Vous êtes sur la page 1sur 254

NPTEL – Mechanical Engineering – Continuum Mechanics

Module-1: Tensor Algebra


Exercises

(1) Prove the following:


(i) (a × b) · (c × d) = (a · c)(b · d) − (a · d)(b · c).
(ii) (a × b) × (c × d) = (d · (a × b))c − (c · (a × b))d.
(iii) (a × b) × (a × c) = (a ⊗ a)(b × c).
(2) Prove the following:
(i) (a ⊗ b)T = (b ⊗ a).
(ii) (a ⊗ b)(c ⊗ d) = (b · c)a ⊗ d.
(iii) (a ⊗ b) : (c ⊗ d) = (a · c)(b · d).
(iv) T (a ⊗ b) = (T a) ⊗ b.
(v) (a ⊗ b)T = a ⊗ (T T b).
(vi) T : (a ⊗ b) = a · (T b).

(3) If S and W are symmetric and skewsymmetric tensors, respectively, and T is an


arbitrary second-order tensor then show that
1
  
S : T = S : TT = S :T + TT ,
2 
T 1 T

W : T = −W : T = W : T −T ,
2
S : W = 0.

(4) Prove the following:


(i) R : (ST ) = (S T R) : T = (RT T ) : S = (T RT ) : S T .
(ii) If S : W = 0 holds for every symmetric tensor S, then show that W is a skewsym-
metric tensor.
(iii) If S : W = 0 holds for every skewsymmetric tensor W , then show that S is a
symmetric tensor.
(5) Let T be a second-order tensor. Then show that Tmi mjk +Tmj imk +Tmk ijm = Tmm ijk .

(6) If W and w are skewsymmetric tensor and its axial vector, respectively, then show
that
(i) cof W = w ⊗ w,
(ii) W (u × v) = (W u) × v + u × (W v), ∀u, v ∈ V.
(iii) cof (I − W ) = I − W + cof W .
(iv) det (I − W ) = 1 + w · w.
(7) If W is skewsymmetric tensor the show that Q = (I − W )−1 (I + W ) is an orthogonal
tensor.

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

(8) If Q is rotation tensor then show that W = (I − Q)(I + Q)−1 is a skewsymmetric


tensor.

(9) If Q is improper orthogonal tensor, i.e., det (Q) = -1, then show that I + Q is not
an invertible tensor.
(10) Let u, v and v be linearly independent vectors, i.e., [u, v, v] 6= 0. Then prove that
(i) u × v, v × w and w × u are also linearly independent,
(ii) (u × v) ⊗ w + (v × w) ⊗ u + (w × u) ⊗ v = [u, v, v]I.
(11) Let R and S be second-order tensors. Then show that

det(R + S) = det R + cof R : S + R : cof S + det S,


cof(R + S) = cof R + cof S + ((tr R)(tr S) − tr(RS)) I − (tr R)S T − (tr S)RT
+(RS)T + (SR)T .

(12) Let T be a second-order tensor. Then show that


2
1h
 i
cof T = T 2 − (tr T )T + (tr T )2 − tr(T 2 ) I .
2
(13) Let {u, v, w} be orthonormal basis to the vector space V. Then find the eigenvalues
and also linearly independent eigenvectors of a second-order tensor S = u ⊗ u.
(14) Let p be a non-zero vector and S be a symmetric second-order tensor. Then show
that p is a eigenvector of S if and only if p ⊗ p commutes with S.
(15) Let Ts and Tss be symmetric and skewsymmetric parts of a second-order tensor T ,
respectively. Then show that the additive decomposition of second-order tensor T =
Ts + Tss is unique.

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-2: Tensor Calculus


Exercises

(1) Let T be a second-order tensor and U be a second-order tensor such that ||U || = 1.
Then show that the directional derivative of G(T ) = cof T along the direction of U ,
   
DG(T )[U ] = ((tr T )(tr U ) − tr(T U )) I + T T − (tr T )I U T + U T − (tr U )I T T .

(2) Let φ1 (T ) = tr(T −1 T −1 ) = T −T : T −1 , φ2 (T ) = (det T )T −1 : T −1 , and φ3 (T ) =


1 
(cof T ) : (cof T ) = (T : T )2 − (T T T ) : (T T T ) . Then show that the Fréchet deriva-
2
tives
∂φ1  3
= −2 T −T ,
∂T
∂φ2  
= T −1 : T −1 cof T − 2(det T )T −T T −1 T −T ,
∂T
∂φ3
= 2(T : T )T − 2T T T T .
∂T

(3) Prove the following:


(i) ∇(φv) = φ∇v + v ⊗ ∇φ.
(ii) ∇ · (φv) = φ(∇ · v) + v · (∇φ).
(iii) ∇ × (u × v) = (∇ · v)u − (∇v)u − (∇ · u)v + (∇u)v.
(iv) ∇(u · v) = (∇u)T v + (∇v)T u.
(v) ∇ · (u ⊗ v) = (∇ · v)u + (∇u)v.
(vi) ∇ · (φT ) = φ(∇ · T ) + T ∇φ.
(vii) ∇(φT ) = φ(∇T ) + T ⊗ ∇φ.
(viii) ∇ · (T T v) = T : ∇v + v · ∇ · T .
(ix) ∇2 (u · v) = u · ∇2 v + v · ∇2 u + 2∇u : ∇v.
(x) ∇ · [(∇u)u] = ∇u : (∇u)T + u · [∇(∇ · u)].
(xi) ∇u : (∇u)T = ∇ · [(∇u)u − (∇ · u)u] + (∇ · u)2 .
(4) Let Ω and Γ be domain in <3 and its boundary. Let v and T be vector and tensor
fields. Let n be unit normal field to Γ. Then show that
Z Z
T
[u ⊗ (∇ · T ) + (∇u)T ] ∂Ω = u ⊗ (T T n) ∂Γ
Ω Γ

(5) Let Γ be a surface bounded by a contour C. Let n be a unit outward normal to Γ.


Let v be a vector field that follow v · n = 0 on Γ. Then using Stokes’ theorem show that
Z Z
(n × v) · dx = (I − n ⊗ n) : ∇v ∂Γ.
C Γ

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

(6) Let Γ be a surface bounded by a closed curve C. Let n be unit normal to Γ. Let u
and v be vector fields. Then show that
I Z
(u ⊗ v)dx = [(∇u)(v × n) + (n · (∇ × v)) u] ∂Γ.
C Γ

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-3: Kinematics
Exercises

(1) Let θ0 be an angle between two line segments, dX1 and dX2 , at a point in the
reference configuration. Let θ be an angle between the corresponding line segments in
deformed configuration. Then show that the angles are related by
2dX1 · EdX2 + cos θ0 |dX1 ||dX2 |
cos θ = √ √
dX1 · CdX1 dX2 · CdX2
where E is the Green strain tensor and C is the right Cauchy-Green strain.
(2) The Lagrangian description of motion of continuum medium is given by
x1 = X1 e−t + X3 (e−t − 1)
x2 = X2 et + X3 (1 − e−t )
x3 = X 3 e t
find out the Eulerian description of the motion; velocity and acceleration in both La-
grangian and Eulerian description.

(3) The motion of continuum is defined by the velocity components v1 = −αx2 , v2 = αx1 ,
and v3 = β where α and β are non-zero constants. Then
(i) show that the motion is isochoric,
(ii) Find out the components of acceleration.

(4) Let J be determinant of deformation gradient F and C be right Cauchy-Green strain


tensor. Then show that
∂J 1
= JC −1
∂C 2
(5) Show that the acceleration field in a rigid motion has the following form:

a(x1 , t) = a(x2 , t) + ω̇ × (x1 − x2 ) + ω × (ω × (x1 − x2 ))

where ω is angular velocity vector which is an axis of skewsymmetric tensor Q̇QT .

(6) Let v, D and W be velocity field, rate of deformation tensor and spin tensor, respec-
tively. Then show that
Dv D
 
∇· = (∇ · v) + D : D − W : W .
Dt Dt
(7) Let W be a spin tensor. Then show that
Z 
v·v
Z 
(2W v + (∇x · v)v) ∂Ω = (v · n)v − n ∂Γ.
Ω Γ 2

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-4: Balance Laws


Exercises

(1) Two symmetric tensors are said to be coaxial if their eigenvectors are same (i.e., if they
have a set of orthonormal eigenvectors common). Show that the second Piola-Kirchhoff
stress S and the right Cauchy-Green strain C are coaxial if and only if the Cauchy stress
τ and left Cauchy-Green strain tensor B.
(2) Let φ and H be scalar and second-order tensor fields. Then prove the following
transport theorems:
!
d I I

φ dx = + φL dx,
dt C C Dt
!
d Z Z
DH
Hn ∂Γ = + (∇ · v)H − H(∇v)T n ∂Γ.
dt Γ Γ Dt

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-5: Constitutive Relations


Exercises

(1) Let p, α and β be constants. Let a, F , L, D and W be acceleration, deforma-


tion gradient, velocity gradient, rate of deformation and spin tensor, respectively. Then
verify whether the following constitutive relations satisfy the principle of material frame-
indifference or not.
(i) τ = −pI.
(ii) τ = α(F + F T ).
(iii) τ = α[∇a + (∇a)T + 2LT L].
(iv) τ̇ = W τ − τ W + α(tr D)I + βD.

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-1: Tensor Algebra


Lecture-1: Introduction to Continuum Mechanics

Continuum mechanics is a branch of mathematical physics and it deals with the deforma-
tion of matter under the action of forces and thermal effects. The treatment is given for
all the forms of matter (solids, liquids, and gasses) in a unified framework. The frame-
work of continuum mechanics is developed by assuming fundamental laws of mechanics
and thermodynamics as axioms.
Structure of matter
Well known fact from elementary physics that the matter is composed of molecules, which
in turn consists of atoms. Further, the atoms are composition of elementary particles (elec-
trons, protons, neutrons, etc.). Thus, the matter is composition of elementary particles
irrespective of the phase of the matter (i.e., solid, liquid or gas state). Consequently, the
material properties of matter are related directly to the molecular structure and also to
the intermolecular forces. These forces not only depends on structure of molecules but
also on intermolecular distances. Although the molecules in the matter always undergo
random vibration/motion, the bulk material exhibits a stable behavior at macroscopic
level. Hence, the study of deformation behavior of matter can be approached fundamen-
tally in two ways: (i) considering the molecules as discrete particles which is known as
statistical mechanics (ii) considering the bulk material as continuous medium which is
known as continuum mechanics.
Statistical mechanics
This approach seems useful in unifying the mechanics as every material in any phase
consists of atoms. In this approach, we need to know the intermolecular potentials to
predict the behavior of matter. It can be noted that the separation distance between
two adjacent atoms (i.e., center to center distance) is approximately in the order of 10−8
cm. Furthermore, there are about 1022 atoms present in a gram of copper. Therefore,
this approach becomes too complicated to predict the desired results in any large scale
problem due to the computational difficulties that arise in evaluating particle interactions.
Apart from the computational complexity, it is also very difficult to know the exact
intermolecular potentials. Also, small error in the intermolecular potentials can lead to
a large error in predicted results. Hence, this approach has limited usage for large scale
problems such as failure of turbine blade, building response under seismic waves and
drag forces on aircraft. Thus, the continuum mechanics approach is preferred over the
statistical mechanics in order to solve large scale problems in physics and engineering.
Continuum assumption
In nature, despite having random vibration/motion of molecules, we observe the definite

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

behavior for the bulk material (i.e., the matter at macroscopic scale). For example there is
definite shape for a wooden block and water in rotating bucket assumes paraboloid shape.
Therefore, the matter can be treated as continuum (i.e., the matter without tiny holes)
at macroscopic scale. In other words, the molecular structure can be disregarded. In fact,
historically, due to stable behavior of matter at macro-scale, the study of mechanics of
solids and fluids came into existence much before the discovery of molecular and atomic
structure of matter.

Continuum mechanics and its objective:


The fluid continuously deform (flow) under action of forces while solids exhibit the finite
deformation. This distinct quality of fluids and solids lead to the development of two
diverse branches of mechanics called solid mechanics and fluid mechanics. Although the
fluids and solids exhibits entirely different behavior, it should be noted that both of
them follow conservation of momenta and energy. Hence, we expect common governing
laws for both fluids and solids. Furthermore, the continuum mechanics brings common
framework based on fundamental laws of mechanics and thermodynamics. Though both
fluids and solids possess molecules as building block, the continuum mechanics framework
is developed without considering the molecular nature of matter. This unified framework
also helps in study of some complex material which exhibit both solid and fluid like
behaviors known as viscoelastic solids and viscoelastic fluids.
Comparison of continuum and statistical mechanics:
We develop the concept of stress in continuum mechanics which is equivalent to inter-
molecular forces in statistical mechanics. We define geometric quantities called strain and
strain rate in continuum mechanics and they are equivalent to intermolecular distance in
statistical mechanics. The constitutive relations in continuum mechanics assume the role
of intermolecular potentials in statistical mechanics.

Need for tensor analysis


As pointed out in previous discussion, we define two new quantities called stress and
strain. These quantities can not be represented either by scalars or vectors. Therefore, a
new mathematical object tensor is introduced to represent these new physical quantities
in continuum mechanics. There are two advantages with the definition of tensor:

• The more general definition of tensor accounts scalars and vectors as special cases.
In other words the tensor unifies the definition of physical quantities.

• The physical quantities defined using concept of tensor can be independent coor-
dinate frame. Hence, tensors are ideally suited for describing basic (fundamental)
laws of continuum physics.

Continuum nature of real numbers


We need numbers not only for counting but also for quantifying any physical quantities
such as density and temperature. We know that the rational numbers are sufficient to
quantify all quantities that are encountered in our practice up to a desired accuracy.
However, it can be noted that there are gaps (tiny holes) in the rational numbers. In
fact, however close may be two distinct rational number, we can always find an irrational
number between them. Although the rational numbers serves all the practical purpose

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

of quantification, there is a difficulty arises in the field of rational numbers while taking
limiting process. Hence, it is essential to fill the gaps with irrational numbers. The ratio-
nal numbers and irrational number put together forms a gap-less number system called
real numbers. Therefore, the real number system can be thought of as one dimensional
continuum since there are no gaps. We take the advantage of the limiting process of real
numbers in continuum mechanics.
Three-dimensional Euclidean space and continuum body:
Cartesian product of three sets of real numbers, i.e. the set of all triplets (x, y, z), where
x, y and z are real numbers, is sufficient to represent every point in our real world. Fur-
thermore, the Cartesian product of three sets of real numbers is called three-dimensional
(3-D) Euclidean space. The 3-D Euclidean space is a continuum as the real number set
is continuum. Consequently, it is convenient to represent continuum bodies as bounded
domain in the 3-D space.
The concept of continuum or continuous media allow us to define piecewise continuous
functions on material domain (continuum body) and also permit us to take limits and
differentiation at a point in order to define the stress. Therefore, this approach enables us
to use powerful calculus of several variables and thereby facilitate the study of nonuniform
distribution of stresses in the material body. The quantitative predictions by continuum
theories closely agree with the wide range of experimental observations. Therefore, this
powerful theory can be utilized in complicated engineering problems.

Limitations of continuum mechanics:


The assumptions of continuum mechanics do not allow us to create new cracks in solids
during the process of deformation. Furthermore, the continuum theory is also not ap-
plicable at high altitudes from the earth as presence of air is in the form of rarefied gas
(i.e., molecules of air sparsely distributed). However, these cases can be modeled using
continuum mechanics in combination with some empirical information.

References

1. M. N. L. Narasimhan, Principles of Continuum Mechanics, 1993, Wiley, New York.

2. E. B. Tadmor, R. E. Miller and R. S. Elliott, Continuum Mechanics and Thermo-


dynamics from Fundamental Concepts to Governing Equations, 2012, Cambridge
University Press, UK.

3. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-1: Tensor Algebra


Lecture-2: Vector and Vector Spaces

We develop the tensor algebra that is required to study the continuum mechanics. In this
lecture, we introduce the concept of vector, vector space, basis, and inner product.
Elementary concept of vector
The vector in elementary physics is defined as the quantity which has magnitude and
direction. The vector is graphically represented by an arrow shown in Fig. (1a) while
the algebraic representation (in 3-D space) is done with three numbers. For example the
vector u is represented as {u1 , u2 , u3 } or u1 e1 + u2 e2 + u3 e3 , q
where e1 , e2 , e3 are unit
vectors along coordinate axes. The length of vector, |u| = u21 + u22 + u23 , represents
the magnitude while direction cosines (cos θ1 , cos θ2 , cos θ3 ) represent the direction of
vector as shown in Fig. (1a). Two vectors can be added and subtracted as shown in
Figs. (1b) and (1c). Two vectors are said to be equal if they have same magnitude and
direction. Examples for the vector quantities are position, velocity, acceleration etc. It
can be observed that the coordinate transformation can alter the components but cannot
affect the vector. For example somebody throws a stone then the velocity vector of stone
does not depend on the choice of coordinate system, i.e., the existence of the vector
is independent of the coordinate frame. However, the components of velocity vector do
depend on the coordinate frame, i.e., the column matrix that represents the velocity vector
does change with coordinate frame. The foregoing discussion on vectors is intuitive from
the elementary physics.

1_
2
|u| = (u12 + u22 + u32 )

cos θ1 = |_
u1
u|
u+ v y u = y-x
u cos θ2 = |u_2
u|
v
θ2
cos θ3 = |_
u3
u|
θ1 u x
θ3

Figure 1: Vector in Euclidean three-dimensional space: (a) vector representation, (b)


vector addition (c) vector subtraction

From the previous discussion, it is clear that the definition of elementary concept of
vector requires unit vectors along coordinate axes. But the definition of unit vectors along
coordinates cannot have an independent definition as they are also vectors. Therefore,

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

the elementary definition of vector is running into a cyclic argument that is similar to the
egg and chicken paradox. We now state the following axiomatic framework for the vector
space to avoid this ambiguity.

Vector space
We consider only the vector spaces over the field of real numbers. Let us denote the set
of real numbers by < and vector space by V. The set, V, equipped with an addition
operation (V × V → V) and a scalar multiplication operation (< × V → V), is called
vector space if it follows:
(i) Associativity: (u + v) + w = u + (v + w), for all u, v, w ∈ V.
(ii) Commutativity: u + v = v + u, for all u, v ∈ V.
(iii) Existence of a zero element: There exist 0 ∈ V such that u + 0 = u, for all u ∈ V.
(iv) Existence of negative elements: For each u ∈ V there exist a unique v ∈ V such that
u + v = 0 and v is denoted as −u.
(v) Associativity in scalar multiplication (or the compatibility of multiplication defined
between the field elements (real numbers) and scalar multiplication operation): α(βu) =
(αβ)u, for all u, ∈ V, and α, β ∈ <.
(vi) Identity in scalar multiplication: There exist a unique element 1 ∈ < such that
1u = u, for all u ∈ V.
(vii) Distributivity with respect to addition operation on vectors: α(u + v) = αu + αv,
for all u, v ∈ V, and α ∈ <.
(viii) Distributivity with respect to scalar addition: (α + β)u = αu + βu, for all u ∈ V,
and α, β ∈ <.
The elements of set V are called vectors. This axiomatic framework generalizes the
notion of vector. We now present few examples for the vector space.

Examples of vector spaces


Example-1: Let us consider the set of all n-tuples,

<n = {(u1 , u2 , · · · , un ) : ui ∈ <}

Addition operation:
u + v = (u1 + v1 , u2 + v2 , · · · , un + vn )

Scalar multiplication operation:

αu = (αu1 , αu2 , · · · , αun )

It is easy to verify that the set, <n , obey all the axioms of vector space under the addition
and scalar multiplication. Therefore, the set, <n , is a vector space. It can be observed
that the special case n = 3 is our usual three-dimensional (3-D) Euclidean space.
Example-2: Let us consider the set of all m × n matrices over the real numbers,
 
· · · a1n


 a11 a12 

· · · a2n
 

 a21 a22  

<m×n =
 
 .. .. .. ..  : aij ∈ <




 . . . .

 


 
am1 am2 · · · amn
 

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Addition operation:
···
 
a11 + b11 a12 + b12 a1n + b1n

 a21 + b21 a22 + b22 ··· a2n + b2n 

A+B =  .. .. .. .. 

 . . . .


am1 + bm1 am2 + bm2 · · · amn + bmn
Scalar multiplication:
· · · αa1n
 
αa11 αa12

 αa21 αa22 · · · αa2n 

αA =  .. .. .. .. 

 . . . .


αam1 αam2 · · · αamn
It is easy to verify that the given set, <m×n , with defined operations is a vector space.
Example-3: Let us consider set of all Lebesgue measurable functions over the domain of
[0,1], as stated in the following expression,
 Z 1 
p p
L [0, 1] = f : |f | dx < ∞ ,
0

where p is a positive real number.


Addition operation:
(f + g)(x) = f (x) + g(x), x ∈ [0, 1]
Scalar multiplication:
(αf )(x) = α(f (x)), x ∈ [0, 1]
It is also easy to verify that the set, Lp [0, 1], is a vector space over the defined operations
and this vector space is known as Lp -space.
From these examples, it is clear that the definition of vector space is not limited to the
2-D or 3-D Euclidean spaces. This general definition of vector space has many advantages
in the study of linear algebra and operator theory. We now state the formal definition of
coordinate axes or coordinate frame.
Coordinate frame and basis:
We know from the discussion on elementary concept of vector that the definition of coordi-
nate vectors are useful in the representation of other vectors. Therefore, this necessitates
the definition of a basis which acts as coordinate frame. We can visualize the coordinate
axes as the geometric lines/curves along the basis vectors. The definition of basis require
set of linearly independent vector which is stated next.

Linear independence:
A subset {u1 , u2 , · · · , un } of V is said to be linearly dependent if and only if there exist
set of scalars α1 , α2 , · · · , αn , not all zero, such that

α1 u 1 + α2 u 2 + · · · + αn u n = 0

If such nonzero scalars do not exist, i.e, α1 = α2 = · · · = αn = 0, then the set vectors,
{u1 , u2 , · · · , un }, are said to be linearly independent.
Basis vectors:
A subset {u1 , u2 , · · · , un } of V is said to be the basis if the subset is linearly indepen-
dent and linear combination of this subset spans the total set V, i.e. there exist some

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

α1 , α2 , · · · , αn ∈ <, such that x = α1 u1 +α2 u2 +· · ·+αn un for every x ∈ V. The number of


vectors in the subset is called dimension of vector space. The basis vectors are the tangents
to the coordinate axes. It is clear that the basis vectors are not necessarily orthogonal and
also not unique but linearly independent. For example, {(1, 0, 0), (0, 1, 0), (0, 0, 1)} is one
set of orthogonal basis to 3-D Euclidean space whereas the set {(1, 0, 0), (1, 1, 0), (1, 1, 1)}
is also another basis but non-orthogonal. Every vector in 3-D space can be generated by
the linear combination from any one of the set of basis vectors.

Length of the vector:


None of the axioms of vector space gives neither an idea of length of the vector nor angle
between vectors. But, we require length of vector and angle between vectors to study
the quantities such as strain and strain rates in continuum mechanics. We know from
the elementary physics that the dot product involves length of vectors as well as angle
between vectors. Therefore, we introduce the length of vector through the dot product.
Dot product on 3-D Euclidean space
Let us consider 3-D Euclidean space with canonical basis {e1 , e2 , e3 }, i.e., e1 = (1, 0, 0),
e2 = (0, 1, 1) and e3 = (0, 0, 1). Any vector in this space can have three components along
{e1 , e2 , e3 } directions. For example the vector u is represented as either (u1 , u2 , u3 ) or
u1 e1 + u2 e2 + u3 e3 with the meaning of u1 , u2 , u3 are components of u along {e1 , e2 , e3 }
directions (see Fig. (2a)). We now recall the definition of dot product (in 3-D space) from
the elementary physics,
u · v = |u||v| cos(θ), (1)
where |u| and |v| are length of vectors u, v, respectively and θ is the angle between the
vectors u and v as shown in Fig. (2b).

Figure 2: Vector representation in 3-D: (a) components of vector (b) two vectors separated
by an angle θ.

The length of the vector is defined by

u · u = |u|2 = u21 + u22 + u23 (2)

Now, an alternative definition is stated for dot product which is equivalent to the one
defined in Eq. (1).
u · v = u1 v1 + u2 v2 + u3 v3 (3)
We note that the definition of dot product given in Eq. (3) accounts for both length of
vector shown in Eq. (2) and also angle between vectors shown in Eq. (1). We will prove
the equivalence of definitions stated in Eq. (1) and Eq. (3).

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Problem 1. In 2-D Euclidean space with canonical basis ({e1 = (1, 0), e2 = (0, 1)}) the
definition of dot product defined in Eq. (1) and Eq. (3) are equivalent, i.e. |u||v| cos(θ) =
u1 v1 + u2 v2

Proof-1. Let us consider two vectors in two dimensional space as shown in the Fig. (3).
We note that e1 is unit vector along horizontal axis and e2 is unit vector along vertical
axis.

u2 u

v
v2 θ
θv θu
u1
v1

Figure 3: Two vectors in 2-D space

The length of vectors can be represented by


1 1
|u| = (u21 + u22 ) 2 , |v| = (v12 + v22 ) 2 . (4)

Let θu and θv be angles made by vectors u and v with respect to horizontal axis as shown
in Fig. (3). Therefore, we have
u1 u2 v1 v2
cosθu = , sinθu = , cosθv = , sinθv = . (5)
|u| |u| |v| |v|

Combining these relations, we get


u1 v1 u2 v2
+ = cosθu cosθv + sinθu sinθv = cos(θu − θv ) (6)
|u| |v| |u| |v|

It is easy to observe from Fig. (3) that θ = θu − θv . Therefore, we have

u1 v1 + u2 v2 = |u||v|cosθ (7)

The result can be generalized to 3-D as any two vectors from two-dimensional subspace.
We can have the following alternative proof for the same result based on cosine rule.
Proof-2. We have two vectors u and v with an angle θ between them. Therefore, the
vectors u, v and u − v forms
q a triangle
q as shown q
in Fig. (4). Three sides of triangle are
2 2 2 2
|u|, |v| and |u − v|, i.e., u1 + u2 , v1 + v2 and (u1 − v1 )2 + (u2 − v2 )2 .

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

u u-v

θ
v

Figure 4: Two vectors with an angle θ

We now apply the cosine rule to the triangle shown in Fig. (4), i.e.,
|u − v|2 = |u|2 + |v|2 − 2|u||v|cosθ (8)
We can rewrite Eq. (8) as
1
|u||v|cosθ = (|u|2 + |v|2 − |u − v|2 )
2
1 2 1 1
= (u1 + u22 ) + (v12 + v22 ) − ((u1 − v1 )2 + (u2 − v2 )2 )
2 2 2
= u1 v1 + u2 v2
Thus, the result is proved. This can be easily generalized to the vectors in 3-D space.
Inner product:
We saw the general definition of vector space in the previous discussion. Now, we gener-
alize the definition of dot product over the vector space, which is known as inner product
or scalar product.
Let V be a vector space. The inner product is a function from V × V → <, denoted by
(u, v), satisfying the following properties:
(i) Linearity: (αu + βv, w) = α(u, w) + β(v, w) ∀α, β ∈ < and ∀u, v, w ∈ V
(ii) Symmetry: (u, v) = (v, u) ∀u, v ∈ V
(iii) Positive-definiteness: (u, u) ≥ 0, ∀u ∈ V and (u, u) = 0 if and only if u = 0.
The definition of inner product brings out the geometric quantities such as length and
orthogonality. The vector space equipped with inner product is called inner product space
or Hilbert space.
Example-4: There is an inner product on the vector space <n . It is defined as,
n
X
u · v = u1 v1 + u2 v2 + · · · + un vn = ui vi (9)
i=1

This definition satisfies all the properties of inner product. It is known as standard inner
product on <n .
Example-5: The choice of inner product is not unique. For any given positive definite
matrix S, another choice of inner product on <n is
n X
X n
u·v = ui Sij vj (10)
i=1 j=1

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

where Sij is ij th component of S.


This choice of inner product has application in measuring length of vector and angle
between two vectors when the basis of <n are non-orthogonal. The application of inner
product involving positive definite matrix is explained through the following numerical
example.
Example-6: Let us consider two vectors, u and v, in 2-D vector space (<2 ) as shown in Fig.
(5). As stated earlier the basis are not unique to the vector space. Hence, consider two
different basis to the vector space. Let first basis be the canonical basis, i.e. B1 = {e1 , e2 }
where e1 = (1, 0), e2 = (0, 1) and other basis be a non-orthogonal basis B2 = {e01 , e02 }
 1 1 
where e01 = (1, 0), e02 = √ , √ . It is trivial to find length of vectors and angle between
2 2
vector in canonical basis. We now present the way to find the length and angle between
vectors in non-orthogonal basis.

u = e1 + 2e2

e2 e'2 v = 2e1 + e2

e 1 e'1

Figure 5: Two vectors in the space of <2

Solution: Given vectors, u = e1 + 2e2 and v = 2e1 + e2 are in canonical basis. Length of
√ √
vectors, |u| = 5 and |v| = 5.
Let θ be the angle between two vectors. Then
u·v u1 v1 + u2 v2 4
cos θ = = =
|u||v| |u||v| 5
Both vectors u and v have the following representation in the basis B1 and B2 .

u = u1 e1 + u2 e2 = e1 + 2e2 in the basis B1



= u01 e01 + u02 e02 = −e01 + 2 2e02 in the basis B2
v = v1 e1 + v2 e2 = 2e1 + e2 in the basis B1

= v10 e01 + v20 e02 = e01 + 2e02 in the basis B2

Let us consider a positive definite matrix,


1
 
 1 √
2

S :=  1
 


√ 1

2
The inner product in the basis B2 ,
2 X
2
u0i Sij vj0
X
(u, v) =
i=1 j=1
= u01 S11 v10 + u01 S12 v20 + u02 S21 v10 + u02 S22 v20
= 4

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Length of vectors,
 1  1
q 2 X
2 2
√ q 2 X
2 2

u0i Sij u0j  = 0 0
X X
|u| = (u, u) =  5 and |v| = (v, v) =  v Sij v
i j = 5
i=1 j=1 i=1 j=1

The angle between two vectors,


(u, v) 4
cos θ = =
|u||v| 5
This example shows that a positive definite matrix is required to get physically meaningful
measure length and angle in non-orthogonal basis. Furthermore, if the positive definite
matrix is identity then we can recover usual inner product. Thus, the inner product with
respect to a positive definite matrix not only provide useful measure of length and angle
but generalizes the concept of inner product.
Example-7: The vector space L2 [0, 1] is an inner product space. The inner product over
L2 is defined by Z 1
(f, g) = f (x)g(x)dx, ∀f, g ∈ L2 [0, 1].
0
It can be observed that this definition follows the set of axioms defined for the inner
product space.

Orthogonality:
If the inner product between two non-zero vectors is zero then two vectors are said to be
orthogonal. Let u and v be orthogonal vectors in an inner product space V. Then

(u, v) = 0.

In two-dimensional or three-dimensional spaces if the vectors are orthogonal then the


angle between them would be π/2. Thus, the inner product brings out the concept of
orthogonality. Furthermore, vectors in inner product space follow an important inequality
called Cauchy-Schwartz inequality.
Cauchy-Schwartz inequality:
Let V be an inner product space. Then

(u, v)2 ≤ (u, u)(v, v), ∀u, v ∈ V (11)

and the equality holds if and only if u and v are linearly dependent.

Proof. If (u, v) = 0 then the inequality is trivial. Let us assume both vectors u and v
are non-zero. The positive definite property of inner product implies

f (α) = (u − αv, u − αv) ≥ 0 ∀α ∈ <

Using linearity, i.e., property (i) of inner product, we get

f (α) = (u, u) − 2α(u, v) + α2 (v, v) ≥ 0

The quadratic function f (α) is minimum at α = (u, v)/(v, v). Thus, we obtain the
Cauchy-Schwartz inequality by substituting α = (u, v)/(v, v) in previous equation. If
equality holds in Eq. (11) then we get (u − αv, u − αv) = 0 where α = (u, v)/(v, v). The
relation (u − αv, u − αv) = 0 implies u = αv, i.e., u and v are linearly dependent.

Joint initiative of IITs and IISc – Funded by MHRD 8


NPTEL – Mechanical Engineering – Continuum Mechanics

The application of the Cauchy-Schwartz inequality to <n and L2 [0, 1] spaces yields

n
!2 n
! n
!
u2i vi2 ,
X X X
ui vi ≤
i=1 i=1 i=1
Z 1 2 Z 1  Z 1 
f (x)g(x)dx ≤ f (x)2 dx g(x)2 dx .
0 0 0

The triangle inequality follows from the Cauchy-Schwartz inequality. The proof is pre-
sented next.

Triangle inequality:
Let V be an inner product space. Then

|u + v| ≤ |u| + |v| ∀u, v ∈ V

Proof:

|u + v|2 = (u + v, u + v)
= |u|2 + 2(u, v) + |v|2
≤ |u|2 + 2|(u, v)| + |v|2 (Since (u, v) ≤ |(u, v)|)
≤ |u|2 + 2|u||v| + |v|2 (By Cauchy-Schwartz inequality)
2
≤ (|u| + |v|)

We can get the triangle inequality by taking square-root on both sides. The triangle
inequality is essential in defining the normed vector space where the definition of length
or norm of vector is the main task.

Normed vector space:


Let V be a vector space. If there is a function from V → <, denoted by kuk where u ∈ V,
obey the following properties then V is called normed vector space.
(i) kuk ≥ 0 ∀u ∈ V with kuk = 0 ⇐⇒ u = 0
(ii) kαuk = |α|kuk ∀α ∈ < and ∀u ∈ V
(iii) ku + vk ≤ kuk + kvk ∀u, v ∈ V
It is clear that all inner product spaces are normed vector spaces equipped with the norm

1
kuk := |u| = (u, u) 2 .

Converse is not true, i.e., normed vector space is not necessarily an inner product space.
Consequently, the normed space is subset of general vector space and the inner product
space is a subset of normed vector space. This fact is depicted in Fig. (6). Furthermore,
the concept of length is defined for the normed space whereas the length and also angle
between vectors is defined for the inner product spaces. In continuum mechanics, we
require the definition of both length of vector and angle between vectors. Thus, all the
analysis are being done in the inner product spaces.

Joint initiative of IITs and IISc – Funded by MHRD 9


NPTEL – Mechanical Engineering – Continuum Mechanics

Normed space Vector space

Inner product space

Figure 6: Venn diagram for vector spaces, normed spaces, and inner product spaces

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 10


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-1: Tensor Algebra


Lecture-3: Index Notation and Conventions

The concept of the length of vector was discussed in previous lecture. We now introduce
the definitions of area and volume. To introduce these new quantities, we require the
definition of cross product between two vectors. We also introduce the index notation
along with some conventions to handle the vector algebra. These conventions are useful
in handling cumbersome algebraic operations between the vectors. Hereon we consider all
the vectors belonging to Euclidean three-dimensional space with orthonormal basis. The
orthonormal basis to the Euclidean three-dimensional space is known as Cartesian frame.
The summation convention:
Let us consider the sum,
s = u1 v1 + u2 v2 + u3 v3
We, usually, write the preceding equation using summation sign
3
X
s= ui vi . (1)
i=1

If the index i replaced with j or k in above expression the meaning would not alter, i.e.,
3
X 3
X 3
X
s= ui vi = uj vj = uk vk (2)
i=1 j=1 k=1

Therefore, the index i in Eq. (1) is called dummy index as it can be changed to any other
letter as shown in Eq. (2). Of course the dummy index appears twice in a term. We now
simplify the way of expressing Eq. (1) by dropping the explicit summation symbol, i.e.,

s = ui vi .

Although the summation symbol is not written explicitly, the summation is taken over
the index when it appears twice in a term. Of course the summation can also be expressed
s = uj vj as the index is dummy. This convention, i.e., taking the summation over dummy
index, is called Einstein’s summation convention. We follow this convention throughout
the tensor algebra and calculus.
We further make a note that no index should appear more than twice in a term. In
other words, the summation is not implied in a term ui vi wi over index i. If we want to
obtain summation then summation symbol should be written explicitly, i.e.,
3
X
ui vi wi .
i=1

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Example-1: We can have Aii = A11 + A22 + A33 . In this sum i is the dummy index. Since
the dimension of space is three, we have taken the summation over index i from 1 to 3.
If dimension of space is n, the summation should be taken over i from 1 to n.

Example-2: We can represent the summations


3 X
X 3 3 X
X 3 X
3
α= Aij ui vj , β= Bijk ui vj wk
i=1 j=1 i=1 j=1 k=1

with α = Aij ui vj and β = Bijk ui vj wk . The indices i and j are dummy indices in first
expression while i, j and k are dummy indices in second expression.
Free index:
If the index appear once in a term is called free index. Let us consider the system of linear
equations,
y1 = A11 x1 + A12 x2 + A13 x3
y2 = A21 x1 + A22 x2 + A23 x3
y3 = A31 x1 + A32 x2 + A33 x3
Using the summation convention, we get
y1 = A1j xj , y2 = A2j xj , y3 = A3j xj
These equations can also be written as,
yi = Aij xj , i = 1, 2, 3 (3)
The index j is dummy as it appears twice in a term on the right hand side of equation. On
the other hand, the index i is called free index since it appears once in a term. Of course
the index i appears once on left hand side and again once on the right hand side but once
in a term. In this example, clearly, index j is the dummy index and index i is the free
index. Usually the free index takes values 1, 2 or 3, unless stated explicitly. In conclusion,
the short form of three system of linear equations can be written as yi = Aij xj .
Example-1 For a given vector u and orthonormal basis {e1 , e2 , e3 }, the component along
basis vector
ui = u · ei , (4)
where the index i is free index. Furthermore, for the given components and also basis,
the vector can be represented by
u = uj ej , (5)
where j is dummy index. Upon substitution of Eq. (4) in Eq. (5), we get
u = (u · ei )ei (6)
Example-2: Consider the following set of equations,
A11 = B111 x1 + B112 x2 + B113 x3 , A12 = B121 x1 + B122 x2 + B123 x3 ,
A13 = B131 x1 + B132 x2 + B133 x3 , A21 = B211 x1 + B212 x2 + B213 x3 ,
A22 = B221 x1 + B222 x2 + B223 x3 , A23 = B231 x1 + B232 x2 + B233 x3 ,
A31 = B311 x1 + B312 x2 + B313 x3 , A32 = B321 x1 + B322 x2 + B323 x3 ,
A33 = B331 x1 + B332 x2 + B333 x3 .

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

The set of equations is represented by Aij = Bijk xk in the index notation. Clearly, i and
j are free indices while k is a dummy index.

Example-3: Consider the following set of equations,

x1 = B111 A11 + B112 A12 + B113 A13 + B121 A21 + B122 A22 + B123 A23 + B131 A31 +
B132 A32 + B133 A33
x2 = B211 A11 + B212 A12 + B213 A13 + B221 A21 + B222 A22 + B223 A23 + B231 A31 +
B232 A32 + B233 A33
x3 = B311 A11 + B312 A12 + B313 A13 + B321 A21 + B322 A22 + B323 A23 + B331 A31 +
B332 A32 + B333 A33 .

The set of equations in index notation is represented by xi = Bijk Ajk . In this expression,
i is free index while j and k are dummy indices.
Example-4: Consider the following set of equations,

x1 = A11 y1 + A12 y2 + A13 y3 + B11 z1 + B12 z2 + B13 z3


x2 = A21 y1 + A22 y2 + A23 y3 + B21 z1 + B22 z2 + B23 z3
x3 = A31 y1 + A32 y2 + A33 y3 + B31 z1 + B32 z2 + B33 z3 .

The set of equations in index notation is represented by xi = Aij yj + Bij zj or xi =


Aij yj + Bik zk .
From all the examples, it is easy to see the free index should appear once in a term and
never taken a summation over free index. On the other hand, the dummy index appears
twice in a term. In Example-4 Aij yj , Bij zj and xi represent separate terms. Clearly, the
index i appears once in every term. Therefore, the free index should appear once but in
every term.
Rules of indicial or index notation:

• The index that appear twice in a term is called dummy index and summation is
taken over all such dummy indices.

• The index that appear once but in every term is called free index and no summation
is taken over such free indices.

• No index should appear more than twice in a term.

The Kronecker delta:


Let us consider {e1 , e2 , e3 } be an orthonormal basis of the three-dimensional Euclidean
space <3 . From the definition of standard inner product (dot product), we have
(
0 if i 6= j
ei · ej = (7)
1 if i = j

The Kronecker delta, δij , is defined by


(
0 if i 6= j
δij = ei · ej = (8)
1 if i = j

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

It is clear from the definition of Kronecker delta that they are components of the identity
matrix. Following are the properties of Kronecker delta:
(i) δij = δji
(ii) δii = δ11 + δ22 + δ33 = 3
(iii) xi = δij xj , τij = τik δkj , and so on
Because of property (iii) in the preceding list, the Kronecker delta is known as substitution
operator.

The permutation symbol and cross product:


The Kronecker delta is defined through the dot product of basis vectors. Similarly, we
now define the permutation symbol or alternate tensor through the cross product between
basis vectors. Let V be a three-dimensional Euclidean vector space and the set {e1 , e2 , e3 }
be a orthonormal basis. We now define the cross product of basis vectors ej and ek by

ej × ek = ijk ei (9)

where ijk is permutation symbol or ijk th component the alternate tensor ε, and it is
given by

123 = 231 = 312 = 1


132 = 321 = 213 = −1
ijk = 0 otherwise.

Let us take dot product on both sides of Eq. (9) with em then we get

em · (ej × ek ) = ijk (ei · em ) = ijk δim = mjk (10)

We note that the dot product is taken with em instead of ei to avoid repetition of index
more than twice. Replacing the index i in place of m in the preceding equation, we get

ijk = ei · (ej × ek ). (11)

The distributivity of cross product is needed along with the definition stated in Eq. (9)
to perform the cross product between any two vectors.

Distributive property of cross product:


The distributive property of cross product is defined by

(αu) × (βv) = αβ(u × v) ∀α, β ∈ < and u, v ∈ V

Example-5: Let w be the resultant vector of cross product between two vectors u and v.
Then express w in terms of components of u, v.

Solution: Consider the given relation

w = u×v
= (uj ej ) × (vk ek ) (from Eq. (5))
= uj vk (ej × ek ) (using distributive property)
= ijk uj vk ei

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

It is clear from preceding equation that


u × v = −v × u (12)
If v = αu then we get u × v = 0. Therefore, the cross product between any linearly
dependent vectors is zero.
Area and cross product:
Geometrically, the magnitude of cross product represents the area of parallelogram formed
by two vectors. The direction represents the orientation of area which is given by right-
hand screw rule. The geometrical meaning of cross product is depicted in Fig. (1) which
is consistent with algebraic result shown in Eq. (12).

Figure 1: Geometrical interpretation of cross product

Scalar triple product and volume:


We need quantities such as length, angle, area and volume in continuum mechanics. Till
now the length and angle is taken care by the dot product whereas the area is taken care
by the cross product. We now introduce the scalar triple product between three vectors
to account for volume.

Let V be a vector space. Let u, v and w be three vectors in V. Then the scalar triple
product, denoted by [u, v, w], is defined by
[u, v, w] := u · (v × w) (13)
Clearly, the scalar triple product is a function from V × V × V to <. The absolute value
of scalar triple product gives the volume of parallelepiped formed by three vectors. The
scalar triple product in indicial notation,
[u, v, w] := ijk ui vj wk . (14)
Properties of scalar triple product:
(i) [u, v, w] = [w, u, v] = [v, w, u] = −[v, u, w] = −[w, v, u] = −[u, w, v] ∀u, v, w ∈ V
(ii) Multi-linear function or in other words linear in every variable, i.e.,
[αu + βx, v, w] = α[u, v, w] + β[x, v, w] ∀x, u, v, w ∈ V and ∀α, β ∈ <
[u, αv + βy, w] = α[u, v, w] + β[u, y, w] ∀y, u, v, w ∈ V and ∀α, β ∈ <
[u, v, αw + βz] = α[u, v, w] + β[u, v, z] ∀z, u, v, w ∈ V and ∀α, β ∈ <

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

(iii) The scalar triple product can be written in determinant form as,
 
u1 u2 u3
[u, v, w] = det  v1 v2 v3  . (15)
 

w1 w2 w3

We know from matrix algebra that the determinant1 for above 3×3 matrix can be written
as u1 (v2 w3 − w2 v3 ) − u2 (w3 v1 − v3 w1 ) + u3 (v1 w2 − w1 v2 ). Therefore, this form can be easily
verified using the definition shown in Eq. (14).
(iv) The scalar triple product [u, v, w] is zero if and only if the vectors u, v and w are
linearly dependent, i.e., all three vectors lie on a plane.
The epsilon-delta identity:
The permutation symbol (ijk th component of alternate tensor ε) can be represented using
scalar triple product shown in Eq. (11). Using property (iii) of scalar triple product shown
in Eq. (15) and Eq. (4), we get
   
ei · e1 ei · e2 ei · e3 δi1 δi2 δi3
ijk = [ei , ej , ek ] = det  ej · e1 ej · e2 ej · e3  = det  δj1 δj2 δj3  (16)
   

ek · e1 ek · e2 ek · e3 δk1 δk2 δk3

Multiplication of two permutation symbols can be written as,


   
δi1 δi2 δi3 δp1 δp2 δp3
ijk pqr = det  δj1 δj2 δj3  det  δq1 δq2 δq3 
  

δk1 δk2 δk3 δr1 δr2 δr3
   
δi1 δi2 δi3 δp1 δq1 δr1   
= det  δj1 δj2 δj3  det  δp2 δq2 δr2  since det(T ) = det T T
   

δk1 δk2 δk3 δp3 δq3 δr3


  

 δi1 δi2 δi3 δp1 δq1 δr1 
= det  δj1 δj2 δj3   δp2 δq2 δr2 

 (since(det R)(det S) = det(RS))
 

δk1 δk2 δk3 δp3 δq3 δr3 
   
δim δmp δim δmq δim δmr δip δiq δir
= det  δjm δmp δjm δmq δjm δmr  = det  δjp δjq δjr  (17)
   

δkm δmp δkm δmq δkm δmr δkp δkq δkr

Using Eq. (17) and a property of Kronecker delta, δii = 3, we get the following famous
-δ identity
ijk iqr = δjq δkr − δjr δkq (18)
As a consequence of -δ identity, we also have

ijk ijl = 2δkl and ijk ijk = 6 (19)

We now state another important identity


1
ei = ijk ej × ek (20)
2
This identity can be verified using first part of Eq. (19) and Eq. (9).
1
G. Strang, “Linear Algebra and Its Applications”, 2005, Fourth edition, Cengage Learning, New
Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

Example-6: Show that the vector triple product, in general, is not associative i.e., u ×
(v × w) is different from (u × v) × w.

