &
Diffusion Ion source
c diffusion & J D ˜ ’ x& c
coefficient J
Prandtl stress
Saint-Venant function
1 Hooke’s law
torsion 2T
T G

5.15 Initial Boundary Value Problem (IBVP) and


Computational Mechanics

An Initial Boundary Value Problem (IBVP) is defined by the governing equations (a set of
partial differential equations-PDEs) and by boundary and initial conditions. Said conditions are
restrictions on the governing equations. The IBVP solution is the one that is given by the
solution of the equations and which also satisfies the boundary and initial conditions. The
IBVP solution will be unique if the problem is well posed, i.e. given a boundary and initial
conditions, there is only one solution to the problem. Then, the governing equations are
defined by the fundamental and constitutive equations. In subsequent chapters of this
publication we will fundamentally deal with the constitutive models (constitutive
equations), which are used to complete the IBVP.
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS 339

It must be stressed that until the fundamental equations were established, we did not
discuss the type of material that the continuum is made up of. When we begin to specify
this is the time when the concept of the constitutive equation appears. Then, we can
automatically believe that every class of material has its particular constitutive equations.
In order to represent the real behavior of the material, the constitutive equation has to be
calibrated with macroscopic parameters, i.e. these parameters, which are obtained in the
laboratory, represent the material behavior at the macroscopic level. Remember that the scale
of study of continuum mechanics is macroscopic, and then we need to obtain some
representative macroscopic parameters of the phenomena that occur at the microscopic
scale. We can consider this to be the Achilles’ heel of Continuum Mechanics, because in
some cases we are not able to obtain a macroscopic parameter that characterize
phenomena that are taking place at the microscopic level. In fact, the evolution of the
constitutive equation is directly linked to the precision of the instrumentation and new
techniques used in laboratory testing of such materials. So in summary we can state that the
constitutive equations used to characterize a material must be capable of simulating any
phenomena that arise in the material during the loading/unloading/loading process, or at
least the most significant.
As discussed earlier, the constitutive equations complete the set of equations that govern a
particular physical problem. That is, they complement the IBVP. The solution of the
problem can be analytical (the exact solution) or numerical (an approximate solution). In
most cases it is impossible to obtain the analytical solution, so we turn to computational
mechanics to obtain the numerical solution of the physical problem.
Computational Mechanics resolve specific problems by using numerical simulation tools
incorporated in the computer. In general we can state that computational mechanics is not
an independent block, i.e., for its complete implementation it is directly dependent on three
areas: Theoretical Analysis (IBVP establishment), Experimental Analysis (Lab), and Numerical
Analysis (a numerical methodology incorporated into the computer to obtain the numerical
solution for IBVP), (see Figure 5.18). From a very general point of view, we can also
appreciate in Figure 5.18 how the constitutive models are embedded within the field of
Computational Mechanics.
340 NOTES ON CONTINUUM MECHANICS

COMPUTATIONAL MECHANICS

STRUCTURE

Proposition for a
LABORATORY CONSTITUTIVE
MODEL

Initial Boundary Value Problem


Testing proposal
IBVP

If possible

NUMERICAL SOLUTION

Input data

Does the proposed model NO


accurately simulate lab testing?
New testing Option 3:
proposal New numerical
method
YES
Numerical Simulation

Option 1

Is the simulation
realistic?
NO
Option 4:
New IBVP proposal.
Option 2 The Continuum Theory
is not suitable.

Figure 5.18: Role of the constitutive model in Computational Mechanics.

You might also like