Solution:

u × (v × w) = ijk uj (v × w)k ei
= ijk uj kmn vm wn ei
= kij kmn uj vm wn ei
= (δim δjn − δin δjm )uj vm wn ei
= (δim δjn uj vm wn − δin δjm uj vm wn )ei
= (uj vi wj − uj vj wi )ei
= (uj wj )vi ei − (uj vj )wi ei
= (u · w)v − (u · v)w

(u × v) × w = ijk (u × v)j wk ei
= ijk jmn um vn wk ei
= jki jmn um vn wk ei
= (δkm δin − δkn δim )um vn wk ei
= (δkm δin um vn wk − δkn δim um vn wk )ei
= (uk vi wk − ui vk wk )ei
= (uk wk )vi ei − (vk wk )ui ei
= (u · w)v − (v · w)u

It is clear that the vector triple product of non-zero vectors u, v and w is associative if
either u and w are linearly dependent or all three vectors are mutually orthogonal. In
general, the vector triple product does not follow associative property.

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. K. Dullemond and K. Peeters, Introduction to Tensor Calculus, 2010.


(website: http://www.ita.uni-heidelberg.de/ dullemond/lectures/tensor/tensor.pdf)

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-1: Tensor Algebra


Lecture-4: The Second-Order Tensor

We know from elementary physics that the quantities such as density and temperature
are represented by scalar functions while velocity and acceleration are represented by
vectors. But the stress and strain cannot be represented by a scalar or a vector. Thus, we
introduce the concept of tensor to represent these new physical quantities. In this lecture,
we specialize our discussion to second-order tensors after defining the tensor.
Tensor:
Let V and W be two vector spaces. Then the linear function L that is defined from
V → W is called a tensor1 , i.e., the function L : V → W with the following property is
called tensor.

L(αu + βv) = αLu + βLv, ∀u, v ∈ V and ∀α, β ∈ <.

We note that L(αu + βv), Lu and Lv are vectors in the vector space W. Furthermore,
the linear function L is also known as linear transformation. It can be observed from the
following examples that both the scalar and vector are special cases of tensor.
Example-1: We know from Example-1 in Lecture-2 that the set of real numbers < forms
a vector space. Let us choose V = W = < in above definition. Let f be a linear function
defined from V → W, i.e., f (αx+βy) = αf (x)+βf (y) for all α, β, x, y ∈ <. Then there is
a scalar c such that f (x) = cx for every x ∈ <. In other words f assumes a representation
of a scalar. It is easy to verify that the function follows the linearity property defined
above. Thus, all scalars are tensors.

Example-2: Let us choose V = <3 and W = < in above definition of tensor. Let f : V →
W be a linear function, i.e., f (αx + βy) = αf (x) + βf (y) for all x, y ∈ <3 and for all
α, β ∈ <. Then there exist a vector v ∈ <3 such that f (u) = v · u. It is easy to see that
the function f follows the above stated linearity. Thus, every vector in vector space <3 is
also a tensor.
Clearly, the scalars and vectors are included in the definition of tensor. It is possible to
define various tensors by choosing the vector spaces V and W. We know that the scalars
have one component whereas the vectors have three components in a given coordinate
frame. We will see that the second-order tensor possess nine components. In general nth
order tensors have 3n components. Therefore, the scalars are zeroth-order tensors and the
vectors are first-order tensors.
We now discuss the algebra of second-order tensors.
1
Linear transformation is a synonym for tensor

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Equality of two tensors:


Let R and S be two tensors defined from V → W. Then the tensors R and S said to be
equal if
Ru = Su ∈ W, ∀u ∈ V. (1)
This equality property is utilized to show many results in continuum mechanics.

We know from Example-1 in Lecture-1 that set of scalars forms one dimensional vector
space and set of vectors forms an n-dimensional vector space. We now show that the
collection tensors also forms a new vector space (space of second-order tensors). We
denoted the new vector space formed by second order tensors with Lin.

Algebra of tensors:
Let us collect the set of all tensors defined from V to W and denote by Lin. We now
define sum, R + S, of two tensors R and S and the scalar product, αT , of tensor T by

(R + S)u = Ru + Su, (2)


(αT )u = α(T u), ∀u ∈ V and α ∈ <. (3)

We note that Ru, Su and T u are vectors in the vector space W. As a consequence of
definitions of sum and scalar product, the set of all tensors forms a new vector space over
the field of real numbers. This fact can be verified using vector space axioms stated in
Lecture-2. Analogous to zero vector, there is a zero tensor, O in the vector space Lin.
The zero tensor is defined by

Ou = 0 ∈ W, ∀u ∈ V.

It is easy to see that

(T + O)u = T u + Ou = T u, ∀u ∈ V and ∀T ∈ Lin.

The definition of equality of tensors implies,

T + O = T ∀T ∈ Lin.

All other axioms of vector space are easy to verify. Thus, with the definition of addition
and scalar multiplication given in Eq. (2) and (3), the set Lin is a vector space over <.
We will see in later lectures that stress, strain and rate of deformation are second-order
tensors. Therefore, second-order tensors play an important role in continuum mechanics.
Thus, we specialize our discussion more towards the second-order tensors than other
higher-order tensors. Here on tensor without specifying order would be treated as second-
order tensor.

Definition of second-order tensor (SOT):


Let V be a vector space then the second-order tensor, T , is defined as a linear transfor-
mation from V → V, i.e. the function T : V → V with the following property,

T (αu + βv) = αT u + βT v, ∀u, v ∈ V and α, β ∈ <.

It is easy to see that T (0) = 0 by choosing α = 1, β = −1, and u = v.

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Equality of two second-order tensors:


Here we now invoke the previously stated general definition of equality of tensors. Two
second-order tensors R and S are said to be equal if

Ru = Su, ∀u ∈ V. (4)

The condition of equality of tensors shown in Eq. (4) is also equivalent to

(v, Ru) = (v, Su), ∀u, v ∈ V. (5)

It is easy to see that the condition shown in Eq. (4) implies the condition shown in Eq.
(5). Using linearity of inner product, we can write Eq. (5) as,

(v, (Ru − Su)) = 0, ∀u, v ∈ V.

If we choose v = Ru − Su and substitute in preceding equation then positive definite


property of inner product implies the condition shown in Eq. (4).
Algebra of second-order tensors:
As pointed out in algebra of tensors that set of all tensors defined from V → W forms
a vector space. Therefore, the set of all second-order tensors is a vector space. Since
V = W = <3 , we denote the space of second-order tensors by V 2 or <3×3 .
Zero tensor:
The zero tensor is a zero element in vector space of second-order tensors V 2 . If the O is
zero tensor then
T + O = T ∀T ∈ V 2 .
Identity tensor:
There is a special tensor called identity tensor in the collection of second-order tensors.
The identity tensor is defined by

Iu = u, ∀u ∈ V.

The product of two second-order tensors:


The product of two second-order tensors R and S, denoted by RS, defined by

(RS)u = R(Su) ∀u ∈ V. (6)

This product or composition is similar to the usual product of two matrices. We note
that any tensor, T , on pre-multiplication or post multiplication with identity tensor I
does not alter the action of T on vectors, i.e.,

T I = IT = T ∀T ∈ V 2 .

Transpose of tensor:
Let T T be a transpose of tensor T . Then the transpose of tensor T , using inner product,
is defined by
(T T u, v) = (u, T v), ∀u, v ∈ V (7)
The transpose of tensor has the following properties:

(T T )T = T
(αT )T = αT T
(R + S)T = RT + S T

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Symmetric tensor:
The tensor S is said to be symmetric if S T = S.

Skewsymmetric tensor:
The tensor W is said to be skewsymmetric (antisymmetric) if WT = −W.
Additive decomposition of tensor:
Every second-order tensor T can be uniquely decomposed into a symmetric and a skew-
symmetric part as
T = Ts + Tss (8)
where
1 1
Ts = (T + T T ), Tss = (T − T T )
2 2
Components of second-order tensor in a coordinate frame:
Let e1 , e2 and e3 be an orthonormal basis to the Euclidean three-dimensional vector space
V. Let T be a second-order tensor from V to V. Then, we have

Tu = v (9)

where u and v are in V. Note that u and v are linear combination of basis vectors as
they are in V. Choosing e1 , e2 , and e3 for u in Eq. (9), we get

T e1 = α1 e1 + α2 e2 + α3 e3
T e2 = α4 e1 + α5 e2 + α6 e3
T e3 = α7 e1 + α8 e2 + α9 e3 ,

where αi , i = 1 to 9, are scalars. Renaming αi as Tij , i = 1, 2, 3; j = 1, 2, 3, we get

T ej = Tij ei . (10)

Clearly, i is dummy index and j is the free index. The scalars Tij are called components
of the tensor T with respect to the base vectors ei , i = 1, 2, 3. Taking the dot product
both sides of Eq. (10) with ek , we get

ek · T ej = Tij ek · ei = Tij δik = Tkj

Replacing k with i, we can write


Tij = ei · T ej . (11)

Components of identity tensor:


Let I be an identity tensor then applying the component form (see Eq. (11)) to the
tensor, we get
Iij = ei · Iej = ei · ej = δij . (12)
Thus, the Kronecker delta represents the components of identity tensor. Furthermore,
the components of identity tensor are fixed in the Cartesian basis, i.e., the components
are independent of a set of orthonormal basis.

Components of tensor-vector product:


In general, the action of tensor T on a vector u shown in Eq. (9). Applying Eq. (5) in
Lecture-3 along with linearity property of tensor to (9), we get

vi ei = T (uj ej ) = uj T ej = uj Tij ei .

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Thus, the components of resulting tensor-vector product can be written as

vi = Tij uj (13)

Components of transpose tensor:


Let Tij be ij th component of the tensor T . Then the components of T T are related to the
component of T by the following relation.

(T T )ij = ei · T T ej (using Eq. (11))


= T ei · ej (by definition of transpose, i.e., Eq. (7))
= Tji (again using Eq. (11)) (14)

Tensor components under change of basis:


We mentioned in elementary concept of vector (recall from Lecture-2) that the components
of vector do change with change of coordinate frame although the vector is invariant.
Similarly, the components of tensor are also altered with change of coordinate frame
although the tensor is invariant. We now discuss the transformation of vector and second-
order tensor components under the change of coordinate frame. Since the basis vectors
are assumed to be orthonormal in both coordinate frames, the transformation of basis
indicates the rotation from one coordinate frame to another coordinate frame. Let two sets
of orthonormal vectors {e1 , e2 , e3 } and {e01 , e02 , e03 } represent original and new coordinate
frames. We now define the direction cosines between new and original basis vectors as
lij = e0i · ej . Representation of the new basis vectors in original basis yields

e0i = (e0i · ej )ej (since u = (u · ei )ei )


= lij ej . (15)

Similarly, we can represent the original basis vectors in new basis by

ei = (ei · e0j )ej = lji e0j . (16)

Let us consider a vector v. Then the transformation law for the component of the vector
is given by,
vi0 = v · e0i = v · (lij ej ) = lij v · ej = lij vj (17)
Similar to transformation of vector components, using Eq. (11) and Eq. (15), we get the
following transformation law for the second-order tensor.

Tij0 = e0i · T e0j = lim ljn Tmn . (18)

The tensor product of two vectors:


We already discussed two types of products between a pair of vectors, i.e., an inner product
as a function from V × V → < and cross product as V × V → V. We now introduce new
product between two vectors is called dyadic or tensor product which is a function from
V × V → V 2 . Let a and b are vectors in the vector space V. Then the tensor product of
two vectors is defined by
(a ⊗ b)c = (b · c)a, ∀c ∈ V. (19)
Clearly, a vector is an input and also output to the dyadic product of two vectors. There-
fore, if linearity property holds then we can conclude that a ⊗ b is a second-order tensor.

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Let us consider operation of a ⊗ b on a vector αx + βy.

(a ⊗ b)(αx + βy) = (b · (αx + βy))a


= (αb · x + βb · y)a
= α(b · x)a + β(b · y)a
= α(a ⊗ b)x + β(a ⊗ b)y.

It is easy to see from above equation that a ⊗ b is a linear function from V to V. Thus,
a ⊗ b is a second-order tensor.
Representation of second-order tensor:
Every second-order tensor T assumes the representation

T = Tij ei ⊗ ej . (20)

Considering a vector u and the representation of tensor shown in Eq. (20), we get

(Tij ei ⊗ ej )u = Tij (ei ⊗ ej )u = Tij (ej · u)ei = Tij uj ei = (T u)i ei = T u

This holds for all u ∈ V. Therefore, any second-order tensor admits the representation
shown in Eq. (20). We already showed that the set of second-order tensors V 2 is a real
vector space. We now state that the set of tensors {ei ⊗ ej }, i = 1, 2, 3 and j = 1, 2, 3
forms a basis to the space of second-order tensors. In fact the linear combination can
span every second-order tensor which is shown as representation in Eq. (20).

Representation of identity tensor:


Since components of the identity tensor are represented by Kronecker delta (see Eq. (12)),
we can have the following representation for the identity tensor.

I = ei ⊗ ei . (21)

Equality of tensors based on components:


Two second-order tensors T and S are equal if and only if corresponding components are
equal in a given basis of vector space V.

Proof. Let {e1 , e2 , e3 } be an orthonormal basis of vector space V. Then we have

Tij = ei · T ej and Sij = ei · Sej .

We know that two tensors are equal if and only if

(u, T v) = (u, Sv) ∀u, v ∈ V.

Choosing u = ei and v = ej , we get

(ei , T ej ) = (ei , Sej ) =⇒ Tij = Rij .

Therefore, equality of tensors implies respective components are equal.

Conversely, let Tij = Rij then we want to show that (u, T v) = (u, Sv) ∀u, v ∈ V.
Consider the dyadic product representation of two tensors in a given basis, i.e,

T = Tij ei ⊗ ej and R = Rmn em ⊗ en .

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

We can have the following scalar product of two vectors u and v with respect to tensors,
(u, T v) − (u, Sv) = (ui ei , T (vj ej )) − (ui ei , S(vj ej )) (Since u = ui ei and v = vj ej )
= ui vj (ei , T ej ) − ui vj (ei , Sej )
= ui vj Tij − ui vj Sij
= ui vj (Tij − Sij )
= 0 (Since Tij = Sij ).
Thus, two second-order tensors are equal if and only if respective components are equal
in a given basis.

Representation of transpose using basis tensors:


Let T be a tensor then components of transpose are given by
(T T )ij = Tji .
Therefore, the transpose can be represented by
T T = Tji ei ⊗ ej . (22)
Components of tensor product of two vectors:
Let u and v be two vectors in vector space V. Then, using the definition of components
of tensor, we get
(u ⊗ v)ij = ei · (u ⊗ v)ej
= ei · (v · ej )u
= vj (ei · u)
= ui vj (23)
Invariance of vector and tensor:
We made a remark that the vector does not depend on the coordinate frame but compo-
nents of vector do depend on the coordinate frame. We now prove the statement using
transformation law presented previously. Let {e1 , e2 , e3 } and {e10 , e20 , e30 } are set of orig-
inal and new basis respectively. Let u = ui ei and u0 = ui0 ei0 represent a vector in original
and new basis. Then
u = ui ei
= ui (lji e0j ) (by Eq. (16))
= lji ui e0j
= u0j e0j (By transformation law)
0
= u
Similar to vectors, the second-order tensor also does not depend on the coordinate frame
but the components do depend on the frame. Let T 0 = Ti0 j 0 ei0 ⊗ ej 0 and T = Tij ei ⊗ ej
be the tensor representations in new and original coordinate frames respectively. Then
T = Tij ei ⊗ ej
= Tij (lmi e0m ) ⊗ (lnj e0n ) (by Eq. (16))
= lmi lnj Tij e0m ⊗ e0n
0
= Tmn e0m ⊗ e0n (By transformation law)
0
= T

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

The vector and tensor components follow the transformation law then the physical vector
and tensor are independent of coordinate frame. This is true for the higher order tensors
too. Therefore, some authors consider the transformation law as definition for the tensor.

Motivation for inner product of second-order tensors:


Similar to inner product of vectors in three dimensional Euclidean space, we can define
inner product on the space of second-order tensors. This definition is useful for quantifying
the stress. Otherwise, the comparison of stress state is very difficult as it consists of nine
components.
Definition of inner-product:
Let R and S be second-order tensors. Then the inner product of two second-order tensors
R and S, denoted by R : S or (R, S), is defined by

R : S = tr(RT S) = tr(RS T ) = Rij Sij (24)

Norm of second-order tensor:


The inner product induces the norm or magnitude. The norm of second-order tensor T
is given by
|T | = (T : T )1/2

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 8


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-1: Tensor Algebra


Lecture-5: The Principal Invariants and
the Inverse of Tensor

As noted in Lecture-4, the components of tensor are altered with the change of basis
although the tensor is invariant. We now ask a question that are there any relations
among the components such that they are invariant under change of basis for a given
tensor. The answer is affirmative and in fact there are three such kind of invariant
relations exist for a second-order tensor. These three relations are known as principal
invariants. Of course, any other invariant relation can be expressed as function of these
three principal invariants. The relations can be obtained from the following equations.

Let {u, v, w} and {a, b, c} be two set of linearly independent vectors and let T be a
second-order tensor. Then
[T u, v, w] + [u, T v, w] + [u, v, T w] [T a, b, c] + [a, T b, c] + [a, b, T c]
= (1)
[u, v, w] [a, b, c]
[T u, T v, w] + [u, T v, T w] + [T u, v, T w] [T a, T b, c] + [a, T b, T c] + [T a, b, T c]
= (2)
[u, v, w] [a, b, c]
[T u, T v, T w] [T a, T b, T c]
= (3)
[u, v, w] [a, b, c]

Problem 1. Let {u, v, w} and {a, b, c} are two sets of linearly independent vectors.
Then prove Eq. (1).

Proof. Since the set of vectors {u, v, w} are linearly independent they form a basis to the
vector space V. Let T be a second-order tensor. Then

T u = α1 u + β1 v + γ1 w
T v = α2 u + β2 v + γ2 w
T w = α3 u + β3 v + γ3 w (4)

Using properties of scalar triple product (see Lecture-3 for properties of scalar triple
product) and above relations, we get

[T u, v, w] + [u, T v, w] + [u, v, T w]
= [α1 u + β1 v + γ1 w, v, w] + [u, α2 u + β2 v + γ2 w, w] + [u, v, α3 u + β3 v + γ3 w]
= α1 [u, v, w] + β2 [u, v, w] + γ3 [u, v, w]
= (α1 + β2 + γ3 )[u, v, w] (5)

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Let {e1 , e2 , e3 } be orthonormal basis. Then the vectors u, v and w have the representa-
tion u = ui ei , v = vj ej and w = wk ek . Therefore, the relation shown in Eq. (5) can also
be expressed as

[T u, v, w] + [u, T v, w] + [u, v, T w]
= [T ui ei , vj ej , wk ek ] + [ui ei , T vj ej , wk ek ] + [ui ei , vj ej , T wk ek ]
= ui vj wk {[T ei , ej , ek ] + [ei , T ej , ek ] + [ei , ej , T ek ]} (6)

Choosing u = ei , v = ej and w = ek in Eq. (5), we get

[T ei , ej , ek ] + [ei , T ej , ek ] + [ei , ej , T ek ] = (T11 + T22 + T33 )[ei , ej , ek ]


= (T11 + T22 + T33 )ijk (7)

Let us define the invariant of tensor IT by

IT = T11 + T22 + T33 . (8)

We note that T11 = e1 · T e1 , T22 = e2 · T e2 and T33 = e3 · T e3 . It is easy to verify Eq.


(7) for all possible indices of i, j and k. Substitution of Eq. (7) in Eq. (6) yields

[T u, v, w] + [u, T v, w] + [u, v, T w] = (T11 + T22 + T33 )ijk ui vj wk


= (T11 + T22 + T33 )[u, v, w] (9)

The arbitrariness of u, v and w implies Eq. (1).

We now prove that [T x, y, z] + [x, T y, z] + [x, y, T z] = 0 if x, y and z are linearly


dependent. There are two ways to get linearly dependent set of x, y and z. In first case
there is only one independent vector among three vectors. Whereas, in the second case
two vectors are independent among three vectors.
Case-i: Let one vector be independent among three, i.e., two vectors can be written as
scalar multiple of one vector. Without loss of generality, we can write

x = αy, x = βz.

It is easy to see from the definition of triple scalar product that

[T x, y, z] + [x, T y, z] + [x, y, T z] = 0

Case-ii: There are two linearly independent vectors among three, i.e., one vector can be
written as linear combination of other two vectors. Without loss of generality, we can
write
z = αx + βy
Using properties of scalar triple product, we get

[T x, y, z] + [x, T y, z] + [x, y, T z] = [T x, y, αx + βy] + [x, T y, αx + βy]


+[x, y, T (αx + βy)]
= α[T x, y, x] + β[x, T y, y] + [x, y, αT x + βT y]
= α[T x, y, x] + β[x, T y, y] + α[x, y, T x] + β[x, y, T y]
= 0.

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Thus, we have the more general result


[T u, v, w] + [u, T v, w] + [u, v, T w] = IT [u, v, w], (10)
where IT is the first invariant. This result holds for all vectors u, v and w irrespective of
whether the vectors are dependent or independent.
The first invariant of tensor associated with Eq. (1) is called the trace of tensor. If u,
v and w are linearly independent vectors then the trace of tensor is defined by

[T u, v, w] + [u, T v, w] + [u, v, T w]
tr(T ) =
[u, v, w]

Problem 2. Let {u, v, w} and {a, b, c} are two set of linearly independent vectors. Then
prove Eq. (2).

Proof. This can be proved similar to the previous problem. Let u, v, w be independent
vectors and let T be a tensor. Then we have Eq. (4). Using properties of scalar triple
product (see Lecture-3 for properties of scalar triple product) and Eq. (4), we get
[T u, T v, w] + [u, T v, T w] + [T u, v, T w] = [α1 u + β1 v + γ1 w, α2 u + β2 v + γ2 w, w]
+[u, α2 u + β2 v + γ2 w, α3 u + β3 v + γ3 w]
+[α1 u + β1 v + γ1 w, v, α3 u + β3 v + γ3 w]
= α1 β2 [u, v, w] + β1 α2 [v, u, w]
+β2 γ3 [u, v, w] + γ2 β3 [u, w, v]
+α1 γ3 [u, v, w] + γ1 α3 [w, v, u]
= {(α1 β2 − β1 α2 ) + (β2 γ3 − γ2 β3 )
+(α1 γ3 − γ1 α3 )}[u, v, w]. (11)
Let {e1 , e2 , e3 } be orthonormal basis to the vector space. Then we have u = ui ei , v = vj ej
and w = wk ek . An alternative form of Eq. (11) can be written as
[T u, T v, w] + [u, T v, T w] + [T u, v, T w]
= [T ui ei , T vj ej , wk ek ] + [ui ei , T vj ej , T wk ek ] + [T ui ei , vj ej , T wk ek ]
= ui vj wk {[T ei , T ej , ek ] + [ei , T ej , T ek ] + [T ei , ej , T ek ]}. (12)
Choosing u = ei , v = ej and w = ek and substituting in Eq. (11), we get
[T ei , T ej , ek ] + [ei , T ej , T ek ] + [T ei , ej , T ek ]
= {(T11 T22 − T12 T21 ) + (T22 T33 − T23 T32 ) + (T33 T11 − T31 T13 )}[ei , ej , ek ]
= ijk {(T11 T22 − T12 T21 ) + (T22 T33 − T23 T32 ) + (T33 T11 − T31 T13 )} (13)
Let IIT be second invariant of tensor T . Then we have
IIT = (T11 T22 − T12 T21 ) + (T22 T33 − T23 T32 ) + (T33 T11 − T31 T13 ). (14)
Substituting Eq. (13) in Eq. (12), we get
[T u, T v, w] + [u, T v, T w] + [T u, v, T w] = IIT ijk ui vj wk
= IIT [u, v, w] (15)
The arbitrariness of u, v and w implies Eq. (2).

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

If the vectors x, y and z are linearly dependent then we get [T x, T y, z]+[x, T y, T z]+
[T x, y, T z] = 0. This can be proved by following similar steps that were adopted in
previous cases of first invariant.

Problem 3. If {u, v, w} and {a, b, c} are two set of linearly independent vectors then
prove Eq. (3).

Proof. Similar to previous problems, substituting Eq. (4), we get

[T u, T v, T w] = [α1 u + β1 v + γ1 w, α2 u + β2 v + γ2 w, α3 u + β3 v + γ3 w]
= α1 β2 γ3 [u, v, w] + α1 β3 γ2 [u, w, v] + β1 α2 γ3 [v, u, w]
+β1 γ2 α3 [v, w, u] + γ1 α2 β3 [w, u, v] + γ1 α3 β2 [w, v, w]
= {α1 (β2 γ3 − β3 γ2 ) − β1 (α2 γ3 − γ2 α3 ) + γ1 (α2 β3 − α3 β2 )}[u, v, w] (16)

Let {e1 , e2 , e3 } be orthonormal basis. Then we can obtain

[T u, T v, T w] = [T ui ei , T vj ej , T wk ek ] = ui vj wk [T ei , T ej , T ek ] (17)

Choosing u = ei , v = ej and w = ek in Eq. (16), we get

[T ei , T ej , T ek ] = ijk {T11 (T22 T33 − T23 T32 ) − T12 (T21 T33 − T23 T31 ) + T13 (T21 T32 − T31 T22 )}
(18)
Let IIIT be third invariant which is defined by

IIIT = T11 (T22 T33 − T23 T32 ) − T12 (T21 T33 − T23 T31 ) + T13 (T21 T32 − T31 T22 ). (19)

Substituting Eq. (18) in Eq. (17), we get

[T u, T v, T w] = IIIT ijk ui vj wk = IIIT [u, v, w]. (20)

Arbitrariness of u, v and w implies Eq. (3).

The third invariant of tensor associated with Eq.(3) is known as determinant of tensor.
If u, v and w are linearly independent vectors then the determinant

[T u, T v, T w]
det(T ) =
[u, v, w]

Invariants using indicial notation:


We now express the invariants shown in Eqs. (8), (14) and (19) using indicial notation.

tr(T ) = IT = Tii (21)


1
IIT = imn ipq Tmp Tnq (22)
2
1
det(T ) = IIIT = ijk pqr Tip Tjq Tkr (23)
6
Properties of trace:
(i) tr(T T ) = tr(T ) ∀T ∈ V 2
(ii) tr(αT ) = α tr(T ) ∀T ∈ V 2 and ∀α ∈ <

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

(iii) tr(RS) = tr(SR) ∀R, S ∈ V 2


(iv) tr(I) = 3, where I is identity tensor.

Properties of determinant:
(i) det(T ) =det(T T ) ∀T ∈ V 2
(ii) det(αT ) = α3 det(T ) ∀T ∈ V 2 and ∀α ∈ <
(iii) det(RS) = det(S)det(R) = det(SR) ∀R, S ∈ V 2
(iv) det(I) = 1, where I is identity tensor.

Cofactor of tensor:
We know how vectors are being transformed under the action of tensor. We now want
to know how the area is being transformed under the action of tensor. Recall from
Lecture-3 that area of parallelogram formed with two vectors can be obtained by the
cross-product. We now define a new tensor called cofactor of tensor that takes care of
areal transformation.
Definition of cofactor:
The cofactor of tensor T is defined by

cof T (u × v) = T u × T v, ∀u, v ∈ V. (24)

We note that the cofactor of tensor cof T is also a second-order tensor as input and
output are vectors.
Problem 4. Let R and S are two second-order tensors. Then show that

cof (RS) = (cof R)(cof S) (25)

Proof. Let u, v be arbitrary vectors. Then

(cof R)(cof S)(u × v) = cof R(Su × Sv)


= (RSu) × (RSv) (Since Su, Sv ∈ V)
= cof (RS)(u × v) (Since RS is a second-order tensor)

Arbitrariness of u, v implies Eq. (25).

The second invariant and trace of cofactor:


We now show that the second invariant of tensor T is equal to the trace of cofactor of
tensor T . Let us consider second invariant shown in Eq. (15), i.e.,

IIT [u, v, w] = [T u, T v, w] + [T u, v, T w] + [u, T v, T w]


= w · (T u × T v) + v · (T w × T u) + u · (T v × T w)
= w · cof T (u × v) + v · cof T (w × u) + u · cof T (v × w)
= (cof T )T w · (u × v) + (cof T )T v · (w × u) + (cof T )T u · (v × w)
= [u, v, (cof T )T w] + [u, (cof T )T v, w] + [(cof T )T u, v, w]
= tr((cof T )T )[u, v, w]
= tr(cof T )[u, v, w] (26)

The arbitrariness of vectors u, v and w imply the trace of cofactor tensor, tr(cof T), is
equal to the second invariant of the tensor T .

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Problem 5. Let T be a second-order tensor. Then prove that the components of cofactor
1
(cof T )ij = imn jpq Tmp Tnq . (27)
2

Proof. Let us consider the ij th component of cofactor,


(cof T )ij = ei · (cof T )ej (By definition of ij th component of tensor)
1 1
= ei · (cof T ( jpq ep × eq )) (Using an identity, ek = kmn em × en )
2 2
1
= jpq ei · (T ep × T eq ) (Using the definition of cofactor)
2
1
= jpq ei · (Tmp em × Tnq en )
2
1
= jpq ei · (Tmp em × Tnq en )
2
1 1
= jpq Tmp Tnq ei · (em × en ) = [ei , em , en ]jpq Tmp Tnq
2 2
1
= imn jpq Tmp Tnq
2
Therefore, the second invariant of tensor T is equal to the trace of cofactor, i.e., (cof T )ii =
1
  T T .
2 imn ipq mp nq

Relation between cofactor and determinant:


We have studied how the area transforms under the action of tensor. We now want to
study the transformation of volume. Recall from Lecture-3 that the volume of paral-
lelepiped formed with three vectors can be defined by the absolute value of triple scalar
product. Therefore, the volume can be defined under the action of tensor T as
[T u, T v, T w] = det(T )[u, v, w] (28)
where u, v, w are linearly independent vectors.

Thus, det(T ) represents the scale factor for the volume transformation under the action
of tensor T . Using the definition of scalar triple product in Eq. (28), we get
det(T )[u, v, w] = T u · (T v × T w)
=⇒ [det(T )u, v, w] = T u · ((cof T )(v × w)) (By definition of cofactor)
T
=⇒ det(T )Iu · (v × w) = (cof T ) T u · (v × w) (Using definition of transpose)
where I is the identity tensor. Arbitrariness of u, v and w implies
(cof T )T T = det(T )I (29)
Using Eq. (27) and Eq. (29), we can obtain the following expression for determinant in
indicial notation.
tr((cof T )T T ) = tr(det(T )I)
=⇒ (cof T )ij Tij = det(T )tr(I)
1
=⇒ det(T ) = imn jpq Tij Tmp Tnq (Since tr(I) = 3)
6
Summary of principal invariants:
In summary, we have the following principal invariants.

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

• First invariant or trace


IT = tr(T ) = Tii .

• Second invariant or trace of cofactor


1
IIT = tr(cof T ) = imn ipq Tmp Tnq .
2

• Third invariant or determinant


1
IIIT = det(T ) = imn jpq Tij Tmp Tnq .
6

The inverse of tensor:


The inverse of second-order tensor T , denoted by T −1 , is defined by

T −1 T = T T −1 = I (30)

where I is the identity tensor. If the determinant of tensor T is non-zero, i.e., det(T ) 6= 0
then there exists a unique inverse. Using Eq. (29), the inverse can be written as
1
T −1 = (cof T )T (31)
det(T )
Properties of inverse:
(i) If T is invertible and T u = v then u = T −1 v
(ii) T −1 (αu + βv) = αT −1 u + βT −1 v
The above properties show that the inverse is a second-order tensor.

Problem 6. Let u and v be arbitrary vectors. Then show that

cof (I + u ⊗ v) = (1 + u · v)I − v ⊗ u. (32)

Proof. Let x and y be two vectors. Then

cof (I + u ⊗ v)(x × y) = ((I + u ⊗ v)x) × ((I + u ⊗ v)y)


= (x + (u ⊗ v)x) × (y + (u ⊗ v)y)
= x × y + ((u ⊗ v)x) × y + x × ((u ⊗ v)y)
+((u ⊗ v)x) × ((u ⊗ v)y)

Since ((u ⊗ v)x) × ((u ⊗ v)y) = ((v · x)u) × ((v · y)u) = 0, we can write

cof (I + u ⊗ v)(x × y) = x × y + (v · x)u × y + x × (v · y)u


= x × y + u × ((v · x)y − (v · y)x)
= x × y − u × (v × (x × y))
= x × y − (u · (x × y))v + (u · v)x × y
= x × y + (u · v)x × y − (v ⊗ u)x × y
= (1 + (u · v))x × y − (v ⊗ u)x × y
= ((1 + (u · v))I − v ⊗ u)(x × y)

Arbitrariness of x and y implies Eq. (32).

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Problem 7 (Sherman-Morrison formula). Let T be an invertible tensor and let u


and v be two vectors such that 1 + v · T −1 u 6= 0. Then show that

T −1 (u ⊗ v)T −1
(T + u ⊗ v)−1 = T −1 − (33)
1 + u · T −1 v

Proof. Since given tensor T is invertible, we get


 −1
(T + u ⊗ v)−1 = T (I + T −1 (u ⊗ v)
 −1
= I + T −1 (u ⊗ v) T −1 (Since (RS)−1 = S −1 R−1 )
   −1
= I + T −1 u ⊗ v T −1 (Since R(x ⊗ y) = (Rx) ⊗ y)
 
T
[cof (I + (T −1 u) ⊗ v)]  −1
=  T (Using the definition of inverse)
det(I + (T −1 u) ⊗ v)

It is easy to see from the definition of determinant that det(I + x ⊗ y) = 1 + x · y. Using


the determinant formula and previous problem, we get

(T −1 u) ⊗ v
!
−1
(T + u ⊗ v) = I− T −1
1 + v · T −1 u
T −1 (u ⊗ v)
!
= I− T −1
1 + v · T −1 u
T −1 (u ⊗ v)T −1
= T −1 −
1 + v · T −1 u
This update-formula for inverse is known as Sherman-Morrison formula. Since the rank
of u ⊗ v is one, some authors refer this formula as rank-one update. This formula is also
applicable to higher dimensional vector space.

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 8


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-1: Tensor Algebra


Lecture-6: The Eigenvalues and Eigenvectors

We shall see in forthcoming lectures that the eigenvalues and eigenvectors play an impor-
tant role in obtaining the principal stresses and principal directions. Furthermore, it helps
in defining generalized strain measures. We note that the eigenvalues and eigenvectors
are also known as characteristic values and characteristic vectors or proper values and
proper vectors. We now define the eigenvalue and eigenvector.
Let T be a second-order tensor. If there exists a non-zero vector n and a scalar λ such
that
T n = λn, (1)
then the vector n is said to be an eigenvector and the scalar λ is said to be an eigenvalue.
Equation (1) is known as eigenvalue problem. We can also express Eq. (1) as (T −λI)n =
0. We now prove that there exists a non-zero vector n such that (T − λI)n = 0 if and
only if det(T − λI) = 0.

Problem 1. Let R be a second-order tensor. Then there is a non-zero vector u such that
Ru = 0 if an only if det(R) = 0.

Proof. If det(R) = 0 then from the definition of third invariant (see Lecture-5) we can
have [Rx, Ry, Rz] = 0 for arbitrary vectors x, y, z. Therefore, the vectors Rx, Ry, Rz
are linearly dependent i.e., there exist scalars α, β, γ, not all zero such that

αRx + βRy + γRz = R(αx + βy + γz) = Ru = 0 (2)

where u = αx + βy + γz is a non-zero vector.


Conversely, if there is a non-zero vector u such that Ru = 0. We now choose two
vectors v and w such that u, v and w are linearly independent. Using the definition of
third invariant, we get

det(R)[u, v, w] = [Ru, Rv, Rw] = [0, Rv, Rw] = 0 (3)

Equation (3) implies det(R) = 0 as the vectors u, v and w are linearly independent.
Thus, Ru = 0 for non-zero u ∈ V ⇐⇒ det(R) = 0.

Let T be a second-order tensor then preceding problem guarantees the existence of


eigenvector if
det(T − λI) = 0 (4)
This is known as characteristic equation and roots of this equation are called eigenvalues.

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Let u, v and w be linearly independent vectors. Then, using the definition of third
invariant, we can express Eq. (4) as

[T u − λu, T v − λv, T w − λw] = 0.

Applying properties of scalar triple product, we obtain

−λ3 {[u, v, w]} + λ2 {[T u, v, w] + [u, T v, w] + [u, v, T w]}


−λ{[T u, T v, w] + [u, T v, T w] + [T u, v, T w]} + [T u, T v, T w] = 0
=⇒ λ3 {[u, v, w]} − λ2 I1 {[u, v, w]} + λI2 {[u, v, w]} − I3 {[u, v, w]} = 0,

where I1 , I2 and I3 are principal invariants of tensor T . Since u, v, and w are linearly
independent, we get
λ3 − I1 λ2 + I2 λ − I3 = 0. (5)
The polynomial of λ shown in preceding equation is known as characteristic polynomial.
We note that all principal invariants are real as tensor components are real in any co-
ordinate frame. In other words, the coefficients of characteristic polynomial are real.
Thus, the roots of characteristic equation are either real or complex conjugates. Conse-
quently, the cubic equation shown in Eq. (5) has at least one real root. In other words,
every second-order tensor defined over three dimensional space possess at least one real
eigenvalue.

Relation between principal invariants and eigenvalues:


Let λ1 , λ2 , λ3 be roots of the characteristic equation. Recall the definition of principal
invariants from Lecture-5. Then the principal invariants can be written as

I1 = tr(T ) = λ1 + λ2 + λ3
I2 = tr(cof T ) = λ1 λ2 + λ2 λ3 + λ3 λ1 (6)
I3 = det(T ) = λ1 λ2 λ3

Numerically, these principal invariants do agree with our earlier definition of principal
invariants.

Let λ an n be an eigenvalue and corresponding eigenvector of tensor T . Then f (λ) is an


eigenvalue of f (T ) corresponding to the eigenvector n. For example, λ2 is an eigenvalue
to T 2 .
T 2 n = T (T n) = T (λn) = λT n = λ2 n.
Similarly, λn is an eigenvalue of T n corresponding to the eigenvector n. We now present
an important theorem called Cayley-Hamilton Theorem.

Problem 2 (Cayley-Hamilton Theorem). Let P (λ) = 0 is the characteristic equation


of a second-order tensor T then the tensor satisfies its own characteristic equation i.e.,

P (T ) = O,

where P (λ) = λ3 −I1 λ2 +I2 λ−I3 is the characteristic polynomial and I1 , I2 , I3 are principal
invariants of tensor T .

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Proof. Let u1 , u2 and u3 be linearly independent eigenvectors of given second-order tensor


T and corresponding eigenvalues are λ1 , λ2 , and λ3 , respectively. Using the definition of
eigenvalue problem, we have

T u1 = λ1 u1 , T u2 = λ2 u2 , T u3 = λ3 u3 . (7)

Let v be a vector. Then it is a linear combination of eigenvectors, i.e.,

v = α1 u1 + α2 u2 + α3 u3 , α1 , α2 , α3 ∈ <.

We now examine the following product using Eq. (7)

(T − λ1 I)(T − λ2 I)(T − λ3 I)v = (T − λ1 I)(T − λ2 I)(T v − λ3 v)


= (T − λ1 I)(T − λ2 I)(T (α1 u1 + α2 u2 + α3 u3 ) − λ3 (α1 u1 + α2 u2 + α3 u3 ))
= (T − λ1 I)(T − λ2 I)((α1 T u1 + α2 T u2 + α3 T u3 ) − (λ3 α1 u1 + λ3 α2 u2 + λ3 α3 u3 ))
= (T − λ1 I)(T − λ2 I)((α1 λ1 u1 + α2 λ2 u2 + α3 λ3 u3 ) − (λ3 α1 u1 + λ3 α2 u2 + λ3 α3 u3 ))
= (T − λ1 I)(T − λ2 I)(α1 (λ1 − λ3 )u1 + α2 (λ2 − λ3 )u2 )
= (T − λ1 I)(T (α1 (λ1 − λ3 )u1 + α2 (λ2 − λ3 )u2 ) − λ2 (α1 (λ1 − λ3 )u1 + α2 (λ2 − λ3 )u2 ))
= (T − λ1 I)((α1 λ1 (λ1 − λ3 )u1 + α2 λ2 (λ2 − λ3 )u2 ) − (λ2 α1 (λ1 − λ3 )u1 + λ2 α2 (λ2 − λ3 )u2 ))
= (T − λ1 I)(α1 (λ1 − λ2 )(λ1 − λ3 )u1 )
= (α1 (λ1 − λ2 )(λ1 − λ3 ))(T u1 − λ1 u1 )
= α1 (λ1 − λ2 )(λ1 − λ3 )(λ1 − λ1 )u1
= 0

Since the eigenvectors u1 , u2 and u3 are linearly independent and form a basis to vector
space. Consequently, (T − λ1 I)(T − λ2 I)(T − λ3 I)v = 0 is true for every vector v ∈ V.
Therefore, we have

(T − λ1 I)(T − λ2 I)(T − λ3 I) = O
3 2
T − (λ1 + λ2 + λ3 )T + (λ1 λ2 + λ2 λ3 + λ3 λ1 )T − (λ1 λ2 λ3 )I = O
T 3 − I1 T 2 + I2 T − I3 I = O (using Eq. (6)).

Hence, the theorem is proved if eigenvectors form a basis to the vector space irrespective
of whether eigenvalues are repetitive or not. This proof can be extended easily to any
second-order tensor defined over n-dimensional vector space where eigenvectors form a
basis to the vector space.
The tensor is said to be “defective” or “non-diagonalizable” if the eigenvectors of tensor
do not form basis to the vector space. The Cayley-Hamilton theorem still holds for the
defective tensors but the proof presented above does not take care of defective tensors.
Reader can refer any standard book on linear algebra for the general proof.

Geometrical interpretation of eigenvectors and eigenvalues:


The action of any second-order tensor on a vector can result in rotation, scaling or both
rotation and scaling. Let us consider the eigenvalue problem for a tensor T i.e., T n = λn
where n and λ are eigenvector and eigenvalue, respectively. It is easy to see from the
problem definition (see Eq. 1) that we are looking for the directions those not undergoing

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

rotation but scaling. We note that scaling includes multiplication with negative number
too. In general, any resultant vector that is collinear with original vector is considered as
scaling. For example u and −u are scaled vectors and not considered as rotated vectors.
Thus, geometrically, the eigenvectors or eigen-directions are the directions that do not
experience rotation under the action of tensor. Furthermore, the scale factors of unro-
tated directions represents corresponding eigenvalues. We present two simple examples
to understand this geometric interpretation.
Example 1. Let us consider the following matrix form of a rotation tensor R in canonical
basis. Show that the eigenvalues and eigen vectors are complex as every direction is
undergoing rotation. " √ √ #
1/√2 −1/√ 2
R=
1/ 2 1/ 2

Solution: Given tensor rotates every vector in the plane by π/4 angle as shown in Fig.
(1). Therefore, geometrically no vector exists without undergoing rotation. Let λ be an
eigenvalue. Then we get
det(R − λI) = 0
√ " √ #
1/ 2√− λ −1/ √ 2
=⇒ det = 0
1/ 2 1/ 2 − λ

=⇒ λ2 − 2λ + 1 = 0
1 1 1 1
=⇒ λ = √ + i √ or √ − i √
2 2 2 2

where i is the complex number −1. We now have following complex conjugate pair of
eigenvalues (λ1 and λ2 ) and corresponding eigenvectors (v1 and v2 ):
√i − √i2
( ) ( )
1 1 1 1
λ1 = √ + i √ , v1 = √1
2 , λ2 = √ − i √ , v2 = √1
2 2 2 2 2 2

Since we are getting complex eigenvectors it means that no real eigenvector exists. In
other words, there is no direction which is not undergoing rotation. It is consistent with
the geometric depiction of tensor action in Fig. (1).

e1 = (1,0)
e2 = (0,1) Ru1
u2 u1
R Re 2 e2 Re 1
e2
-e 1 e1
Ru2 e1
-e 2
-Re 1 -Re 2

Figure 1: Rotation transformation in 2-D space

Example 2. Consider the following shear transformation tensor in canonical basis. Show
that the eigenvectors make an angle of π/4 with basis.
" √ #
3/2 √1/2
S=
1/2 3/2

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Solution: Geometrically given shear transformation tensor and canonical basis e1 = (1, 0)
and e2 = (0, 1) are shown in Fig. (2). We get the following characteristic equation by
considering the eigenvalue problem of given tensor.

e1 = (1,0)
u2 e2 = (0,1)
u1 e2 Se2 Su1
e2 S Su2
-e 1 e1 Se1

-e 2 e1
-Se1

-Se2

Figure 2: Shear transformation in 2-D space


3/2 − λ √ 1/2
det(S − λI) = =0

1/2 3/2 − λ


q
2
=⇒ λ − (3)λ + 1/2 = 0
√ √
3+1 3−1
=⇒ λ = or
2 2
Eigenvalues (λ1 and λ2 ) and corresponding eigenvectors (u1 and u2 ) are
√ ( 1 ) √
√1
( )
3+1 √ 3 − 1 −
λ1 = , u1 = √12 , λ2 = , u2 = √1
2
2 2 2 2

The basis vectors and eigenvectors are shown in Fig. (2). The angles between eigenvectors
and basis are as follows:

cos(θu1 e1 ) = u1 · e1 = 1/ 2
=⇒ θu1 e1 = π/4

cos(θu2 e2 ) = u2 · e2 = 1/ 2
=⇒ θu2 e2 = π/4

It is clear from geometric picture shown in Fig. (2) and also algebraic procedure that the
angle between eigenvectors and canonical basis is π/4.
Newton identities:
Newton identity is a recursive formula for obtaining principal invariants of any second-
order tensor T defined over n-dimensional vector space. Let principal invariants are
Ik , k = 1, 2, · · · , n then recursive formula is given by

(−1)k+1
Ik = [trT k − I1 trT k−1 + · · · + (−1)k−1 Ik−1 trT ]
k
k−1
(−1)k+1
" #
tr T k + (−1)i Ii T k−i
X
= (8)
k i=1

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Applying this formula to second-order tensor in three dimensions, we get

I1 = tr(T )
1 1
I2 = − (tr(T 2 ) − I1 tr(T )) = ((trT )2 − tr(T 2 )) (9)
2 2
1 1
I3 = (tr(T ) − I1 tr(T ) + I2 tr(T )) = ((trT )3 − 3tr(T )tr(T 2 ) + tr(T 3 ))
3 2
3 6
Exponential tensor:
Exponential tensor can be defined in terms of series expansion or in terms of solution
to an ordinary differential equation. Here we define exponential of tensor through the
series expansion. In forthcoming discussion on tensor calculus, we show an alternative
definition. The exponential of tensor A is defined by
1 22 1 33
eAt := I + At + A t + A t + ··· (10)
2! 3!
We can easily verify by considering an example that taking exponential to every com-
ponent of tensor does not give the exponential of tensor. Furthermore, just computing
sufficiently large powers of tensor and taking summation as shown in Eq. (10) can provide
an approximation but not the exact components of tensor. We now show using eigenval-
ues and eigenvectors that closed form can be obtained for the exponential tensor. If the
eigenvectors of tensor form a basis to the vector space then we have,

A = P ΛP −1

where P is an invertible matrix with linearly independent eigenvectors as its columns and
Λ is a diagonal matrix with corresponding eigenvalues on the diagonal. From preceding
equation we get
A2 = P Λ2 P −1 , A3 = P Λ3 P −1 , · · ·
Substituting preceding terms in Eq. (10), we get
1 1
eAt = P P −1 + P ΛP −1 t +P Λ2 P −1 t2 + P Λ3 P −1 t3 + · · ·
2! 3!
1 22 1 33
= P (I + Λ + Λ t + Λ t + · · · )P −1
2! 3!
Λt −1
= Pe P

Since the product of diagonal matrix is diagonal, we can get following form for the eΛt ,

e λ1 t 0 ···
 
0
Λt

 0 e λ2 t ··· 0 

e =  .. .. .. .. 

 . . . .


0 0 · · · eλn t

Therefore, if we know eigenvalues and eigenvectors then the exponential of tensor can
be computed exactly using above relations. In forthcoming discussion, we show that
exponential of a skewsymmetric is rotation.
One can define exponential transformation for defective tensors where eigenvectors are
not sufficient to form basis to the vector space. Here we do not present the exponential

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

of defective tensors as they require the definition of Jordan form. We note that the eigen-
vectors and eigenvalues not only play an important role in exponential of diagonalizable
tensor but also in exponential of defective tensors.

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. C. Lanczos, Applied Analysis, 1989, Dover Publications Inc., New York.

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-1: Tensor Algebra


Lecture-7: The Skewsymmetric Tensor

Definition of skewsymmetric tensor:


A second-order tensor W is said to be skewsymmetric if W T = −W . Let u and v be
arbitrary vectors in vector space V then

(u, W v) = (W T u, v) = −(W u, v) = −(v, W u) (1)

Choosing u = v, we have

(u, W u) = −(u, W u)
=⇒ (u, W u) = 0, ∀u ∈ V (2)

From Eq. (2), we conclude that the vector W u is either orthogonal to the vector u or
W u = 0. Choosing u = ei and v = ej in Eq. (1), we get

Wij = −Wji .

Clearly, Wij = 0 if i = j. Let {e1 , e2 , e3 } be canonical basis. Then the skewsymmetric


tensor can be represented by
 
0 W12 W13
W =  −W12 0 W23  (3)


−W13 −W23 0

Thus, any skewsymmetric tensor defined over three dimensional space has three indepen-
dent components.

Problem 1. Given any skewsymmetric tensor W ∈ V 2 , there exist a unique w ∈ V


corresponding to W such that

W u = w × u, ∀u ∈ V (4)

Conversely, every vector w ∈ V has association with a unique skewsymmetric second-order


tensor W such that Eq. (4) holds. The vector w is known as axial vector corresponding
to W .

Proof. Consider the determinant of skewsymmetric tensor W ,

det(W ) = det(W T ) = det(−W ) = (−1)3 det(W ) = −det(W )


=⇒ det(W ) = 0 (5)

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Since the determinant of tensor W is zero there must exist at least one zero eigenvalue.
Let p be a unit eigenvector corresponding to zero eigenvalue i.e., W p = 0. Consider
q and r are two mutually orthogonal unit vectors such that they are orthogonal to the
vector p. Thus, the set {p, q, r} forms an orthonormal basis to the vector space V. Since
W is a second-order tensor we can have the following representation.

W = W11 p ⊗ p + W12 p ⊗ q + W13 p ⊗ r + W21 q ⊗ p + W22 q ⊗ q + W23 q ⊗ r +


W31 r ⊗ p + W32 r ⊗ q + W33 r ⊗ r (6)

Since the set {p, q, r} is an orthonormal basis, the components of W can be written as

W11 = p · W p, W12 = p · W q, W13 = p · W r, W21 = q · W p, W22 = q · W q,


W23 = q · W r, W31 = r · W p, W32 = r · W q, W33 = r · W r.

Using Eq. (2), we get (p, W p) = (q, W q) = (r, W r) = 0. The relation W p = 0 give
rise to (q, W p) = (r, W p) = 0. In addition, using skewsymmetric nature of tensor W ,
we have

(p, W q) = (W T p, q) = −(W p, q) = −(q, W p) = 0


(p, W r) = (W T p, r) = −(W p, r) = −(r, W p) = 0.

In summary, the components of tensor in orthonormal basis {p, q, r}: W11 = (p, W p) =
0, W12 = −W21 = (p, W q) = 0, W13 = −W31 = (p, W r) = 0, W22 = (q, W p) = 0,
W33 = (r, W r) = 0, W23 = (q, W r) = −(r, W q) = −W32 . If W is non-zero then W23
and W32 are only non-zero components. Furthermore, W23 = −W32 . Substituting these
components in Eq. (6), we obtain

W = W23 (q ⊗ r − r ⊗ q)

Consider the following product,

W u = W23 (q ⊗ r − r ⊗ q)u
= W23 ((r · u)q − (q · u)r)
= W23 ((r × q) × u)

Let the basis {p, q, r} be a right hand coordinate frame, i.e., p = q × r, q = r × p,


r = p × q. Let us choose W23 = −γ then we have

W u = γ((q × r) × u) = (γp) × u. (7)

Let w = γp then
W u = w × u, ∀u ∈ V.
Thus, there is an axial vector w corresponding to every skewsymmetric tensor W .

Conversely, if the vector w is given then we need to show that there exists a corre-
sponding skewsymmetric tensor W . For every vector u ∈ V there is a vector v such
that
w × u = v.

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

There is an output vector for every input vector and hence, the operation with given
vector is a transformation from vector space to vector space. The linearity of the preceding
operation follows from the distributive property of cross product.
w × (αu + βv) = α(w × u) + β(w × v), ∀u, v ∈ V and ∀α, β ∈ <
Therefore, the cross product with a given vector w is a linear transformation. Conse-
quently, there should exist a second-order tensor corresponding to the operation. Let W
be a second-order tensor such that
W u = w × u, ∀u ∈ V.
Writing the preceding equation in indicial notation, we get
(W u)i = (w × u)i
Wij uj = ikj wk uj
Wij uj = −ijk wk uj
Arbitrariness of vector u implies Wij = −ijk wk . It is easy to see that Wij = −Wji . There-
fore, the second-order tensor W is a skewsymmetric tensor. Thus, there is a skewsym-
metric tensor W associated with a given vector w.

Component wise relations:


In summary, we have the following relations between the skewsymmetric tensor W and
its axial vector w.
Wij = −ijk wk (8)
1
wi = − ijk Wjk (9)
2
W = |w|(r ⊗ q − q ⊗ r) (10)
where q and r are two orthogonal unit vectors such that
w
=q×r
|w|
Eigenvalues of skewsymmetric tensor:
Let W and w be skewsymmetric tensor and its axial vector, respectively. Let I1 , I2 and
I3 be principal invariants of W . Then
I1 = Wii = ei · W ei = 0
1
I2 = imn ipq Wmp Wnq
2
1
= (δmp δnq − δmq δnp )Wmp Wnq (Using epsilon-delta identity)
2
1
= (Wmm Wnn − Wmn Wnm )
2
1
= Wmn Wmn (Since Wmm = Wnn = 0 and Wnm = −Wmn )
2
1
= (mni wi )(mnj wj ) (Using the relation Wpq = pqr wr )
2
1
= (2δij )wi wj (Using pqi pqj = 2δij )
2
= wi wi = w · w = |w|2
I3 = det(W ) = 0

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Using principal invariants, we get the following characteristic equation for the tensor W .

λ3 + |w|2 λ = 0 (11)

where λ is the eigenvalue. This cubic equation has one zero root and two purely imaginary

roots. If i is imaginary number −1 then the eigenvalues can be expressed as λ1 = 0,
λ2 = i|w| and λ3 = −i|w|. It can be verified that eigenvalues satisfy the characteristic
equation. Geometrically, there are no real eigenvectors for the skewsymmetric tensor
except the axial vector. Furthermore, the axial vector also maps to the zero vector as it
is corresponding to zero eigenvalue. In conclusion, the skewsymmetric tensor W assumes
a diagonal form if the field is extended to the complex numbers.
Skewsymmetric tensor in higher dimensional vector space:
Let us consider a skewsymmetric tensor in n-dimensional vector space. The following
property is true for any dimension,

(u, W u) = 0, ∀u ∈ V

From the geometrical interpretation, all eigenvalues are either complex numbers or zeros.
In other words either the directions undergo orthogonal rotation or the directions have
zero scale factors.
It is easy to see that any skewsymmetric tensor defined over n-dimensional vector space
has n(n − 1)/2 independent components. Consequently, the skewsymmetric tensor can
not be associated with a vector in the vector space other than three dimensional space.
Thus, three dimensional space has special property that every skewsymmetric tensor can
be associated with a unique vector and vice versa.

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-1: Tensor Algebra


Lecture-8: Representation of the Orthogonal Tensor

The orthogonal tensors appear in rotation transformation and also in transformation rule
of tensors. Hence, orthogonal tensors play an important role in continuum mechanics.
We now discuss the representation and properties of orthogonal tensors.
Definition of orthogonal tensor:
A second-order tensor Q is said to be orthogonal if QT = Q−1 , i.e.,
QT Q = QQT = I, (1)
where I is the identity tensor. We show in the following problem that length of vector
and angle between any two vectors are preserved under the action of orthogonal tensor,
i.e., inner product is preserved.
Problem 1. A tensor Q is orthogonal if and only if it preserves the inner product, i.e.,
(Qu, Qv) = (u, v), ∀u, v ∈ V. (2)

Proof. Assuming that Q is orthogonal,


(Qu, Qv) = (QT Qu, v) (Using the definition of transpose of tensor)
= (Iu, v) (Since Q is orthogonal tensor)
= (u, v).
Conversely, if (Qu, Qv) = (u, v), ∀u, v ∈ V then we want to show that Q is orthogonal
tensor.
(u, v) = (Qu, Qv) = (QT Qu, v)
=⇒ (Iu, v) = (QT Qu, v)
Arbitrariness of u and v implies I = QT Q. Therefore, Q is orthogonal tensor.

It is easy to see that the orthogonal tensor preserves the length of vector by choosing
u = v in Eq. (2), i.e.,
|Qu| = |u|. (3)
Using Eq. (1), we get
det(QT Q) = det(I)
=⇒ det(QT )det(Q) = 1 (Since det(RS) = det(R)det(S))
=⇒ (det(Q))2 = 1 (Since det(T T ) = det(T ))
det(Q) = ±1 (4)

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

If the determinant of orthogonal tensor is +1 then it is called proper orthogonal or ro-


tation. The set of all proper orthogonal tensors are denoted by Orth+ . Geometrically,
the rotation preserves any right-handed coordinate frame, i.e., if the set {e1 , e2 , e3 } is
a right-handed orthonormal frame (i.e., e1 × e2 = e3 , e2 × e3 = e1 and e3 × e1 = e2 )
then {Qe1 , Qe2 , Qe3 } is also a right-handed orthonormal frame. If the determinant of
orthogonal tensor is −1 then it is known as improper orthogonal tensor and it is not a
rotation. An improper orthogonal tensor can be converted to rotation by multiplying
with −I when the vector space dimension is odd. Whereas in even dimensional space it is
not possible to convert improper to proper orthogonal. For example, in three dimensional
space as det(−Q) = (−1)3 det(Q) = −det(Q). Hence, in two dimensional space it is not
possible to convert an improper orthogonal to a rotation by a simple multiplication with
−I.

The rotation tensor preserves the length of vector and angle between vectors as shown in
Eqs. (2) and (3). We now show that the rotation tensor preserves the area of parallelogram
formed by two vectors.

Problem 2. Show that Q(u × v) = Qu × Qv, ∀u, v ∈ V, where Q is a rotation tensor.

Proof. det(Q) = 1 for the given rotation tensor. Using the relation between cofactor and
determinant, we get
cof Q = (det(Q))Q−T = Q.
Since the inverse of transpose is equal to the transpose of inverse for any arbitrary in-
vertible second-order tensor, i.e., (QT )−1 = (Q−1 )T , we use the notation to indicate Q−T
either cases. Note that this notation used in the above equation. From the definition of
cofactor, we have

Qu × Qv = (cof Q)(u × v) = Q(u × v), ∀u, v ∈ V. (5)

Problem 3. Show that any rotation tensor Q defined over three dimensional space always
possesses a unit eigenvalue.

Proof. Since Q is rotation tensor, we have QT Q = I and also det(Q) = 1. Using the
relation between orthogonal and identity tensors, we get

QT (Q − I) = I − QT = −(QT − I) = −(Q − I)T


=⇒ det(QT (Q − I)) = det(−(Q − I)T )
=⇒ det(QT )det(Q − I) = (−1)3 det((Q − I)T )
=⇒ det(Q)det(Q − I) = −det(Q − I)
=⇒ det(Q − I) = −det(Q − I)
=⇒ det(Q − I) = 0

Comparing with eigenvalue problem, we conclude that there is a unit eigenvalue to the
rotation tensor. The eigenvector corresponding to the unit eigenvalue is indeed the axis
of rotation as the direction and length of vector are preserved.

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Geometrically, we know that rotation of any (rigid) body takes place about a fixed
axis known as axis of rotation (see Fig. 1). As previously mentioned, the axis of rotation
indeed the eigenvector corresponding to a unit eigenvalue. In fact, the rotation tensor can
be represented using axis of rotation which is discussed in the following problem.

Rigid body

Axis of rotation

Figure 1: Rotation of a body about an axis

Problem 4. Show that any rotation tensor Q can be represented by the following form,

Q = p ⊗ p + α(q ⊗ q + r ⊗ r) + β(r ⊗ q − q ⊗ r) (6)

where p is the eigenvector corresponding to the unit eigenvalue, q and r are two orthonor-
mal vectors such that the set {p, q, r} is an orthonormal basis to the vector space V, and
α, β are two scalars such that α2 + β 2 = 1.

Proof. As shown in Problem-3, rotation tensor possess a unit eigenvalue. Let p be a


unit eigenvector corresponding to the unit eigenvalue. We now choose two mutually
orthonormal vectors q and r such that the set {p, q, r} forms an orthonormal basis to
the vector space V. Since the orthogonal tensor Q is a second-order tensor, we have the
following representation.

Q = Q11 p ⊗ p + Q12 p ⊗ q + Q13 p ⊗ r + Q21 q ⊗ p + Q22 q ⊗ q + Q23 q ⊗ r +


Q31 r ⊗ p + Q32 r ⊗ q + Q33 r ⊗ r. (7)

Since the set {p, q, r} is an orthonormal basis, we can write components

Q11 = p · Qp, Q12 = p · Qq, Q13 = p · Qr, Q21 = q · Qp, Q22 = q · Qq,
Q23 = q · Qr, Q31 = r · Qp, Q32 = r · Qq, Q33 = r · Qr. (8)

Since p is an eigenvector corresponding to unit eigenvalue and also orthogonal to q and


r, we can have

Qp = p, q · p = 0 and r · p = 0 =⇒ p · Qp = 1, q · Qp = 0 and r · Qp = 0. (9)

Using inner product preserving nature of orthogonal tensor (i.e., using Eq. (2)), we get

p·q =0 =⇒ Qp · Qq = 0 =⇒ p · Qq = 0, (10)
p·r =0 =⇒ Qp · Qr = 0 =⇒ p · Qr = 0. (11)

The above relations (i.e., p·Qq = 0 and p·Qr = 0) show that Qq and Qr are orthogonal
to the vector p. Therefore, the vectors Qq and Qr are linear combination of q and r,
i.e.,
Qq = α1 q + β1 r and Qr = α2 q + β2 r, (12)

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

where α1 , β1 , α2 , β2 are scalars.


We get the following relation from the orthogonality of q and r,

q · r = 0 =⇒ Qq · Qr = 0,
α1 α2 + β1 β2 = 0. (13)

Using determinant of rotation, we get

[Qp, Qq, Qr] = det(Q)[p, q, r]


=⇒ [p, α1 q + β1 r, α2 q + β2 r] = 1
α1 β2 [p, q, r] + β1 α2 [p, r, q] = 1
α1 β2 − β1 α2 = 1 (14)

Since rotation tensor preserves the length, we get

Qq · Qq = 1 =⇒ α12 + β12 = 1 (15)


Qr · Qr = 1 =⇒ α22 + β22 = 1 (16)

Equations (13) and (14) imply

α2 (α1 α2 + β1 β2 ) + β2 (α1 β2 − β1 α2 ) = β2
α1 α22 + α2 β1 β2 + α1 β22 − α2 β1 β2 = β2
α1 (α22 + β22 ) = β2
α 1 = β2 (Using Eq. (16)) (17)
β2 (α1 α2 + β1 β2 ) − α2 (α1 β2 − β1 α2 ) = −α2
α1 α2 β2 + β1 β22 − α1 α2 β2 + β1 α22 = −α2
β1 (α22 + β22 ) = −α2
β1 = −α2 (Using Eq. (16)). (18)

Let us choose α1 ≡ α and β1 ≡ β. Then Eqs. (12), (17) and (18) imply

q · Qq = α, r · Qr = α, q · Qr = −β and r · Qq = β. (19)

Substitution of Eqs. (9), (10), (11) and (19) in Eq. (8), yields

Q11 = 1, Q12 = 0, Q13 = 0, Q21 = 0, Q22 = α, Q23 = −β, Q31 = 0, Q32 = β, Q33 = α.
(20)
With substitution of these components (i.e., Eq. (20)) in Eq. (7) give rise the final form
of rotation tensor shown in Eq. (6).

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. P. Chadwick, Continuum Mechanics: Concise Theory and Problems, 1999, Dover


Publications Inc., New York.

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-1: Tensor Algebra


Lecture-9: Relation between
Skewsymmetric and Orthogonal Tensors

In this lecture, we present an important relation between skewsymmetric and orthogonal


tensors known as Rodrigues formula.
Problem 1 (Rodrigues formula). Let the axis of rotation tensor Q and the axial vector
of skewsymmetric tensor W be equal. If the magnitude of axial vector corresponding to
the skewsymmetric is unity then we have the following relation between Q and W .

Q = I + sin θ W + (1 − cos θ) W 2 , (1)

where θ is the angle of rotation.

Proof. We have the following representation for rotation tensor Q from Problem-4 of
Lecture-8.
Q = p ⊗ p + α(q ⊗ q + r ⊗ r) + β(r ⊗ q − q ⊗ r), (2)
where α2 + β 2 = 1, p is the eigenvector corresponding to unit eigenvalue (i.e., the axis
of rotation) and the set {p, q, r} is the orthonormal basis. Since p is the axial vector of
skewsymmetric tensor W , we have the following form (see Eq. (10) in Lecture-7).

W = r ⊗ q − q ⊗ r. (3)

Taking square to the tensor W , we get

W 2 = (r ⊗ q − q ⊗ r)(r ⊗ q − q ⊗ r)
= (r ⊗ q)(r ⊗ q) − (q ⊗ r)(r ⊗ q) − (r ⊗ q)(q ⊗ r) + (q ⊗ r)(q ⊗ r)
= (q · r)r ⊗ q − (r · r)q ⊗ q − (q · q)r ⊗ r + (r · q)q ⊗ r

To obtain the last step we use the relation (a ⊗ b)(c ⊗ d) = (b · c)(a ⊗ d)) (see Module-1:
Tensor Algebra exercise 2(ii)). Since q and r are orthonormal vectors, we can write

W 2 = −(q ⊗ q + r ⊗ r) (4)

Since the identity tensor I = p ⊗ p + q ⊗ q + r ⊗ r (see Eq. (21) in Lecture-4), we can


have
p ⊗ p = W 2 + I. (5)
Let θ be an angle of rotation. Then the angle between vectors q and Qq is θ. Furthermore,
the angle between vectors r and Qr is also θ. Therefore, Eq. (2) implies α = q · Qq =
r · Qr = cos θ and β = sin θ as α2 + β 2 = 1. Upon substituting Eqs. (3), (4) and (5)
along with α = cos θ and β = sin θ in Eq. (2), we get Eq. (1).

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Problem 2. Show that eW θ is an orthogonal tensor, where W is a skewsymmetric tensor


corresponding to a unit axial vector w and θ is a scalar.

Proof. Since W is a second-order tensor, using definition of exponential tensor (see Eq.
(10) in Lecture-6), we get
1 2 2 1 3 3 1 4 4 1 5 5 1 6 6
eW θ = I + W θ + W θ + W θ + W θ + W θ + W θ + ··· (6)
2! 3! 4! 5! 6!
We know from Lecture-7 that the characteristic equation for the skewsymmetric tensor is
λ3 + |w|λ = 0. Using Cayley-Hamilton theorem, we get

W 3 = −W

as the magnitude of axial vector is unit, i.e., |w| = 1. Therefore, we can have the higher
powers of W
W 4 = −W 2 , W 5 = W , W 6 = W 2 , · · ·
Substituting higher powers in exponential tensor, we obtain
1 3 1 5 1 7 1 1 1 1
eW θ = I + (θ − θ + θ − θ + · · · )W + ( θ2 − θ4 + θ6 − θ8 + · · · )W 2
3! 5! 7! 2! 4! 6! 8!
1 2 1 4 1 6 2
= I + sin θ W + (1 − (1 − θ + θ − θ + · · · ))W
2! 4! 6!
2
= I + sin θ W + (1 − cos θ)W
 T
To show orthogonality of exponential tensor, it is sufficient to prove eW θ eW θ = I.
 T
eW θ eW θ = (I + sin θ W + (1 − cos θ)W 2 )(I + sin θ W + (1 − cos θ)W 2 )T
= (I + sin θ W + (1 − cos θ)W 2 )(I − sin θ W + (1 − cos θ)W 2 )
= I + sin θ W + (1 − cos θ)W 2 − sin θ W − sin2 θ W 2 − (1 − cos θ)sin θ W 3
+(1 − cos θ)W 2 + sin θ (1 − cos θ)W 3 + (1 − cos θ)2 W 4
= I + 2(1 − cos θ)W 2 − sin2 θ W 2 − (1 − cos θ)2 W 2 (since W 4 = −W 2 )
= I + 2(1 − cos θ)W 2 − sin2 θ W 2 − (1 + cos2 θ − 2cos θ)W 2
= I + (1 − sin2 θ)W 2 − cos2 θ W 2
= I

Hence, eW θ is an orthogonal tensor. It can be shown that the determinant of exponential


tensor is positive. Thus, eW θ is a rotation tensor where θ is the angle of rotation.

Problem 3. Show that first invariant (trace) and second invariant (trace of cofactor) of
rotation tensor Q are equal to 1 + 2cos θ if the angle of rotation is θ.

Proof. Let p be the axis of rotation tensor Q. Let the set of vectors {p, q, r} be an
orthonormal basis. Then, using Eq. (2), we can write the rotation tensor

Q = p ⊗ p + cos θ (q ⊗ q + r ⊗ r) + sin θ (r ⊗ q − q ⊗ r), (7)

where θ is an angle of rotation. Using the definition of trace (see Lecture-5), we can write

tr(Q) = [Qp, q, r] + [p, Qq, r] + [p, q, Qr].

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Using the representation of rotation tensor shown in Eq. (7), the following relations can
be obtained.

Qp = p (Since (u ⊗ v)w = (v · w)u, q · p = r · p = 0)


Qq = cos θ q + sin θ r
Qr = cos θ r − sin θ q

Substituting Qp, Qq, Qr in trace of rotation, we get

tr(Q) = [p, q, r] + [p, cos θ q + sin θ r, r] + [p, q, cos θ r − sin θ q]


= [p, q, r] + [p, cos θ q, r] + [p, q, cos θ r]
= [p, q, r] + cos θ[p, q, r] + cos θ[p, q, r]
= 1 + 2 cos θ

Since the second invariant is the trace of cofactor, we can write


 
I2 = tr(cof Q) = tr (det(Q)) Q−T
 
−1 T

= tr Q (Since det(Q) = 1)
 T 
= tr QT (Since Q−1 = QT )
= tr(Q) = 1 + 2 cos θ

Thus, the lower and upper bounds on first and second invariants are -1 and 3, i.e., −1 ≤
I1 = I2 ≤ 3.

The orthogonal tensor not only useful in representing the rotation but also useful in
tensor transformation law. In fact the direction cosines that appear in tensor transforma-
tion are components of an orthogonal tensor. Following problem addresses the orthogonal
tensor that can transform one orthogonal basis to another orthogonal basis.

Problem 4. Let {e01 , e02 , e03 } and {e∗1 , e∗2 , e∗3 } be two sets of orthonormal basis to the vector
space V. In addition, both sets represents right hand coordinate system, i.e., e01 × e02 = e03
and e∗1 × e∗2 = e∗3 . Then the following tensor Q is a proper orthogonal tensor.

Q = e01 ⊗ e∗1 + e02 ⊗ e∗2 + e03 ⊗ e∗3 . (8)

This form of tensor is useful in transformation of tensors under change of basis.

Proof. {e01 , e02 , e03 } and {e∗1 , e∗2 , e∗3 } are given sets of two orthonormal basis to the vector
space V. Then we want to show that the following second-order is an orthogonal tensor.

Q = e01 ⊗ e∗1 + e02 ⊗ e∗2 + e03 ⊗ e∗3 .

It is sufficient to show that QT Q = I. Using properties of tensor product, we have

QT = e∗1 ⊗ e01 + e∗2 ⊗ e02 + e∗3 ⊗ e03 .

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Using above two relations, we get

QT Q = (e01 ⊗ e∗1 + e02 ⊗ e∗2 + e03 ⊗ e∗3 )(e∗1 ⊗ e01 + e∗2 ⊗ e02 + e∗3 ⊗ e03 )
= (e01 ⊗ e∗1 )(e∗1 ⊗ e01 ) + (e01 ⊗ e∗1 )(e∗2 ⊗ e02 ) + (e01 ⊗ e∗1 )(e∗3 ⊗ e03 )
+(e02 ⊗ e∗2 )(e∗1 ⊗ e01 ) + (e02 ⊗ e∗2 )(e∗2 ⊗ e02 ) + (e02 ⊗ e∗2 )(e∗3 ⊗ e03 )
+(e03 ⊗ e∗3 )(e∗1 ⊗ e01 ) + (e03 ⊗ e∗3 )(e∗2 ⊗ e02 ) + (e03 ⊗ e∗3 )(e∗3 ⊗ e03 )
= e01 ⊗ e01 + e02 ⊗ e02 + e03 ⊗ e03
= I.

Thus, the tensor Q is an orthogonal tensor.

We can also show {e01 , e02 , e03 } = {Qe∗1 , Qe∗2 , Qe∗3 }. Since,

Q = QI
= Q(e∗1 ⊗ e∗1 + e∗2 ⊗ e∗2 + e∗3 ⊗ e∗3 )
= (Qe∗1 ) ⊗ e∗1 + (Qe∗2 ) ⊗ e∗2 + (Qe∗3 ) ⊗ e∗3 )
= e01 ⊗ e∗1 + e02 ⊗ e∗2 + e03 ⊗ e∗3 .

The given basis {e01 , e02 , e03 } and {e∗1 , e∗2 , e∗3 } are right hand coordinate system, i.e., [e01 , e02 , e03 ] =
1 and [e∗1 , e∗2 , e∗3 ] = 1. Since [e01 , e02 , e03 ] = [Qe∗1 , Qe∗2 , Qe∗3 ], by definition of determinant,
we get det(Q) = 1. Thus, Q is a rotation tensor or proper orthogonal tensor.

Transformation law:
Let {e1 , e2 , e3 } and {e01 , e02 , e03 } be two orthonormal basis. Then the following proper
orthogonal tensor can be constructed with help of both basis.

Q = e1 ⊗ e01 + e2 ⊗ e02 + e3 ⊗ e03 .

We can represent vector e0i in the basis {e1 , e2 , e3 } as

e0i = (e0i · ej )ej = Qij ej . (9)

On the other hand, we can also represent ei in the basis {e01 , e02 , e03 } as

ei = (ei · e0j )e0j = Qji e0j . (10)

Let vi0 be ith component of vector v in the basis {e01 , e02 , e03 } and vj be j th component of
vector v in the basis {e1 , e2 , e3 }. Then we have the following transformation law for the
components of vector.

vi0 = v · e0i = v · (Qij ej ) = Qij v · ej = Qij vj .

Similarly, let Tij0 be ij th component of second-order tensor T in the basis {e01 , e02 , e03 } and
Tmn be mnth component of tensor T in the basis {e1 , e2 , e3 }. Then the transformation
law for the components of tensor

Tij0 = e0i · T e0j = (Qim em ) · T (Qjn en ) = Qim Qjn em · T en = Qim Qjn Tmn

Let [v 0 ] and [T 0 ] be components of vector v and second-order tensor T in the basis


{e0i , i = 1, 2, 3}, respectively. Similarly, let [v] and [T ] be components of vector and
tensor in the basis {ei , i = 1, 2, 3}. Then we have the transformation law

[v 0 ] = Q[v] and [T 0 ] = Q[T ]QT . (11)

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Since QT = Q−1 , we can also have

[v] = QT [v 0 ] and [T ] = QT [T 0 ]Q. (12)

Thus, the proper orthogonal tensor play an important role in transformation of tensors.
We will be discussing another important class of second-order tensors called symmetric
tensors in the next lecture.

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-1: Tensor Algebra


Lecture-10: The Symmetric Tensor

The second-order symmetric tensors play an important role in continuum mechanics as


they represent stress, strain and rate of deformation.
Definition of symmetric tensor:
A second-order tensor S is said to be symmetric if

(u, Sv) = (Su, v), ∀u, v ∈ V.

We will now show two important properties of the symmetric tensors:

• All eigenvalues of symmetric tensor are real.

• The eigenvectors corresponding to distinct eigenvalues are orthogonal.

Problem 1. Show that all eigenvalues of symmetric tensor S are real numbers.

Proof. Extending the inner product over complex number field, we can define (u, v) =
ui v̄i , where v̄i is the complex conjugate of vi (ith component of vector v), and ui is the
ith component of vector u. It is easy to see that the positive definite property of inner
product is followed by the new definition, i.e., if u = v then (u, u) ≥ 0 for all vectors
in vector space. The new definition of inner product follows the conjugate symmetry
property, i.e., (u, v) = (v, u). It can also be easily verified that the inner product follows
linear property. Furthermore, if there is no imaginary part in the vectors, i.e., vectors
are defined over real field, then it reduces to the standard dot product. Using the new
definition of inner product, we get

(αu, v) = α(u, v) and (u, βv) = β̄(u, v), ∀u, v ∈ V

where β̄ is the complex conjugate of β. Since S is a real symmetric tensor, we have

(Su, v) = (u, S T v) = (u, Sv).

Let λ and n be an eigenvalue and its corresponding eigenvector of given symmetric tensor
S, respectively. Then

λ(n, n) = (λn, n) = (Sn, n) = (n, S T n) = (n, Sn) = (n, λn) = λ̄(n, n).

λ = λ̄, since n is a non-zero vector, i.e., (n, n) > 0. If complex number and its conjugate
are equal then the imaginary part is zero. Thus, eigenvalues of symmetric tensor are real
numbers.

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Problem 2. Show that eigenvectors corresponding to distinct eigenvalues are orthogonal


for the symmetric tensor.

Proof. Let λ1 and λ2 be two distinct eigenvalues of second-order symmetric tensor S. Let
n1 and n2 be eigenvectors corresponding to λ1 and λ2 . Then we have

Sn1 = λ1 n1 ,
Sn2 = λ2 n2 .

Using the definition of transpose, we can write

(Sn1 , n2 ) = (n1 , S T n2 ) = (n1 , Sn2 ).

Rewriting preceding equations, we get

λ1 (n1 , n2 ) = (λ1 n1 , n2 ) = (Sn1 , n2 ) = (n1 , Sn2 ) = (n1 , λ2 n2 ) = λ2 (n1 , n2 )


=⇒ (λ1 − λ2 )(n1 , n2 ) = 0.

We know that eigenvalues are distinct, i.e., λ1 − λ2 6= 0. Therefore, we have (n1 , n2 ) = 0


from the preceding equation. Thus, the eigenvectors n1 and n2 are orthogonal.

The symmetric tensor can be divided into three classes based on the eigenvalues: (i) all
three eigenvalues are distinct (ii) two eigenvalues are equal and other one is distinct (iii)
all three eigenvalues are equal. There is convenient representation for all the cases of
symmetric tensor known as spectral decomposition. We now discuss the spectral decom-
position for all the cases.

Problem 3. If all three eigenvalues are distinct then the symmetric tensor S can be
represented by the following form,

S = λ1 n1 ⊗ n1 + λ2 n2 ⊗ n2 + λ3 n3 ⊗ n3 , (1)

where λ1 , λ2 and λ3 are eigenvalues and n1 , n2 and n3 are corresponding eigenvectors.

Proof. We have shown in previous problem that the eigenvectors corresponding to dis-
tinct eigenvalues are orthogonal. Let n1 , n2 and n3 be unit eigenvectors then the set
{n1 , n2 , n3 } forms an orthonormal basis to the vector space. Since S is a second-order
tensor, it has the following representation,

S = S11 n1 ⊗ n1 + S12 n1 ⊗ n2 + S13 n1 ⊗ n3 + S21 n2 ⊗ n1 + S22 n2 ⊗ n2


+S23 n2 ⊗ n3 + S31 n3 ⊗ n1 + S32 n3 ⊗ n2 + S33 n3 ⊗ n3 . (2)

where Sij = ni · Snj .


Since n1 , n2 and n3 are eigenvectors of S, we have Sn1 = λ1 n1 , Sn2 = λ2 n2 and
Sn3 = λ3 n3 . Using these relations and orthogonality of n1 , n2 and n3 , we get ni ·Snj = 0
if i 6= j, n1 · Sn1 = λ1 , n2 · Sn2 = λ2 and n3 · Sn3 = λ3 . Substituting, coefficients Sij in
Eq. (2), we get
S = λ1 n1 ⊗ n1 + λ2 n2 ⊗ n2 + λ3 n3 ⊗ n3 .

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Problem 4. Let λ1 , λ2 and λ3 be eigenvalues of symmetric tensor S. Let λ2 = λ3 (i.e.,


two eigenvalues are equal) and λ1 6= λ2 . Let p be a unit eigenvector corresponding to λ1 .
Then show that the symmetric tensor can be represented by

S = λ1 p ⊗ p + λ2 (I − p ⊗ p). (3)

Proof. Since p is an eigenvector corresponding to an eigenvalue λ1 , we have Sp = λ1 p.


Let q be another eigenvector corresponding to repeated eigenvalue λ2 = λ3 . We choose a
set of orthonormal basis {n1 , n2 , n3 } such that n1 = p and n2 = q. We note that p and
q are orthogonal as they are eigenvectors corresponding to distinct eigenvalues. We have
the relations p · q = 0, p · n3 = 0 and q · n3 = 0.

The symmetric tensor S can have the general representation in the basis {n1 , n2 , n3 } as
shown in Eq. (2). Since n2 is an eigenvector corresponding λ2 , we can write Sn2 = λ2 n2 .
It is easy to see the relations S11 = n1 · Sn1 = λ1 , S22 = n2 · Sn2 = λ2 , S12 = n1 · Sn2 =
n1 · λ2 n2 = 0 and S21 = n2 · Sn1 = S T n2 · n1 = Sn2 · n1 = S12 . Similarly, we also have
S13 = S31 = n3 · Sn1 = n3 · λ1 n1 = 0 and S23 = S32 = n3 · Sn2 = n3 · λ2 n2 = 0

Substituting these coefficients in general form (i.e., in Eq. (2)), we get

S = λ1 p ⊗ p + λ2 q ⊗ q + S33 n3 ⊗ n3 .

Since the first invariant is equal to sum of eigenvalues. Hence, we get S33 = λ2 . Substi-
tuting S33 in above equation, we obtain

S = λ1 p ⊗ p + λ2 q ⊗ q + λ2 n3 ⊗ n3

We can easily see that not only p and q are eigenvectors but also n3 is an eigenvector.
Furthermore, the linear combination of q and n3 is also an eigenvector. Therefore, we
have two eigensubspaces: (i) subspace generated by vector p, (ii) orthogonal plane to p.
Replacing q with n2 in preceding equation, we get

S = λ1 p ⊗ p + λ2 (n2 ⊗ n2 + n3 ⊗ n3 ).

Substituting of n2 ⊗ n2 + n3 ⊗ n3 = I − n1 ⊗ n1 in above expression yields S =


λ1 p ⊗ p + λ2 (I − p ⊗ p).

Problem 5. If all eigenvalues are equal then we have following representation for sym-
metric tensor S.
S = λI, (4)
where λ is the eigenvalue and I is the identity tensor.

Proof. Let λ1 , λ2 and λ3 be eigenvalues. Since eigenvalues are equal, we choose a scalar λ
such that λ ≡ λ1 = λ2 = λ3 . Let p be an unit eigenvector corresponding to the eigenvalue
λ, i.e., Sp = λp. Let {n1 , n2 , n3 } be an orthonormal basis such that n1 = p. The general
representation of S in the basis {n1 , n2 , n3 } is shown in Eq. (2).

The coefficients in general representation S11 = n1 · Sn1 = p · Sp = p · λp = λ,


S21 = n2 · Sn1 = n2 · λn1 = 0 and S12 = n1 · Sn2 = S T n1 · n2 = Sn1 · n2 = S21 ,

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Similarly, S13 = S31 = n3 · Sn1 = n3 · λn1 = 0 and S23 = S32 . Substituting these
coefficients in general form, we get

S = λn1 ⊗ n1 + S22 n2 ⊗ n2 + S33 n3 ⊗ n3 + S23 (n2 ⊗ n3 + n3 ⊗ n2 ).

Using the definition of first and second principal invariants, we obtain

S22 + S33 = 2λ (from the first invariant)


2
S22 S33 − S23 = λ2 (from the third invariant)

Eliminating λ from preceding equations, we get

(S22 − S33 )2 = −4S23


2
.

We know that all components of tensor S are real. Therefore, equation implies S23 = 0
and S22 = S33 . From the trace of tensor, we conclude that S22 = S33 = λ. Substitution
of the coefficients in the general form yields

S = λn1 ⊗ n1 + λn2 ⊗ n2 + λn3 ⊗ n3 = λI.

We note that vectors n1 , n2 and n3 are eigenvectors. Furthermore, linear combination


of n1 , n2 and n3 is also an eigenvector. In conclusion, every vector in the vector space is
an eigenvector for the symmetric tensor if all the eigenvalues are equal.

Above three problems can be summarized into the following spectral decomposition of
symmetric tensor.

Problem 6 (Spectral decomposition). Every symmetric tensor S admits following


form,
S = λ1 e∗1 ⊗ e∗1 + λ2 e∗2 ⊗ e∗2 + λ3 e∗3 ⊗ e∗3 , (5)
where e∗1 , e∗2 and e∗3 are orthonormal eigenvectors.

Proof. Problems (3), (4) and (5) show the existence of at least an orthonormal set of
vectors such that they are eigenvectors of symmetric tensor S. Let e∗1 , e∗2 and e∗3 be
three orthonormal eigenvectors. Then we have the following result from the preceding
problems.
S = λ1 e∗1 ⊗ e∗1 + λ2 e∗2 ⊗ e∗2 + λ3 e∗3 ⊗ e∗3 ,
The unique representation of S is shown in previous three problems for the three different
cases.

Problem 7. There is an orthogonal tensor Q such that every symmetric tensor S assumes
diagonal form (i.e., matrix form of S is diagonal). The orthogonal transformation is given
by Q = ei ⊗e∗i , where {ei } is canonical basis and {e∗i } is another set of orthonormal basis.

Proof. Let us consider the form of S as shown in preceding problem of spectral decom-
position, i.e.,
3
λi e∗i ⊗ e∗i ,
X
S=
i=1

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

where {e∗i }, i = 1, 2, 3, is a set of three orthonormal eigenvectors of S. For a given any


orthogonal tensor Q, we can have
!
T
λi e∗i e∗i QT
X
QSQ = Q ⊗
i
  
λi (Qe∗i ) ⊗ (Qe∗i ) since T (a ⊗ b) R = (T a) ⊗ RT b
X
= (6)
i

Let {ei }, i = 1, 2, 3, be the canonical basis. Then we can define an orthogonal tensor (see
Problem-4 in Lecture-9)
Q = ei ⊗ e∗i .
such that it transforms the orthogonal basis {e∗1 , e∗2 , e∗3 } to the other orthogonal basis
{e1 , e2 , e3 }. It can also be represented by Qe∗i = ei . Upon applying this transformation
rule to Eq. (6), we get

QSQT = λi (Qe∗i ) ⊗ (Qe∗i ) =


X X
λi ei ⊗ ei = Λ,
i i

where Λ is the diagonal matrix with eigenvalues of S on its diagonal, i.e.,


 
λ1 0 0
 
Λ =  0 λ2 0 

. (7)
0 0 λ3

Conversely, for given eigenvalues and eigenvectors, we can recover symmetric matrix S.
Let Λ be the diagonal matrix with eigenvalues on the diagonal. Let Q be the orthogonal
matrix with orthonormal eigenvectors of S as its rows. Then the symmetric tensor is
S = QT ΛQ. Thus, recovering the symmetric tensor.

Problem 8 (Cayley-Hamilton theorem for symmetric tensor). Prove the Cayley-


Hamilton theorem for symmetric tensor using spectral decomposition.

Proof. Let S be a symmetric tensor. Then it is enough to show that the symmetric tensor
S satisfies the following (characteristic) equation.

S 3 − I1 S 2 + I2 S − I3 = O (8)

where O is zero tensor, and I1 , I2 , I3 are principal invariants of S.


We already proved the Cayley-Hamilton theorem for the case of diagonalizable tensors
(see Problem-2 in Lecture-6). The previous proof is directly applicable to symmetric
tensors as they are diagonalizable. Here, we present an easy proof using the spectral
decomposition of symmetric tensors. The spectral decomposition of S, S 2 and S 3 are
given by

S = λ1 e∗1 ⊗ e∗1 + λ2 e∗2 ⊗ e∗2 + λ3 e∗3 ⊗ e∗3


S 2 = λ21 e∗1 ⊗ e∗1 + λ22 e∗2 ⊗ e∗2 + λ23 e∗3 ⊗ e∗3
S 3 = λ31 e∗1 ⊗ e∗1 + λ32 e∗2 ⊗ e∗2 + λ33 e∗3 ⊗ e∗3

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Substituting these relations on the left hand side of Eq. (8), we get
P (S) = (λ31 − I1 λ21 + I2 λ1 − I3 )e∗1 ⊗ e∗1 + (λ32 − I1 λ22 + I2 λ2 − I3 )e∗2 ⊗ e∗2
+(λ33 − I1 λ23 + I2 λ3 − I3 )e∗3 ⊗ e∗3
= O.
The last step follows from the characteristic equation. We note that the eigenvalues are
roots of characteristic equation, i.e., λ3i − I1 λ2i + I2 λi − I3 = 0 for i = 1, 2, 3. Thus, the
theorem is proved.

We can prove an interesting result that eigenvector extremizes the quantity n · Sn,
where n is a unit vector and S is a symmetric tensor. Furthermore, the extremum value
is an eigenvalue. The proof of this result is presented in the properties of stress tensor
which will be discussed later.
Example 1. Consider following matrix representation of shear transformation S. Find
the eigenvalues and eigenvectors of shear transformation. Show that the eigenvectors are
mutually orthogonal and graphically depict the extremum directions.
" √ #
3/2 1/2
S= √
1/2 3/2

Solution: Given shear transformation is a symmetric tensor and it can be easily observed
from above matrix representation. The given shear transformation can be graphically
represented as shown in Fig. (1).

e1 = (1,0)
u2 e2 = (0,1)
u1 e2 Se2 Su1
e2 S Su2
-e 1 e1 Se1

-e 2 e1
-Se1

-Se2

Figure 1: Shear transformation in 2-D space

Using the characteristic equation, we get



3/2 − λ 1/2
det(S − λI) = √
=0
1/2 3/2 − λ
q
=⇒ λ2 − (3)λ + 1/2 = 0
√ √
3+1 3−1
=⇒ λ = or
2 2
Let λ1 and λ2 be eigenvalues and let u1 and u2 be corresponding eigenvectors. Then we
can write
√ 
 √1 
 √ 
 − √1 

3+1 3 − 1
λ1 = , u1 =  12  , λ2 = , u2 =  1 2 
2 √
2
2 √
2

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

It is easy to see that eigenvectors are mutually orthogonal as u1 · u2 = 0. This can also
be observed from the Fig. (1) that the eigenvectors are mutually orthogonal and u1 · Su1
is maximum and u2 · Su2 is minimum. We note that the saddle point may also exist as
an extremum in three-dimensional case along with maximum and minimum.

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-1: Tensor Algebra


Lecture-11: Positive Definite Tensor and
the Polar Decomposition

Here, we discuss the positive definiteness and polar decomposition of second-order tensors.
The polar decomposition plays an important role in defining generalized strain and also
in the applications of principle of material frame-indifferent.
Definition of positive definite tensor:
A second-order tensor S is said to be positive definite if

(u, Su) ≥ 0, ∀u ∈ V with (u, Su) = 0 if and only if u = 0. (1)

Any second-order tensor T can be uniquely decomposed into a symmetric part Ts , and a
skew-symmetric part Tss , i.e., T = Ts + Tss where Ts = 12 (T + T T ) and Tss = 21 (T − T T )
(see Lecture-4). Therefore, we have

(u, T u) = (u, Ts u) + (u, Tss u)


= (u, Ts u) (Since (u, Tss u) = 0 ∀u ∈ V)

Thus, it is sufficient to study the positive definiteness of symmetric tensors.

Problem 1. Given a symmetric tensor S is positive definite if and only if all eigenvalues
are positive.

Proof. Let us assume that the given tensor S is positive definite tensor. Then we have

(u, Su) > 0, for every non-zero vector u ∈ V

Let λ and n be an eigenvalue and its corresponding eigenvector of S. Then we have the
relation Sn = λn. Taking inner product with n on both sides give rise to

(n, Sn)
λ= (2)
(n, n)

Since (n, n) > 0 and S is positive definite tensor (i.e., (n, Sn) > 0), Eq. (2) implies
λ > 0. Thus, positive definite property of tensor S implies positive nature of eigenvalues.
Conversely, let all eigenvalues of symmetric tensor S be positive. Then using spectral
decomposition (see Problem-6 in Lecture-10), we can have

S = λ1 e∗1 ⊗ e∗1 + λ2 e∗2 ⊗ e∗2 + λ3 e∗3 ⊗ e∗3

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

where λ1 , λ2 , λ3 are eigenvalues and e∗1 , e∗2 , e∗3 are corresponding orthonormal eigenvectors.
Let u be a non-zero vector. Then

Su = (λ1 e∗1 ⊗ e∗1 + λ2 e∗2 ⊗ e∗2 + λ3 e∗3 ⊗ e∗3 )u


= λ1 (e∗1 ⊗ e∗1 )u + λ2 (e∗2 ⊗ e∗2 )u + λ3 (e∗3 ⊗ e∗3 )u
= λ1 (e∗1 · u)e∗1 + λ2 (e∗2 · u)e∗2 + λ3 (e∗3 · u)e∗3
= λ1 u∗1 e∗1 + λ2 u∗2 e∗2 + λ3 u∗3 e∗3

where u∗1 , u∗2 , u∗3 are components vector in the basis {e∗1 , e∗2 , e∗3 }.

Taking inner product of u and Su, we get

(u, Su) = u · (λ1 u∗1 e∗1 + λ2 u∗2 e∗2 + λ3 u∗3 e∗3 )


= λ1 u∗1 (u · e∗1 ) + λ2 u∗2 (u · e∗2 ) + λ3 u∗3 (u · e∗3 )
= λ1 (u∗1 )2 + λ2 (u∗2 )2 + λ3 (u∗3 )2

The positive property of eigenvalues implies (u, Su) > 0. Suppose, (u, Su) = 0 then
from preceding equation u∗i = 0, which implies that u = 0. Thus, S is positive definite
tensor if and only if all eigenvalues are positive.

Problem 2. All principal invariants of symmetric tensor S are positive if and only if all
eigenvalues are positive.

Proof. Let λ1 , λ2 and λ3 be eigenvalues of S and I1 , I2 and I3 be three principal invariants.


Then we have the following relations between eigenvalue and principal invariants.

I1 = λ1 + λ2 + λ3
I2 = λ1 λ2 + λ2 λ3 + λ3 λ1
I3 = λ1 λ2 λ3

It is easy to see that if eigenvalues are positive then principal invariants are also positive.
Conversely, let principal invariants I1 , I2 and I3 be positive. The characteristic equation
of tensor S is given by
λ3 − I1 λ2 + I2 λ − I3 = 0,
where λ is an eigenvalue of tensor S. We know that eigenvalues of symmetric tensor are
real. Hence, if eigenvalue is negative then it cannot satisfy the characteristic equation as
the term λ3 − I1 λ2 + I2 λ − I3 is negative. Therefore, eigenvalue must be positive since the
principal invariants are positive. Thus, all principal invariants are positive if and only if
eigenvalues are positive.

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Problem 3 (Square root tensor). Let S be a positive definite symmetric tensor. Then
2
there exist a unique positive definite symmetric tensor U such
√ that U = S. The tensor
U is called square root tensor of S. We usually write U = S.

Proof. If U is not positive definite symmetric (P sym) tensor then there can be many U
satisfying U U = S such that S is P sym. For example, the matrices diag[1, -1, 1] (i.e.
diagonal matrix with entities 1, -1, 1 on the diagonal) and diag[1, -1, -1] are non-positive
definite square roots of the identity tensor I. Therefore, in order to ensure uniqueness,
we stated in the problem that U is the positive definite symmetric tensor. We now show
the relation between the given positive definite symmetric tensor and its positive definite
square root. Consider the spectral decomposition of given symmetric tensor S.

S = λ1 e∗1 ⊗ e∗1 + λ2 e∗2 ⊗ e∗2 + λ3 e∗3 ⊗ e∗3

where λ1 , λ2 and λ3 are eigenvalues and e∗1 , e∗2 and e∗3 are corresponding orthonormal
eigenvectors.
The eigenvalues are positive as S is a positive definite symmetric tensor. Therefore,
we can define a tensor
q q q
U= λ1 e∗1 ⊗ e∗1 + λ2 e∗2 ⊗ e∗2 + λ3 e∗3 ⊗ e∗3 . (3)
√ √
Since the set {e∗1 , e∗2 , e∗3 } is orthonormal, it is easy to see U U = S. Clearly, λ1 , λ2

and λ3 are eigenvalues and e∗1 , e∗2 and e∗3 are corresponding orthonormal eigenvectors
of U . Thus, there is a positive definite square root of S. We now show the uniqueness of
positive definite square root.
Let λ and n be the eigenvalue and corresponding eigenvector of S. Then we have

0 = (S − λI)n
= (U 2 − λI)n
√ √
= (U + λI)(U − λI)n.


Suppose (U − λI)n = ñ then we have

(U + λI)ñ = 0.

This equation√implies ñ = 0. Otherwise, the tensor U violates the positive definite


property as − λ is an eigenvalue. Therefore,

(U − λI)n = 0.

clear that the vector n is eigenvector corresponding to an eigenvalue λ. Thus,
It is √
U = S is unique as the spectral decomposition of S is unique.

We have discussed the additive decomposition of arbitrary tensor into symmetric and
skew-symmetric parts. We now discuss the multiplicative decomposition of invertible
tensors given by the polar decomposition theorem.

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

The following polar decomposition is motivated by a representation of complex num-


bers. Every non-zero complex number, z = x + iy, is expressible uniquely as z = reiθ

where r > 0 and θ ∈ [0, 2π). We note that |eiθ | = 1 and r = z̄z where z̄ is complex
conjugate of z.
Problem 4 (Polar decomposition theorem). Every invertible tensor F can be uniquely
factored into the following forms

F = QU = V Q, (4)

where Q is an orthogonal tensor and U , V are positive definite symmetric tensors.

Proof. Since given tensor F is invertible, F T F is positive definite symmetric tensor, i.e.,

(u, F T F u) = (F u, F u) > 0 for every non-zero vector u ∈ V

and (F u, F u) = 0 if and only if u = 0.



We can define a unique square root tensor, U = F T F (see Problem-3). Furthermore,
the square root tensor U is invertible as it is positive definite tensor. Therefore, we can
have
Q = F U −1 . (5)
It is clear from the following steps that Q is an orthogonal tensor.

QT Q = (F U −1 )T (F U −1 )
= U −T F T F U −1
= U −1 (F T F )U −1 (since U is symmetric)
−1 −1
= U (U U )U
= I

Thus, Eq. (5) implies first part of polar decomposition F = QU , where Q is orthogonal
tensor and U is positive definite tensor.
We now show the uniqueness of the decomposition F = QU . Suppose F has second
polar decomposition such that F = Q̆Ŭ where Q̆ is also orthogonal tensor. We can
obtain F T F = Ŭ T Q̆T Q̆Ŭ = Ŭ 2 and we know F T F = U 2 . Hence, uniqueness of square
root tensor implies Ŭ = U and Q̆ = Q. Furthermore, The det(Q) and det(F ) must have
same sign as det(U ) > 0. Clearly, the tensor Q is rotation as det(F ) is positive.
T
We note that invertibility of F implies invertibility
√ of F . Therefore, similar to pre-
vious case we can define a square root tensor V = F F T since F F T is positive defi-
nite symmetric tensor. We now define an orthogonal tensor Q̃ = V −1 F and it implies
F = V Q̃. It is easy to see that Q̃ is an orthogonal tensor, i.e., Q̃T Q̃ = I. We can have
the relation F = Q̃(Q̃T V Q̃) since Q̃Q̃T = I. We now show that Q̃ = Q. Since F = QU
and F = V Q̃, we can write

QU = V Q̃ =⇒ U = QT V Q̃.

Since U and V are symmetric tensors, taking transpose on both sides of above equation,
we get
U = Q̃T V Q =⇒ Q̃U = V Q

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

The uniqueness of F = QU factorization implies Q̃ = Q. Thus, we have following unique


decomposition
F = QU = V Q.
We note that the tensors U and V have same eigenvalues since U = QT V Q. If λ1 , λ2 ,
λ3 are eigenvalues and e∗1 , e∗2 , e∗3 are eigenvectors of U respectively then λ1 , λ2 , λ3 are
eigenvalues and Qe∗1 , Qe∗2 , Qe∗3 eigenvectors of V .

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-1: Tensor Algebra


Lecture-12: The Isotropic Functions

Isotropic functions are useful in bringing material symmetry in constitutive relations of


solids and fluids. Here, we define isotropic functions in general form over the domain of
scalar, vector and tensor (< × V × V 2 ).
Definition of isotropic functions:
Let Orth+ be a set of all proper orthogonal tensors and let any α ∈ <, v ∈ V, T ∈ V 2
then definitions are as follow:
(i) A scalar-valued function φ : < × V × V 2 → < is said to be isotropic if

φ(α, Qv, QT QT ) = φ(α, v, T ) ∀Q ∈ Orth+ . (1)

(ii) A vector-valued function g : < × V × V 2 → V is said to be isotropic if

g(α, Qv, QT QT ) = Qg(α, v, T ) ∀Q ∈ Orth+ . (2)

(iii) A tensor-valued function G : < × V × V 2 → V 2 is said to be isotropic if

G(α, Qv, QT QT ) = QG(α, v, T )QT ∀Q ∈ Orth+ . (3)

It is clear from the definitions that there are no restrictions on the scalar variables of
isotropic functions. Hence, the representation of isotropic functions of vectors and tensors
are presented in the following discussion.

Example 1. Let φ be scalar-valued isotropic function of vector v. Then show that the
function can be represented by φ(v) = φ̃(v · v).

Solution: Let V be an inner product space. If there exists a vector u ∈ V then the inner
product guaranties the existence of a scalar function such that ψ(v) = u · v, ∀v ∈ V. We
now decompose the vector u into two orthogonal parts: one part is along the direction of
the vector v and other part is orthogonal to v. Hence, the vector u can be written as

(u · v)
u= v + (u · v ⊥ )v ⊥ ,
(v · v)

where v ⊥ is unit orthogonal vector to the vector v.


Let α = (u · v)/(v · v) and β = u · v ⊥ then u = αv + βv ⊥ . Substituting u in the scalar
function ψ, we get

ψ(v) = (αv + βv ⊥ ) · v = αv · v (Since v ⊥ · v = 0)

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

For any constant α the function ψ(v) is an isotropic function, i.e., ψ(Qv) = αQv · Qv =
α(QT Qv) · v = αv · v = ψ(v). Since ψ is an isotropic function, we now define a more
general isotropic function ψ̃(ψ(v)) = φ̃(v · v). It is easy to see that the function φ̃ is an
isotropic function. Thus, the function φ(v) = φ̃(v · v) is isotropic.

Example 2. Let φ is the scalar-valued isotropic function of symmetric tensor S. Then


show that the function can be represented by φ(S) = φ̃(I1 , I2 , I3 ), where I1 , I2 and I3 are
principal invariants of symmetric tensor S.

Solution Let {e∗1 , e∗2 , e∗3 } be orthonormal eigenvectors of tensor S and λ1 , λ2 , λ3 be the
corresponding eigenvalues. Then we have the spectral representation

S = λ1 e∗1 ⊗ e∗1 + λ2 e∗2 ⊗ e∗2 + λ3 e∗3 ⊗ e∗3 .

Let ψ(S) be a scalar-valued function then


3
ψ(QSQT ) = ψ(Q( λi e∗i ⊗ e∗i )QT )
X

i=1
3
λi (Qe∗i ) ⊗ (Qe∗i ))
X
= ψ(
i=1

If the function ψ(S) is isotropic then Q is arbitrary in the previous relation. Hence, we
can write ψ(QSQT ) = ψ̃(λ1 , λ2 , λ3 ) = ψ(S). We know that eigenvalues and principal
invariants have one to one relation. Therefore, we can define a function ψ̃(λ1 , λ2 , λ3 ) =
φ̃(I1 , I2 , I3 ). Furthermore, it is easy to verify that the function φ̃(I1 , I2 , I3 ) is isotropic as
det(S − λI) = det(QSQT − λI). Thus, the function φ(S) = φ̃(I1 , I2 , I3 ) is an isotropic
function.

Example 3. Prove that the scalar-valued isotropic function has representation φ(v, S) =
φ̃(v · v, v · Sv, v · S 2 v, I1 , I2 , I3 ) where I1 , I2 , I3 are invariants of symmetric tensor S and
v is a vector.

Solution: Examples (1) and (2) show that v · v is an invariant of vector v; I1 , I2 and I3
are invariants of tensor S. We can also create the following invariants as combination of
the tensor S and the vector v under every orthogonal transformation.

v · Sv, v · S 2 v, v · S 3 v · · · , v · S n v, · · ·

It is evident from Cayley-Hamilton theorem that not all invariants are independent. Using
Cayley-Hamilton theorem, we have

S 3 − I1 S 2 + I2 S − I3 I = O,

where I is identity tensor and O is zero-tensor.


Multiplying with the vector v on both side of terms, we get

v · S 3 v − I1 v · S 2 v + I2 v · Sv − I3 v · Iv = 0
=⇒ v · S 3 v = I1 v · S 2 v − I2 v · Sv + I3 v · v

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

It is clear that v · S n v for n ≥ 3 can be expressed as v · S 2 v, v · Sv, v · v and principal


invariants of S. Any function of the following set of independent invariants is isotropic
function.
v · v, v · Sv, v · S 2 v, I1 , I2 , I3 .
Therefore, the isotropic function can be represented as

φ(v, S) = φ̃(v · v, v · Sv, v · S 2 v, I1 , I2 , I3 ).

It is easy to verify that the function φ̃ is isotropic function, i.e., φ(Qv, QSQT ) = φ(v, S)
for every orthogonal tensor Q.

Example 4. Show that there is no non-zero vector-valued isotropic function of scalar.

Solution: Let g(α) be a vector-valued isotropic function of scalar. Then, using the defini-
tion of isotropic function, we can have

g(α) = Qg(α) ∀Q ∈ Orth+ .

The above relation shows that the vector g is an axis of the rotation tensor Q. Arbitrari-
ness of Q implies g(α) = 0. Thus, the isotropic vector-valued function of scalar variable
is zero.

Example 5. Let g be a vector-valued isotropic function of vector v. Then show that g


has the representation, g(v) = φ̃(v · v)v.

Solution: If v = 0 then the isotropic function g(0) = Qg(0) for all orthogonal tensors Q.
This implies g(0) is an axis to every orthogonal tensor and hence, g(0) = 0. It is easy to
see that g(0) is isotropic function.
We now consider the vector v 6= 0 then we can write the vector g(v) as

g(v) = α(v)v + β(v)v ⊥ , (4)

where α(v) and β(v) are scalar functions and v ⊥ orthogonal vector to v. We now choose
the orthogonal tensor
(v ⊗ v)
Q=2 − I, (5)
(v · v)
where I is the identity tensor. Using Eqs. (4) and (5), we can write

g(Qv) = α(Qv)Qv + β(Qv)v ⊥ = α(v)v + β(v)v ⊥


Qg(v) = α(v)Qv + β(v)Qv ⊥ = α(v)v − β(v)v ⊥ .

Since g(v) is an isotropic function, equating above relations it is easy to see that β(v) = 0.
Furthermore, isotropy of g(v) implies

α(Qv) = α(v) ∀Q ∈ Orth+ .

The above relation implies that α(v) is an isotropic function. Hence, based on Example-
1 the function α(v) can be represented by φ̃(v · v). Thus, the vector-valued isotropic
function of vector admits the representation g(v) = φ̃(v · v)v.

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Example 6. Show that there is no non-zero vector-valued isotropic function of symmetric


tensor.

Solution: Let p, q and r be orthonormal eigenvectors of tensor S and corresponding


eigenvalues be λ1 , λ2 and λ3 . Choose an orthogonal tensor Q = 2p ⊗ p − I as it commutes
with S, i.e.,

QS = (2p ⊗ p − I)S = 2(p ⊗ p)S − S


= 2p ⊗ (S T p) − S = 2p ⊗ (Sp) − S (Since S is symmetric tensor)
= 2p ⊗ (λ1 p) − S = 2(λ1 p) ⊗ p − S
= 2(Sp) ⊗ p − S = SQ
T
=⇒ QSQ = SQQT = S

Let g(S) be a vector-valued isotopic function of second-order tensor S. Then, using the
definition of isotropic function, we have

g(QSQT ) = Qg(S)
=⇒ g(S) = Qg(S)
=⇒ g(S) = αp (Since p is the axis of Q)

Similarly, choosing an orthogonal tensor Q = 2q ⊗ q − I we can show that

g(S) = βq.

Any non-zero vector cannot be collinear with two orthogonal vectors. Therefore, the
isotropic function g(S) = 0. Conversely, it is easy to see that g(S) = 0 is an isotropic
function. Thus, the vector-valued isotropic function of symmetric tensor is zero.

Example 7. Let G is symmetric tensor-valued isotropic function of scalar α. Then show


that G(α) = λI where I is identity tensor and λ is some scalar.

Solution: Since given function G(α) is isotropic, using the definition of tensor valued
isotropic function, we can write

G(α) = QG(α)QT
=⇒ QG = GQ

Preceding relation shows that G commutes with every proper orthogonal tensor Q. Let
p be an axis of orthogonal tensor Q. Then we can have

QGp = GQp = Gp.

We conclude from the preceding equation that Gp is a vector along the axis of Q and
hence, Gp = λp where λ is some scalar. Arbitrariness of Q implies Gp = λp for
every vector p ∈ V. Since whole vector space is an eigenspace for G, we can have the
representation G = λI where I is identity tensor.
Thus, the symmetric tensor-valued isotropic function of scalar has the representation
G(α) = λI.

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Example 8. Let G be a symmetric tensor-valued, isotropic function of vector v. Then


show that the function has representation G(v) = φ̃0 (v · v)I + φ̃1 (v · v)v ⊗ v where φ̃0
and φ̃1 are scalar-valued isotropic functions and I is the identity tensor.

Solution: If v = 0 then the case is similar to the previous example because G(0) =
QG(0)QT . This implies φ̃0 = λ where λ is some scalar and φ̃1 = 0 in the representation.
Let us consider v 6= 0. Let us define a vector valued function g(v) = G(v)v where
G(v) is given isotropic tensor valued function. We now show that isotropy of G implies
isotropy of g.

g(Qv) = G(Qv)Qv
= QG(v)QT Qv (Since Q is isotropic)
= QG(v)v
= Qg(v).

Clearly, g(v) is isotropic function. Using Example-5, the vector-valued isotropic g is


represented by g(v) = α(v)v where α(v) is a scalar-valued isotropic function. Therefore,
we have
Gv = αv
This shows that v is an eigenvector of G(v). Let q and r be unit eigenvectors of G such
that the set {v/|v|, q, r} forms orthonormal basis to the vector space V. Let β and γ be
eigenvalues of G corresponding to q and r, respectively. Then we have
α(v)
G(v) = v ⊗ v + β(v)q ⊗ q + γ(v)r ⊗ r,
v·v
where β(v) and γ(v) are not only eigenvalues of G but also scalar-valued isotropic func-
tions of v because G is isotropic function of v.
We now choose proper orthogonal tensor
1
Q= v ⊗ v + r ⊗ q − q ⊗ r.
v·v
Using tensor Q, we can obtain following relations.

Q (q ⊗ q) QT = (Qq) ⊗ (Qq) = r ⊗ r, (6)


T
Q (r ⊗ r) Q = (Qr) ⊗ (Qr) = (−q) ⊗ (−q) = q ⊗ q. (7)

Furthermore, using spectral form of function G(v) and previously defined orthogonal
tensor Q, we get
α(Qv)
G(Qv) = (Qv) ⊗ (Qv) + β(Qv)q ⊗ q + γ(Qv)r ⊗ r
(Qv) · (Qv)
α(v)
= Q(v ⊗ v)QT + β(v)q ⊗ q + γ(v)r ⊗ r (using isotropy of α, β and γ)
(v · v)
α(v)
QGQT = Q(v ⊗ v)QT + β(v)Q (q ⊗ q) QT + γ(v)Q (r ⊗ r) QT
(v · v)
α(v)
= Q(v ⊗ v)QT + β(v)r ⊗ r + γ(v)q ⊗ q. (using Eqs. (6) and (7))
(v · v)

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

It is clear from the expressions G(Qv) and QGQT that isotropy of G implies β(v) = γ(v).
Therefore, we can write,
α(v)
G(v) = v ⊗ v + β(v)(q ⊗ q + r ⊗ r)
(v · v)
α(v) − β(v) 1
= v ⊗ v + β(v)( v ⊗ v + q ⊗ q + r ⊗ r)
(v · v) v·v
α(v) − β(v)
= β(v)I + v⊗v (8)
(v · v)
Since α(v) and β(v) are scalar-valued isotopic functions, using Example-1, we can have
the following representation.
α(v) − β(v)
β(v) = φ̃0 (v · v) and = φ̃1 (v · v).
(v · v)
Substituting the representation for β and (α − β)/(v · v) in Eq. (8), yields following final
form for symmetric tensor-valued isotropic function of vector

G(v) = φ̃0 (v · v)I + φ̃1 (v · v)v ⊗ v.

This, completes the proof.


We now present a central result in this lecture which is about representation of symmet-
ric tensor-valued isotropic function of symmetric tensor. This result plays an important
role in constitutive equations of both fluids and solids. The following representation for
symmetric tensor-valued isotropic function is known as famous Rivlin-Ericksen represen-
tation theorem.

Problem 1 (Rivlin-Ericksen representation theorem). Let G be a symmetric tensor-


valued function of positive definite symmetric tensor S. Then show that the function has
following representation,

G(S) = φ0 (IS )I + φ1 (IS )S + φ2 (IS )S 2 , (9)

where IS is set of principal invariants of S and φ0 , φ1 , and φ2 are scalar-valued isotropic


functions of S.

Proof. We first show that eigenvectors of S and G are same if G(S) is an isotropic
function. Let {e∗1 , e∗2 , e∗3 } and {λ1 , λ2 , λ3 } be a set of orthonormal eigenvectors and cor-
responding eigenvalues of symmetric tensor S. Then the spectral decomposition of S is
given by
S = λ1 e∗1 ⊗ e∗1 + λ2 e∗2 ⊗ e∗2 + λ3 e∗3 ⊗ e∗3 .
Choose a proper orthogonal tensor Q = 2e∗1 ⊗ e∗1 − I. The rotation tensor Q commutes
with S, i.e.,

QS = (2e∗1 ⊗ e∗1 − I)S = 2(e∗1 ⊗ e∗1 )S − S = 2e∗1 ⊗ (S T e∗1 ) − S


= 2e∗1 ⊗ (Se∗1 ) − S = 2e∗1 ⊗ (λ1 e∗1 ) − S = 2(λ1 e∗1 ) ⊗ e∗1 − S
= 2(Se∗1 ) ⊗ e∗1 − S = 2S(e∗1 ⊗ e∗1 ) − S = S(2e∗1 ⊗ e∗1 − I)
= SQ (10)

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

Since QQT = I, with post-multiplication of QT on both sides of Eq. (10), we get


QSQT = S. (11)
The isotropic nature of G(S) implies
G(QSQT ) = QG(S)QT
=⇒ G(S) = QG(S)QT (using Eq. (11))
=⇒ QG(S) = G(S)Q (Since QQT = I)
=⇒ QG(S)e∗1 = G(S)Qe∗1
=⇒ QG(S)e∗1 = G(S)e∗1 (Since Qe∗1 = e∗1 )
It is clear that G(S)e∗1 is a vector along the axis of Q. Therefore, G(S)e∗1 = µ1 e∗1 . This
implies that e∗1 is an eigenvector of tensor G. Similarly, choosing another orthogonal
tensor Q = 2e∗2 ⊗ e∗2 − I we can show that G(S)e∗2 = µ2 e∗2 and symmetric nature of
G implies G(S)e∗3 = µ3 e∗3 . Therefore, the symmetric tensor-valued function has the
following spectral decomposition,
G(S) = µ1 e∗1 ⊗ e∗1 + µ2 e∗2 ⊗ e∗2 + µ3 e∗3 ⊗ e∗3 . (12)
We now show that all eigenvalues of S are distinct then G admits the general form of
representation shown in Eq. (9). Furthermore, if there are any repeated eigenvalues then
it admits the form that is a special case of the general form.

Case-1: Let all eigenvalues of S be distinct, i.e., λ1 6= λ2 , λ2 6= λ3 and λ1 6= λ3 ,


then eigenvectors associated with eigenvalues are orthogonal. Since e∗1 , e∗2 and e∗3 are
orthonormal eigenvectors of S, we show that the tensors e∗1 ⊗ e∗1 , e∗2 ⊗ e∗2 and e∗3 ⊗ e∗3
are linearly independent in the space of tensors. If tensors are linearly independent then
αe∗1 ⊗ e∗1 + βe∗2 ⊗ e∗2 + γe∗3 ⊗ e∗3 = O should imply α = β = γ = 0. From the linear
combination of tensors, we get
e∗1 · (αe∗1 ⊗ e∗1 + βe∗2 ⊗ e∗2 + γe∗3 ⊗ e∗3 )e∗1 = e∗1 · Oe∗1
=⇒ αe∗1 · (e∗1 · e∗1 )e∗1 + βe∗1 · (e∗2 · e∗1 )e∗2 + γe∗1 · (e∗3 · e∗1 )e∗3 = e∗1 · 0
=⇒ α = 0
e∗2 · (αe∗1 ⊗ e∗1 + βe∗2 ⊗ e∗2 + γe∗3 ⊗ e∗3 )e∗2 = e∗2 · Oe∗2
=⇒ β = 0
e∗3 · (αe∗1 ⊗ e∗1 + βe∗2 ⊗ e∗2 + γe∗3 ⊗ e∗3 )e∗3 = e∗3 · Oe∗3
=⇒ γ = 0
Therefore, the set {e∗1 ⊗ e∗1 , e∗2 ⊗ e∗2 , e∗3 ⊗ e∗3 } is linearly independent and basis to the three
dimensional subspace of tensors. The linear combination results in a symmetric tensor
as the basis tensors are symmetric. We now show that the set {I, S, S 2 } is also another
basis to the subspace spanned by {e∗1 ⊗ e∗1 , e∗2 ⊗ e∗2 , e∗3 ⊗ e∗3 }. Since e∗1 , e∗2 and e∗3 are
orthonormal eigenvectors of S, we can have the following representation for I, S and S 2
(see spectral decomposition in Lecture-10).
I = e∗1 ⊗ e∗1 + e∗2 ⊗ e∗2 + e∗3 ⊗ e∗3 ,
S = λ1 e∗1 ⊗ e∗1 + λ2 e∗2 ⊗ e∗2 + λ3 e∗3 ⊗ e∗3 ,
S 2 = λ21 e∗1 ⊗ e∗1 + λ22 e∗2 ⊗ e∗2 + λ23 e∗3 ⊗ e∗3 .

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Clearly, the tensors I, S and S 2 belong to the subspace spanned by {e∗1 ⊗ e∗1 , e∗2 ⊗ e∗2 , e∗3 ⊗
e∗3 }. We further know that following determinant of Vandermonde matrix is non-zero.
 
1 1 1
 
det  λ 1 λ 2 λ 3
 = (λ1 − λ2 )(λ2 − λ3 )(λ3 − λ1 ) 6= 0 (since λ1 6= λ2 , λ2 6= λ3 , λ3 6= λ1 )
 
λ21 λ22 λ23

Therefore, the set {I, S, S 2 } is linearly independent and hence, forms another basis to
the space generated by {e∗1 ⊗ e∗1 , e∗2 ⊗ e∗2 , e∗3 ⊗ e∗3 }. In other words, any tensor in the space
of {e∗1 ⊗ e∗1 , e∗2 ⊗ e∗2 , e∗3 ⊗ e∗3 } can be represented by linear combination of set {I, S, S 2 }.
Consequently, the tensor-valued function shown in Eq. (12) admits the following form.

G(S) = α0 (S)I + α1 (S)S + α2 (S)S 2 ,

where α0 (S), α1 (S) and α2 (S) are scalar-valued isotropic functions of S. It is easy to
see G is isotropic function. Furthermore, Example-2 shows that α0 (S), α1 (S) and α2 (S)
can have the following representation.

α0 (S) = φ0 (I1 , I2 , I3 ), α1 (S) = φ1 (I1 , I2 , I3 ), α3 (S) = φ3 (I1 , I2 , I3 ),

where I1 , I2 and I3 are principal invariants of S.


Case-2: Let two eigenvalues of S be distinct, i.e., λ1 6= λ2 = λ3 and e∗1 , e∗2 and e∗3 be
orthonormal eigenvectors corresponding to eigenvalues. Then the symmetric tensor has
the spectral decomposition S = λ1 e∗1 ⊗ e∗1 + λ2 (e∗2 ⊗ e∗2 + e∗3 ⊗ e∗3 ). We note that the
subspace spanned by linear combination of e∗2 and e∗3 are also eigenvectors of S. As it
was shown earlier, every eigenvector of S is also eigenvector of G(S) and hence

Ge∗2 = µ2 e∗2 , Ge∗3 = µ3 e∗3 , G(e∗2 + e∗3 ) = µ(e∗2 + e∗3 ).

Combining preceding equations, we get

G(e∗2 + e∗3 ) = µ(e∗2 + e∗3 )


=⇒ Ge∗2 + Ge∗3 = µe∗2 + µe∗3
=⇒ µ1 e∗2 + µ2 e∗3 = µe∗2 + µe∗3 .

Taking dot product with e∗2 and e∗3 on both sides of preceding equation, we get µ2 = µ
and µ3 = µ, respectively. Therefore, G(S) has the following representation

G(S) = µ1 e∗1 ⊗ e∗1 + µ(e∗2 ⊗ e∗2 + e∗3 ⊗ e∗3 ).

Similar to previous case, we now show that {e∗1 ⊗ e∗1 , e∗2 ⊗ e∗2 + e∗3 ⊗ e∗3 } and {I, S} are
two sets of basis to two-dimensional subspace of tensors.

I = e∗1 ⊗ e∗1 + (e∗2 ⊗ e∗2 + e∗3 ⊗ e∗3 ),


S = λ1 e∗1 ⊗ e∗1 + λ2 (e∗2 ⊗ e∗2 + e∗3 ⊗ e∗3 )

and " #
1 1
det = (λ1 − λ2 ) 6= 0 (since λ1 6= λ2 )
λ1 λ2

Joint initiative of IITs and IISc – Funded by MHRD 8


NPTEL – Mechanical Engineering – Continuum Mechanics

Therefore, {e∗1 ⊗ e∗1 , e∗2 ⊗ e∗2 + e∗3 ⊗ e∗3 } and {I, S} are two sets of basis to the two-
dimensional subspace of tensors. Consequently, we can have,

G(S) = α0 (S)I + α1 (S)S,

where α0 (S) and α1 (S) are scalar-valued isotropic function of S. Furthermore, Example-2
shows that α0 and α1 can have the following representation

α0 (S) = φ0 (I1 , I2 , I3 ), α1 (S) = φ1 (I1 , I2 , I3 ),

where I1 , I2 and I3 are principal invariants of S.


Case-3: Let all eigenvalues of S be equal, i.e., λ1 = λ2 = λ3 . Then every vector is an
eigenvector of S which also implies every vector is also an eigenvector of G. Let us choose
λ ≡ λ1 = λ2 = λ3 . Then S can be representation by S = λI. Since {e∗1 , e∗2 , e∗3 } are a
set of orthonormal eigenvectors of S, they are also eigenvectors of G(S). Therefore, we
have Ge∗1 = µ1 e∗1 , Ge∗2 = µ2 e∗2 and Ge∗3 = µ3 e∗3 , where µ1 , µ2 and µ3 are eigenvalues of
G. Since every vector in vector space is an eigenvector of G(S), we can have another
relation G(e∗1 + e∗2 + e∗3 ) = µ(e∗1 + e∗2 + e∗3 ) where µ is eigenvalue corresponding to the
vector e∗1 + e∗2 + e∗3 . Similar to the previous case, combining these relations, we get

G(e∗1 + e∗2 + e∗3 ) = µ(e∗1 + e∗2 + e∗3 )


=⇒ Ge∗1 + Ge∗2 + Ge∗3 = µe∗1 + µe∗2 + µe∗3
=⇒ µ1 e∗1 + µ2 e∗2 + µ3 e∗3 = µe∗1 + µe∗2 + µe∗3

Taking dot product with e∗1 , e∗2 and e∗3 on both sides of preceding equation, we get µ1 = µ,
µ2 = µ and µ3 = µ. Therefore, the tensor-valued function has the following representation,

G(S) = µ(e∗1 ⊗ e∗1 + e∗2 ⊗ e∗2 + e∗3 ⊗ e∗3 ).

Let us choose µ = α0 (S). Then we can write,

G(S) = α0 (S)I (Since I = e∗1 ⊗ e∗1 + e∗2 ⊗ e∗2 + e∗3 ⊗ e∗3 )

We note that α0 (S) is a scalar-valued isotropic function of S. Therefore, the scalar func-
tion has the representation α0 (S) = φ0 (I1 , I2 , I3 ), where I1 , I2 , I3 are principal invariants
of S.
Conversely, it is easy to see that the function G(S) = φ0 (IS )I + φ1 (IS )S + φ2 (IS )S 2 is
an isotropic function. Thus, Eq. (9) shows the most general representation of symmetric,
isotropic tensor-valued function.

Problem 2 (Representation theorem for linear isotropic tensor-valued function). Let G


be a symmetric, tensor-valued, isotropic linear function of symmetric tensor S. Then the
tensor-valued function has the representation G(S) = λ(tr S)I + 2µS, where λ and µ are
scalars and I is the identity tensor.

Proof. Though it is possible to specialize the previous result (i.e., Rivlin-Ericksen repre-
sentation theorem), we proceed directly to prove the result as it is easy. Let V1 be the
set of all unit vectors. We now choose a tensor e ⊗ e where e ∈ V1 . It is easy to see that

Joint initiative of IITs and IISc – Funded by MHRD 9


NPTEL – Mechanical Engineering – Continuum Mechanics

eigenvalues of tensor e ⊗ e are λ1 = 1, λ2 = 0 and λ3 = 0 and eigenvector corresponding


to unit eigenvalue is e. Furthermore, any two linearly independent vectors that are or-
thogonal to e are eigenvectors of the tensor e ⊗ e. Similar to Case-2 in previous problem,
we get
G(e ⊗ e) = λ(e)I + 2µ(e)e ⊗ e ∀e ∈ V1 .
If we choose a vector f ∈ V1 then there exist an orthogonal tensor Q such that Qe = f .
Clearly, the axis of Q is orthogonal to f and e. Using the orthogonal tensor Q, we get
Q(e ⊗ e)QT = (Qe) ⊗ (Qe) = f ⊗ f .
Since the given function G is isotropic, we have
O = QG(e ⊗ e)QT − G(Q(e ⊗ e)QT ) = QG(e ⊗ e)QT − G(f ⊗ f )
= Q(λ(e)I + 2µ(e)e ⊗ e)QT − λ(f )I + 2µ(f )f ⊗ f
= (λ(e) − λ(f )) I + 2 (µ(e) − µ(f )) f ⊗ f
Let g ∈ V1 be a vector orthogonal to f . Then we get
g · Og = g · ((λ(e) − λ(f )) I + 2 (µ(e) − µ(f )) f ⊗ f ) g
=⇒ 0 = λ(e) − λ(f ) (since g is orthogonal to f )
=⇒ λ(f ) = λ(e)
Substituting λ(f ) = λ(e) in (λ(e) − λ(f )) I + 2 (µ(e) − µ(f )) f ⊗ f = O we get µ(e) =
µ(f ). Therefore, λ and µ are scalar constants. Therefore, we have
G(e ⊗ e) = λI + 2µe ⊗ e.
We now consider the spectral decomposition of symmetric tensor S,
S = α1 e∗1 ⊗ e∗1 + α2 e∗2 ⊗ e∗2 + α3 e∗3 ⊗ e∗3 ,
where α1 , α2 , α3 are eigenvalues of S and corresponding orthonormal eigenvectors are e∗1 ,
e∗2 , e∗3 respectively. Using linearity of given tensor-valued function G, we get
G(S) = G(α1 e∗1 ⊗ e∗1 + α2 e∗2 ⊗ e∗2 + α3 e∗3 ⊗ e∗3 )
= α1 G(e∗1 ⊗ e∗1 ) + α2 G(e∗2 ⊗ e∗2 ) + α3 G(e∗3 ⊗ e∗3 )
= α1 (λI + 2µe∗1 ⊗ e∗1 ) + α2 (λI + 2µe∗2 ⊗ e∗2 ) + α3 (λI + 2µe∗3 ⊗ e∗3 )
= λ(α1 + α2 + α3 )I + 2µ(α1 e∗1 ⊗ e∗1 + α2 e∗2 ⊗ e∗2 + α3 e∗3 ⊗ e∗3 )
= λ(α1 + α2 + α3 )I + 2µS
= λ(tr S)I + 2µS (Since tr S = α1 + α2 + α3 )
Conversely, the function G(S) = λ(tr S)I+2µS is isotropic as it is easy to see QG(S)QT =
G(QSQT ).

We can use the following identity in proving the preceding isotropic relation.
tr(QSQT ) = tr((QS)QT ) = tr((QS)T Q) (Since tr(RT T ) = tr(RT T ))
= tr((S T QT )Q) = tr(S T (QT Q)) = tr(S T )
= tr(S)
Thus, the linear symmetric tensor-valued isotropic function of symmetric tensor has the
representation G(S) = λ(tr S)I + 2µS, where λ and µ are constants.

Joint initiative of IITs and IISc – Funded by MHRD 10


NPTEL – Mechanical Engineering – Continuum Mechanics

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. I-S. Liu, Continuum Mechanics, 2002, Springer, Berlin.

3. M. E. Gurtin, An Introduction to Continuum Mechanics, 1981, Academic Press,


New York.

Joint initiative of IITs and IISc – Funded by MHRD 11


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-1: Tensor Algebra


Lecture-13: Higher-Order Tensors

Here, we discuss the representation of third-, fourth- and nth -order tensors. The fourth-
order tensors play an important role in constitutive relations as they relate stress with
strain or stress with rate of deformation. Similar to symmetry in case of second-order
tensors, some symmetries can be defined for fourth-order tensors that are discussed in
this lecture. Finally, we discuss an interesting concept called isotropic tensor.
Third-order tensors:
A third-order tensor B is a linear transformation from the vector space V to the space of
second-order tensors V 2 , i.e.,

Bv = T , where v ∈ V and T ∈ V 2 . (1)

Since the third-order tensor is a linear map, we have

B(αv1 + βv2 ) = αBv1 + βBv2 ∀α, β ∈ < and v1 , v2 ∈ V.

Third-order tensor using dyadic product:


Similar to second-order tensor, we can represent any third-order tensor using tensor prod-
uct or dyadic product. Let u, v and w be three vectors. Then the dyadic product of
third-order tensor is defined by

(u ⊗ v ⊗ w)x = (u ⊗ v)(w · x) ∀x ∈ V. (2)

It is easy to see that the input is a vector and output is a second-order tensor in the
operation. Furthermore, it is also easy to see the linearity property. Therefore, the object
u ⊗ v ⊗ w is a third-order tensor. This representation is useful in extracting components
of tensor once basis is known.
Representation of third-order tensor in component form:
Let {e1 , e2 , e3 } be a set of orthonormal basis to the vector space V. Then the third-order
tensor B is represented by
B = Bijk ei ⊗ ej ⊗ ek , (3)
where Bijk = ei · (Bek )ej .
The representation can be verified by the following procedure. We know from Eq. (1)
that the action of third-order tensor result in a second-order tensor. We now consider the
action of B on an arbitrary vector a, then we can write

Ba = (Ba)ij (ei ⊗ ej ) (4)

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Ba = (Ba)ij (ei ⊗ ej )
= {ei · ((Ba)ej )} (ei ⊗ ej )
= {ei · (B(ak ek )ej )} (ei ⊗ ej )
= ak {ei · ((Bek ) ej )} (ei ⊗ ej ) (Since B is linear transformation)
= {ei · ((Bek )ej )} (ek · a)(ei ⊗ ej ) (Since ak = ek · a)
= {ei · ((Bek )ej )} ((ei ⊗ ej ⊗ ek )a)

Arbitrariness of a implies the representation shown in Eq. (3).

Clearly, there are 27 components in third-order tensor. In other words, the dimension
of space of third-order tensors is 27. Similar to third-tensor, the representation using
dyadic product can be extended to nth -order tensor.

nth -order tensor:


An nth -order tensor is defined as a linear transformation from vector space V to (n − 1)th -
order tensor. Similar to second- and third-order tensor, we can have dyadic product or
tensor product representation for nth -order tensor. Let T be a nth -order tensor. Then we
can have
T = Ti1 i2 i3 ···in ei1 ⊗ ei2 ⊗ ei3 · · · ⊗ ein , (5)
   
where Ti1 i2 i3 ···in = ei1 · (T ein ) ein−1 · · · ei2 .
Algebra of nth -order tensors:
Similar to second-order tensors, we can define an addition operation between two nth -order
tensors T and S by
(T + S)x = T x + Sx ∀x ∈ V.
The scalar multiplication of nth -order tensor T is defined by

(αT )x = α(T x) ∀x ∈ V,

where α is scalar, i.e., α ∈ <.

Using these two operation (addition and scalar multiplication), we can easily show that
the collection of all nth -order tensors forms a vector space over the field of real numbers.
Furthermore, 3n components appear in nth -order tensor. Thus, the dimension of space of
nth -order tensors is 3n .

Transformation law for higher-order tensors:


Let {e1 , e2 , e3 } and {e01 , e02 , e03 } be two sets of orthonormal basis to the vector space V.
Then we can define transformation law using component form of tensor. Furthermore, let
Q be a proper orthogonal tensor relating these two basis i.e., Qe0i = ei .

Third-order tensor:
Let Bijk ei ⊗ ej ⊗ ek and B 0 e0i ⊗ e0j ⊗ e0k are two representations of third-order tensor B.

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Then the transformation law is given by


0
Bijk = e0i · (Be0k )e0j
= (Qei ) · (B(Qek ))(Qej ) (Since e0m = QT em )
= (Qil el ) · (B(Qkn en ))(Qjm em )
= Qil Qkm Qjn el · (Ben )em
= Qil Qkm Qjn Blmn
The alternate tensor  that was introduced in the definition of cross product is an example
for third-order tensor. We now prove that the alternate tensor follows the transformation
law. Let 0ijk be ijk th component of alternate tensor in the basis {e01 , e02 , e03 }. Then
0ijk = e0i · (e0j × e0k )
= (QT ei ) · ((QT ej ) × (QT ek )) (Since Qe0p = ep =⇒ e0p = QT ep )
= (Qil el ) · ((Qjm em ) × (Qkn en ))
= Qil Qjm Qkn el · (em × en )
= Qil Qjm Qkn lmn .
Clearly, the components of alternate tensor are related by transformation law. Thus,
alternate tensor is a third-order tensor.
Fourth-order tensor:
Similar to the third-order tensor, the transformation law for fourth-order tensor C can be
stated as.
C0ijkl = Qip Qjq Qkr Qls Cpqrs . (6)
nth -order tensor:
Following the previous cases, we can have the following generalization of the transforma-
tion law for the nth -order tensor T .
Ti01 i2 i3 ···in = Qi1 j1 Qi2 j2 Qi3 j3 · · · Qin jn Tj1 j2 j3 ···jn . (7)

Fourth-order tensors:
In continuum mechanics, fourth-order tensors play an important role along with second-
order tensors. For example, the tensor that relates stress and stain or stress and rate of
deformation can be represented by fourth-order tensors. We now present a convenient
definition of fourth-order tensor.

Definition of fourth-order tensor:


In previous discussion, we defined fourth-order tensor as a linear transformation from
vectors to third-order tensors. However, an alternative definition of fourth-order tensor
can be stated by considering second-order tensors as domain.

The fourth-order tensor C is defined as a linear transformation from second-order


tensors to second-order tensors, i.e.,
CS = R where S, R ∈ V 2 . (8)
Since C is linear transformation, we have
C(αS + βT ) = αCS + βCT , ∀α, β ∈ < and ∀S, T ∈ V 2 .

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Zero tensor: We can define zero fourth-order tensor O as


OT = O ∀T ∈ V,
where O is the zero second-order tensor.

Identity tensor: We can define identity fourth-order tensor I as


IT = T ∀T ∈ V.
Addition and composition of fourth-order tensors is defined as
(C + D)T = CT + DT ∀T ∈ V 2 ,
(CD)T = C(DT ) ∀T ∈ V 2 .
Equality of tensors: Two fourth-order tensors C and D is said to be equal if
CT = DT ∀T ∈ V 2 .
Using inner product of second-order tensors, it is possible to show that two fourth-order
tensors are equal if and only if
(T , CS) = (T , DS) ∀T , S ∈ V 2 .

Components of fourth-order tensor:


Recall the action of second-order tensor on the basis of vector space {e1 , e2 , e3 }, i.e.,
T ej = Tij ei . Since the set {ei ⊗ ej , i = 1, 2, 3 and j = 1, 2, 3} is basis to the space
of second-order tensors, we define the the action of fourth-order tensor on the basis of
second-order tensors by
Cek ⊗ el = Cijkl ei ⊗ ej . (9)
Clearly, the definition of fourth-order tensor on basis is motivated by second-order tensor.
Using inner product between second-order tensors, we get
Cijkl = (ei ⊗ ej , Cek ⊗ el ) = (ei ⊗ ej ) : C(ek ⊗ el ) (10)
Example 1. If T and R are two second-order tensor and they are related by fourth-order
tensor C such that CT = R then show that Rij = Cijkl Tkl .

Solution: Since R and T are second-order tensors, we have


R = Rij ei ⊗ ej and T = Tkl ek ⊗ el .
Consider a component of R,
Rij = ei · Rej
= ei · (CT )ej
= ei · (CTkl ek ⊗ el )ej
= ei · (Tkl C(ek ⊗ el ))ej
= ei · (Tkl Cmnkl em ⊗ en )ej
= Cmnkl Tkl ei · (en · ej )em
= Cmnkl Tkl δnj ei · em
= Cmnkl Tkl δnj δmi
= Cijkl Tkl

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

The symmetries of fourth-order tensors are important in constitutive relations. There-


fore, we now present three types of symmetries for the fourth-order tensor.

Symmetry properties of fourth-order tensor:


Three types of symmetries are defined for the fourth-order tensors, namely, major sym-
metry, right minor symmetry and left minor symmetry.
Major symmetry:
A fourth-order tensor C is said to be symmetric if

(T , CS) = (CT , S) ∀T , S ∈ V 2 .

This symmetry property is also known as major symmetry.


Right minor symmetry:
A fourth-order tensor C is said to have right minor symmetry if

(T , CS) = (T , Csym(S)) ∀T , S ∈ V 2 .
1 
where sym(S) = S + ST .
2
Left minor symmetry: A fourth-order tensor C is said to have left minor symmetry if

(T , CS) = (sym(T ), CS) ∀T , S ∈ V 2 .


1 
where sym(T ) = T + TT .
2
These symmetries impose restrictions on components of fourth-order tensor. The im-
plications of these symmetries on components of tensor are shown below.

• If a fourth-order tensor C has major symmetry then

Cijkl = Cklij .

• If C has right minor symmetry then

Cijkl = Cijlk .

• If C has left minor symmetry then

Cijkl = Cjikl .

We note the major and right minor symmetries of fourth-order tensor imply the left
minor symmetry. Similarly, the major and left minor symmetries also imply right minor
symmetry.

Isotropic tensors:
An isotropic tensor is defined as a tensor whose components are invariant under the change
of orthonormal basis, i.e.,

T = Ti1 i2 ···in ei1 ⊗ ei2 · · · ⊗ ein = Ti1 i2 ···in e0i1 ⊗ ie0i2 · · · ⊗ e0in ,

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

where e0i = Qei and Q is an arbitrary rotation tensor.


Isotropic tensors of order 0: Every scalar is a zeroth-order isotropic tensor as all
scalars are invariant under the transformation of coordinate frame.
Isotropic tensors of order 1: There is no non-zero first-order isotropic tensor as every
vector undergoes transformation under change of frame. The proof is similar to Example-4
in Lecture-12.

Isotropic tensor of order 2: The most general second-order isotropic tensor is λI


where λ is some scalar and I is the identity tensor. The proof is similar to Example-7 in
Lecture-12.
Isotropic tensor of order 3: The most general third-order isotropic tensor is λ where
λ is some scalar and  is alternate tensor. We now recall the definition of alternate tensor.
 = ijk ei ⊗ ej ⊗ ek where 123 = 231 = 312 = 1, 213 = 132 = 321 = −1 and ijk = 0
otherwise.
proof: Let {ei } and {e0i } be two sets of orthonormal basis. Let Q be a transformation
from one basis to another basis. In other words Qei = e0i . Using triple scalar product of
basis vectors, we get

0ijk = e0i · (e0j × e0k )


= (Qei ) · (Qej ) × (Qek )
= (Qei ) · (cof Q)(ej × ek )
= (Qei ) · Q(ej × ek ) (Since cof(Q) = Q)
= QT Qei · ej × ek
= ei · ej × ek (Since QT Q = I)
= ijk

Therefore, the alternate tensor  is a third-order isotropic tensor. Of course, scalar mul-
tiple of alternate tensor is also isotropic. We now want to show that any third-order
isotropic tensor should be scalar multiple of alternate tensor.
Let B = Bijk ei ⊗ ej ⊗ ek be an isotropic third-order tensor. The components of tensor
B are given by Bijk = ei ·(Bek )ej . Consider a rotation tensor Q = ek ⊗ek +ej ⊗ei −ei ⊗ej .
Since B is isotropic tensor, we get
0
Bijk = Bijk = (Qei ) · (BQek )(Qej )
= (ej ) · (Bek )(−ei ) (Since Qei = ej , Qej = −ei , Qek = ek )
= −Bjik

Similarly, considering another rotation tensor Q = ej ⊗ ej + ei ⊗ ek − ek ⊗ ei , we get

Bijk = −Bkji

Using preceding relations, we get Bjik = Bkji . Combining all the relations, we get Bijk =
Bjki = Bkij = −Bjik = −Bikj = −Bkji . Furthermore, we get Bijk = 0 if i = j 6= k or
i 6= j = k or j 6= i = k or i = j = k from the relations Bijk = −Bjik and Bijk = −Bkji .

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

Let B123 = λ then B = λ. Thus, third-order isotropic tensor should be in the form λ
where λ is a scalar.

Isotropic tensor of order 4: The most general fourth-order isotropic tensor is

(λδij δkl + µδik δjl + γδil δjk )ei ⊗ ej ⊗ ek ⊗ el .

where λ, µ and γ are scalars. The proof for this result can be found in a book by
Chandrasekharaiah and Debnath (1994).
References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. D. S. Chandrasekharaiah and L. Debnath, Continuum Mechanics, 1994, Academic


Press Inc., London.

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-2: Tensor Calculus


Lecture-14: The Directional Derivative and
the Fréchet Derivative

In tensor calculus, we present differentiation and integration of scalar, vector and second-
order tensors. We will show in later discussion that the concept of differentiation and
integration play an important role in obtaining governing equations. We assume all func-
tions are sufficiently smooth to perform differentiation. In this lecture, we discuss the
concept of directional derivative (Gâteaux derivative) and Fréchet derivative.
The scalar, vector and tensor fields:
The term field is used to denote a function defined over the three-dimensional Euclidean
space <3 . The scalar field refers to a scalar-valued function φ : <3 → <. Let V be a vector
space. Then the vector field refers to a vector-valued function g : <3 → V. Similarly, the
second-order tensor field refers to a tensor-valued function G : <3 → V 2 .

The directional derivative or Gâteaux derivative:


One of the main task in tensor calculus is to take the derivative to tensor fields. Although
the derivative of tensor field is not a straight forward operation, we want to generalize the
differentiation of scalar-valued function to the tensor function. This extension leads to
the definition of directional derivative or Gâteaux derivative. We start with the definition
of differentiation for a scalar-valued function.
Let f (t) be a function from < to <. Then the derivative Df (t) is defined by
d f (t + ) − f (t)
Df (t) = f (t) = lim .
dt →0 

The directional derivative for a scalar field:


Let φ : <3 → < be a scalar field and u be a unit vector in three-dimensional Euclidean
space. Then the directional derivative at a point x ∈ <3 along the direction u is defined
by
φ(x + u) − φ(x)
Dφ(x)[u] = lim . (1)
→0 
Alternatively, we can have
d
Dφ(x)[u] = φ(x + u)|=0 . (2)
d
Clearly, the directional derivative reduces to ordinary differentiation if x is a scalar vari-
able and u = 1, i.e., Eq. (1) reduces to ordinary differentiation for the scalar-valued
function of scalar variable. Thus, the definition stated in Eq. (1) can be treated as
generalization to the ordinary derivative.

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

The directional derivatives indeed known as partial derivatives if the directions are
chosen to be basis vectors. If the directions along the basis in rectangular three dimen-
sional Cartesian space, i.e., u1 = (1, 0, 0), u2 = (0, 1, 0) and u3 = (0, 0, 1), then the partial
∂φ ∂φ ∂φ
derivatives are denoted by , and . These partial derivatives can be obtained
∂x1 ∂x2 ∂x3
∂φ ∂φ ∂φ
by the relation = Dφ(x)[u1 ], = Dφ(x)[u2 ], = Dφ(x)[u3 ].
∂x1 ∂x2 ∂x3
The directional derivative for a vector field:
Similar to the scalar field, we can define directional derivative of vector fields. Let g be a
given vector field and u be a unit vector in <3 . Then the directional derivative of g along
u at a position x ∈ <3 is defined by

g(x + u) − g(x) d


Dg(x)[u] = lim = g(x + u)|=0 . (3)
→0  d

The directional derivative for a tensor field:


Similar to scalar and vector field, we define direction derivative of second-order tensor field.
Let G be a tensor field and u be a unit vector in <3 . Then the directional derivative of
G along u at a position x ∈ <3 is defined by

G(x + u) − G(x) d


DG(x)[u] = lim = G(x + u)|=0 . (4)
→0  d

It can be observed that the directional derivatives of scalar-, vector- and tensor-valued
functions are scalar, vector and tensors, i respectively.

Directional derivative of scalar-, vector- and tensor-valued functions of tensor:


Similar to scalar-, vector- and tensor fields, we now define the directional derivative to
scalar-, vector- and tensor-valued function of second-order tensor. It can be noted that
a unit vector is used in defining the directional derivative of fields. Similarly, we need a
unit tensor in order to define the direction derivative for the functions whose domain is
tensors. Since the norm of second-order tensor is a measure of magnitude, a unit tensor
U is defined by U : U = 1.
Scalar-valued function of tensor:
Let φ be a scalar-valued function of second-order tensor T . Then the directional derivative
is defined by

φ(T + U ) − φ(T ) d
Dφ(T )[U ] = lim = φ(T + U )|=0 . (5)
→0  d

Vector-valued function of tensor:


Let g be a vector-valued function of second-order tensor T . Then the directional derivative
is defined by

g(T + U ) − g(T ) d
Dg(T )[U ] = lim = g(T + U )|=0 . (6)
→0  d

Tensor-valued function of tensor:


Let G be a vector-valued function of second-oder tensor T . Then the directional derivative

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

is defined by
G(T + U ) − G(T ) d
DG(T )[U ] = lim = G(T + U )|=0 . (7)
→0  d

Examples for directional derivative:


We now present few examples to understand the directional derivative.

Example 1. Show that the direction derivative of a scalar-field φ(x) = x · x along a


direction of unit vector u is 2x · u.

Solution: Given scalar field φ(x) = x · x. Let u be a unit vector. Then we get

φ(x + u) = (x + u) · (x + u)


= x · x + (x · u + u · x) + 2 u · u
= φ(x) + 2x · u + 2 u · u (Since x · u = u · x)
φ(x + u) − φ(x)
=⇒ lim = 2x · u
→0 
Thus, the directional derivative Dφ(x)[u] = 2x · u.

Example 2. Show that the directional derivative of vector field g(x) = x × v along the
direction of unit vector u is u × v where v is a given vector.

Solution: Given vector field g(x) = x × v. Let u be a unit vector. Then we get

g(x + u) = (x + u) × v


= x × v + u × v = g(x) + u × v
=⇒ g(x + u) − g(x) = u × v
g(x + u) − g(x)
=⇒ lim = u×v
→0 
Thus, the directional derivative along Dg(x)[u] = u × v.

Example 3. Show that the directional derivative of G(x) = x ⊗ x along a unit vector u
is x ⊗ u + u ⊗ x.

Solution: Given function G(x) = x ⊗ x. Let u be a unit vector. Then we get,

G(x + u) = (x + u) ⊗ (x + u)


= x ⊗ x + x ⊗ u + u ⊗ x + 2 u ⊗ u
= G(x) + (x ⊗ u + u ⊗ x) + 2 u ⊗ u
G(x + u) − G(x)
=⇒ lim = x ⊗ u + u ⊗ x.
→0 
Thus, the directional derivative DG(x)[u] = x ⊗ u + u ⊗ x.

Example 4. If the scalar-valued function of second-order tensor φ(T ) = α(tr T )2 +


βtr(T 2 ), where α and β are constants, then show that the directional derivative along unit
tensor U is α(tr T )(tr U ) + 2βtr(T U ).

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Solution: Consider given function, φ(T ) = α(tr T )2 + βtr(T 2 ). Let U be a tensor with
unit norm, i.e., U : U = 1. Then we get

φ(T + U ) = α(tr(T + U ))2 + βtr((T + U )2 )


= α(tr T + tr U )2 + βtr(T 2 + T U + U T + 2 U 2 )
= φ(T ) + (α(tr T )(tr U ) + βtr(T U + U T ))
+2 (α(tr U )2 + βtr(U 2 )).

φ(T + U ) − φ(T )
=⇒ lim = α(tr T )(tr U ) + βtr(T U + U T )
→0 
= α(tr T )(tr U ) + 2βtr(T U )

Thus, the directional derivative Dφ(T )[U ] = α(tr T )(tr U ) + 2βtr(T U ).

Example 5. If the tensor-valued function G(T ) = αI + βT + γT 2 where I is identity


tensor and T is second-order tensor then show that the directional derivative along unit
magnitude tensor U is βU + γ(T U + U T ).

Solution: Given function G(T ) = αI + βT + γT 2 . Let U be a tensor with unit norm,


i.e., U : U = 1. Then we get,

G(T + U ) = αI + β(T + U ) + γ(T + U )2


= αI + β(T + U ) + γ(T 2 + T U + U T + 2 U 2 )
= G(T ) + βU + γ(T U + U T ) + 2 γU 2
G(T + U ) − G(T )
=⇒ lim = βU + γ(T U + U T ).
→0 
Thus, the directional derivative DG(T )[U ] = βU + γ(T U + U T ).
The Fréchet derivative:
We now want to define a new derivative for a smooth function such that it is independent
of the direction unlike directional derivative. This motivation leads us to define a new
derivative called Fréchet derivative. This new derivative allows us to define the gradient
of a field which is useful in continuum mechanics. Here, we want to state an important
fact from differential geometry that every smooth and differentiable function is locally a
vector space, i.e., the tangent space to any smooth function is a linear space or vector
space at a point. This fact makes the Fréchet derivative to be linear transformation at a
point. Of course Gâteaux derivative can be recovered by operating the Fréchet derivative
on the direction. Similar to previous derivative, this can also be defined to all fields and
also to the tensor-valued functions.

We require the following notion of limit to define Fréchet derivative. This notion of
limit furthermore elaborated in the following examples. Let f (u) be a function. We say
that f (u) approaches zero faster than u, write f (u) = O(u), if

kf (u)k
lim = 0,
kuk→0 kuk

where kuk is norm of u and kf (u)k is norm of f (u).

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Fréchet derivative of scalar field:


Let φ : <3 → < be a scalar field. If there exists a vector h ∈ <3 such that
φ(x + u) − φ(x) − h · u
lim = 0, (8)
|u|→0 |u|
where u ∈ <3 . Then h is called Fréchet derivative of φ at a point x ∈ <3 . The Fréchet
derivative of field φ is denoted by Dφ(x), i.e., h ≡ Dφ(x). Furthermore, the dot product
between Fréchet derivative and unit vector u can give directional derivative, i.e.,

Dφ(x)[u] = Dφ(x) · u.

Fréchet derivative of vector field:


Let g : <3 → V be a vector field. If there exists a linear transformation H : <3 → V such
that
|g(x + u) − g(x) − Hu|
lim = 0, (9)
|u|→0 |u|
where u ∈ <3 . Then H is called Fréchet derivative of g at a point x ∈ <3 . The Fréchet
derivative is denoted by Dg(x), i.e., H ≡ Dg(x). Furthermore, the action of Fréchet
derivative on unit vector u can give rise to the directional derivative, i.e.,

Dg(x)[u] = Dg(x)u.

Fréchet derivative of tensor field:


Let G : <3 → V 2 be a second-order tensor field. If there exists a linear transformation
B : <3 → V 2 such that
kG(x + u) − G(x) − Buk
lim = 0, (10)
|u|→0 |u|
where u ∈ <3 . Then B is called Fréchet derivative of G at a point x ∈ <3 . The Fréchet
derivative is denoted by DG(x), i.e., B ≡ DG(x). Furthermore, the action of Fréchet
derivative on unit vector u can give rise to the directional derivative

DG(x)[u] = DG(x)u.

Fréchet derivative of scalar- vector- and tensor-valued tensor functions:


Let V and V 2 be vector space and second-order tensor space, respectively. Let φ : V 2 → <,
g : V 2 → V and G : V 2 → V 2 be scalar-, vector- and tensor-valued tensor functions,
respectively. Then, similar to previous cases, we can have the following definitions of
Fréchet derivative for these functions.
If there exists a second-order tensor H ∈ V 2 such that
φ(T + U ) − φ(T ) − H : U
lim = 0, (11)
kU k→0 kU k
where U ∈ V 2 , then H is called Fréchet derivative of φ at T . The Fréchet derivative is
denoted by Dφ(T ).

If there exists a linear transformation (third-order tensor) B : V 2 → V such that


|g(T + U ) − g(T ) − BU |
lim = 0, (12)
kU k→0 kU k

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

where U ∈ V 2 then B is called Fréchet derivative of g at T . The Fréchet derivative is


denoted by Dg(T ).

If there exists a linear transformation (fourth-order tensor) H : V 2 → V 2 such that

kG(T + U ) − G(T ) − HU k
lim = 0, (13)
kU k→0 kU k

where U ∈ V 2 then H is called Fréchet derivative of G at T . The Fréchet derivative is


denoted by DG(T ).
We now present few examples to understand the concept of Fréchet derivative.

Example 6. Show that the Fréchet derivative of function φ(v) = v · T v is (T T + T )v


where T is a given second-order tensor.

Solution: Given function, φ(v) = v · T v. Therefore, we get

φ(v + u) = (v + u) · T (v + u)
= v · Tv + v · Tu + u · Tv + u · Tu
= φ(v) + (T T + T )v · u + o(u).

Taking limit on u, we have

φ(v + u) − φ(v) − (T T + T )v · u
=⇒ lim =0
|u|→0 |u|

Thus, the Fréchet derivative, Dφ(v) = (T T + T )v follows from Eq. (8).

Example 7. Show that the Fréchet derivative of function φ(T ) = v · T v is v ⊗ v where


v is a given vector.

Solution: φ(T ) = v · T v. Therefore, we get

φ(T + U ) = v · (T + U )v = v · T v + v · U v = φ(T ) + (v ⊗ v) : U
φ(T + U ) − φ(T ) − (v ⊗ v) : U
=⇒ lim = 0.
kU k→0 kU k

Thus, the Fréchet derivative, Dφ(T ) = v ⊗ v from Eq. (11).

Example 8. Show that the Fréchet derivative of function φ(T ) = u · T v is u ⊗ v where


v and u are given vectors.

Solution: φ(T ) = u · T v. Therefore, we get

φ(T + U ) = u · (T + U )v = u · T v + u · U v = φ(T ) + (u ⊗ v) : U
φ(T + U ) − φ(T ) − (u ⊗ v) : U
=⇒ lim = 0.
kU k→0 kU k

Thus, the Fréchet derivative, Dφ(T ) = u ⊗ v from Eq. (11).

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

Example 9. Show that the Fréchet derivatives of functions φ(S) = u · Sv and ψ(W ) =
u · W v are 21 (u ⊗ v + v ⊗ u) and 12 (u ⊗ v − v ⊗ u) respectively if S belongs to the space
of symmetric tensors and W belongs to a space of skew-symmetric tensors.

Solution: Similar to previous example, we get


φ(S + U ) − φ(S) − (u ⊗ v) : U
lim = 0.
kU k→0 kU k
We note that U is a symmetric tensor in the preceding equation as the domain of φ is
symmetric tensor. Furthermore, U : M = 0 if M is any skew-symmetric tensor. Using
additive decomposition, we have Fréchet derivative
1
Dφ(S) = (u ⊗ v + v ⊗ u).
2
Similarly,
ψ(W + U ) − ψ(W ) − (u ⊗ v) : U
lim = 0.
kU k→0 kU k
We note that U is a skew-symmetric tensor in the preceding equation as the domain of ψ
is skew-symmetric tensor. Furthermore, U : N = 0 if N is any symmetric tensor. Using
additive decomposition, we have Fréchet derivative
1
Dψ(W ) = (u ⊗ v − v ⊗ u).
2
Example 10. Show that the Fréchet derivative of function φ(T ) = det(T ) is (det T )T −T .

Solution: Given function, φ(T ) = det(T ). Let {p, q, r} be orthonormal set of vectors
then we have,
φ(T + U ) = det(T + U )
= det(T + U )[p, q, r] (Since [p, q, r] = 1)
= [(T + U )p, (T + U )q, (T + U )r]
= [T p + U p, T q + U q, T r + U r]
= [T p, T q, T r] + [U p, T q, T r] + [T p, U q, T r] + [T p, T q, U r]
+[T p, U q, U r] + [U p, T q, U r] + [U p, U q, T r] + [U p, U q, U r]
= det(T )[p, q, r] + [(cof T )T U p, q, r] + [p, (cof T )T U q, r]
+[p, q, (cof T )T U r] + o(U )
= φ(T ) + tr(cof T )T U ) + o(U )
φ(T + U ) − φ(T ) − (cof T ) : U
=⇒ lim = 0.
kU k→0 kU k
Thus, the Fréchet derivative, Dφ(T ) = cof T = det(T )T −T .

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. I-S. Liu, Continuum Mechanics, 2002, Springer, Berlin.

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-2: Tensor Calculus


Lecture-15: The Gradient, Divergence and Curl

Here we discuss the gradient, divergence and curl. These definitions play an important
role in continuum mechanics.
Gradient of a function:
The Fréchet derivative is synonym for the gradient of a function. If the gradient of a
function exists then it is unique. Since the gradient is a linear transformation, we can
represent it in the component form for a known basis. We now present gradient of scalar,
vector and tensor fields in canonical basis.
Gradient of scalar field:
Let φ : <3 → < be a scalar field and the set {e1 , e2 , e3 } be an orthonormal basis to the
three-dimensional Euclidean space. Then the gradient of scalar field is represented by
∂φ(x)
∇φ(x) = ei , (1)
∂xi
where x is a position vector in <3 .
Proof: Let ∇φ(x) be the gradient (Fréchet derivative) of given function φ. As stated in
Lecture-14, we can have the following directional derivative along a unit vector u

Dφ(x)[u] = ∇φ(x) · u

Choosing u = ei in above relation, we get ith component of gradient


φ(x + ei ) − φ(x) ∂φ
(∇φ)i = ∇φ(x) · ei = Dφ(x)[ei ] = lim =
→0  ∂xi
Clearly, ith component of gradient implies the representation shown in Eq. (1).
Gradient of vector field:
Let g : <3 → V be a vector field and the set {e1 , e2 , e3 } be an orthonormal basis to
the three-dimensional Euclidean space. We consider <3 as the vector space V. Then the
gradient of vector field is represented by
∂gi
∇g(x) = ei ⊗ ej , (2)
∂xj
where x is a position vector in <3 and gi is ith component of vector g.
Proof: Let ∇g(x) is the gradient (Fréchet derivative) of vector field. Then ∇g(x) is a
second-order tensor and the directional derivative along unit vector u is given by

Dg(x)[u] = (∇g)u.

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Choosing u = ej , we get

g(x + ej ) − g(x)


(∇g)ej = lim
→0 
gi (x + ej )ei − gi (x)ei
= lim (Since g is vector)
→0 
gi (x + ej ) − gi (x)
= lim ei
→0 
∂gi
= ei
∂xj

∂gi
Therefore, ij th component of gradient of vector field can be written as (∇g)ij = .
∂xj
Clearly, ij th component of gradient implies the representation shown in Eq. (2).
Gradient of second-order tensor field:
Let G : <3 → V 2 be a second-order tensor field and the set {e1 , e2 , e3 } be an orthonormal
basis to the three-dimensional Euclidean space. Let <3 be the vector space V. Then the
gradient of tensor field has the representation,
∂Gij
∇G(x) = ei ⊗ ej ⊗ ek , (3)
∂xk

where x is a position vector in <3 and Gij is ij th component of tensor G.

Proof: Since the gradient is indeed Fréchet derivative, the directional derivative of G
along unit vector u is given by

DG(x)[u] = ∇Gu.

Choosing u = ek , we get

Gij (x + ek ) − Gij (x) ∂Gij


(∇G)ek = lim ei ⊗ ej = ei ⊗ ej .
→0  ∂xk
∂Gij
Therefore, ijk th component of gradient (∇G)ijk = .
∂xk
Clearly, ijk th component of gradient implies the representation shown in Eq. (3).
Derivative of scalar-valued function of tensor:
Let φ : V 2 → < be a scalar-valued function of second-order tensor and the set {e1 , e2 , e3 }
be an orthonormal basis to the vector space V. Then the derivative of scalar function has
the representation,
∂φ
Dφ(T ) = ei ⊗ ej , (4)
∂Tij
where T is a variable in the space of second-order tensors.
Proof: The Fréchet derivative Dφ(T ) is known as derivative. Therefore, the directional
derivative along unit tensor U can be written as

Dφ(T )[U ] = Dφ(T ) : U .

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Since Dφ(T ) ∈ V 2 and hence, we have the representation

Dφ(T ) = (ei · (Dφ(T ))ej )ei ⊗ ej


= (Dφ(T ) : (ei ⊗ ej ))ei ⊗ ej
= (Dφ(T )[ei ⊗ ej ])ei ⊗ ej
!
φ(T − ei ⊗ ej ) − φ(T )
= lim ei ⊗ ej
→0 
∂φ
= ei ⊗ ej
∂Tij

Derivative of tensor-valued function of tensor:


Let G : V 2 → V 2 be a second-order tensor-valued function of second-order tensor and
the set {e1 , e2 , e3 } be an orthonormal basis to the vector space V. Then the derivative of
tensor-valued function has the representation,
∂Gij
DG(T ) = ei ⊗ ej ⊗ ek ⊗ el , (5)
∂Tkl
where T is a variable in the space of second-order tensors.

Proof: The Fréchet derivative Dφ(T ) is known as derivative. Therefore, the directional
derivative along unit tensor U can be written as

DG(T )[U ] = DG(T )U .

Since DG(T ) is a fourth-order tensor, hence, we have the representation

DG(T ) = ((ei ⊗ ej ) : (DG(T )(ek ⊗ el ))ei ⊗ ej ⊗ ek ⊗ el


!!
G(T + ek ⊗ el ) − G(T )
= (ei ⊗ ej ) : lim ei ⊗ ej ⊗ ek ⊗ el
→0 
!
(ei ⊗ ej ) : G(T + ek ⊗ el ) − (ei ⊗ ej ) : G(T )
= lim ei ⊗ ej ⊗ ek ⊗ el
→0 
!
Gij (T + ek ⊗ el ) − Gij (T )
= lim ei ⊗ ej ⊗ ek ⊗ el
→0 
∂Gij
= ei ⊗ ej ⊗ ek ⊗ el .
∂Tkl

Divergence:
Here, we define divergence of a vector and second-order tensor fields.
The divergence of vector field:
Let g : <3 → V be a vector field. Then the divergence of a vector field is defined by

∇ · g = tr(∇g). (6)

It is clear from the definition that the divergence of a vector field is scalar-valued function.

The divergence of a vector field can be expressed in the basis {e1 , e2 , e3 } by


∂gi
∇·g = . (7)
∂xi

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

It is easy to see from Eq. (1) that tr(∇g) = ∇ · g.


The divergence of second-order tensor field:
Let G : <3 → V 2 be a field of second-order tensor then the divergence of tensor field is
defined by
(∇ · G) · v = ∇ · (GT v) ∀ constant v ∈ V. (8)
It is clear from the definition that ∇ · G is a vector-valued function. The divergence of
second-order tensor field can be expressed in the basis {e1 , e2 , e3 } by
∂Gij
∇·G= ei . (9)
∂xj

Curl:
We now define new quantities called curl of vector field and second-order tensor field. The
curl of vector field appears in Stokes’ theorem which is discussed later.
The curl of vector field:
Let g : <3 → V be a vector field. Then the curl of vector field, denoted as ∇ × g, is
defined by
(∇ × g) × v = (∇g − ∇g T )v ∀ constant v ∈ V. (10)
It is easy to see that ∇g − ∇g T is a skew-symmetric tensor and ∇ × g is corresponding
axial vector. Using Eq. (9) in Lecture-7, the curl of vector can be written as
1 1
∇ × g = − ijk (∇g − ∇g T )jk ei = − ijk ((∇g)jk − (∇g)kj )ei = ijk ∇gkj ei
2 2
∂gk
= ej × ek
∂xj

The curl of second-order tensor field:


Let G : <3 → V 2 be a second-order tensor field then the curl of tensor field, denoted as
∇ × G, is defined by

(∇ × G)v = ∇ × (GT v) ∀ constant v ∈ <3 . (11)

It is easy to see from the definition that ∇ × G is a second-order tensor. The curl of
tensor can be expressed in the basis {e1 , e2 , e3 } by

∇ × G = (ei · (∇ × G)ej )ei ⊗ ej = (ei · ∇ × (GT ej ))ei ⊗ ej


∂(em · GT ej ) ∂gm
= (ei · en × em )ei ⊗ ej (Since ∇ × g = en × em )
∂xn ∂xn
!
∂Gjm
= (ei · (en × em ))ei ⊗ ej
∂xn
∂Gjm
= (en × em ) ⊗ ej
∂xn

Laplacian:
We now define Laplacian of scalar, vector and tensor field. These quantities involve
second-order partial differentials and arise in differential form of governing equations of
solids and fluids.

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

The Laplacian of scalar field:


Let φ : <3 → < be a scalar field. Then the Laplacian, denoted as ∇2 φ, is defined by

∇2 φ = ∇ · (∇φ). (12)

It is clear from the definition that the Laplacian of scalar field is a scalar function which
can be expressed as
∂ 2φ
∇2 φ = .
∂xi ∂xi

The Laplacian of vector field:


Let g : <3 → V be a vector field then the Laplacian of vector field, denoted as ∇2 g, is
defined by
∇2 g = ∇ · (∇g) (13)
It is clear from the definition that the Laplacian of vector field is a vector-valued function
which can be expressed as

∂(∇g)ij ∂ 2 gi
∇2 g = (∇ · (∇g))i ei = ei = ei .
∂xj ∂xj ∂xj

The Laplacian of second-order tensor field:


Let G : <3 → V 2 be a second-order tensor field then the Laplacian of tensor field, denoted
as ∇2 G, is defined by

(∇2 G) : H = ∇2 (G : H) ∀ constant H ∈ V 2 .

We can have the following component form of Laplacian of tensor field.

∇2 G = (ei · (∇2 G)ej )ei ⊗ ej


= ((∇2 G) : (ei ⊗ ej ))ei ⊗ ej
= (∇2 (G : (ei ⊗ ej ))ei ⊗ ej
= (∇2 (ei · Gej )ei ⊗ ej
= (∇2 (Gij ))ei ⊗ ej
∂ 2 Gij
= ei ⊗ ej .
∂xk ∂xk
There are many products available in tensor functions. For example, the product of
scalar and vector, the inner product, the cross product, the tensor product, the action of
tensor on a vector. The products that are defined here are bilinear map. Therefore, we
frequently require to have differentiation of these products. We now define product rule
of differentiation.
Product rule:
Let f and g be two differentiable functions at a point x. Then their product h = π(f , g)
is also differentiable at x and directional derivative of product is given by

Dh(x)[u] = π(f (x), Dg(x)[u]) + π(Df (x)[u], g(x)). (14)

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

We now see that this is similar to product rule what we encounter in the case of one
variable calculus. Let x ∈ < then it represents one variable case. Replacing x with t in
product rule, we get
! !
dh(t) dg(t) df (t)
= π f (t), +π , g(t) .
dt dt dt

Clearly, this formula is consistent with one variable case. The proof of this result can be
found in a book by Gurtin (1981).
We now present an example to understand the concept of product rule.

Example 1. Let vector valued function g = φh where φ(x) be a scalar-valued function


and h(x) be a vector valued function then show that ∇(φh) = h⊗∇φ+φ∇h using product
rule of differentiation.

Solution: Given function g = φh. Let u be a unit vector. Then, using product rule, we
get

∇g[u] = (∇g)u = Dg(x)[u]


= (Dφ(x)[u])h + φ(Dh(x)[u])
= (∇φ · u)h + φ(∇hu)
= (h ⊗ ∇φ)u + (φ∇h)u
= (h ⊗ ∇φ + φ∇h)u

Since unit vector u is arbitrary, we get ∇g = h ⊗ ∇φ + φ∇h.


We now discuss another important rule of differentiation called chain rule. If we have
composition of functions then this rule is useful to take directional derivative.

Chain rule:
Let h(x) be a composite function of f and g, i.e., h = f ◦ g. In addition, let g : Ω1 → Ω2
be a differentiable function at x ∈ Ω1 and, f : Ω2 → Ω3 be a differentiable function at
y = g(x). If the function h is differentiable at x then the directional derivative of h
along u ∈ Ω1 is given by

Dh(x)[u] = D(f ◦ g)(x)[u] = Df (g(x))[Dg(x)[u]. (15)

Similar to product rule, this can also be deduced to one variable case. Let x ∈ < to
represent one variable case. Replacing x with t in chain rule, we get

dh(t) df (g(t)) dg(t)


= = Df (g(t))
dt dt dt
Clearly, this rule is also consistent with one variable case. We now present an example to
understand the chain rule.

Example 2. Show that the directional derivative of tr(T −1 ) along the direction of unit
tensor U is −tr(T −1 U T −1 ).

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

Solution: Given function can be written as φ(G(T )) = tr(T −1 ) and G(T ) = T −1 . There-
fore, we can write, φ(G) = tr(G).

The directional derivative of function φ(G(T )) can be obtained using chain rule. There-
fore, we first want to get directional derivative of G(T ). Using the relation T −1 T = I
and product rule, we get

D(T −1 T )[U ] = O (Since DI[U ] = O)


 
−1 −1
DT [U ] T + T (DT [U ]) = O
DT −1 [U ] = −T −1 U T −1 (Since DT [U ] = U )
−1 −1
=⇒ DG(T )[U ] = −T UT .

Let H be a second order tensor then the directional derivative of φ(G) along H can be
written as
tr(G + H) − tr(G)
φ(G)[H] = tr(G)[H] = lim = tr(H).
→0 
Using chain rule, we get directional derivative of function φ(G(T )) along U as

Dtr(T −1 )[U ] = Dφ(G(T ))[U ]


= Dφ(G)[DG(T )[U ]] (Using chain rule)
−1 −1
= Dφ(G)[−T UT ] (Since DG[U ] = −T −1 U T −1 )
= tr(−T −1 U T −1 ) (Since φ(G)[H] = tr(H))
= −tr(T −1 U T −1 )

Thus, the directional derivative of a given function Dtr(T −1 )[U ] = −tr(T −1 U T −1 ).

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. I-S. Liu, Continuum Mechanics, 2002, Springer, Berlin.

3. M. E. Gurtin, An Introduction to Continuum Mechanics, 1981, Academic Press,


New York.

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-2: Tensor Calculus


Lecture-16: The Time Derivative and
Some Integral Theorems

In continuum mechanics, scalar, vector and tensor functions are dependent not only on
space but also on time. Therefore, we need to define the time derivative of scalar, vector
and tensor functions. Since the time t ∈ <+ , the definition of time derivative is easier
than the spatial derivatives that are discussed earlier.
Time derivative of scalar-, vector- and tensor-valued functions:
Let φ : < → <, g : < → V and G : < → V 2 be scalar, vector and tensor functions of time
t, respectively. Then the time derivatives of these functions, denoted as φ̇, ġ and Ġ, are
defined by
dφ φ(t + ) − φ(t)
φ̇ = = lim
dt →0 
dg g(t + ) − g(t)
ġ = = lim
dt →0 
dG G(t + ) − G(t)
Ġ = = lim .
dt →0 
The product rule and the chain rule can be extended the time derivative. We present an
example to understand these rules.
Example 1. Show that time derivative of function tr(T )det(T ) is
! !
dT dT
tr(T ) (cof T ) : + tr det(T ).
dt dt

Solution: Given function φ(T ) = tr(T )det(T ). Then we can apply product rule to get
the derivative. We now calculate derivative of tr(T ) and det(T ).
d(tr T ) tr(T (t + )) − tr(T (t)) tr(T (t + ) − T (t))
= lim = lim
dt →0  !
→0
! 
T (t + ) − T (t) dT
= tr lim = tr ,
→0  dt
alternatively, using chain rule
" # !
d(tr T ) dT dT dT
= Dtr(T ) =I: = tr .
dt dt dt dt
Using chain rule, we can get time derivative of det(T ) as
" #
d(det T ) dT dT
= D(detT ) = (cof T ) : (see Example-10 in Lecture-14)
dt dt dt

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Using product rule, we get


! !
dφ d(tr T ) d(det T )
= det(T ) + (tr T )
dt dt dt
! !
dT dT
= tr(T ) (cof T ) : + tr det(T ).
dt dt

As discussed in Lecture-9, the exponential of a skewsymmetric tensor is rotation tensor.


Here we define the exponential tensor as a solution of initial value problem.
Exponential of tensor:
The exponential of a tensor can be defined in two ways: (i) in terms of series representation
as discussed earlier,
1
eT t = I + T t + T 2 t2 + · · · ,
2!
(ii) in terms of a solution of the initial value problem

Ẋ(t) = T X(t) t > 0


X(0) = I,

where X(t) is a tensor function, 0 ≤ t < ∞. The existence theorem for linear differential
equations guarantees a unique solution to the initial value problem, which we write in the
form
X(t) = eT t .
It is easy to see from the series expansion that
Tt
 T
eT = eT t ∀T ∈ V 2 ,

and if A is an invertible second-order tensor then


−1 BA)t
e(A = A−1 eBt A ∀B ∈ V 2 .
 
Problem 1. Let T be a second-order tensor. Then for each t ≥ 0, det eT t > 0 and
 
det eT t = e(tr T )t
.

Proof. Let X(t) = eT t and X(0) = I then we want to show that det X(t) > 0 for every
t ∈ [0, ∞).
Suppose that det X(τ ) = 0 for some finite time τ . Since X is continuous and
det X(0) = 1, there is some nonempty interval [0, τ ) such that det X(t) > 0 for every
t ∈ [0, τ ). Consequently, X(t) is invertible in nonempty interval and we get

T = Ẋ(t) (X(t))−1 (Since Ẋ(t) = T X(t))


T !
(cof X(t))
=⇒ T = Ẋ(t)
det X(t)
1  
=⇒ tr(T ) = tr Ẋ(t) (cof X(t))T
det X(t)
!
1 d
=⇒ tr(T ) = (det X(t))
det X(t) dt

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Using initial condition det X(0) = 1, we have the following unique solution
T )t
det X(t) = e(tr ,

for 0 ≤ t < τ . The equation shows that det X(t) > 0 for every finite t ≥ 0 because X(t) is
continuous. Therefore, this is contradicting the initial assumption that τ is finite. Thus,
there is no finite τ such that det X(τ ) = 0 and hence, det X(t) > 0 for every t ∈ [0, ∞).
Furthermore,  
det eT t = e(tr T )t (Since X(t) = eT t ).

Problem 2. Let A and be B are two second-order tensors. Then e(A+B)t = eAt eBt if and
only if A and B are commuting tensors.

Proof. Let XA = eAt , XB = eBt , and XA+B = e(A+B)t . Then we get


d (XA XB ) dXA dXB
= XB + XA = AXA XB + XA BXB
dt dt dt
If A commutes with B, i.e., AB = BA, then we get XA B = BXA since XA is
polynomial of A. Therefore, preceding equation can be written as
d (XA XB )
= (A + B)XA XB .
dt
Since initial condition (XA XB )(0) = XA (0)XB (0) = eO eO = I, the differential equation
has the following unique solution.

XA XB = e(A+B)t .

Therefore, commutativity of A and B implies XA XB = XA+B .


Conversely, XA XB = XA+B then we get
dXA+B d (XA XB )
=
dt dt
dXA dXB
=⇒ (A + B)XA+B = XB + X A
dt dt
=⇒ AXA+B + BXA+B = AXA XB + XA BXB
=⇒ BXA+B = XA BXB (Since XA+B = XA XB )
=⇒ BXA XB = XA BXB
=⇒ BXA = XA B (Since det XB > 0)

Differentiating the above equation with respect to time, we get


d (BXA ) d (XA B)
=
dt dt
=⇒ BAXA = AXA B
=⇒ BAXA = ABXA (Since XA B = BXA )
=⇒ BA = AB (Since det XA > 0)

Thus, e(A+B)t = eAt eBt if and only if A and B are commuting tensors.

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Using series definition of exponential (see Problem-2 in Lecture-9), we showed that


Wt
e is a rotation tensor when W is skew-symmetric tensor. We now show the same result
using initial value problem.
Problem 3. If W is skew-symmetric tensor then eW t is a rotation tensor for every t ≥ 0.

Proof. Let X(t) = eW t for t ≥ 0 then, by definition of previously stated initial value
problem, we have
Ẋ = W X, X(0) = I.
Let Z = XX T . Then we get
Ż = ẊX T + X Ẋ T = W XX T + XX T W T .
Since W is skew-symmetric tensor, we can write
Ż = W XX T − XX T W .
Therefore, Z satisfies
Ż = W Z − ZW , Z(0) = I.
This initial value problem has unique solution, Z(t) = I, ∀t ≥ 0. Thus, X(t)X(t)T = I
implies X(t) is an orthogonal tensor. Furthermore, det X(t) > 0 and hence, X(t) is a
rotation tensor.
Problem 4. If Q(t) is an orthogonal tensor then show that Q̇QT is a skew-symmetric
tensor.

Proof. If Q is an orthogonal tensor then we have


QQT = I.
Differentiating above equation with respect to time, we get
Q̇QT + QQ̇T = 0
=⇒ Q̇QT = −QQ̇T
 T
=⇒ Q̇QT = − Q̇QT

Thus, Q̇QT is a skew-symmetric tensor.

The integral theorems play an important role in the balance laws. We now discuss
some integral theorems that are useful for our discussion in continuum mechanics.
Integral theorems:
Here we discuss divergence, Stokes’ and localization theorems.
Divergence theorem or Gauss divergence theorem:
Let Γ be a boundary of regular domain Ω in three-dimensional Euclidean space <3 . Let
φ : <3 → <, v : <3 → V, S : <3 → V 2 be scalar, vector, tensor fields, respectively. Then
Z Z
∇φ ∂Ω = φn ∂Γ (1)
Z Ω ZΓ
∇ · v ∂Ω = v · n ∂Γ (2)
ZΩ ZΓ
∇ · S ∂Ω = Sn ∂Γ (3)
Ω Γ

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

where n is the outward unit normal to Γ.


Equations (1) and (2) are well known classical results in vector calculus. Here, we
prove the divergence theorem for tensor fields.

Proof. Let u be an arbitrary constant vector. Then


Z Z Z
u· Sn ∂Γ = u · Sn ∂Γ = S T u · n ∂Γ
Γ Γ Γ

Since S T u is a vector field, using Eq. (2), we get


Z Z   Z Z
S T u · n ∂Γ = ∇ · S T u ∂Ω = u · (∇ · S) ∂Ω = u · ∇ · S ∂Ω.
Γ Ω Ω Ω

Arbitrariness of u implies Eq. (3).

We note that the divergence theorem is an extension of fundamental theorem of calculus


to higher dimensions. Furthermore, physical meaning of divergence theorem for the vector
field represents the flux (for example fluid) that is leaving normal to the bounded surface
Γ (i.e., flowing into or out-of the boundary) is equal to the flux generated or disappeared
inside the domain Ω. Physically, it can be observed that the flux that is moving along
tangential direction to the boundary does not leave the domain.
Problem 5. Show that Ω ∇ × v ∂Ω = Γ n × v ∂Γ, where Ω is a regular domain, Γ is
R R

boundary of the domain, and n is outward unit normal to Γ.

Proof. Let u be an arbitrary constant vector. Then


Z Z Z Z
u· ∇ × v ∂Ω = u · ∇ × v ∂Ω = ∇ · (v × u) ∂Ω = (v × u) · n ∂Γ
Ω ZΩ Ω Z Γ

= u · (n × v) ∂Γ = u · (n × v) ∂Γ.
Γ Γ

∇ × v ∂Ω = n × v ∂Γ.
R R
Arbitrariness of u implies Ω Γ

Problem 6. Show that Ω ∇v ∂Ω = Γ v ⊗ n ∂Γ, where Ω is a regular domain, Γ is


R R

boundary of the domain, and n is outward unit normal to Γ.

Proof. Let u be an arbitrary constant vector. Then


Z  Z Z Z
∇v ∂Ω u = (∇v) u ∂Ω = ∇ · (v ⊗ u) ∂Ω = (v ⊗ u) n ∂Γ

ZΩ ZΩ Z Γ

= (u · n)v ∂Γ = (n · u)v ∂Γ = (v ⊗ n) u ∂Γ
Γ Γ Γ
Z 
= v ⊗ n ∂Γ u
Γ

∇v ∂Ω = v ⊗ n ∂Γ.
R R
arbitrariness of u implies Ω Γ

Stokes’ theorem:
Let Γ be a surface in three-dimensional Euclidean space <3 with piecewise smooth bound-
ary curve C and v be a vector field. Then
Z I
(∇ × v) · n ∂Γ = v · dx (4)
Γ C

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

where n is unit normal field on Γ and dx is infinitesimal segment of the curve C.


This is also a classical result in vector calculus. Any standard book on vector calculus
can be referred for the proof.
R  
Problem 7. Show that Γ (∇ · v)n − (∇v)T n ∂Γ = C v × dx, where Γ is a surface
H

with boundary contour C, n is unit normal field to Γ, and dx is infinitesimal segment of


the curve C.

Proof. Let u be an arbitrary constant vector. Then


I  I
u· v × dx = u · (v × dx)
C
IC
= (u × v) · dx
ZC
= (∇ × (u × v)) · n ∂Γ (Using Stokes’ theorem)

= ((∇ · v)u − (∇v)u) · n ∂Γ
ZΓ  
= u · (∇ · v)n − u · (∇v)T n ∂Γ
Γ
Z   
T
= u· (∇ · v)n − (∇v) n ∂Γ .
Γ

R  
(∇ · v)n − (∇v)T n ∂Γ = v × dx.
H
Arbitrariness of u implies Γ C

Potential theorem:
Let Ω be a simply-connected region, and u = ∇φ be a vector field over Ω. Then

∇ × u = 0.

Conversely, if ∇ × u = 0, then there exists a scalar potential φ such that u = ∇φ.

Proof. Let u = ∇φ where φ is a scalar function. Then,

(∇ × u)i = (∇ × (∇φ))i
!
∂ ∂φ
= ijk
∂xj ∂xk
∂ 2φ
= ijk
∂xj ∂xk
∂ 2φ
= ikj (interchanging indices j and k)
∂xk ∂xj
∂ 2φ
= −ijk
∂xj ∂xk

It is clear that
∂ 2φ ∂ 2φ
(∇ × u)i = ijk = −ijk = −(∇ × u)i .
∂xj ∂xk ∂xj ∂xk

Thus, ∇ × u = ∇ × (∇φ) = 0.

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

Conversely, we want to show that if ∇ × u = 0 then there is a function φ such that


u = ∇φ. Let C be a closed curve in Ω and Γ be a surface in Ω with C as its boundary.
Then, using stokes theorem, we get
I Z
u · dx = ∇ × u ∂Γ.
C Γ

The condition ∇ × u = 0 implies


I
u · dx = 0.
C

Thus, the value of integral xx0 u(y) · dy along any curve in Ω depends only on x if we
R

assume x0 to be a fixed point. Therefore, we can define a function φ(x) as


Z x
φ(x) = u(y) · dy.
x0

Taking the gradient, we get


u = ∇φ.
Thus, ∇ × u = 0 if and only if there is a function φ such that u = ∇φ.

Localization theorem:
Let φ be a continuous scalar, vector or tensor field, and B be an arbitrary open subset of
domain Ω. If Z
φ(x) ∂Ω = 0, ∀B ⊆ Ω,
B
Then
φ(x) = 0, ∀x ∈ Ω.

Proof. Let x0 be an arbitrary point in the domain Ω and B be a -neighborhood of x0 ,


i.e., B is set of all y ∈ |Ω such that |y − x0 | < . Then

1 Z 1 Z


φ(x0 ) − lim φ(x) ∂Ω = (φ(x0 ) − φ(x)) ∂Ω

→0 vol(B) B vol(B) B
1 Z
≤ |φ(x0 ) − φ(x)| ∂Ω
vol(B) B
≤ sup |φ(x0 ) − φ(x)| ,
x∈B

which tends to zero as  → 0, since φ is continuous.

We refer the localization theorem in continuum mechanics to obtain the differential


form of governing equations from the integral form.

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. M. E. Gurtin, An Introduction to Continuum Mechanics, 1981, Academic Press,


New York.

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-3: Kinematics
Lecture-17: The Lagrangian and
Eulerian Descriptions

In previous lectures, we presented a detailed treatment of tensor algebra and calculus that
is necessary for the study of continuum mechanics. Since this course is about understand-
ing the beahviour of continuum under action of forces and thermal effects, the content
is divided into three major parts: (i) Kinematics, (ii) Balance laws, (iii) Constitutive
equations.
Kinematics refers to the study of the continuous medium under purely geometric con-
siderations. In other words, we study the displacement, velocity, acceleration, strain, and
rate of deformation fields of continuum body. Application of fundamental laws of me-
chanics and thermodynamics to the continuum is discussed in the topic balance laws. An
important physical quantity called stress is introduced via the balance of linear momen-
tum. Finally, the kinematical quantities such as strain and rate of deformation are related
with the quantities such as stress and rate of stress through the constitutive relations.
The task of continuum mechanics finishes with the understanding of restrictions imposed
by second law of thermodynamics, principle of material frame-indifference and material
symmetry on constitutive relations.
In this module, kinematics of continuous medium is presented. We make the following
assumptions in the study of continuous media.

• The boundary of continuum body should not make self contact or penetrate.

• Topology of the body should be invariant under the deformation, i.e., new holes
cannot be generated or existing holes cannot be removed in process of deformation.

Consider a continuum body as occupying a region in the three-dimensional Euclidean


space. The body can undergo the deformation which is characterized by change of the
region with time (see Fig. (1)). Therefore, the continuum body can be defined as reg-
ular three-dimensional region in the three-dimensional space. Let Ω0 be a geometry of
continuum body at initial time t = 0. Then Ω0 is referred as undeformed or reference
configuration. Let Ω be a geometry of continuum body at a time t. Then Ω referred
as deformed configuration. The assumptions stated earlier ensure that the existence of
a smooth function from the reference configuration Ω0 to the deformed configuration Ω.
Furthermore, the mapping function from Ω0 to Ω is one-one function. Let χ be a mapping
function then for every X ∈ Ω0 we can find a unique x ∈ Ω, i.e,
x = χ(X, t) (1)

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

where t is the time.


We will show in Lecture-19 (see the section on “volume element” in Lecture-19) that
det(∇χ) represents volume transformation, i.e., the ratio of volumes after and before the
deformation, locally. Therefore, it is physically meaningful to assume that det ∇χ 6= 0.
Furthermore, the mapping χ is continuous one parameter family of time known as motion.
These conditions imply det ∇χ > 0 through the motion, i.e., volume cannot be zero at
any instant of time during the motion of continuum. We note that continuity of χ and
nonsingular gradient of χ makes the existence of smooth inverse of χ. Thus, we have the
inverse map
X = χ−1 (x, t). (2)
The mapping function χ and its inverse χ−1 make two different ways of describing the

Ω0 Ω
u

e3 X x

e2

e1

Figure 1: Reference and deformed configurations

deformation of a body during its motion. These two are known as Lagrangian and Eulerian
descriptions.
Lagrangian and Eulerian descriptions:
Since the map χ and its inverse χ−1 are smooth, every field defined over the body can
be expressed either in terms of reference variable X ∈ Ω0 or in terms of spatial variable
x ∈ Ω. Expressing all quantities in terms of reference variable X is known as Lagrangian
formulation. Similarly, expressing all quantities in terms of spatial variable x is known
as Eulerian formulation. The Lagrangian and Eulerian formulations are also known as
material description and spatial description, respectively. Of course both formulations
are equivalent as it is just a change of variable. However, depending on convenience
and available experimental observations, we choose the suitable formulation. Usually,
Lagrangian framework is preferred for the solid mechanics problem since the reference
configuration of continuum body is known. On the other hand, Eulerian framework is
preferred for the fluid mechanics as fluid motion is observed in a fixed region of space. For
example the study of channel flow is done by observing water flow in a fixed region of a
channel and the study of air flow is done by observing air flow in a wind-tunnel. We now
present an example of bar with temperature distribution to understand the Lagrangian
and Eulerian formulations.
Example 1 (Temperature distribution of bar under uniaxial motion). Consider a bar of
initial length of four units, i.e., the reference domain Ω0 = [0, 4]. Let the bar is extending

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

with time t and the extension is given by the mapping function x = (1 + t)X. The bar is
experiencing the temperature distribution in Lagrangian description θ = X(4 − X)t2 or
in the Eulerian description θ = x(4 + 4t − x)t2 /(1 + t)2 (Fig. (2)).

X=1 X=2 X=3


θ = 27 θ = 36 θ = 27
t=3
X=1 X=2 X=3
θ =12 θ =16 θ =12
t=2
X=1 X=2 X=3
θ =3 θ =4 θ =3
t=1
x = (1 + t)X
X=1 X=2 X=3
θ=0 θ=0 θ=0
t=0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 x

Figure 2: Temperature distribution in moving bar

It is clear from Fig. (2) that the material description X is always associated with
particle in reference configuration. We can visualize material description in this problem
as thermometer fixed to the bar and measuring the temperature under the extension of
bar. It can be noted that thermometer moves with the bar in material description as
shown in Fig. (3). On the other hand, spatial description can be visualized as measuring
the temperature by fixing the thermometer to the axis x, i.e., thermometers are not moving
with the bar as shown in Fig. (4). It can be noted that the particles of bar occupy different
spatial positions at different time frames.

T1 , T2 , T3 Thermometers
T1 T2 T3
t=3

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 x
T1 T2 T3
t=2

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 x
T1 T2 T3
t=1
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 x
T1 T2 T3

t=0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 x

Figure 3: Lagrangian description

We now define displacement, velocity and acceleration of particles of continuum.

Displacement, velocity and acceleration:


Let Ω0 and Ω be reference and deformed configurations shown in Fig. (1). Then the
displacement of the particle located at X in reference configuration is defined by
u = x − X, ∀X ∈ Ω0 , (3)

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

T₁, T₂, T₃, T₄, T₅, T₆, T₇, T₈ Thermometers

t=3

1
T₁
2
T₂
3
T₃
4
T₄
5 6
T₅
7 8
T₆
9
T₇
10 11 12
T₈
13 14 15 16 x

t=2

1
T₁
2
T₂
3
T₃
4
T₄
5 6
T₅
7 8
T₆
9
T₇
10 11 12
T₈
13 14 15 16 x

t=1
1
T₁
2
T₂
3
T₃
4
T₄
5 6
T₅
7 8
T₆
9
T₇
10 11 12
T₈
13 14 15 16 x

t=0
1
T₁
2
T₂
3
T₃
4
T₄
5 6
T₅
7 8
T₆
9
T₇
10 11 12
T₈
13 14 15 16 x

Figure 4: Eulerian description

where x is the spatial position of particle at instant of time t. If the χ is transformation


from Ω0 to Ω then x = χ(X, t).

We observed from the example that material point X do not vary with time under the
deformation of continuum body. Therefore, the velocity of material point X is defined by

∂(x − X) ∂(χ(X, t) − X) ∂χ
u̇ = = = . (4)
∂t ∂t ∂t

Similarly, the acceleration is defined by

∂ 2χ
ü = . (5)
∂t2
We will present more detailed discussion on spatial description of velocity and acceleration
in Lecture-20. We now define an important tensor field called deformation gradient which
plays a critical role in defining the measure of strain.

The deformation gradient:


The deformation gradient F is defined as the gradient of mapping function x = χ(X, t).
Hence, we can write using Taylor’s expansion

χ(X + U ) − χ(X) = F U + o(U ),

where o(U ) represents higher-order terms, i.e.,

o(U )
lim = 0.
kU k→0 kU k

The deformation gradient F (X) characterizes the local behavior of χ(X, t) at X. Since
χ is the smooth vector-valued function of vector, the deformation gradient F is a second-
order tensor. The components of deformation gradient are given by
 ∂χ1 ∂χ1 ∂χ1 
" # ∂X ∂X2 ∂X3
∂χ  ∂χ21 ∂χ2 ∂χ2 
[F ] = = 
 ∂X1 ∂X2 ∂X3
 (6)
∂X ∂χ3 ∂χ3 ∂χ3

∂X1 ∂X2 ∂X3

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

We note that the deformation with constant gradient F (i.e., F is independent of material
variable X) is known as homogeneous deformation.

Using the definition of displacement field u(X, t) = χ(X, t) − X, we can write the
deformation gradient as
F = I + ∇u, (7)
where ∇u is known as displacement gradient. The relation can be expressed in the
component form as
∂ui
Fij = δij + . (8)
∂Xj

We will see in the next lecture that the deformation gradient play an important role
in defining the strain measures.

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. J. Bonet and R. D. Wood, Nonlinear Continuum Mechanics for Finite Element


Analysis, 1997, Cambridge University Press, Cambridge.

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-3: Kinematics
Lecture-18: The Strain and Stretch Tensors

Here, we discuss the strain measure that is related to the local change in length. Since
the generalized strain measure require the definition of stretch tensors, so their definition
also discussed in this lecture.
Strain tensors:
Let Ω0 and Ω be the reference and deformed configurations respectively (recall from
Lecture-17). Let χ(X, t) be a map from Ω0 to Ω and χ−1 (x, t) be its inverse map.
Let dX be a small fiber in the reference configuration and its transformation during
deformation be dx. Then we have the relation

dx = F dX, (1)

where F is the deformation gradient. As pointed out previously (see Lecture-17), the
deformation gradient is invertible tensor. Therefore, we have

dX = F −1 dx. (2)

The right and left Cauchy-Green strain tensors:


We now introduce two symmetric tensors using deformation gradient called right and left
Cauchy-Green strain tensors C and B. These strain tensors are defined by

C = FTF (3)
B = FFT. (4)

Green strain tensor:


Let dS be the length of small fiber dX in reference configuration and ds be the length of
same fiber dx after deformation. Then we can write

ds2 − dS 2 = dx · dx − dX · dX
= F dX · F dX − dX · dX
= dX · F T F dX − dX · dX
= dX · (C − I)dX
= 2dX · EdX.

The Green strain tensor E is defined by


1 1
E = (C − I) = (F T F − I). (5)
2 2

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

It can be noted that the Green strain tensor is a strain measure in Lagrangian description.
Almansi strain tensor:
We now define another measure of strain known as Almansi strain tensor. Expressing dX
in terms of dx in ds2 − dS 2 yields

ds2 − dS 2 = = dx · dx − dX · dX
= dx · dx − F −1 dx · F −1 dx
= dx · dx − dx · F −T F −1 dx
= dx · (I − B −1 )dx
= 2dx · Ẽdx.

The Almansi strain tensor is defined by


1 1
Ẽ = (I − B −1 ) = (I − F −T F −1 ). (6)
2 2
The Almansi strain tensor is an Eulerian measure of strain.

We can observe that the Green strain and the Almansi strain are related by

E = F T ẼF (7)

Clearly, both the Green strain and the Almansi strain measures are zero whenever there
is no change in length. Similar to Green and Almansi strain tensors, there are many
definitions for strain measure are possible such that the measure is zero when there is no
change in length. Of course the strain is non-zero when there is deformation experienced
by the body. We now define stretch tensors as they play an important role in definition
of generalized strain.

Stretch tensors:
The deformation gradient F is not only invertible but also det(F ) > 0 (see discussion in
Lecture-17). Therefore, using polar decomposition theorem, we can write the deformation
gradient as
F = RU = V R, (8)
where U and V are positive-definite symmetric tensors and R is the rotation tensor.
The positive-definite symmetric tensor U is known as right stretch tensor as it is the
square-root tensor of right Cauchy-Green strain C. Similarly, V is known as left stretch
tensor as it is the square-root tensor of left Cauchy-Green strain B. Using Eqs. (3), (4)
and (8), the left and right Cauchy-Green strains can be expressed by the stretch tensors,

C = U2
B = V 2,

Since RT R = RRT = I where R is rotation.

Let λ1 , λ2 , λ3 be eigenvalues of U and r1 , r2 , r3 be corresponding orthonormal eigen-


vectors. Then we have the following representation:
3
X
U= λi r i ⊗ r i . (9)
i=1

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Equation (8) implies V = RU RT and hence, the eigenvalues of U and V are same. But
eigenvectors are related by the rotation R.

Let l1 , l2 , l3 be orthonormal eigenvectors of V corresponding to eigenvalues λ1 , λ2 , λ3 .


Then we have the representation
3
X
V = λi li ⊗ li . (10)
i=1

Since V = RU RT , the eigenvectors of U and V , i.e., {r1 , r2 , r3 } and {l1 , l2 , l3 }, are


related by
li = Rri . (11)
Using Eqs. (8), (9), (11) and (10), the deformation gradient can be represented by
3 3 3 3
!
X X X X
F = RU = R λi r i ⊗ r i = λi R (ri ⊗ ri ) = λi (Rri ) ⊗ ri = λi l i ⊗ r i .
i=1 i=1 i=1 i=1

The Green and Almansi strain tensors can have the following representation in terms of
principal values and principal directions of stretches.
3
1 1 1 2
(C − I) = (U 2 − I) =
X
E = (λi − 1)ri ⊗ ri (12)
2 2 i=1 2
3
1 1 1
(I − B −1 ) = (I − V −2 ) = (1 − λ−2
X
Ẽ = i )li ⊗ li . (13)
2 2 i=1 2

We now show that Green and Almansi strain tensors are particular cases of generalized
strain defined in the following section.

Generalized strain tensors:


The generalized strain measure is motivated by the representation of Green and Almansi
strain tensors shown in Eqs. (12) and (13). Any measure is said to be qualified as
strain if it is zero when there is no change in length between any two points in the body.
Furthermore, it should provide, locally, change of length in any direction. These basic
requirements of strain are satisfied by the following measures.
3
1 n 1 n
E (n) =
X
(U − I) = (λi − 1)ri ⊗ ri (14)
n i=1 n
3
1 1
Ẽ (n) = (I − V −n ) = (1 − λ−n
X
i )li ⊗ li (15)
n i=1 n
Ẽ (−n) = RE (n) RT . (16)

The strain tensor E (n) is called material strain tensor and it is used in Lagrangian de-
scription. On the other hand, Ẽ (n) is called spatial strain tensor and it is used in Eulerian
description.

For every integer n the measure qualifies as strain measure. Therefore, these measure
of strains are known as generalized strains. The special cases of n = 0, 1 and 2 lead to
known strain measures.

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

The case n → 0 gives material and spatial logarithmic strain tensors.


3
E (0) =
X
(ln λi ) ri ⊗ ri (17)
i=1
3
Ẽ (0) =
X
(ln λi ) li ⊗ li . (18)
i=1

The logarithmic strain tensors E (0) and Ẽ (0) are also known as Hencky strain tensors.
Furthermore, the case n = 1 is known as Biot strain, i.e.,
3
E (1) = U − I =
X
(λi − 1) ri ⊗ ri . (19)
i=1

For the case n = 2, Eqs. (14) and (15) reduces to Green and Almansi strains shown in
Eqs. (12) and (13)
Problem 1 (Characterization of rigid motions-I). Let Ω0 and Ω be the reference and
deformed configurations. Then show that the motion of the body is rigid if and only if the
mapping function χ : Ω0 → Ω is in the form

χ(X, t) = Q(t)X + c(t), ∀X ∈ Ω0 , (20)

where Q(t) is a rotation tensor and c(t) is a vector. (Physically, the rigid motion of
reference configuration consists of only rotation Q(t) and translation c(t).)

Solution: The rigid motion is defined as a motion of body in which the length between
any two points is preserved. Let X1 and X2 be two points in the body. Since χ(X, t) is
given mapping between reference configuration Ω0 and the deformed configuration Ω, the
rigid motion can be expressed by

(χ(X1 , t) − χ(X2 , t)) · (χ(X1 , t) − χ(X2 , t)) = (X1 − X2 ) · (X1 − X2 ) ∀X1 , X2 ∈ Ω0 .

Taking the differentiation with respect to X2 , we get


1
 
T
F (X2 , t) (χ(X1 , t) − χ(X2 , t)) = X1 − X2 Since ∇(u · u) = (∇u)T u .
2
Again taking differentiation with respect X1 , we get

F (X2 , t)T F (X1 , t) = I, ∀X1 , X2 ∈ Ω0 . (21)

Since X1 and X2 are arbitrary, choosing X1 = X2 , Eq. (21) implies the orthogonality of
F (X, t) at every point X ∈ Ω0 . Since det F (X, t) > 0 ∀t ∈ [0, ∞), F (X, t) is a rotation.
We now assume that Q(t) = F (X, t) as F (X, t) is independent of position X. Inte-
grating the relation ∇χ = Q(t), we get

χ(X, t) = Q(t)X + c(t). (22)

Conversely, if the motion is in the form χ(X, t) = Q(t)X + c(t), we can write following
expression for every X1 and X2 in Ω0 .

(χ(X1 , t) − χ(X2 , t)) · (χ(X1 , t) − χ(X2 , t)) = Q(t)(X1 − X2 ) · Q(t)(X1 − X2 )


= (X1 − X2 ) · Q(t)T Q(t)(X1 − X2 ).

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Since Q(t) is a rotation tensor, we get

(χ(X1 , t) − χ(X1 , t)) · (χ(X1 , t) − χ(X1 , t)) = (X1 − X2 ) · (X1 − X2 ).

Thus, the motion is rigid if and only if χ(X, t) = Q(t)X + c(t).

Problem 2. Let Ω0 and Ω be reference and deformed configurations. The Green strain
E(X, t) = O, i.e., C(X, t) = I for all X ∈ Ωo if and only if the motion is rigid.

Solution: If the motion is rigid then it can be represented by (see Problem-1)

χ(X, t) = Q(t)X + c(t),

where Q(t) is the rotation tensor and c(t) is the translation vector.
Using the definition of deformation gradient, we get F (X, t) = Q(t). Therefore, this
implies the Green strain tensor,
1
E(X, t) = (F T F − I) = O, ∀X ∈ Ω0 .
2
Conversely, we now show that E(X, t) = O, ∀X ∈ Ω0 implies rigid motion. Let dS be
infinitesimal length of vector dX at a point X in the reference configuration and ds be
the transformed infinitesimal length in deformed configuration at x = χ(X, t). Locally,
dS = ds if the Green strain E(X, t) = O. We now consider a straight line between two
points X1 and X2 in the reference configuration such that the length is L. Suppose the
straight line is transformed into a curve in the deformed configuration with end points
x1 = χ(X1 , t) and x2 = χ(X2 , t). At an instant of time t, the condition dS = ds implies

L = |X1 − X2 | ≥ |x1 − x2 |.

We now consider a curve in reference configuration which gets transformed into a straight
line. At an instant of time t, the condition dS = ds implies

|X1 − X2 | ≤ |x1 − x2 |.

Using both inequalities shown above, we get

|x1 − x2 | = |χ(X1 , t) − χ(X2 , t)| = |X1 − X2 |.

Thus, the motion is rigid if and only if E(X, t) = O, ∀X ∈ Ω0 .

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. J. Bonet and R. D. Wood, Nonlinear Continuum Mechanics for Finite Element


Analysis, 1997, Cambridge University Press, Cambridge.

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-3: Kinematics
Lecture-19: Area and Volume Transformation

In Lecture-18, transformation of infinitesimal length and strain measures are discussed


when continuum is experiencing the deformation. We now present the transformation
of infinitesimal area and volume elements when deformation is taking place. We also
introduce an important transformation known as Piola transform that relates the material
and spatial integrals.

Area element and the Piola transform:


Let Ω0 be a reference configuration with the boundary Γ0 and Ω be a deformed configu-
ration with the boundary Γ at an instant of time t. Let T (X, t) be a second-order tensor
field over Ω0 . Then the divergence of T with respect to material coordinate X (i.e., in
material description) is given by
∂Tij
∇X · T = ei .
∂Xj
Similarly, let τ (x, t) be a second-order tensor field defined over Ω. Then the divergence
of τ with respect to spatial coordinate x (i.e., in spatial description) is given by
∂τij
∇x · τ = ei .
∂xj
Using divergence theorem for tensor fields, we get
Z Z
∇X · T ∂Ω0 = T n0 ∂Γ0 , (1)
Ω0 Γ0
Z Z
∇x · τ ∂Ω = τ n ∂Γ, (2)
Ω Γ

where n0 and n denote unit outward normals to Γ0 and Γ, respectively.


Let χ(X, t) be a mapping from Ω0 to Ω and F (X, t) be the deformation gradient, i.e.,
F (X, t) = ∇χ. Then the Piola transformation is defined by

T (X) = Jτ (x)F −T = τ (x)cof F , (3)

where J is det F .

If two second-order tensor fields, T (X, t) and τ (x, t) stated above are related by the
Piola transformation then we have the following results:

• Relation between divergence of tensor fields,

∇X · T (X, t) = J∇x · τ (x, t) ∀x = χ(X, t). (4)

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

• Relation on the reference and deformed boundaries,

T (X, t)n0 ∂Γ0 = τ (x, t)n ∂Γ ∀x = χ(X, t). (5)

where ∂Γ0 is infinitesimal area element on the boundary of reference domain Γ0 and
∂Γ is the corresponding infinitesimal area element on the boundary of deformed
domain Γ.

• Relation between area elements

∂Γ = J|F −T n0 | ∂Γ0 = |(cof F )n0 | ∂Γ0 . (6)

This relation is known as Nanson’s formula.

• Relation between outward normals


(cof F )n0
n= . (7)
|(cof F )n0 |

We now prove the relations in steps.

Problem 1 (Piola identity). Let χ(X, t) be a transformation from reference configura-


tion Ω0 to the deformed configuration Ω. Let F (X, t) be the deformation gradient. Then
show that
∇X · (JF −T ) = ∇X · cof F = 0, (8)
where J is det F . The relation shown Eq. (8) is known as Piola identity.

Solution: Let u and v be two arbitrary constant vectors. Then


 
(∇X · cof(F )) · (u × v) = ∇X · (cof F )T (u × v)
 
= ∇X · cof(F T )(u × v) (since (cof F )T = cof(F T ))
 
= ∇X · (F T u) × (F T v)
= (F T v) · (∇X × (F T u)) − (F T u) · (∇X × (F T v))
(since ∇ · (a × b) = b · (∇ × a) − a · (∇ × b))
= (F T v) · ((∇X × F )u)) − (F T u) · ((∇X × F )v))
= 0 (since ∇X × F = ∇X × (∇X (χ)) = O) .

Arbitrariness of u and v implies ∇X · cof F = 0. Since cof F = JF −T , we get

∇X · (JF −T ) = ∇X · cof F = 0,

Problem 2. Let χ(X, t) be a transformation from reference configuration Ω0 to the de-


formed configuration Ω. Let F (X, t) be the deformation gradient. Let T (X, t) and τ (x, t)
be two second-order tensors related by Piola transformation, i.e., T (X, t) = τ (x, t)cof F .
Then show that
∇X · T = J∇x · τ , (9)
where J is det F .

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Solution: Let u be an arbitrary constant vector. Then

(∇X · T ) · u = ∇X · (T T u) (Using the definition of divergence)


 
T
= ∇X · (τ cof F ) u (Using Piola transformation)
 
= ∇X · (cof F )T τ T u

Applying the relation ∇ · (S T v) = (∇ · S) · v + S : ∇v, we get

(∇X · T ) · u = (∇X · cof F ) · (τ T u) + (cof F ) : (∇X (τ T u))


= (cof F ) : (∇X (τ T u)) (Since ∇X · cof F = 0)
  
T
= (cof F ) : ∇x (τ u) F (Using chain rule)
 
= (cof F )F T : ∇x (τ T u) (Since R : (SG) = (RGT ) : S)
= JI : ∇x (τ T u) (Since (cof S)S T = det(S)I)
= J∇x · (τ T u)
= J(∇x · τ ) · u

Arbitrariness of u implies ∇X · T = J∇x · τ .

Problem 3. Let χ(X, t) be a transformation from reference configuration Ω0 to the de-


formed configuration Ω. Let Γ0 and Γ be the boundary of Ω0 and Ω, respectively. Let
F (X, t) be the deformation gradient. Let T (X, t) and τ (x, t) be two second-order ten-
sors related by the Piola transformation T (X, t) = τ (x, t)cof F . Then show that

T n0 ∂Γ0 = τ n ∂Γ
∂Γ = |(cof F )n0 | ∂Γ0
(cof F )n0
n = .
|(cof F )n0 |

where n0 and n are outward normal fields to Γ0 and Γ, respectively.

Solution: We use the relation ∂Ω = J∂Ω0 where J is det F in following proof. The proof
for the relation ∂Ω = J∂Ω0 is presented in the following section. We can obtain the
following relation
Z Z
0
T n ∂Γ0 = ∇X · T ∂Ω0 (Using divergence theorem)
Γ0 Ω0
Z
= ∇x · τ ∂Ω (Using Eq. (9) and ∂Ω = J∂Ω0 )
ZΩ
= τ n ∂Γ. (Using divergence theorem)
Γ

Arbitrariness of Γ implies T n0 ∂Γ0 = τ n ∂Γ. Choosing τ = I, we get

n ∂Γ = (cof F )n0 ∂Γ0 (Since T = τ cof F ).

Since n is unit vector, we get

∂Γ = |(cof F )n0 | ∂Γ0 .

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Substituting the expression for ∂Γ in the relation for n ∂Γ, we get


(cof F )n0
n= .
|(cof F )n0 |
This completes the proof of all three relations stated in the problem.
Volume element:
Let Ω0 and Ω be reference and deformed configurations. Let x = χ(X, t) be a mapping
function from Ω0 to Ω. Let F (X, t) be the deformation gradient. Then the volume
elements in two configurations are related by

∂Ω = J ∂Ω0 , (10)

where J is det F .
Proof: The volume can be defined as the absolute value of triple scalar product. Let dX1 ,
dX2 and dX3 are linearly independent infinitesimal vectors that form a parallelepiped in
the reference configuration as shown in the Fig. (1). Let us consider infinitesimal vectors
dX1 , dX2 and dX3 which are being transformed into dx1 , dx2 and dx3 , respectively. The
infinitesimal vectors dx1 , dx2 and dx3 in deformed configuration form a parallelepiped in
the deformed configuration as shown in Fig. (1).

x = χ(X,t)
dx₃ Ω
e3 dx₂ dx₁

Ω0

e1 e2

Figure 1: Volume transformation

Let ∂Ω be the volume of parallelepiped in the deformed configuration. Then using


third invariant of second-order tensor (see Problem-3 in Lecture-5) we get the following
relation between volume elements.

∂Ω = |[dx1 , dx2 , dx3 ]| = |[F dX1 , F dX2 , F dX3 ]| = |det(F )||[dX1 , dX2 , dX3 ]|
= J ∂Ω0 ,

where ∂Ω0 is the volume of parallelepiped in undeformed configuration. Thus, the volume
elements are related by the determinant of the deformation gradient.

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-3: Kinematics
Lecture-20: Material and Spatial Time Derivatives

The velocity, acceleration and velocity gradient are important quantities of kinematics.
Here, we discuss the description of these kinematic quantities in both Lagrangian and
Eulerian framework.
Material and spatial description of velocity field:
Let Ω0 be a reference configuration and Ω be a deformed configuration at an instant of
time t. Let x = χ(X, t) be a mapping that represents the motion from Ω0 to Ω. Since
χ is invertible for all X ∈ Ω0 at every instant of time t, the inverse map X = χ−1 (x, t)
exists from Ω to Ω0 which associates every spatial point x to a material point X.
The material coordinate X is fixed to a particle which is independent of time (see
Example-1 in Lecture-17). On the other hand, the spatial position of particle x can vary
with time. The displacement of particle is defined by the relation u = x − X. Therefore,
the velocity of particle
∂(x − X) ∂x
u̇ = = (1)
∂t ∂t
Substituting the map x = χ(X, t) in u̇, we get velocity ṽ in terms of X as
∂x ∂χ(X, t)
ṽ(X, t) = = (2)
∂t ∂t
This represents the velocity field in Lagrangian description (material description) as it is
a function of material variable X.
Using the inverse map X = χ−1 (x, t), we can obtain the following Eulerian description
(spatial description) of velocity field
v(x, t) = ṽ(χ−1 (x, t), t). (3)
In general, the function forms of ṽ and v are different while the value of the functions
are equal at a material point X and corresponding spatial point x = χ(X, t). We now
present an example to understand the material and spatial description of velocity fields.
Example 1. Let the mapping function for motion x = χ(X, t) is given by
x1 = X1 (1 + t2 ), x2 = X2 (1 + t), x3 = 0.
Then find the velocity in material and spatial descriptions.

Solution: Consider given motion x1 = X1 (1 + t2 ), x2 = X2 (1 + t), x3 = 0. Since material


∂χ
description of velocity ṽ = , we get
∂t
∂x1 ∂x2 ∂x3
ṽ1 = = 2X1 t, ṽ2 = = X2 , ṽ3 = = 0.
∂t ∂t ∂t

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Using the motion representation, we can obtain X1 = x1 /(1 + t2 ) and X2 = x2 /(1 + t).
Substituting in the material description, we get the following spatial velocity field
2x1 t x2
v1 = , v2 = , v3 = 0.
1 + t2 1+t
It is clear from the example that ṽ and v are having different forms whereas both the
forms give same value of the velocity corresponding material and spatial coordinates, i.e.,
ṽ(X, t) = v(x, t)
For example, material description of velocity ṽ = (4, 2, 0) at a material point X =
(1, 2, 3) at an instant of time t = 2. We now see that x = (5, 6, 0) is a spatial coor-
dinate corresponding to the material point X = (1, 2, 3) at an instant of time t = 2.
Substituting x in spatial description of velocity, we get v = (4, 2, 0). Therefore, the
values of velocity are equal both in material or spatial at corresponding points, i.e.,
ṽ(X, t)|(1,2,3) = v(x, t)|(5,6,0)
We now introduce the concept of material and spatial time derivatives of more general
fields. Later, we apply this result to get the acceleration in both descriptions.
Material and spatial time derivative of general fields:
There is a possibility of confusion in taking material and spatial time derivatives of general
fields. Hence, we introduce the following notations to avoid the confusion in taking the
differentiation of fields.
Let φ̃(X, t), g̃(X, t), G̃(X, t) be scalar, vector and tensor fields in the Lagrangian
(material) description, respectively. Then the material gradient of scalar, vector and
tensor fields are denoted by ∇X φ̃, ∇X g̃, ∇X G̃, the material divergence of vector and
tensor field are denoted by ∇X · g̃, ∇X · G̃, and the material time derivative of scalar,
Dφ̃ ˙ Dg̃ ˙ DG̃
vector and tensor fields are denoted by φ̃˙ or , g̃ or , G̃ or . In component
Dt Dt Dt
form
∂ φ̃ ∂ g̃i ∂ G̃ij
(∇X φ̃)i = , (∇X g̃)ij = , (∇X G̃)ijk = ,
∂Xi ∂Xj ∂Xk
∂ g̃i ∂ G̃ij
∇X · g̃ = , (∇X · G̃)i = ,
∂Xi ∂Xj
Dφ̃ ∂ φ̃ ˙ Dg̃ ∂ g̃ ˙ = DG̃ = ∂ G̃ .
φ̃˙ = = , g̃ = = , G̃
Dt ∂t Dt ∂t Dt ∂t
Let φ(x, t), g(x, t), G(x, t) be scalar, vector and tensor fields in the Eulerian or spatial
description. Then the spatial gradient of scalar, vector and tensor fields are denoted by
∇x φ, ∇x g, ∇x G, the spatial divergence of vector and tensor field are denoted by ∇x · g,
∇x · G, and the spatial time derivative of scalar, vector and tensor fields are denoted by
φ0 , g 0 , G0 . In component form
∂φ ∂gi ∂Gij
(∇x φ)i = , (∇x g)ij = , (∇x G)ijk = ,
∂xi ∂xj ∂xk
∂gi ∂Gij
∇x · g = , (∇x · G)i = ,
∂xi ∂xj
∂φ ∂g ∂G
φ0 = , g0 = , G0 = .
∂t ∂t ∂t

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

We now derive the relation between the material and spatial gradients.
Relation between material and spatial gradients:
Let Ω0 be a reference configuration and Ω be a deformed configuration. Let x = χ(X, t)
be a map from Ω0 to Ω and F = ∇X χ be the deformation gradient. Let φ(x, t) and
g(x, t) be scalar and vector fields defined over Ω. Then we have
∂φ ∂φ ∂χj ∂φ
(∇X φ)i = = = Fji
∂Xi ∂xj ∂Xi ∂xj
∂gi ∂gi ∂χk ∂gi
(∇X g)ij = = = Fkj
∂Xj ∂xk ∂Xj ∂xk
or, equivalently,

∇X φ = F T ∇x φ (4)
∇X g = ∇x gF . (5)

We now use the relation between material and spatial gradients to represent Green and
Almansi strain tensors in terms of displacement gradients. We know that the displacement
u = x − X. Furthermore, the deformation gradient ∇X u = F − I. Therefore, the Green
strain tensor,
1 T  1 
E= F F −I = ∇X u + ∇X uT + ∇X uT ∇X u . (6)
2 2
The Green strain tensor in component form,
!
1 ∂ui ∂uj ∂uk ∂uk
Eij = + + . (7)
2 ∂Xj ∂Xi ∂Xi ∂Xj

Using Eq. (5), we get spatial gradient of displacement

∇x u = (∇X u) F −1 = (F − I)F −1 = I − F −1 . (8)

Substituting F −1 = I − ∇x u in Almansi strain tensor (see Eq. (6) in Lecture-18), we get


1  1 
Ẽ = I − F −T F −1 = ∇x u + ∇x uT − ∇x uT ∇x u . (9)
2 2
The Almansi strain tensor in component form,
!
1 ∂ui ∂uj ∂uk ∂uk
Ẽij = + − . (10)
2 ∂xj ∂xi ∂xi ∂xj

The relations between spatial and material gradients are shown in Eqs. (4) and (5). We
now want to derive relation between material and spatial time derivatives. In order to
get the relation between time derivatives, we need to have the following chain rule.

Chain rule:
Let ξ(t) be a scalar valued function of time t. Let ψ(ξ(t), t) be a scalar valued function
which depends explicitly and implicitly, through ξ, on t. Then the total derivative with
respect to time t can be written as
dψ ∂ψ ∂ψ dξ
= + . (11)
dt ∂t ∂ξ dt

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Proof: Consider Taylor’s expansion of ψ(ξ(t + ∆t), t + ∆t).


!
∂ψ
ψ(ξ(t + ∆t), t + ∆t) = ψ(ξ(t + ∆t), t) + + o(∆t)


∂t
ξ(t+∆t)
! !
dξ ∂ψ
= ψ ξ(t) + ∆t + o(∆t), t + + o(∆t)


dt ∂t
ξ(t+∆t)
! !
∂ψ dξ ∂ψ
= ψ(ξ(t), t) + ∆t + + o(∆t)


∂ξ dt ∂t
ξ(t+∆t)
 ! ! 
∂ψ dξ ∂ψ
= ψ(ξ(t), t) +  +  ∆t + o(∆t)


∂ξ dt ∂t
ξ(t+∆t)

Using definition of differentiation, we get


dψ ψ(ξ(t + ∆t), t + ∆t) − ψ(ξ(t), t)
= lim
dt ∆t→0 ∆t
∂ψ ∂ψ dξ
= + .
∂t ∂ξ dt
∂ψ ∂ψ
we should treat ξ as fixed while computing . Similarly, t is fixed for evaluating .
∂t ∂ξ
This formula works for scalar function, we now extend this to vector fields.
Let g(t) be a vector valued function of time t. Let Φ(g(t), t) be a vector valued function
which depends explicitly and implicitly, through g, on t. Then the following chain rule
can be obtained for the total derivative of Φ.
dΦi ∂Φi ∂Φi dgj
= + . (12)
dt ∂t ∂gj dt
This result can be proved, similar to the previous case of scalar valued function.
Relation between material and spatial time derivatives:
Let φ̃(X, t), g̃(X, t) and G̃(X, t) be scalar, vector and tensor valued function over Ω0 ,
respectively. Let φ(x, t), g(x, t) and G(x, t) be scalar, vector and tensor valued function
over Ω, respectively. Let these functions are related by the map x = χ(X, t), i.e.,
φ̃(X, t) = φ(x, t), g̃(X, t) = g(x, t) and G̃(X, t) = G(x, t) for x = χ(X, t). Then

Dφ̃
• the material time derivative of scalar field and the spatial time derivative of
Dt
∂φ
scalar field are related by
∂t
Dφ̃(X, t) Dφ(x, t) ∂φ ∂φ ∂χi ∂φ
= = + = + (∇x φ) · v.
Dt Dt ∂t ∂xi ∂t ∂t

Dg̃
• the material time derivative of vector field and the spatial time derivative of
Dt
∂g
vector field are related by
∂t
Dg̃(X, t) Dg(x, t) ∂g ∂g ∂χi ∂g
= = + = + (∇x g)v.
Dt Dt ∂t ∂xi ∂t ∂t

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

DG̃
• the material time derivative of tensor field and the spatial time derivative of
Dt
∂G
vector field are related by
∂t
DG̃(X, t) DG(x, t) ∂G ∂G ∂χi ∂G
= = + = + (∇x G)v.
Dt Dt ∂t ∂xi ∂t ∂t

We note that the material time derivative in reference domain can be obtained by

Dφ̃(X, t) ∂ φ̃ Dg̃(X, t) ∂ g̃ DG̃(X, t) ∂ G̃


= , = , and = .
Dt ∂t Dt ∂t Dt ∂t
We now present an example to understand the relation between material and spatial time
derivatives. Recall from Lecture-17 about the example of temperature distribution of bar
under uniform motion presented in Lagrangian and Eulerian description. We now use the
same example to explain the relation between material time derivative and spatial time
derivative.

Example 2. Consider a bar of initial length of four units, i.e., the reference domain
Ω0 = [0, 4]. Let the bar is extending with time t and the extension is given by the mapping
function x = (1+t)X. The bar is experiencing the temperature distribution in Lagrangian
description θ̃(X, t) = X(4 − X)t2 or in the Eulerian description θ(x, t) = x(4 + 4t −
x)t2 /(1 + t)2 shown in Fig. (1).

X=1 X=2 X=3


θ = 27 θ = 36 θ = 27
t=3
X=1 X=2 X=3
θ =12 θ =16 θ =12
t=2
X=1 X=2 X=3
θ =3 θ =4 θ =3
t=1
x = (1 + t)X
X=1 X=2 X=3
θ=0 θ=0 θ=0
t=0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 x

Figure 1: Temperature distribution in moving bar

It is easy to see that θ̃(X, t) = θ(x, t) for every x = χ(X, t) = (1 + t)X.


Consider the temperature distribution in Lagrangian framework, θ̃(X, t). Using the
definition of material derivative, we get

Dθ̃(X, t) ∂ θ̃
= = 2X(4 − X)t.
Dt ∂t
We now consider the spatial description of temperature, θ = x(4 + 4t − x)t/(1 + t)2 . Using
the definition, we get the spatial time derivative as
∂θ 2xt
= 3
(4 + 6t − x + 2t2 ).
∂t (1 + t)

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

We need the following terms to get total derivative in spatial description.


∂θ 2t2 ∂x x
= (2 + 2t − x) and = .
∂x (1 + t)2 ∂t (1 + t)
Substituting above terms, we get total derivative
Dθ ∂θ ∂θ ∂x 2xt
= + = (4 + 4t − x).
Dt ∂t ∂x ∂t (1 + t)2
It can be observed that for x = (1 + t)X

Dθ̃ Dθ
= .
Dt Dt
Thus, the material time derivative is equal to the total derivative in spatial description.
In general, material time derivative and spatial time derivative are not equal.

Material and spatial description of acceleration field:


Let Ω0 be a reference configuration and let Ω be a deformed configuration. Let x =
χ(X, t) be a mapping from Ω0 to Ω. Let ṽ(X, t) and v(x, t) be velocity fields described
in the Lagrangian and Eulerian framework, i.e., v(x, t) = ṽ(X, t) for every x = χ(X, t).
Then the acceleration field is defined by
Dv ∂v
a(x, t) = = + (∇x v)v. (13)
Dt ∂t
The second term in the equation, (∇x v)v, is known as convective acceleration. Therefore,
the acceleration is sum of the spatial time derivative and the convective acceleration.
The material description of acceleration is given by
Dṽ ∂ ṽ
ã(X, t) = = . (14)
Dt ∂t
The material description of acceleration field is equivalent to spatial description, i.e.,
ã(X, t) = a(x, t) for x = χ(X, t).
Velocity gradient:
The velocity gradient L is defined by

L = ∇x v, (15)

where v is the velocity field. Using L, the acceleration can be written as


∂v
a(x, t) = + Lv. (16)
∂t
The velocity gradient L is a second-order tensor whose action on velocity gives the con-
vective acceleration.

Relation between F and L:


The velocity gradient L is defined as the spatial gradient of velocity. On the other hand,
the material time derivative of deformation gradient gives the material gradient of velocity
i.e.,
DF ∂ ∂χ(X, t)
= (∇X χ(X, t)) = ∇X = ∇X v.
Dt ∂t ∂t

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

Using Eq. (5), we can obtain the following relation between spatial and material gradient
of velocity.
DF
= ∇X v = (∇x v) F = LF .
Dt
Since the tensor F is invertible, we have F F −1 = I. Taking the material time derivative
on both sides, we get

D (F F −1 ) DF −1 DF −1
= F +F =O
Dt Dt Dt
DF −1 DF −1
=⇒ = −F −1 F = −F −1 L.
Dt Dt
The material derivative of J (i.e., det(F )) in can be written as

DJ DF
= (cof F ) : (see an epression in Example-1 of Lecture-16)
Dt Dt
= (cof F ) : (LF )
= F (cof F )T : LT
= JI : LT
= J tr(L)
= J ∇x · v.

Since J represents locally the ratio of deformed volume to reference volume, the isochoric
motion (J = 1) can be characterized by

1 DJ
= ∇x · v = 0. (17)
J Dt
In summary, we have the relations
DF
= LF (18)
Dt
DF −1
= −F −1 L (19)
Dt
1 DJ
= tr(L) = ∇x · v (20)
J Dt

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. M. E. Gurtin, E. Fried and L. Anand, The Mechanics and Thermodynamics of


Continua, 2010, Cambridge University Press, New York.

3. J. Bonet and R. D. Wood, Nonlinear Continuum Mechanics for Finite Element


Analysis, 1997, Cambridge University Press, Cambridge.

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-3: Kinematics
Lecture-21: The Rate of Deformation and Spin

In kinematics, the rate of deformation and spin tensors are also useful in addition to
previously defined stretch and strain tensors. In particular, these are important in fluid
mechanics, visco-elasticity and plasticity. Therefore, we now define the rate of deformation
and spin tensors.
Rate of deformation and spin tensors:
Let Ω0 and Ω be reference and deformed configurations, respectively. Let v be a velocity
field over Ω at an instant of time t. Let y be a point in Ω and x ∈ Ω be a neighborhood
of y, i.e., |y − x| <  for a given  ∈ <. Let L be the spatial velocity gradient. Then,
using the definition of gradient, we can write
v(x, t) − v(y, t) = L(y, t)(x − y) + o(|x − y|), (1)
where o(|x − y|) approaches zero faster than |x − y|. As |x − y| → 0, the second-order
tensor L(y, t) represents the velocity gradient at a point y at an instant of time t. Let
D and W be the symmetric and skew-symmetric parts of L, i.e.,
1 1
D = (L + LT ) = (∇x v + ∇x v T ), (2)
2 2
1 1
W = (L − L ) = (∇x v − ∇x v T ).
T
(3)
2 2
Substituting D and W in the above expansion, we get
v(x, t) − v(y, t) = W (y, t)(x − y) + D(y, t)(x − y) + o(|x − y|).
Since W is skew-symmetric tensor, we show in the following problem that v(y, t) +
W (y, t)(x − y) represents the velocity field of rigid motion. Therefore, W is called spin
tensor. The other part of velocity field
D(y, t)(x − y)
represents the deformation with time, i.e., the rate of stretch of fibers of continuum body
locally at an instant of time t. Therefore, the tensor D is called rate of deformation
tensor or stretching tensor. It should be noted that the stretch tensor is different from
the stretching tensor as stretch is not rate of deformation (see definition of stretch tensor
in Lecture-18).
We now derive the relation between right and left stretch tensors of the deformation
gradient F , i.e., U and V , and the velocity gradient L. Using the polar decomposition
theorem (see Lecture-18), we have
F = RU = V R

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

where U is right stretch tensor, V is left stretch tensor and R is the rotation tensor. We
have the following relation between L and F (see Lecture-20)

DF −1 DF
L= F (Since = LF ).
Dt Dt
Substituting right polar decomposition F = RU , we get

D(RU )
L = (RU )−1
Dt !
DR DU
= U +R U −1 RT
Dt Dt
!
DR T DU −1
= R +R U RT .
Dt Dt

Since RRT = I, we have


!T
DR T DR T
R =− R .
Dt Dt
DR T
Clearly, R is a skewsymmetric tensor. Hence, the symmetric and skewsymmetric
Dt
parts of L can be written as
!!
DU −1
D = R sym U RT (4)
Dt
!!
DR T DU −1
W = R + R skw U RT . (5)
Dt Dt

where
! !
DU −1 1 DU −1 DU −1
sym U = U + U , and
Dt 2 Dt Dt
! !
DU −1 1 DU −1 DU −1
skw U = U − U .
Dt 2 Dt Dt

It is interesting to see that W has two contributions: one due to rotation R denoted by
Wrot and other due to stretch tensor denoted by Wstr , i.e.,

W = Wrot + Wstr , (6)

where
DR T
Wrot = R
Dt !!
DU −1
Wstr = R skw U RT .
Dt

Relation between velocity and spatial position in rigid motion:


We now derive the rigid motion in terms of velocity field. Let X1 and X2 be two material
points in Ω0 . Then the distance is defined by

d(t) = |x1 − x2 | = |χ(X1 , t) − χ(X2 , t)|.

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

The motion is said to be rigid if the distance between two arbitrary material points is
preserved during the motion, i.e.,

|χ(X1 , t) − χ(X2 , t)| = 0, ∀X1 , X2 ∈ Ω0 .
∂t
Taking differentiation, we get

(x1 − x2 ) · (v(x1 , t) − v(x2 , t)) = 0, ∀x1 , x2 ∈ Ω. (7)

We now prove a result in differential equations that relates skewsymmetric and rotation
tensors. Note that following result is converse to Problem-4 of Lecture-16.

Problem 1. Let Q(t) be a second-order tensor which is continuous with respect to time.
Let Ŵ (t) be a skewsymmetric tensor related to Q(t) by
DQ
= Ŵ Q, (8)
Dt
and Q(0) = I. Then show that Q(t) is rotation tensor 1 .

Proof. The tensor Ŵ is skewsymmetric in Eq. (8). Therefore, taking transpose, yields

DQT
= −QT Ŵ . (9)
Dt
Taking material time derivative of QT Q, we get

D DQT DQ
(QT Q) = Q + QT
Dt Dt Dt
= −QT Ŵ Q + QT Ŵ Q (Substituting Eqs. (8) and (9))
= O

The unique solution Q(t)T Q(t) = I, ∀ t ∈ [0, ∞) to the differential equation follows from
the initial condition Q(0) = I.
The continuity of Q implies det(Q(t)) = 1 as Q(0) = I. Thus, Q(t) is not just an
orthogonal tensor but a rotation tensor.

Problem 2 (Characterization of rigid motion-II). Let x = χ(X, t) and v be a motion


and corresponding velocity field of continuum, respectively. Then the motion is rigid if
and only if the velocity v admits the representation

v(x1 , t) = v(x2 , t) + Ŵ (t)(x1 − x2 ) ∀x1 , x2 ∈ Ω, (10)

where Ŵ is a skewsymmetric tensor.

Proof. We showed in Lecture-18 (see Problem-1 in Lecture-18) that the rigid motion
should have the form
χ(X, t) = Q(t)X + c(t), (11)
1
The proof for this problem is taken from Bar-Itzhack, I. Y., and Markley, F. L., “Minimal parameter
solution of the orthogonal matrix differential equation”, IEEE Transactions on Automatic Control, 1990,
35(7), 314-317

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

where Q(t) is rotation tensor and c(t) is a vector. We now show that Eq. (11) implies
Eq. (10). Let X1 and X2 be two material points in Ω0 . Then we have

χ(X1 , t) = Q(t)X1 + c(t) and χ(X2 , t) = Q(t)X2 + c(t).

Combining above equations, yields

χ(X2 , t) − χ(X1 , t) = Q(t)(X2 − X1 ),

or, alternatively,
x2 − x1 = Q(t)(X2 − X1 ). (12)
Taking material derivative, we get
D DQ
(χ(X1 , t) − χ(X2 , t)) = (X1 − X2 )
Dt Dt
DQ −1
=⇒ v(x1 , t) − v(x2 , t) = Q (χ(X1 , t) − χ(X2 , t))
Dt
DQ T
=⇒ v(x1 , t) − v(x2 , t) = Q (x1 − x2 ).
Dt
DQ T DQ T
Since QQT = I implies Q is a skewsymmetric tensor. Let us denote Q with
Dt Dt
Ŵ . Then we have
v(x1 , t) − v(x2 , t) = Ŵ (t)(x1 − x2 ).
Clearly, rigid motion implies Eq. (10).
Conversely, we now show that Eq. (10) implies Eq. (11), i.e., Eq. (10) also represents
rigid motion. Applying Eq. (11) to both material points X1 and X2 yields Eq. (12).
Therefore, it is enough to show that Eq. (10) implies Eq. (12).
Let z(t) = x1 − x2 = χ(X1 , t) − χ(X2 , t). Then we have
Dz
= Ŵ (t)z, z(0) = X1 − X2 . (13)
Dt
Integrating above differential equation, we get
Z t
z(t) = z(0) + Ŵ (τ 0 )z(τ 0 )dτ 0 (14)
0

From above equation, we can obtain


Z τ0
0
z(τ ) = z(0) + Ŵ (τ 00 )z(τ 00 )dτ 00 (15)
0

Substitution of Eq. (15) in Eq. (14), yields


Z t  Z t Z τ0 !
0 0 0 00 00 00
z(t) = z(0) + Ŵ (τ )dτ z(0) + Ŵ (τ ) Ŵ (τ )z(τ )dτ dτ 0 . (16)
0 0 0

Repeating this process, we get the following solution.


" Z t Z t Z τ0 !
0 0 0 00 00
z(t) = I+ Ŵ (τ )dτ + Ŵ (τ ) Ŵ (τ )dτ dτ 0
0 0 0
Z t (Z 0 Z τ 00 ! ) #
τ
0 00 000 000 00 0
+ Ŵ (τ ) Ŵ (τ ) Ŵ (τ )dτ dτ dτ + · · · z(0). (17)
0 0 0

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Let
" Z t Z t Z τ0 !
0 0 0 00 00
R(t) = I+ Ŵ (τ )dτ + Ŵ (τ ) Ŵ (τ )dτ dτ 0
0 0 0
Z t (Z 0 Z τ 00 ! ) #
τ
0 00 000 000 00 0
+ Ŵ (τ ) Ŵ (τ ) Ŵ (τ )dτ dτ dτ + · · · . (18)
0 0 0

Then the solution to Eq. (13) is given by

z(t) = R(t)z(0). (19)

Comparing above equation with Eq. (12), it is enough to show that R(t) is a rotation
tensor.

Taking time derivative on both sides of Eq. (18), we get


" Z t Z t Z τ 00 ! #
DR 00 00 00 000 000 00
= Ŵ (t) + Ŵ (t) Ŵ (τ )dτ + Ŵ (t) Ŵ (τ ) Ŵ (τ )dτ dτ · · ·
Dt 0 0 0

Clearly, we have the relation


DR
= Ŵ (t)R(t). (20)
Dt
Applying the result of Problem-1, it is easy to see that R(t) is a rotation tensor. We also
note that R(0) = I. This completes the proof.

Problem 3. The motion is rigid if and only if the rate of deformation D(x, t) = O, ∀
x ∈ Ω and t ∈ [0, ∞).

Proof. We showed in previous problem (i.e., Problem-2) that the motion is rigid if and
only if the motion follows Eq. (10). Let ĉ(t) be a constant vector. Then Eq. (10) can be
written as

v(x1 , t) − Ŵ (t)x1 = v(x2 , t) − Ŵ (t)x2 = ĉ(t) ∀x1 , x2 ∈ Ω,

Since x1 and x2 are arbitrary in above equation, we can have the following equation of
motion for continuum.
v(x, t) = Ŵ (t)x + ĉ(t) ∀x ∈ Ω.
Taking the gradient on both sides of above equation, we get velocity gradient L = Ŵ .
Since velocity gradient is skewsymmetric tensor, the rate of deformation D(x, t) = O.
Conversely, we now show that D(x, t) = O for all x ∈ Ω implies the motion is rigid.
We have the following relation in indicial form as D = O.
∂vi ∂vj
=− .
∂xj ∂xi

Taking gradient on both sides, we can get


!
∂ 2 vi ∂ 2 vj ∂ 2 vj ∂ ∂vj
=− =− =− .
∂xk ∂xj ∂xk ∂xi ∂xi ∂xk ∂xi ∂xk

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

∂vj ∂vk
Since =− , we can write above equation
∂xk ∂xj
! ! !
∂ 2 vi ∂ ∂vj ∂ ∂vk ∂ ∂vk ∂ 2 vi
=− = = =−
∂xk ∂xj ∂xi ∂xk ∂xi ∂xj ∂xj ∂xi ∂xk ∂xj
∂vk ∂vi
Last step follows from =− . In the above equation, right hand side term is same
∂xi ∂xk
as that of left hand side term with negative sign. Therefore, above equation yields
∂ 2 vi
= 0,
∂xk ∂xj
and it implies ∇x v is a constant skewsymmetric tensor. Thus, the motion is rigid.

Representation of rigid motion using angular velocity vector:


Let w be the axial vector of skewsymmetric tensor Ŵ . Then Eq. (10) can be written as

v(x1 , t) = v(x2 , t) + w(t) × (x1 − x2 ),

which is the classical formula for velocity field in the rigid body mechanics, where w is
known as angular velocity.
Rate of strain tensor:
Usually, in the literature of fluid mechanics, the rate-of-deformation D is referred to as
rate-of-strain. We may infer that D is rate of some strain tensor but it is not true (see
Jog (2007)). Furthermore, if displacements and displacement gradients are very small
then D is approximately equal to the rate of small strain tensor. However, the rate of
deformation measure is usually applied in the cases where the displacements are not small
(for example fluid flow, metal forming, etc.)2 . Hence, the rate of deformation does not
coincide with the rate of strain in most of the applications. It is important to note that
there is no assumption made about displacements and its gradients in the definition of
rate of deformation. Hence, D is applicable for general flow fields. We now show that
rate-of-deformation is not a time derivative of any strain tensor.

Rate of Green strain tensor:


The Green strain tensor in terms of deformation gradient F can be written as
1
E = (F T F − I).
2
Taking the material derivative, we get
DF T
!
DE 1 DF 1 T T 
= F + FT = F L F + F T LF = F T DF .
Dt 2 Dt Dt 2
DE
It is clear that if displacement gradients are small then ≈ D.
Dt
Rate of small strain tensor:
The small strain tensor is given by
1
ε = (∇X u + ∇X uT ).
2
2
Malvern, L.E., ”Introduction to the mechanics of a continuous medium”, 1969, Prentice-Hall, Inc.,
Englewood Cliffs, New Jersey, USA.

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

If displacements are also small then the rate of strain approximately equal to D.
Rate of Almansi strain tensor:
The Almansi strain tensor and the Green strain tensor are related by
1
Ẽ = (I − F −T F −1 ) = F −T EF −1
2
Taking the material time derivative, we get
!
DẼ DE
= D − LT Ẽ − ẼL. Since = F T DF
Dt Dt

Clearly, the rate of Almansi strain is not equal to D.


Rate of logarithmic or Hencky strain tensors 3 :
Let U and V be right and left stretch tensors of deformation gradient F . Then

F = RU = V R,

where R is the rotation tensor.


The material logarithmic strain tensor E (0) and spatial logarithmic strain tensor Ẽ (0)
are defined by
E (0) = ln U and Ẽ (0) = ln V . (21)
If all three principal stretches of U and V are equal then the material time derivatives of
E (0) and Ẽ (0) are given by

DE (0)
= RT DR
Dt
DẼ (0)
= D.
Dt
Clearly, if all three principal stretches of stretch tensors are equal then the material time
derivative of spatial logarithmic strain is equal to rate of deformation. The equality of
all principal stretches is very strong condition and it is not applicable in general motion
of fluid. Thus, in general, the rate of logarithmic strain tensors are also not equal to the
rate of deformation.
In conclusion, none of the rate of strains are equal to the rate of deformation. Although
rate-of-deformation is referred as rate-of-strain in fluid mechanics literature, the usage is
misnomer.

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

3
Hoger, A., “The material time derivative of logarithmic strain”, International Journal of Solids and
Structures, 1986, 22(9), pp. 1019-1032.

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-3: Kinematics
Lecture-22: The Reynolds’ Transport Theorem

Here, we discuss the transformation between material and spatial integrals. We also
present the famous Reynolds’ transport theorem which plays a crucial role in the deriva-
tion of differential form of governing equation from the integral form of balance laws.
Transformation of integrals from spatial to material and vice versa is equivalent to
the change of variable in the case of one-variable calculus. Therefore, we first present an
example to understand the change of variable in the integration.
Example 1. Let {e1 , e2 } be orthonormal basis to the two dimensional vector space and
x = (x1 , x2 ) be a position vector. Let g(x) = (−x1 , x2 ) be a given vector field. Let C be a
quarter circle with unit radius in the first quadrant of plane as shown in Fig. (1). Then
show that the line integral Z
g(x) · dx = −1,
C
on given curve of quarter circular arch using direct method and also change of variables.
Consider the starting and ending points of the curve as (0, 1) and (1, 0) (see Fig. (1)).

x2
(0,1)

x1
(0,0) (1,0)

Figure 1: A quarter circle in 2-D space

Solution: We first present the direct solution. Since the given curve x21 + x22 = 1, taking
the differentiation, we get
x2 dx2 = −x1 dx1 .
Given vector valued function g(x) = (−x1 , x2 ) with (0, 1) and (1, 0) as starting and ending
points, respectively. Therefore, the integration along the curve
Z Z
g(x) · dx = (−x1 dx1 + x2 dx2 )
C C
Z 1
= −2x1 dx1 (Since x2 dx2 = −x1 dx1 )
0
h i1
= −x21
0
= −1

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

We now use change of variables to evaluate the integration. Let x1 = cos ξ and x2 = sin ξ.
Then x1 and x2 satisfy the curve x21 + x22 = 1 automatically. The parameter ξ = π/2 and
ξ = 0 represent the starting and ending points of the curve as (x1 , x2 ) = (0, 1) at ξ = π/2
and (x1 , x2 ) = (1, 0) at ξ = 0. Since both x1 and x2 are functions of ξ, we can have
dx1 = − sin ξ and dx2 = cos ξ. Using change of variables, we can evaluate integral along
the curve
Z 0
Z
dx
g(x) · dx = g(x(ξ)) ·

C π
2

Z 0
dx1 dx2
= π (−x1 (ξ) + x2 (ξ) ) dξ
2
dξ dξ
Z 0
= π
((cos ξ)(sin ξ) + (sin ξ)(cos ξ)) dξ
2
Z 0
= π
sin(2ξ) dξ
2
" #0
cos(2ξ)
= −
2 π
2
= −1

Clearly, the evaluation of integration by direct method and change of variables are
equivalent. Hence, we observe the following points from this simple example:

• The integrand should be transformed using changed variables.

• The differential element in the integration should be transformed.

• The integral limits or domain of integration should be transformed.

These three points need to be accounted in order to transform integration from the refer-
ence domain to spatial domain or vice versa.
In the transformation of the integrals, let Ω0 be a reference configuration and Ω be
a deformed configuration at an instant of time t. Let us also consider x = χ(X, t)
be a transformation from Ω0 to Ω and F (X, t) be the deformation gradient. Then the
transformation of line, surface and volume integrals are stated below.
Transformation of line integrals:
The line integrals are always evaluated along a curve for a given vector field. Let g(x, t)
be a vector field defined over Ω and Ct be a curve in the deformed configuration. Then
the line integral is given by Z
IL = g(x, t) · dx.
Ct

In order to transform the spatial line integral IL , we need to transform spatial vector field
g(x, t), the differential line element dx and the spatial curve Ct into material description.
Let g(x, t) = g̃(X, t) for every x = χ(X, t), i.e.,

g(x, t) = g(χ(X, t), t) = g̃(X, t).

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Since F (X, t) is the deformation gradient, the transformation of differential line element
can be written as
dx = F (X, t) dX.
Let material curve C in Ω0 be a transformation of spatial curve Ct . Then the transfor-
mation of line integral can be written as
Z Z
g(x, t) · dx = g̃(X, t) · F dX
Ct
ZC
= F T g̃(X, t) · dX. (1)
C

Transformation of surface integrals:


Let Γ0 be a boundary of reference domain Ω0 and Γ be a boundary of deformed domain
Ω. Let g(x, t) be a vector field defined over Γ. Then the surface integral is defined by
Z
IS = g(x, t) · n ∂Γ, (2)
Γ

where n is outward unit normal to the surface Γ. Similar to previous case, we can
transform g(x, t) = g̃(X, t) for x = χ(X, t). From our earlier discussion (see Problem-3
in Lecture-19), the transformation to area element is given by

n ∂Γ = (cof F )n0 ∂Γ0 = JF −T n0 ∂Γ0 ,

where n0 is normal to the surface Γ0 and J is the determinant of F .


Substituting the expression for g and n ∂Γ in Eq. (2), we can get the following trans-
formation for surface integrals of vector field.
Z Z Z
g(x, t) · n ∂Γ = g̃(X, t) · (cof F )n0 ∂Γ0 = g̃(X, t) · JF −T n0 ∂Γ0 . (3)
Γ Γ0 Γ0

If G(x, t) is a tensor field defined over Ω and G(x, t) = G̃(X, t) for x = χ(X, t) then we
can have the following relation for surface integral of tensor field.
Z Z
G(x, t)n ∂Γ = G̃(X, t)(cof F )n0 ∂Γ0 . (4)
Γ Γ0

Transformation of volume integrals:


Let φ(x, t), g(x, t) and G(x, t) be scalar, vector and tensor fields. Let J be the de-
terminant of deformation gradient F . Let φ(x, t) = φ̃(X, t), g(x, t) = g̃(X, t) and
G(x, t) = G̃(X, t) for x = χ(X, t). Then we have
Z Z
φ(x, t) ∂Ω = φ̃(X, t)J ∂Ω0 , (5)
Ω Ω0
Z Z
g(x, t) ∂Ω = g̃(X, t)J ∂Ω0 , (6)
Ω Ω0
Z Z
G(x, t) ∂Ω = G̃(X, t)J ∂Ω0 . (7)
Ω Ω0

The transformation of volume element ∂Ω = J∂Ω0 can refer to a section on “volume


element” in Lecture-19.

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

We now want to prove an important result called Reynolds’ transport theorem which
plays an important role in balance laws. However, before going to show the general
result, we want to discuss the Leibniz rule of integration which is one dimensional version
of Reynolds’ transport theorem.
Leibniz rule of integration:
Let f (ξ, t) be a function. Let α(t) and β(t) be direct function of time t. Then we have
the following result known as Leibniz rule of integration.

d Z β(t) Z β(t)
∂f dβ dα
f (ξ, t) dξ = dξ + f (β(t), t) − f (α(t), t) (8)
dt α(t) α(t) ∂t dt dt

Proof. Using the definition of differentiation, we can write


!
d Z β(t) 1 Z β(t+) Z β(t)
f (ξ, t) dξ = lim f (ξ, t + ) dξ − f (ξ, t) dξ .
dt α(t) →0  α(t+) α(t)

Substituting Taylor’s expansion of α(t + ) and β(t + ), we get


dβ !
d Z β(t) 1 Z β(t)+ dt +o() Z β(t)
f (ξ, t) dξ = lim f (ξ, t + ) dξ − f (ξ, t) dξ
dt α(t) →0  α(t)+ dα
dt
+o() α(t)

1 Z α(t) Z β(t)
= lim f (ξ, t + ) dξ + f (ξ, t + ) dξ
→0  α(t)+ dα +o() α(t)
dt
Z β(t)+ dβ +o() Z β(t) !
dt
+ f (ξ, t + ) dξ − f (ξ, t) dξ
β(t) α(t)

1 Z α(t)+ +o()
dt
= lim − f (ξ, t + ) dξ
→0  α(t)
Z β(t)+ dβ +o() Z β(t) !
dt
+ f (ξ, t + ) dξ + f (ξ, t + ) − f (ξ, t) dξ
β(t) α(t)
dβ Z α(t)+ dα +o() !
1 Z β(t)+ dt +o() dt
= lim f (ξ, t + ) dξ − f (ξ, t + ) dξ
→0  β(t) α(t)
!
Z β(t)
f (ξ, t + ) − f (ξ, t)
+ lim dξ
α(t) →0 
! !!
1 dβ dα
= lim f (β(t), t + )  + o() − f (α(t), t + )  + o()
→0  dt dt
Z β(t)
∂f
+ dξ
α(t) ∂t
Z β(t)
∂f dβ dα
= dξ + f (β(t), t) − f (α(t), t)
α(t) ∂t dt dt
This completes the proof.

We now present an example to understand the theorem.

Example 2. Show that


d Z cos t 1
(1 + tξ)dξ = cos(2t) − t sin(2t) − sin t − cos t.
dt sin t 2

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Solution: Let us evaluate integral using direct integration.


" #cos t
d Z cos t d ξ2
(1 + tξ)dξ = ξ+t
dt sin t dt 2 sin t
d t
 
2 2
= (cos t − sin t) + (cos t − sin t)
dt 2
1 t
= (− sin t − cos t) + (cos2 t − sin2 t) + (−2 cos t sin t − 2 sin t cos t)
2 2
1
= cos(2t) − t sin(2t) − sin t − cos t.
2
We now evaluate integral using Leibniz rule and show that it is indeed same as that of
direct evaluation. Let α(t) = sin t, β(t) = cos t and f (ξ, t) = 1 + tξ. Then

∂f dα dβ
= ξ, = cos t, and = − sin t.
∂t dt dt
Substituting in Leibniz rule, we get
d Z cos t Z cos t
(1 + tξ)dξ = ξ dξ + (1 + t cos t)(− sin t) − (1 + t sin t)(cos t)
dt sin t sin t
" #cos t
ξ2
= − sin t − t cos t sin t − cos t − t sin t cos t
2 sin t
1
= (cos2 t − sin2 t) − 2t sin t cos t − − sin t − cos t
2
1
= cos(2t) − t sin(2t) − sin t − cos t.
2
dα dβ
The solution obtained in both ways is same. In case of Leibniz rule and represent
dt dt
velocity of boundaries. Similarly, we now see that the effect of boundary velocity appears
in the following Reynolds’ transport theorem.
Reynolds’ transport theorem:
Let Ω0 be a reference configuration with Γ0 as its boundary and Ω be a deformed con-
figuration with Γ as its boundary. Let φ(x, t) be a scalar field defined over Ω. Then the
R
material time derivative of integral Ω φ(x, t) ∂Ω is given by

d Z Z
∂φ Z
φ ∂Ω = ∂Ω + φ v · n ∂Γ, (9)
dt Ω Ω ∂t Γ

where n is normal to Γ.

Proof. We note that the deformed configuration Ω changes its shape with time. This is
similar to the earlier integral (i.e., the integral on the left hand side of Eq. (8)) where
the integration limits are function of time. Analogous to earlier integration, the time
derivative cannot be taken inside the spatial integration as Ω changes with time. On
the other hand, the undeformed domain that is independent of time. Therefore, we
follow three steps to evaluate integral: (i) transform the spatial integration to material
integral, (ii) take the total time derivative inside the integration as reference domain (i.e.,
undeformed domain) is independent of time (iii) transform the material integration again
back to spatial integration.

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Let x = χ(X, t) be a map from Ω0 to Ω. Let J be a determinant of deformation


gradient F . Let φ̃(X, t) = φ(x, t) for x = χ(X, t), i.e., φ(x, t) = φ(χ(X, t), t) = φ̃(X, t).
Then
d Z d Z
φ(x, t) ∂Ω = φ̃(X, t)J ∂Ω0
dt Ω dt Ω0
Z
∂(φ̃J)
= ∂Ω0
Ω0 ∂t
!
Z
∂ φ̃ ∂J
= J + φ̃ ∂Ω0
Ω0 ∂t ∂t
!
Z
∂ φ̃ φ̃ ∂J
= + J ∂Ω0
Ω0 ∂t J ∂t
!
Z
Dφ φ DJ
= + ∂Ω
Ω Dt J Dt
!
Z
∂φ
= + (∇x φ) · v + φ(∇x · v) ∂Ω
Ω ∂t
Z
∂φ Z
= ∂Ω + ∇x · (φv) ∂Ω
Ω ∂t Ω
Z
∂φ Z
= ∂Ω + φ v · n ∂Γ.
Ω ∂t Γ

Clearly, the second term on the right hand side of equation represents contribution from
velocity of the boundary.

Let w(x, t) be a vector field. Then, following similar steps as that of previous result,
we can prove the following Reynolds’ transport theorem for the vector field.
d Z Z
∂w Z
w ∂Ω = ∂Ω + w(v · n) ∂Γ. (10)
dt Ω Ω ∂t Γ

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-3: Kinematics
Lecture-23: The Vorticity and Circulation

We now introduce vorticity vector and circulation that are important in the study of
fluid mechanics. In addition, we discuss the Cauchy’s vorticity formula and the Kelvin’s
circulation theorem.
Vorticity vector:
Let v be a spatial velocity field defined over the deformed body Ω at an instant of time
t. Then the vorticity vector is defined by

w = ∇x × v. (1)

Since the spin tensor W is a skew-symmetric part of the velocity gradient L, we can write
1
W = (∇x v − ∇x v T ).
2
It is clear from the definition that w/2 is an axial vector of the spin tensor W . Therefore,
the vorticity vector at a point gives twice the rate of rotation or angular velocity. We now
define circulation and also the relation between circulation and vorticity.
Circulation:
Let Ω0 be a reference configuration and Ω be a deformed configuration at an instant of
time t. Let x = χ(X, t) be a mapping from Ω0 to Ω. Let C be a closed curve in Ω0 and
Ct be a corresponding closed curve in Ω. Then the circulation { is defined by
I
{= v · dx. (2)
Ct

Let Γ be a surface that passes through the closed curve Ct . Then, using Stokes’ theorem
(see Lecture-16), we get I Z
v · dx = (∇x × v) · n ∂Γ, (3)
Ct Γ

where n is the outward unit normal to the surface Γ. Thus, the circulation at a point is
directly related to the vorticity vector at that point.
Transport of circulation:
Let C be a closed curve in the reference configuration Ω0 and Ct be a corresponding closed
curve in deformed configuration Ω. Then the material time derivative of circulation is
given by
d I I
Dv
v · dx = · dx. (4)
dt Ct Ct Dt

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Proof. Let x = χ(X, t) be a map from Ω0 to Ω. Let F (X, t) be the deformation gradient.
Let ṽ(X, t) = v(x, t) for x = χ(X, t), i.e., v(x, t) = v(χ(X, t), t) = ṽ(X, t). Then

d I d I
v · dx = ṽ · F dX
dt Ct dt C
I
∂ ṽ ∂F
= · F dX + ṽ · dX.
C ∂t ∂t
Let L be a velocity gradient. Then
d I I
∂ ṽ
v · dx = · F dX + ṽ · LF dX
dt Ct C ∂t
I
Dv I
= · dx + v · L dx
Ct Dt Ct
I
Dv I
= · dx + v · dv
Ct Dt Ct
I
Dv I
1
= · dx + d(v.v)
Ct Dt Ct 2
Dv
I  I 
= · dx since d(v.v) = 0
Ct Dt Ct

This completes the proof.

Kelvin’s circulation theorem:


The motion preserves circulation if there exist a scalar function φ such that the accelera-
Dv
tion = ∇x φ.
Dt

Proof. Let Ω0 be a reference configuration and C be a closed curve in Ω0 . Let Ω be a


deformed configuration at an instant of time t and Ct be a closed curve defined in Ω
corresponding to C. Then the circulation { is defined by
I
{= v · dx.
Ct

The material time derivative of circulation of the transport of circulation (see Eq. (4)),

d{ I Dv
= · dx.
dt Ct Dt

Since the acceleration is gradient of potential function φ, we can write


Dv
= ∇x φ.
Dt
Combining above two equations, we get

d{ I I
= ∇x φ · dx = dφ = 0.
dt Ct Ct

Thus, the circulation { is constant in Ct throughout the motion.

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

The Kelvin’s circulation theorem shows that the motion is irrotational at t = 0, i.e.,
∇x × v(x, 0) = 0, then the motion is irrotational forever, i.e., ∇x × v(x, t) = 0, ∀ t ∈
[0, ∞).

Cauchy’s vorticity formula:


Let Ω0 be a reference configuration and Ω be a deformed configuration at an instant of
time t. Let x = χ(X, t) be a mapping from Ω0 to Ω. Let F be a deformation gradient
and L be a velocity gradient. Let v be a velocity field and a be an acceleration field. Let
ω0 be a vorticity vector at X ∈ Ω0 and ω be a vorticity vector at x ∈ Ω. Then we show
that the vorticity fields are related by

ω = J −1 F ω0 , (5)

where J is determinant of F . This relationship is known as the Cauchy’s vorticity formula.

Proof. The vorticity vector ω(x, t) in Ω is defined by

ω = ∇x × v.

Let ω̃(X, t) be Lagrangian description of vorticity vector, i.e., ω(x, t) = ω(χ(X, t), t) =
ω̃(X, t). Since ω0 is vorticity vector field at time t = 0, we have ω0 = ω̃(X, 0).
Let Ct be a arbitrary closed curve in Ω and C be a corresponding curve in Ω0 . Let Γ
be surface in Ω bounded by Ct and Γ0 be a surface corresponding to Γ in Ω0 . Let n be
a unit normal field to Γ and n0 be a unit normal field to Γ0 . Then consider total time
derivative of circulation,
d I d Z
v · dx = (∇x × v) · n ∂Γ
dt Ct dt Γ
d Z
= ω · n ∂Γ
dt Γ
d Z
= ω̃ · JF −T n0 ∂Γ0
dt Γ0
Z
∂(ω̃ · JF −T n0 )
= ∂Γ0
Γ0 ∂t
Z
∂ ω̃ ∂J −T 0 ∂F −T 0
= · JF −T n0 + ω̃ · F n + ω̃ · J n ∂Γ0
Γ0 ∂t ∂t ∂t
Z
∂ ω̃
= · JF −T n0 + ω̃ · (tr L)JF −T n0 + ω̃ · J(−LT F −T )n0 ∂Γ0
Γ0 ∂t
!
Z
∂ ω̃
= + ω̃(tr L) − Lω̃ · J(−LT F −T )n0 ∂Γ0
Γ0 ∂t
Z 


= + ω(tr L) − Lω · n ∂Γ.
Γ Dt
Applying Stokes’ theorem to Eq. (4), we get
d I I Z
v · dx = a · dx = (∇x × a) · n ∂Γ.
dt Ct Ct Γ

Comparing above two equations, we get


Z 

Z 
(∇x × a) · n ∂Γ = + ω(tr L) − Lω · n ∂Γ. (6)
Γ Γ Dt

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

The arbitrariness of curve C and surface Γ implies



∇x × a = + ω(tr L) − Lω.
Dt
Since the motion is circulation preserving, we have ∇x × a = 0. Therefore, Eq. (6) can
be written as

+ ω(tr L) − Lω = 0.
Dt
We can express the above equation using total derivative

D(JF −1 ω)
= 0.
Dt
Since Ω0 is the configuration at time t = 0, we have F = I, J = 1 and ω = ω0 . We
get following relation by integrating above differential equation and substituting initial
conditions F = I and ω = ω0 .
ω0 = JF −1 ω.
This relation is known as Cauchy’s vorticity formula.

Similar to previous result of Kelvin’s circulation theorem, we can conclude from the
Cauchy’s vorticity formula that the motion is irrotational once then it is irrotational
forever.

References

1. C. S. Jog, Foundations and Applications of Mechanics, Volume-I (Continuum Me-


chanics) & Volume-II (Fluid Mechanics), 2007, Narosa Publishing House Pvt. Ltd.,
New Delhi.

2. P. Chadwick, Continuum Mechanics: Concise Theory and Problems, 1999, Dover


Publications, Inc., New York.

3. M. E. Gurtin, An Introduction to Continuum Mechanics, 1981, Academic Press,


New York.

4. M. E. Gurtin, E. Fried and L. Anand, The Mechanics and Thermodynamics of


Continua, 2010, Cambridge University Press, New York.

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-4: Balance Laws


Lecture-24: Law of Conservation of Mass

Classical continuum mechanics assumes the following laws of mechanics and thermody-
namics as axioms.

• Conservation of mass

• Conservation of linear momentum

• Conservation of angular momentum

• Conservation of energy or first law of thermodynamics

• Second law of thermodynamics

Since the classical continuum mechanics assumes Newtonian space-time framework (i.e.,
non-relativistic setup), the conservation of mass is an axiom.

Except second law of thermodynamics all others are balance laws. We will see in the
forthcoming discussion that the second law of thermodynamics is an inequality which im-
poses restrictions on constitutive relations. These balance laws are applied to continuum
in order to obtain governing equations in integral and differential forms. We now define
three types of regions of continuum body called reference volume, material volume and
control volume to apply balance laws.
Let B0 be the reference configuration and B be the deformed configuration at an
instant of time t. Let x = χ(X, t) be a map from B0 to B that represent the motion.
Let Ω0 be a part of B0 and its associated part in B be Ω, i.e., for every X ∈ Ω0 there is
a unique x ∈ Ω.

• Reference volume Ω0 which is part of reference configuration B0 .

• Material volume Ω convects with body. In this case the material does not flow into
and out of Ω during motion, i.e., for every x ∈ Ω there is a X ∈ Ω0 and vise versa.

• Control volume Ωc is usually a fixed region1 in space that lie in the deformed body
B for some non-zero time interval. In general material flows into and/or out of a
control volume.
1
In general the control volume Ωc need not be fixed in space and it can have the motion and change
its shape with time. The control volume Ωc allows the flow of material into and out of control surface
whereas the material volume Ω does not allow, i.e., Ω consists of fixed set of particles with time.

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

The material volume Ω0 , the spatial volume Ω and the control volume Ωc are depicted in
Fig. (1).

Figure 1: Material volume (Ω0 ), spatial volume (Ω) and control volume (Ωc )

We recall the Reynolds’ transport theorem from Lecture-22.


The Reynolds’ transport theorem (the first transport theorem):
Recall the Reynolds’ transport theorem (RRT) from Lecture-22.

d Z Z
∂φ Z
φ ∂Ω = ∂Ω + φ(v · n) ∂Γ,
dt Ω Ω ∂t Γ

where Γ is the bounding surface of Ω and n is the outward normal of Γ. Similarly, RTT
for vector field w can be written as
d Z Z
∂w Z
w ∂Ω = ∂Ω + w(v · n) ∂Γ,
dt Ω Ω ∂t Γ

We refer the Reynolds’ transport theorem as first transport theorem, since another
convenient form of transport theorem is stated in fourth coming lecture in order to apply
directly to the conservation of momenta and energy.

Mass and density:


In continuum mechanics, we assume that every body posses a mass and it is a positive
quantity. Furthermore, the mass is distributed continuously over domain of the body
whatever may be the motion or deformation. The distribution of the mass is called
density. Of course the integration of the density over the part of the domain gives the
mass of that particular part. Since the mass is positive for arbitrary part, the density is
defined as positive measure over the domain.
Conservation of mass:
In Newtonian framework (non-relativistic setup) the mass of the body cannot be altered
by the deformation, i.e., the mass of a body before and after deformation is constant. We
now consider the mass balance is an axiom and derive the mathematical formulation of
the axiom. As we see in the following discussion that the conservation of mass can be
expressed both in Lagrangian and Eulerian framework.

Lagrangian framework:
Let ρ(x, t) be a density in the deformed configuration B. Let ρ̃(X, t) be a Lagrangian
description of ρ(x, t), i.e., ρ(x, t) = ρ(χ(X, t), t) = ρ̃(X, t). We denote the density field
of reference configuration by ρ0 (X), i.e., ρ0 (X) = ρ̃(X, 0).

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

The axiom of mass conservation states that the material volume Ω0 evolves to Ω with
time t but the mass remains constant (assuming that there are no mass sources within
Ω), i.e., Z Z
ρ0 (X) ∂Ω0 = ρ(x, t) ∂Ω. (1)
Ω0 Ω

Transforming the right hand side of volume integral, we get


Z Z
ρ0 (X) ∂Ω0 = ρ̃(X, t)J ∂Ω0
Ω Ω0
Z0
=⇒ (ρ0 (X) − ρ̃(X, t)J) ∂Ω0 = 0.
Ω0

Since Ω0 is arbitrary, using localization theorem (see Lecture-16), we get

ρ0 (X) = ρ̃(X, t)J. (2)

Equation (2) represents the conservation of mass in Lagrangian description as equation


is expressed using material coordinate X.
Eulerian framework
We now derive the mass conservation in Eulerian (spatial) description. Since the left hand
side of Eq. (1) is independent of time, we can write

d Z
ρ ∂Ω = 0. (3)
dt Ω
Let Γ be the boundary surface of Ω. Then, using Reynolds’ transport theorem, we get
Z
∂ρ Z
∂Ω + ρ(v · n) ∂Γ = 0, (4)
Ω ∂t Γ

where n is normal to Γ and v is velocity of boundary Γ. Applying the divergence theorem


to the second term, we can write
!
Z
∂ρ
+ ∇x · (ρv) ∂Ω = 0. (5)
Ω ∂t

Since Ω is arbitrary, using the localization theorem, we can have


∂ρ
+ ∇x · (ρv) = 0. (6)
∂t
Substituting ∇x · (ρv) = v · (∇x ρ) + ρ(∇x · v) in above equation, we get the following final
form of mass conservation in Eulerian framework.

+ ρ∇x · v = 0. (7)
Dt

Balance of mass for incompressible medium:


In practical problems, we assume some materials as incompressible if negligible compres-
sion is observed experimentally under different loading conditions. For example, rubber
is usually assumed to be incompressible solid material and water is assumed to be incom-
pressible fluid medium. Therefore, we now discuss the mass balance of incompressible
continuum bodies.

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

The medium is said to be incompressible if all possible deformations are isochoric. The
isochoric constraint implies J = 1 (i.e., det(F ) = 1). Taking material time derivative of
J (see Eq. (20) in Lecture-20), we get

∇x · v = 0.

From Eq.(7), the mass conservation in Eulerian framework for incompressible medium

is given by = 0, while in the Lagrangian framework (see Eq. (2)), we get the
Dt
condition ρ(x, t) = ρ0 (X). Thus, the density at any material point remains the same in
an incompressible material. The density of deformed configuration is same at every point
if the density of the reference configuration is same at every point.

Balance of mass in control volume:


Analyzing dynamics of continuous media in control volume is useful for the study of fluid
mechanics. Let the control volume Ωc be a fixed region and Γc be its boundary. Using
mass conservation in Eulerian framework (see Eq. (5)), we can have the following integral
equation, !
Z
∂ρ
+ ∇x · (ρv) ∂Ω = 0.
Ωc ∂t
Applying divergence theorem to the second term, we get
Z
∂ρ Z
∂Ω + ρ(v · n) ∂Γ = 0. (8)
Ωc ∂t Γc

It is clear that the rate of change of mass in control volume Ωc is balanced by the mass
flux into and out of the surface Γc .
Since Ωc is fixed in space, Eq. (8) can also be written as

d Z Z
ρ ∂Ω + ρ(v · n) ∂Γ = 0. (9)
dt Ωc Γc

This integral form of mass conservation in Eulerian framework is useful in implementation


of numerical methods such as finite volume method.

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-4: Balance Laws


Lecture-25: The Laws of Dynamics and
the Cauchy’s Hypothesis

The linear and angular momenta balance are known as laws of dynamics. These laws are
postulated to the continuous medium by Euler (C.E. 1707 - 1783). Therefore, these laws
of dynamics for massive (continuum) bodies are also known as Euler laws of motion. We
require the Cauchy’s hypothesis for the surface forces and second transport theorem in
order to obtain the differential form of equation of momenta balance. Therefore, in this
discussion, we introduce second transport theorem and the Cauchy’s hypothesis.
The second transport theorem:
Let B0 be the reference configuration and B be the deformed configuration at an instant
of time t. Let x = χ(X, t) be a map from B0 to B that represent the motion. Let Ω0
be a part of B0 and Ω be a convection1 of Ω0 with B, i.e., Ω is part of B corresponding
to Ω0 .
Let φ(x, t) be a scalar field defined over B and φ̃(X, t) = φ(x, t) for every x ∈ B.
Let ρ(x, t) be a density defined over B and ρ̃(X, t) be a material description of density,
i.e., ρ(x, t) = ρ(χ(X, t), t) = ρ̃(X, t). Let ρ0 (X) = ρ̃(X, 0) be a density of reference
configuration B0 . Then the time derivative of integral Ω ρφ ∂Ω is given by
R

!
d Z Z
Dφ Z
∂φ
ρφ ∂Ω = ρ ∂Ω = ρ + v · (∇x φ) ∂Ω. (1)
dt Ω Ω Dt Ω ∂t

Proof. Let J be the determinant of deformation gradient F . Then transformation of


integral from spatial to reference (see Lecture-22) can be written as
d Z d Z
ρφ ∂Ω = ρ̃φ̃J ∂Ω0 .
dt Ω dt Ω0
We have the relation ρ0 (X) = J ρ̃(X, t) from the conservation of mass in Lagrangian
framework (see Eq. (2) in Lecture-24). Substituting the mass balance in above relation,
give rise to
d Z d Z
ρφ ∂Ω = ρ0 φ̃ ∂Ω0 .
dt Ω dt Ω0
Clearly, the integral on the right hand side of equation is a function of time t and the
material variable X. Since the material variable is independent of time, we can write
d Z Z
∂(ρ0 φ̃)
ρφ ∂Ω = ∂Ω0 .
dt Ω Ω0 ∂t
1
Convection is referred as part of reference domain Ω0 moving with deformed configuration changes
its shape to Ω

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

We know that the density of reference configuration is not a function of time. Hence, we
have
d Z Z
∂ φ̃
ρφ ∂Ω = ρ0 ∂Ω0
dt Ω Ω0 ∂t
Z
∂ φ̃
= ρ̃ J ∂Ω0
Ω0 ∂t
Z

= ρ ∂Ω
Ω Dt

The material time derivative of scalar field can be expressed as (see Lecture-20)
Dφ ∂φ
= + v · ∇x φ,
Dt ∂t
Therefore, we get
!
d Z Z
Dφ Z
∂φ
ρφ ∂Ω = ρ ∂Ω = ρ + v · (∇x φ) ∂Ω.
dt Ω Ω Dt Ω ∂t

Let w(x, t) be a vector field defined over B. Then the following second transport
theorem for vector field can be proved using the similar steps as that of previous result.
!
d Z Z
Dw Z
∂w
ρw ∂Ω = ρ ∂Ω = ρ + (∇x w)v ∂Ω. (2)
dt Ω Ω Dt Ω ∂t

Laws of dynamics: linear and angular momenta balance:


The axioms of linear and angular momenta balance for continuum body in the present
form, i.e., the integral forms of laws of dynamics, were postulated by Euler.
The balance of linear momentum states that the rate of change of linear momentum
of the material volume is equal to the net external forces acting on it. Let B0 be a
reference configuration and B be a deformed configuration at an instant of time t. Let Ω
be a material volume in B corresponding to a reference volume Ω0 in B0 (see previous
lecture for the definition of Ω and Ω0 ). Let f e be net forces acting on Ω. Then the linear
momentum balance can be written as
d Z
ρv ∂Ω = f e (3)
dt Ω
where v is velocity field defined over B.
Similarly, the balance of angular momentum states that the rate of change of angular
momentum of the material volume is equal to the net moment (couple) acting on it. Let
ce be a net moment acting on Ω. Then the balance of angular momentum can be written
as
d Z
ρ(x − x0 ) × v ∂Ω = ce (4)
dt Ω
where x0 is a fixed point in the space and x is a spatial position in Ω.
If the body is isolated from the external forces, i.e., net force and moment acting on
body are zero (f e = 0 and ce = 0), then the linear and angular momenta are constant.

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

In classical continuum mechanics, we assume that the moment is defined as the moment
of a force. Therefore, we require the definition of force for further study. The detailed
axiomatic set up for forces can be found in a book by Truesdell (1977)2 . Here, we present
a brief introduction to forces.
The forces on continuum bodies:
In classical continuum mechanics, the forces can be classified into two types: (i) body
forces (ii) surface forces.

1. Body forces
The body force acts on every particle. The source of these body forces usually lies
outside the body. The examples for these forces are the gravitational forces and
electromagnetic forces. Let ρ(x, t) be a density of the medium and b(x, t) be a
body force field per unit mass then the body force over the domain Ω is given by
R
Ω ρb ∂Ω.

2. Surface forces
The forces that are acting on the boundary of region are known as surfaces forces.
These forces are usually exerted by one body onto another through the contact. We
assume that these forces are related to the contact surface. Unlike body forces, the
surface forces are independent of mass of the body.

We can again divide the surface forces into two categories:

• Contact force exerted on the boundary of body by its environment.


For example car crashes with wall then the wall exert the force on car.

• Contact force between separate parts of same body.


These contact forces are also known as internal surface forces as they occur within
the body. For example, elastic restoring forces in solids and viscous forces in fluids
are internal surface forces.

In any case the surface force are dependent on the surface that is in contact. We make
the following assumption on surface forces known as Cauchy’s hypothesis.
The Cauchy’s hypothesis:
The distribution of the force over the surface area is known as traction field t. In other
words, the force per unit area on a given surface is known as traction field. The Cauchy’s
hypothesis states that the traction vector field t on the surface Γ in B depends only on
the unit normal to the surface at a point. In other words, the traction field is a function
of position, time and outward normal to the surface, i.e., t(x, t, n) where n is outward
unit normal.

Let ψ(x) be a surface Γ at an instant of time t in the deformed configuration B. Then


the surface is characterized by all gradients (gradient and its higher order gradients),
i.e., ∇x ψ, ∇x (∇x ψ), ∇x (∇x (∇x ψ)), etc. In other words, if we know the information of
2
Truesdell, C.A., “A first course in rational continuum mechanics”, 1977, Academic press, Inc., London,
U.K.

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

all gradients of surface then it is equivalent to knowing the full surface. Among all the
gradients, the first gradient ∇x ψ is known as normal and the second gradient ∇x (∇x ψ) is
known as curvature. Although all gradients contain the information of surface, we assume
that the traction t at a point depends on surface only through the unit normal and not
on higher gradients such as curvature. This assumption is known as Cauchy’s hypothesis
which is depicted in Fig. (1). It can be observed in Fig. (1) that there are three different
surfaces having common normal n but having different curvatures passing through a point
x have the same traction vector.

Figure 1: Depiction of Cauchy’s hypothesis: traction vector t on different surfaces having


common normal n at a point x

Cauchy’s lemma: action and reaction:


Suppose that the traction vector t(x, t, n) is continuous function of x at any time t, and
density field ρ(x, t), acceleration a(x, t) and body force field per unit mass b(x, t) are
bounded at any time t and any position ∀ x ∈ B. Then the linear momentum balance
implies
t(x, t, −n) = −t(x, t, n), (5)
for every x ∈ B and for every unit vector n.

Proof. Let x be an arbitrary point in B and let n be an arbitrary unit vector. We


consider a cylindrical element C of thickness  in B such that n is normal to the flat
surface and x is centroid of cylinder (see Fig. (2)).

Figure 2: A small cylinder in deformed configuration B with centroid at x and a unit


vector n as normal to flat surface

Let y be a position vector in the cylinder C . Then using linear momentum balance to

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

the cylindrical element, we get


d Z Z Z
ρv ∂Ω = ρb ∂Ω + t(y, t, n) ∂Γ
dt C C ΓC
Z Z Z
=⇒ ρa ∂Ω = ρb ∂Ω + t(y, t, n) ∂Γ (Using transport theorem-II)
C C ΓC
Z Z
=⇒ ρ(a − b) ∂Ω = t(y, t, n) ∂Γ.
C ΓC

If cylinder approaches the disc, i.e.,  tends to zero, then the volume integral goes to zero
as ρa and ρb are bounded. Furthermore, the surface integral remains finite. Therefore,
we can write Z
lim t(y, t, n) ∂Γ = 0.
→0 ΓC


Let Γ is curved surface, Γ+ is flat surface with n as outward normal and Γ− is flat surface
with −n as outward normal to the cylinder. Then ΓC = Γ ∪ Γ+ ∪ Γ− . Let D be a disc
passing through x in the cylinder with same radius as that of cylinder (see Fig. (2)).
Then Γ vanishes and Γ+ and Γ− approaches D as  → 0. Therefore, above integral can
be written as Z
t(y, t, n) + t(y, t, −n) ∂Γ = 0.
D
Since the radius of cylinder is arbitrary, the continuity of integrand implies that the
integrand must vanish at x, i.e.,
t(x, t, n) + t(x, t, −n) = 0,
which proves the Cauchy’s lemma.

This result is equivalent to Newton’s law of action and reaction as shown in Fig. (3).

Figure 3: Principle of action and reaction

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. O. Golzalez and A. M. Stuart, A First Course in Continuum Mechanics, 2008,


Cambridge University Press, Cambridge.

3. I-S. Liu, Continuum Mechanics, 2002, Springer, Berlin.

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-4: Balance Laws


Lecture-26: The Linear Momentum Balance and
the Cauchy Stress Tensor

The concept of stress is central to the continuum mechanics. In this lecture, we show the
existence of stress tensor. We also derive the differential form of linear momentum balance
which is known as Cauchy’s equation of motion-I (first Cauchy’s equation of motion).
The Cauchy’s theorem and the stress tensor:
Let the traction field t(x, t, n) be a continuous function of x at any instant of time t for a
given unit vector n. Let the density ρ(x, t), acceleration a(x, t) and body force field per
unit mass b(x, t) be bounded for every x in B at any time t. Then the linear momentum
balance implies the existence of stress tensor τ such that

t(x, t, n) = τ (x, t)n. (1)

Proof. Let {e1 , e2 , e3 } be a set of orthonormal basis to the 3-D Cartesian space. Consider
a tetrahedron shown in Fig. (1) with three faces lying on Cartesian coordinate planes pass
through O and an inclined plane pass through A, B and C. Let x be a position vector
of O. Let Γn , Γ1 , Γ2 and Γ3 be four bounding surfaces ∆ABC, ∆OCB, ∆OAC ∆OBA,
respectively. Outward normals to the surfaces ∆ABC, ∆OCB, ∆OAC and ∆OBA are
n, −e1 , −e2 and −e3 , respectively.

Figure 1: Tetrahedron used in Cauchy’s theorem

Let Ω and Γ be domain and boundary of tetrahedron. Since Γn , Γ1 , Γ2 and Γ3 are


boundary surface of tetrahedron, we have Γ = Γn ∪ Γ1 ∪ Γ2 ∪ Γ3 . Let v(y, t) be a velocity
field where y ∈ B. Then using linear momentum balance, we get
d Z Z Z
ρv ∂Ω = ρb ∂Ω + t(y, t, n) ∂Γ.
dt Ω Ω Γ

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Using the second transport theorem, we obtain


Z
Dv Z Z
ρ ∂Ω = ρb ∂Ω + t(y, t, n) ∂Γ.
Ω Dt Ω Γ

Applying the surface integral to four faces of tetrahedron (see Fig. 1), we can write
Z Z Z Z Z
ρ(a − b) ∂Ω = t(y, t, n) ∂Γ + t(y, t, n) ∂Γ + t(y, t, n) ∂Γ + t(y, t, n) ∂Γ.
Ω Γn Γ1 Γ2 Γ3

Let AS be surface area of Γn (i.e., ∆ABC) and h be the distance from O to Γn . Let n1 ,
n2 and n3 be components of normal n along e1 , e2 and e3 , respectively. Then the surface
area of Γ1 , Γ2 and Γ3 (i.e., area of ∆OCB, ∆OAC and ∆OBA) are given by n1 AS , n2 AS
and n3 AS , respectively. If h is small then, using the mean value theorem1 , we get
1
ρ(a − b)|x∗∗ As h = (t(x∗ , t, n) + t(x∗i , t, ei )ni ) As
3
where x∗∗ is centroid of tetrahedron and x∗ is centroid of Γn and x∗i is centroid of Γi .
Let h tends to zero such that the normal of Γn continues to be n. Then the continuity of
traction t and boundedness of ρ(a − b) implies

t(x, t, n) = −t(x, t, −ei )ni .

Using the Cauchy’s lemma (see Lecture-25), we get

t(x, t, n) = t(x, t, ei )ni . (2)

Following similar steps one can show Eq. (2) for the tetrahedron lying in other octants
of 3-D Cartesian space. We now define the Cauchy stress tensor τ by

τ (x, t) = t(x, t, ei ) ⊗ ei . (3)

Equations (3) and (2) imply τ (x, t)n = t(x, t, n). We note that one should use the tensor
product definition given in Eq. (16) of Lecture-4 to see the equivalence between τ (x, t)n
and t. Thus, there is a spatial tensor field known as Cauchy stress tensor.

The linear momentum balance: Cauchy’s equation of motion-I:


Let B0 be a reference configuration of body. Let B be a deformed configuration at an
instant of time t. Let Ω0 be a arbitrary volume of B0 and Ω be a material volume
corresponding to Ω0 . Let body force field per unit mass b(x, t) and acceleration field
a(x, t) be continuous at every instant of time t. Then the Cauchy’s hypothesis and linear
momentum balance imply the following equation known as first Cauchy’s equation of
motion.
Dv
ρ = ∇x · τ + ρb, (4)
Dt
where τ is the Cauchy stress tensor and v is a spatial velocity field.
1
If the domain of integral Ω is sufficiently small then volume integralR can be approximated by the
product between volume of domain and integrand value at centroid, i.e., Ω g(x)dΩ = g(x∗ )Ω , where
x∗ is centroid of Ω . Of course the integrand is assumed to be a bounded function in the domain. This
result is applicable to scalar, vector and tensor valued function

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Problem 1. Show that linear momentum balance along with Cauchy’s hypothesis imply
the following governing differential equation.
Dv
ρ = ∇x · τ + ρb,
Dt

Proof. Let Γ be a boundary of material volume Ω ∈ B and t(x, t, n) be a traction field


over Γ. Then the Cauchy’s theorem implies the following relation (see Eq. (1)).

t(x, t, n) = τ (x, t)n,

where n is normal to the boundary Γ and τ is Cauchy stress tensor. Since b(x, t) is body
force, v(x, t) is velocity, ρ(x, t) is density and t(x, t) is traction, we can write the linear
momentum balance for material volume Ω as
d Z Z Z
ρv ∂Ω = ρb ∂Ω + t ∂Γ.
dt Ω Ω Γ

Substituting stress tensor and applying second transport theorem to the above equation,
we get Z
Dv Z Z
ρ ∂Ω = ρb ∂Ω + τ n ∂Γ.
Ω Dt Ω Γ
Applying the divergence theorem to the surface integral and grouping terms, yields
Z 
Dv

ρ − ∇x · τ − ρb ∂Ω = 0.
Ω Dt
Since the material volume Ω is arbitrary, application of the localization theorem give rise
to the first Cauchy’s equation of motion shown in Eq. (4). We now show that the Cauchy
stress and differential form of equation of motion (first Cauchy’s equation of motion)
also imply the linear momentum balance. Considering the linear momentum of material
volume Ω, we obtain
d Z Z
Dv
ρv ∂Ω = ρ ∂Ω (Using transport theorem-II)
dt Ω ZΩ Dt
= (∇x · τ + ρb) ∂Ω (Using first Cauchy’s equation of motion)
ZΩ Z
= τ n ∂Γ + ρb ∂Ω (Using divergence theorem)
ZΓ Z Ω

= t ∂Γ + ρb ∂Ω (Since τ n = t).
Γ Ω

Thus, the linear momentum balance along with Cauchy’s hypothesis imply the existence
of Cauchy stress tensor. Conversely, the Cauchy stress tensor with first Cauchy’s equation
of motion imply the linear momentum balance.

Cauchy stress and differential equation of motion in indicial notation:


The index represention of Eqs. (1) and (4) can be written as

ti = τij nj ,
!
∂vi ∂vi ∂τij
ρ + vj = + ρbi .
∂t ∂xj ∂xj

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Equation (3) implies


τij = ei · τ ej = ti (x, t, ej ).
Thus, the component of stress τij represents the i’th component of the traction t (i.e.,
the component of t along ei ) on the surface whose normal is in the j’th direction (i.e.,
the surface with ej as its normal). It can be observed that τ11 , τ22 and τ33 are normal
stresses as they act normal to the surface whereas other components (τ12 , τ13 , τ21 , τ23 , τ31 ,
τ32 ) represent shear stresses as they act tangential to the surface. The stress components
on infinitesimal cube are shown in Fig. (2).

Figure 2: Cauchy stress components on infinitesimal cube

Linear momentum balance in control volume:


The control volume approach is suitable for fluid mechanics. So, we now present the
integral form of linear momentum balance in control volume. Let Ωc be a fixed control
volume, i.e., Ωc is fixed volume in space, and Γc be a boundary of Ωc . Using the first
Cauchy’s equation of motion (differential form of linear momentum balance in Eulerian
framework), we can write
Dv
Z  
ρ − ρb − ∇x · τ ∂Ω = 0.
Ωc Dt
Recall the relation between material and spatial time derivative of vector field from
Lecture-20. Upon substituting the relation between material and spatial time derivative
for velocity field yields
!
Z
∂v Z Z
ρ + (∇x v)v ∂Ω = ρb ∂Ω + ∇x · τ ∂Ω.
Ωc ∂t Ωc Ωc

Applying divergence theorem to the second term on the right hand side, we get
!
Z
∂(ρv) ∂ρ Z Z
− v + (∇x v)ρv ∂Ω = ρb ∂Ω + τ n ∂Γ,
Ωc ∂t ∂t Ωc Γc

where n is normal to the control surface Γc .


∂ρ
Using the conservation of mass in Eulerian framework, i.e., = −∇x · (ρv), we can
∂t
write
Z
∂(ρv) Z Z Z
∂Ω + (∇x · (ρv)) v + (∇x v)ρv ∂Ω = ρb ∂Ω + τ n ∂Γ
Ωc ∂t Ωc Ωc Γc

By substituting the relations ∇x · (v ⊗ (ρv)) = (∇x · (ρv)) v + (∇x v)ρv and t = τ n, we


can obtain
Z
∂(ρv) Z Z Z
∂Ω + ∇x · (v ⊗ (ρv)) ∂Ω = ρb ∂Ω + t ∂Γ.
Ωc ∂t Ωc Ωc Γc

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Applying the divergence theorem to the second term on the left hand side of equation,
we get the following linear momentum balance for the control volume.
Z
∂(ρv) Z Z Z
∂Ω + ρv(v · n) ∂Γ = ρb ∂Ω + t ∂Γ. (5)
Ωc ∂t Γc Ωc Γc

Since Ωc is fixed in space, we can also write the linear momentum balance as
d Z Z Z Z
ρv ∂Ω + ρv(v · n) ∂Γ = ρb ∂Ω + t ∂Γ. (6)
dt Ωc Γc Ωc Γc

The term Γc ρv(v · n) ∂Γ represents the rate of change of linear momentum due to the
R

flux across the control surface Γc .


Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-4: Balance Laws


Lecture-27: The Angular Momentum Balance and
the Properties of Cauchy Stress

We now show the implication of angular momentum balance on stress tensor. Let B0
be a reference configuration and B be a deformed configuration at an instant of time t.
Let Ω0 be a part of B0 and Ω be a corresponding material volume in B. Let x0 be a
fixed point in the space. Then the principle of angular momentum balance states that the
rate of change of angular momentum of material volume Ω is equal to the net moment of
forces acting on Ω with respect to x0 , i.e.,
d Z Z Z
ρ(x − x0 ) × v ∂Ω = ρ(x − x0 ) × b ∂Ω + (x − x0 ) × t ∂Γ, (1)
dt Ω Ω Γ

where Γ is boundary of Ω, ρ is density field, v is spatial velocity field, b is body force


field per unit mass, and t is traction over the surface Γ. Since the continuum is assumed
to be non-polar, i.e., there are no body and surface couples, the moment is generated by
the body forces and traction.
We now choose the fixed point x0 to be origin. Then the above equation reduces to
d Z Z Z
ρx × v ∂Ω = ρx × b ∂Ω + x × t ∂Γ. (2)
dt Ω Ω Γ

The angular momentum balance and symmetry of Cauchy stress tensor:


If classical continuum (i.e., non-polar continuum) follows the linear momentum balance,
the angular momentum balance is equivalent to symmetry of the Cauchy stress tensor,
i.e.,
τ = τT. (3)
Equation (3) is known as second Cauchy’s equation of motion.

Proof. Applying the transport theorem-II to Eq. (2), we get


Z
D(x × v) Z Z
ρ ∂Ω = ρx × b ∂Ω + x × t ∂Γ
Ω Dt Ω Γ
Dx Dv
Z   Z Z
=⇒ ρ ×v+x× ∂Ω = ρx × b ∂Ω + x × t ∂Γ
Ω Dt Dt Ω Γ
Z
Dv Z Z
Dx
=⇒ ρx × ∂Ω = ρx × b ∂Ω + x × t ∂Γ (Since v = ).
Ω Dt Ω Γ Dt
Dv
Upon substituting ρ = ∇x · τ + ρb and t = τ n, we can write
Dt
Z Z
x × (∇x · τ ) ∂Ω = x × (τ n) ∂Γ.
Ω Γ

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Let u be a constant arbitrary vector. Taking dot product on both sides of equation, we
obtain
Z  Z 
x × (∇x · τ ) ∂Ω · u = x × (τ n) ∂Γ · u

Z Γ
= (x × (τ n)) · u ∂Γ.
Γ

Using a property of scalar triple product from Lecture-3, we can have [x, τ n, u] =
[u, x, τ n] = τ T (u × x) · n. Substituting this relation in above equation, yields
Z  Z
x × (∇x · τ ) ∂Ω · u = τ T (u × x) · n ∂Γ.
Ω Γ

Applying the divergence theorem, we get


Z Z  
(x × (∇x · τ )) · u ∂Ω = ∇x · τ T (u × x) ∂Ω
Ω ZΩ
= (τ : ∇x (u × x) + (u × x) · (∇x · τ )) ∂Ω
ZΩ Z
= τ : ∇x (u × x) ∂Ω + (x × (∇x · τ )) · u ∂Ω.
Ω Ω

Cancelling the second term in right hand side, above relation reduces to
Z
τ : ∇x (u × x) ∂Ω = 0

It is easy to verify that ∇x (u × x) is a skewsymmetric tensor with u as its axis, i.e.,


(∇x (u × x))ij = −ijk uk . Arbitrariness of u implies the symmetry of stress tensor τ .
Thus, the Cauchy stress tensor is symmetric for non-polar continuous medium.
Converse can be easily showed by reversing the above steps. Thus, the angular mo-
mentum balance is equivalent to symmetry of the Cauchy stress.

We now derive the angular momentum balance for control volume.


The angular momentum balance for control volume:
Let Ωc be a fixed control volume in space and Γc be its boundary. Let n be unit outward
normal to the surface Γc . Let τ be Cauchy stress tensor field. Let x be a position in
control volume. Let u be an arbitrary constant vector. Then considering the surface
integral Γc (x × τ n) · u ∂Γ with following the relation
R

(x × τ n) · u = (τ n) · (u × x) = τ T (u × x) · n,
we can write
Z Z  
(x × τ n) · u ∂Γ = τ T (u × x) · n ∂Γ
Γc
ZΓc  
= ∇x · τ T (u × x) ∂Ω
ZΩc
= (τ : ∇x (u × x) + (u × x) · (∇x · τ )) ∂Ω
Ωc

Since Cauchy stress τ is a symmetric tensor and ∇x (u × x) is a skewsymmetric tensor,


we have τ : ∇x (u × x) = 0. Therefore, above equation reduces to
Z Z
(x × τ n) · u ∂Γ = (u × x) · (∇x · τ ) ∂Ω
Γc Ωc
Z
= (x × ∇x · τ ) · u ∂Ω.
Ωc

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Since u is an arbitrary constant vector, we have


Z Z
x × τ n ∂Γ = x × ∇x · τ ∂Ω.
Γc Ωc

Let t be a traction vector and v be a velocity field. Then the linear momentum balance
implies
Dv
Z Z  
x × t ∂Γ = x× ρ − ρb ∂Ω
Γc Ωc Dt
Z
Dv Z
= ρx × ∂Ω − ρx × b ∂Ω.
Ωc Dt Ωc

Rearranging terms, we get


Z Z Z
Dv
x × t ∂Γ + ρx × b ∂Ω = ρx × ∂Ω
Γc Ωc Ωc Dt
Z
D
= (x × v) ∂Ω
ρ
Ωc Dt
Z 
D Dρ

= (ρx × v) − (x × v) ∂Ω.
Ωc Dt Dt
Recall the relation between material and spatial time derivative from Lecture-20. The
relation between material time derivative and spatial time derivative along with mass

conservation = ρ(∇x · v), give rise to
Dt
Z Z Z

x × t ∂Γ + ρx × b ∂Ω = (ρx × v) + ∇x (ρx × v)v + ρ(∇x · v)(x × v) ∂Ω
Γc Ωc Ωc ∂t
Z
∂ Z
= (ρx × v) ∂Ω + ∇x · (ρ(x × v) ⊗ v) ∂Ω
Ωc ∂t Ωc

Using divergence theorem to the second term on right hand side of equation, we get
Z Z Z
∂ Z
x × t ∂Γ + ρx × b ∂Ω = (ρx × v) ∂Ω + (ρ(x × v) ⊗ v) n ∂Γ
Γc Ωc Ωc ∂t Γc

We obtain the following form of the conservation of angular momentum for control volume
using the definition of tensor product between two vectors.
Z
∂ Z Z Z
(ρx × v) ∂Ω + ρ(x × v)(v · n) ∂Γ = x × t ∂Γ + ρx × b ∂Ω. (4)
Ωc ∂t Γc Γc Ωc

Since Ωc is fixed region in space, the conservation of angular momentum for control volume
can also be written in the following form.
d Z Z Z Z
ρx × v ∂Ω + ρ(x × v)(v · n) ∂Γ = x × t ∂Γ + ρx × b ∂Ω. (5)
dt Ωc Γc Γc Ωc

Properties of the Cauchy stress tensor:


We now state some important properties of Cauchy stress tensor.

1. The stress tensor τ is symmetric.


Proof follows from balance of angular momentum as discussed previously.

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

2. At any point in continuum body at least a set of three orthogonal planes exist such
that the traction vector is collinear with the normal n, i.e., t = λn.
Proof follows from the symmetry of stress tensor and the relation between traction
and stress tensor, i.e., t = τ n.

3. The plane where unit normal n and traction t are collinear is called principal plane
and the direction n is called principal direction. Since t = λn, the scalar λ is called
principal stress. We note that all principal stresses are real. The principal directions
corresponding to distinct principal stresses are mutually orthogonal.
Proof follows from the symmetry of stress tensor (see Lecture-10).

4. The principal directions are the directions corresponding to extremum normal stress.

Proof. Let n be a unit vector, i.e., n · n = 1. Let τ be a stress tensor which is


symmetric. The normal stress Tn acting on plane with normal n is given by

Tn = n · τ n. (6)

Then the condition for optimality can be written as


∂(nτ · n) ∂(1 − (n · n))
+λ = 0, (7)
∂nk ∂nk
where λ is the Lagrange multiplier and nk is k’th component of n. We can get the
following relations from tensor calculus.
∂(n · τ n) ∂(ni (τ n)i ) ∂(ni τij nj ) ∂ni ∂nj
= = = τij nj + ni τij
∂nk ∂nk ∂nk ∂nk ∂nk
∂(n · n) ∂(ni ni ) ∂ni ∂ni
= = ni + ni
∂nk ∂nk ∂nk ∂nk
∂ni
= δik
∂nk
Upon substituting these three relations in above optimality condition, i.e., in Eq.
(7), we get

δik τij nj + ni τij δjk − λδik ni − λni δik = 0


=⇒ τkj nj + τik ni − 2λnk = 0
=⇒ 2τkj nj − 2λnk = 0 (Since τij = τji )
=⇒ (τ n)k = λnk
=⇒ τ n = λn.

It is easy to see that the extremum of normal stress n · τ n is given by the Lagrange
multiplier as λ = n · τ n. Thus, extremum of the quantity is equal to eigenvalue or
principal stress and the extremum directions are eigenvectors or principal directions.

5. The angle between maximum shear stress plane and maximum normal stress is 45◦
and the angle between maximum shear stress plane and minimum normal stress is
also 45◦ . In other words, let λ1 , λ2 and λ3 be principal stresses such that λ1 ≥ λ2 ≥

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

λ3 and n1 , n2 and n3 be corresponding principal directions. Then the normal to


maximum shear stress plane makes 45◦ angle with both n1 and n3 . The maximum
shear stress is given by (λ1 − λ3 )/2.

Proof. Let {n1 , n2 , n3 } be a set of orthonormal eigenvectors of stress tensor τ . Then


the set forms a basis to the 3-D Euclidean space. Let us consider a plane with a
unit normal ñ = (ñ1 , ñ2 , ñ3 ). Then the components of traction vector t on the plane
is given by (λ1 ñ1 , λ2 ñ2 , λ3 ñ3 ). Therefore, the square of magnitude of shear stress is
given by

Ts2 = t · t − Tn2 (where Tn is normal stress)


2
= t · t − (n · t)
 2
= (λ1 ñ1 )2 + (λ2 ñ2 )2 + (λ3 ñ3 )2 − λ1 (ñ1 )2 + λ2 (ñ2 )2 + λ3 (ñ3 )2
= (ñ1 ñ2 )2 (λ1 − λ2 )2 + (ñ2 ñ3 )2 (λ2 − λ3 )2 + (ñ3 ñ1 )2 (λ3 − λ1 )2

Similar to previous proof, the condition for extremization of shear stress can be
stated as
∂Ts2 ∂(1 − ñ21 − ñ22 − ñ23 )
+γ = 0.
∂ ñk ∂ ñk
Taking partial derivative, we get
 
(λ1 − λ2 )2 ñ22 + (λ3 − λ1 )2 ñ23 − γ ñ1 = 0
 
(λ1 − λ2 )2 ñ21 + (λ2 − λ3 )2 ñ23 − γ ñ2 = 0
 
(λ1 − λ3 )2 ñ21 + (λ2 − λ3 )2 ñ22 − γ ñ3 = 0.

In order to get maximum shear stress, we need to solve above equations along with
constraint equation ñ21 + ñ22 + ñ23 = 1. It is easy to verify that the following sets are
solutions to above equations (Chandrasekharaiah and Debnath, 1994):
(i) ñ1 = ±1, ñ2 = 0, ñ3 = 0.
(ii) ñ1 = 0, ñ2 = ±1, ñ3 = 0.
(iii) ñ1 = 0, ñ2 = 0, ñ3 = ±1
(iv) ñ1 = ñ2 = ± √12 , ñ3 = 0.
(v) ñ1 = ñ3 = ± √12 , ñ2 = 0.
Among all sets of solutions the maximum shear stress planes are given by solution
set (v), i.e., ñ1 = ñ3 = ± √12 , ñ2 = 0. Furthermore, the maximum shear stress is
given by
1
Ts = (λ1 − λ3 ).
2

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. D. S. Chandrasekharaiah and L. Debnath, Continuum Mechanics, 1994, Academic


Press Inc., London.

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-4: Balance Laws


Lecture-28: The Equations of Motion in
Lagrangian Description

It is convenient to work with Lagrangian description for solid mechanics as usually ref-
erence configuration is known. Therefore, we present the transformation of differential
form of linear momentum balance from spatial description to material description. The
new stress measure called first Piola-Kirchhoff stress is defined, in this lecture, which can
take care of transformation of Cauchy stress. To retain the symmetry, we define another
stress measure called second Piola-Kirchhoff stress which requires the concept of power
conjugate and it is discussed. We now start with transformation of linear momentum
balance from spatial description to material description.
The linear momentum balance and the first Piola-Kirchhoff stress tensor:
Let B0 be a reference configuration and B be a deformed configuration at an instant
of time t. Let x = χ(X, t) be a map from B0 to B. Let ρ(x, t), v(x, t), b(x, t) and
τ (x, t) be density, velocity, body force per unit mass and Cauchy stress fields over B,
respectively. Then recall from Lectures 26 and 27 that the equations of motion can be
written as
Dv
ρ = ∇x · τ + ρb ∀x ∈ B, (1)
Dt
and
τ = τT. (2)
In solid mechanics, the goal of the problem is to find the deformation of the body and
the state of stress under the action of applied forces. However, it can be noted that the
governing equations (Cauchy’s equations of motion) consists of the variable x which itself
is unknown. Thus, we transform the equations into the reference configuration B0 . It is
clear from above equations that the following terms need to be transformed in order to
achieve the target.

• Transformation of Cauchy stress τ .

• Transformation of spatial divergence of Cauchy stress (∇x · τ ).

• Transformation of traction vector t(x, t, n).


Dv
• Transformation of body force b(x, t) and acceleration .
Dt
• Transformation of density ρ.

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Let F be a deformation gradient. Then we define the first Piola-Kirchhoff stress tensor
T (X, t) as Piola transformation1 of Cauchy stress, i.e.,

T = Jτ F −T = τ cof F , (3)

where J is determinant of F .
Using Piola identity, i.e., ∇X · cof F = 0 (see Problem-1 in Lecture-19), we can show
that
∇X · T = J∇x · τ , ∀x = χ(X, t). (4)
The proof is presented in Problem-2 of Lecture-19.
Let Ω0 be a part of B0 and Ω be corresponding material volume in B. Let Γ0 and Γ
be boundaries of Ω0 and Ω, respectively. Let n0 and n be field of unit normals to Γ0 and
Γ, respectively. Then, using Eq. (4) and ∂Ω = J∂Ω0 , we get
Z Z
∇x · τ ∂Ω = ∇X · T ∂Ω0
Ω Ω0

Using divergence theorem, we get


Z Z
τ n ∂Γ = T n0 ∂Γ0 .
Γ Γ0

Since the traction field in spatial description is given by t = τ n, using first Piola-Kirchhoff
stress we can have traction vector in material description

t0 (X, t, n0 ) = T n0 . (5)

Therefore, it is easy to see from integral transformation (see Problem-3 of Lecture-19)


Z Z
t ∂Γ = t0 ∂Γ0 .
Γ Γ0

Let b0 (X, t) be a transformation of b(x, t) to the reference configuration, i.e.,

b(x, t) = b(χ(X, t), t) = b0 (X, t) (6)

The relation ρ0 (X) = Jρ(x, t), where ρ0 (X) is the density field of reference configuration
B0 , follow from conservation of mass (see Eq (2) in Lecture-24).

The acceleration description in reference configuration can be written as


Dv ∂ 2χ
= 2.
Dt ∂t
Multiplying with J on both sides of first Cauchy’s equation of motion (i.e., Eq. (1)) and
substituting above terms, we get
∂ 2χ
ρ0 = ∇X · T + ρ0 b0 . (7)
∂t2
The symmetry of Cauchy stress and Eq. (3) implies

T T = F −1 T F T . (8)
1
See the Piola transformation in Lecture-19

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Clearly, the first Piola-Kirchhoff stress is not a symmetric tensor. In summary, the dif-
ferential form of equations of motion in the reference configuration can be stated as

∂ 2χ
ρ = ∇X · T + ρ0 b0
∂t2
FTT = TFT
T n0 = t0 .

Now the goal is to find the stress measure in Lagrangian description such that it is
symmetric.
Mechanical energy balance and the second Piola-Kirchhoff stress tensor:
Let v̄ be a velocity of particle and f be a force exerted on the particle. Then, in classical
mechanics, the power expended on particle f · v̄ is balanced by the rate of change of kinetic
energy of particle. We can have the counterpart of this notion in continuum mechanics
and it can be obtained from the equations of motion.
Taking dot product with velocity field v on both sides of equation of motion, i.e., Eq.
(1), we get
Dv
ρv · = v · (∇x · τ ) + ρv · b.
Dt
Let L be a velocity gradient, i.e., L = ∇x v, and D be a rate of deformation i.e., D =
(∇x v + ∇x v T )/2. Then the above equation can be written as

D v·v
 
ρ = v · (∇x · τ ) + ρv · b
Dt 2
= ∇x · (τ T v) − τ : ∇x v + ρv · b
= ∇x · (τ T v) − τ : L + ρv · b
= ∇x · (τ T v) − τ : D + ρv · b (since τ is symmetric).

Let Ω be a material volume in B. Then taking the integration of above equation over Ω,
we get following mechanical energy balance
D v·v
Z   Z Z Z
T
ρ ∂Ω = ∇x · (τ v) ∂Ω − τ : D ∂Ω + ρv · b ∂Ω.
Ω Dt 2 Ω Ω Ω

Let Γ be a boundary of Ω and n be a unit normal field to Γ. Then using divergence


theorem to first term on right hand side of equation, we get
D v·v
Z   Z Z Z
T
ρ ∂Ω = (τ v) · n ∂Γ + ρv · b ∂Ω − τ : D ∂Ω
Ω Dt 2 ZΓ Z Ω Z Ω
= v · (τ n) ∂Γ + ρv · b ∂Ω − τ : D ∂Ω.
Γ Ω Ω

Using transport theorem-II and the relation t = τ n, we can obtain


d Z v·v
  Z Z Z
ρ ∂Ω + τ : D ∂Ω = v · t ∂Γ + ρv · b ∂Ω. (9)
dt Ω 2 Ω Γ Ω

v · t ∂Γ +
ρv · b ∂Ω represents the net power
R R
Clearly, right hand side of equation Γ Ω
d R v·v

expended by the external forces on the material volume Ω and the term ρ ∂Ω
dt Ω 2

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

R
represents the rate of change of kinetic energy of material volume. The term Ω τ : D ∂Ω
is known as stress power or internal power.

Thus, the external power expended by the force field on material volume is balanced
by the sum of internal power and rate of change of kinetic energy.
R
Since Ω τ : D ∂Ω is stress power (internal power), the Cauchy stress τ is said to
be power conjugate to the rate of deformation tensor D. We now examine the power
conjugate to first Piola-Kirchhoff stress by transforming the stress power to material
description.
Let Ω0 be a volume in reference configuration corresponding to the material volume
Ω. Let T be first Piola-Kirchhoff stress. Let ρ0 be density of reference configuration. Let
b0 be body force per unit mass in material description in and t0 be first Piola-Kirchhoff
traction vector. Then using Eq. (9), we get
!
d Z 1 ∂χ ∂χ Z   Z
∂χ 0 Z
∂χ 0
ρ0 · ∂Ω0 + T F T : L ∂Ω0 = · t ∂Γ0 + ρ0 · b ∂Ω0 .
dt Ω0 2 ∂t ∂t Ω0 Γ0 ∂t Ω0 ∂t

The equation can be written as


!
d Z 1 ∂χ ∂χ Z Z
∂χ 0 Z
∂χ 0
ρ0 · ∂Ω0 + T : (LF ) ∂Ω0 = · t ∂Γ0 + ρ0 · b ∂Ω0 .
dt Ω0 2 ∂t ∂t Ω0 Γ0 ∂t Ω0 ∂t

DF
Substituting the relation, = LF , we obtain the following mechanical energy balance
Dt
in Lagrangian framework.
!
d Z 1 ∂χ ∂χ Z
DF Z
∂χ 0 Z
∂χ 0
ρ0 · ∂Ω0 + T : ∂Ω0 = · t ∂Γ0 + ρ0 · b ∂Ω0 . (10)
dt Ω0 2 ∂t ∂t Ω0 Dt Γ0 ∂t Ω0 ∂t

DF
The term T : represents the stress power per unit mass. Thus, the material time
Dt
DF
derivative of deformation gradient is power conjugate to the first Piola-Kirchhoff
Dt
stress.

We now want to find the stress measure that is power conjugate to the rate of Green
strain tensor. Let E be the Green strain tensor. Then
1
τ :D = τ : (L + LT )
2
1 DF −1 DF T
= τ :( F + F −T )
2 Dt Dt
1 DF DF T
= τ : F −T (F T + F )F −1
2 Dt Dt
1 −1 DF DF T
= (F τ F −T ) : (F T + F)
2 Dt Dt
1 −1 DC
= (F τ F −T ) : (Since the left Cauchy-Green strain C = F T F )
2 Dt
1 −1 −1 DC
= J (F T ) :
2 Dt 
DE 1

= J −1 (F −1 T ) : Since E = (C − I) .
Dt 2

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

We now define the second Piola-Kirchhoff stress tensor S by

S = F −1 T . (11)

Substituting for first Piola-Kirchhoff stress in terms of Cauchy stress (i.e., Eq. (3)), we
get the following relation between second Piola-Kirchhoff stress and Cauchy stress.

S = JF −1 τ F −T . (12)

Upon substituting Eq. (11) in Eq. (10), we get the following mechanical energy balance
in terms of second Piola-Kirchhoff stress.
!
d Z 1 ∂χ ∂χ Z
DE Z
∂χ 0 Z
∂χ 0
ρ0 · ∂Ω0 + S: ∂Ω0 = · t ∂Γ0 + ρ0 · b ∂Ω0 .
dt Ω0 2 ∂t ∂t Ω0 Dt Γ0 ∂t Ω0 ∂t

Clearly, the second Piola-Kirchhoff stress tensor S is power conjugate to rate of Green
strain tensor. The second Piola-Kirchhoff stress tensor is symmetric stress measure in
material description unlike the first Piola-Kirchhoff stress tensor. The constitutive equa-
tions also take simple form due to symmetric nature of this stress measure. Thus, the
equations of motion in reference configuration using the second Piola-Kirchhoff stress can
be written as
∂ 2χ
ρ = ∇X · (F S) + ρ0 b0 , (13)
∂t2
S T = S, (14)
0 0
F Sn = t. (15)

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-4: Balance Laws


Lecture-29: The First Law of Thermodynamics

The energy conservation is known as the first law of thermodynamics which is one of
the basic axioms of continuum mechanics. This brings the effect of heat energy into the
continuum mechanics.
First law of thermodynamics: energy conservation:
Every continuum body always interact with its surroundings and in the process of interac-
tion there can be heat and work exchange between continuum body and its surroundings.
The interaction of heat and work is governed by the first law of thermodynamics also
known as law of conservation of energy. In previous lecture, we discussed the mechanical
energy balance of continuum body by ignoring the heat energy interaction with its sur-
roundings. We now account the heat interaction along with mechanical energy (or work)
in the first law of thermodynamics.
Rate of work transfer:
Let Ω be a material volume in deformed configuration B. Let ρ(x, t), b(x, t) and v be
density, body force per unit mass and velocity fields over Ω, respectively. Let Γ be the
boundary of Ω and n be its normal field. Let t(x, t, n) be a traction on Γ. Then the
kinetic energy of Ω is defined by
1Z
K(Ω) = ρv · v ∂Ω, (1)
2 Ω
and the power expended by external forces on Ω is defined by
Z Z
P (Ω) = ρb · v ∂Ω + t · v ∂Γ. (2)
Ω Γ

dW
We now define the net rate of work transfer as the delivery of expanded power to the
dt
continuum body without spending on the motion, i.e.,
dW dK
= P (Ω) − . (3)
dt dt

Internal energy:
The energy of continuum body excluding the kinetic energy is known as internal energy
which is denoted by U . Let e(x, t) be a specific internal energy (i.e., internal energy per
unit mass). Then the internal energy of material volume is given by
Z
U= ρe ∂Ω. (4)

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

The first law of thermodynamics states that


dU dW dQ
= + , (5)
dt dt dt
where W is work done on the system and Q is heat supplied to the system.
Let Qh be heat generation per unit mass and q be the rate of heat flux. Then heat
generation is given by Ω ρQh ∂Ω, and outflow of heat is given by Γ q · n ∂Γ. Thus, the
R R

net rate of heat flow can be written as


dQ Z Z
= ρQh ∂Ω − q · n ∂Γ. (6)
dt Ω Γ

Substituting Eqs. (3), (6) and Eq. (4) in Eq. (5), we get the following integral form of
first law of thermodynamics:
d Z Z Z
1d Z Z Z
ρe ∂Ω = ρb · v ∂Ω + t · v ∂Γ − ρv · v ∂Ω + ρQh ∂Ω − q · n ∂Γ. (7)
dt Ω Ω Γ 2 dt Ω Ω Γ

Rearranging the terms, we get


d Z v·v
 Z  Z Z Z
ρ + e ∂Ω = ρb · v ∂Ω + t · v ∂Γ + ρQh ∂Ω − q · n ∂Γ. (8)
dt Ω 2 Ω Γ Ω Γ

Let us consider the sum of internal energy and kinetic energy as total energy of Ω. Then
the rate of work done on the body plus rate of heat supplied to the body equal to rate of
change of total energy of the body. We now derive the differential form of conservation of
energy based on the statement of first law of thermodynamics, i.e., the energy expended
(work + heat) by the environment on the body is equal to the energy stored in the body.
Applying the second transport theorem to above equation (i.e., Eq. (8)), we get

D v·v
Z  Z  Z Z Z
ρ + e ∂Ω = ρb · v ∂Ω + t · v ∂Γ + ρQh ∂Ω − q · n ∂Γ. (9)
Ω Dt 2 Ω Γ Ω Γ

Let τ (x, t) be a Cauchy stress tensor field. Then using divergence theorem, we can have
Z Z
q · n ∂Γ = ∇x · q ∂Ω,
ZΓ ZΩ Z   Z  
T
t · v ∂Γ = (τ n) · v ∂Γ = τ v · n ∂Γ = ∇x · τ T v ∂Ω.
Γ Γ Γ Ω

Substituting these relations in Eq. (9), we get

D v·v
Z  Z  Z   Z Z
ρ + e ∂Ω = ρb · v ∂Ω + ∇x · τ T v ∂Ω + ρQh ∂Ω − ∇x · q ∂Ω.
Ω Dt 2 Ω Ω Ω Ω

Using the localization theorem, we obtain


D v·v
   
ρ + e = ρb · v + ∇x · τ T v + ρQh − ∇x · q.
Dt 2
This equation can also be written as
Dv De
ρv · +ρ = ρb · v + (∇x · τ ) · v + τ : ∇x v + ρQh − ∇x · q.
Dt Dt

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

(∇x v + ∇x v T )
Since the Cauchy stress is symmetric, we have τ : ∇x v = τ : .
2
Dv
Substituting the linear momentum balance ρ = ∇x · τ + ρb and the relation τ :
Dt
T
∇x v = τ : (∇x v + ∇x v )/2, in above equation, we get the following differential form of
first law of thermodynamics.
De
ρ = τ : D − ∇x · q + ρQh , (10)
Dt
where D is the rate of deformation, i.e., D = (∇x v + ∇x v T )/2.
Conservation of energy in reference configuration:
Let Ω0 be part of reference configuration and Ω be its corresponding material volume.
Let Γ0 be boundary of Ω0 . Let ρ0 be density of reference configuration. Let e0 (X, t) and
Q0h (X, t) be Lagrangian description of specific internal energy and heat generation per unit
mass, i.e., e(x, t) = e(χ(X, t), t) = e0 (X, t) and Qh (x, t) = Qh (χ(X, t), t) = Q0h (X, t).
Using Piola transform (see Lecture-19), we get

T (X, t) = Jτ (x, t)F −T ,


q 0 (X, t) = JF −1 q(x, t),

where J is determinant of deformation gradient F . Using Piola identity ∇X · cof F = 0)


(see Problem-1 in Lecture-19), we can show that ∇X · q 0 = J∇x · q. Let L be velocity
DF
gradient, i.e., L = ∇x v, then T : = Jτ : L. Substituting these relations in Eq. (10),
Dt
we get the following energy balance in material description.
∂e0 DF
ρ0 =T : − ∇X · q 0 + ρ0 Q0h . (11)
∂t Dt
Let S(X, t) be a second Piola-Kirchhoff stress tensor. Let E be a Green strain tensor.
Then we can also rewrite above equation in reference configuration as
∂e0 DE
ρ0 =S: − ∇X · q 0 + ρ0 Q0h . (12)
∂t Dt

Conservation of energy for control volume:


Let Ωc be a fixed control volume in space and Γc be its boundary. Then, using Eq. (10),
we can write
Z
De Z Z Z
ρ ∂Ω = τ : D ∂Ω − ∇x · q ∂Ω + ρQh ∂Ω.
Ωc Dt Ωc Ωc Ωc
 
Substituting the relation τ : D = τ : L = ∇x · τ T v − (∇x · τ ) · v, we get
Z
De Z 
T
 Z Z
ρ ∂Ω = ∇x · τ v − (∇x · τ ) · v ∂Ω − ∇x · q ∂Ω + ρQh ∂Ω.
Ωc Dt Ωc Ωc Ωc

Dv
Upon substituting the linear momentum balance, i.e., ∇x · τ = ρ − ρb and applying
Dt
divergence theorem, we obtain
De Dv
Z Z   Z  Z Z
ρ ∂Ω = τ T v · n ∂Γ − ρ − ρb · v ∂Ω − q · n ∂Γ + ρQh ∂Ω,
Ωc Dt Γc Ωc Dt Γc Ωc

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

where n is unit normal field to Γc . We can have the following form by rearranging terms.
D v·v
Z  Z  Z Z Z
ρ + e ∂Ω = (τ n) · v ∂Γ + ρb · v ∂Ω − q · n ∂Γ + ρQh ∂Ω.
Ωc Dt 2 Γc Ωc Γc Ωc

Since t(x, t, n) be a traction field, we can write

D v·v
Z  Z  Z Z Z
ρ + e ∂Ω = t · v ∂Γ + ρb · v ∂Ω − q · n ∂Γ + ρQh ∂Ω.
Ωc Dt 2 Γc Ωc Γc Ωc

Consider the term


D v·v D v·v Dρ v · v
      
ρ +e = ρ +e − +e
Dt 2 Dt 2 Dt 2
∂ v·v v·v Dρ v · v
        
= ρ + e + ∇x · ρ +e ·v− +e
∂t 2 2 Dt 2
∂ v·v v·v
      
= ρ + e + ∇x · ρ +e ·v
∂t 2 2
v·v Dρ
   
+ρ(∇x · v) +e Since mass balance = −ρ(∇x · v)
2 Dt
∂ v·v v·v
      
= ρ + e + ∇x · ρ +e v .
∂t 2 2
Substituting this term in above integral equation, we get
∂ v·v v·v
Z       
ρ + e + ∇x · ρ + e v ∂Ω
Z
Ωc ∂t
Z
2 Z
2 Z
= t · v ∂Γ + ρb · v ∂Ω − q · n ∂Γ + ρQh ∂Ω.
Γc Ωc Γc Ωc

Using the divergence theorem, we have the following energy balance for control volume:
∂ v·v v·v
Z   
Z  
ρ +e ∂Ω + ρ + e (v · n) ∂Γ
Ωc ∂t
Z
2 Z Γc
Z
2 Z
= t · v ∂Γ + ρb · v ∂Ω − q · n ∂Γ + ρQh ∂Ω.
Γc Ωc Γc Ωc

Since Ωc is fixed, we can also write


d Z v·v v·v
 
Z  
ρ + e ∂Ω + ρ + e (v · n) ∂Γ
Z
dt Ωc 2Z Γc
Z
2 Z
= t · v ∂Γ + ρb · v ∂Ω − q · n ∂Γ + ρQh ∂Ω.
Γc Ωc Γc Ωc

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-4: Balance Laws


Lecture-30: The Second Law of Thermodynamics

The second law of thermodynamics is another basic axiom of continuum mechanics. Un-
like other basic axioms, the second law is not a conservation law. However, this brings
new information on the direction of heat transfer and also imposes some restrictions on
constitutive relations.
The second law and the entropy:
The first law of thermodynamics states that the energy is conserved and does not put
any restriction on the direction of the process. For example, consider a hot body in
contact with a cold body. Naturally, the hot body releases heat energy and cold body
gains same amount of heat to get thermal equilibrium. The reverse process, i.e., hot
body becoming hotter and cold body becoming colder by exchanging equal amount of
heat energy, is not possible but does not violate the first law. Thus, the second law of
thermodynamics is stated to account the direction of the thermodynamic process. The
second law has an interesting history1 and it has many equivalent statements. Here, we
adopt a form of the second law of thermodynamics called the Clausius-Duhem inequality.
A new thermodynamic quantity called entropy is introduced to account the directionality
of thermodynamic processes.
The entropy physically represents a measure of the disorderness of the system. We know
from basic thermodynamics that any system can possess internal entropy. In addition,
the system can also exchange the entropy with its surroundings through the interaction
of energy and/or mass. Thus, the entropy production of the system can be defined by the
difference between the internal entropy and the entropy influx. The second law asserts that
the entropy production universe (system + surroundings) is nonnegative. Considering
the material volume as system, we can apply the second law of thermodynamics to the
continuous medium.
Mathematical statement of entropy imbalance:
Let Ω be a material volume of continuum body. Let S be net internal entropy of the
material volume Ω. Let J be the rate of entropy flow to the material volume from its
surroundings. Let H be the rate of entropy production. Then
dS
H= −J. (1)
dt
The second law of thermodynamics states that
H ≥ 0. (2)
1
D. Kondepudi and I. Prigogine, “Modern thermodynamics: From heat engines to dissipative struc-
tures”, 1998, John Wiley & Sons.

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

The non-negativity of entropy production is also known as entropy imbalance. The inte-
gral form of entropy imbalance can be stated by introducing field variables.

Integral form of entropy imbalance:


The entropy addition to the material volume can be divided into two parts: (i) entropy
influx/outflux through the boundary (ii) volumetric addition of entropy. Let f be a rate
of entropy outflux. Let ς(x, t) be rate of entropy supply per unit volume. Then the
entropy inflow is defined by
Z Z
J =− f · n ∂Γ + ς ∂Ω, (3)
Γ Ω

where Γ is boundary of material volume Ω and n is unit normal field to Γ. The negative
sign appears for the first term on the right hand side of equation as f is outflux (not
influx).
Let η(x, t) be specific entropy, i.e., entropy per unit mass. Then the net internal
entropy of material volume is defined by
Z
S = ρη ∂Ω. (4)

Using entropy imbalance, i.e., Eq. (2), we get

d Z Z Z
ρη ∂Ω ≥ − f · n ∂Γ + ς ∂Ω. (5)
dt Ω Γ Ω

This inequality is more general inequality of second law of thermodynamics. We now


specialize the inequality to get Clausius-Duhem inequality.

The Clausius-Duhem inequality:


Let θ be the absolute temperature scale, i.e., Kelvin’s scale of temperature. Then, moti-
vated by classical thermodynamics, we define
q ρQh
f= and ς= , (6)
θ θ
where q is rate of heat flux vector and Qh heat generation per unit mass as defined in
first law. Substituting these relation in Eq. (5), we get the following Clausius-Duhem
inequality:
d Z Z
ρQh Z
q·n
ρη ∂Ω ≥ ∂Ω − ∂Γ. (7)
dt Ω Ω θ Γ θ
The equality holds, in the above relation, only for a reversible process.
Using divergence theorem and localization theorem, we get the following differential
form of the Clausius-Duhem inequality:
Dη Qh 1 q
 
≥ − ∇x · . (8)
Dt θ ρ θ
Expansion of above relation yields
Dη Qh 1 1
≥ − ∇x · q + 2 q · ∇x θ. (9)
Dt θ ρθ ρθ

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Free energy imbalance and the dissipation:


We define a scalar field called specific free energy by

ψ(x, t) = e(x, t) − η(x, t) θ(x, t), (10)

where e(x, t) is specific internal energy. Taking material time derivative, we obtain
Dψ De Dη Dθ
= − θ−η . (11)
Dt Dt Dt Dt
De De
Substituting from the fist law of thermodynamics, i.e., ρ = τ : D − ∇x · q + ρQh ,
Dt Dt
we get
Dψ 1 1 Dη Dθ
= τ : D − ∇x · q + Qh − θ−η . (12)
Dt ρ ρ Dt Dt
Rearranging terms, we obtain
Dη 1 Dψ 1 Dθ
− θ + Qh − ∇x · q = − τ :D+η .
Dt ρ Dt ρ Dt
Substituting this relation in Eq. (9), we get the following Clausius-Duhem inequality in
terms of free-energy.
Dψ Dθ 1 q · ∇x θ
+η − τ :D+ ≤ 0. (13)
Dt Dt ρ ρθ
We now define the dissipation δ per unit volume by
!
Dψ Dθ 1
δ =τ :D−ρ +η − q · ∇x θ. (14)
Dt Dt θ
Thus, the dissipation δ ≥ 0.
The Clausius-Duhem inequality in Lagrangian description:
Let ρ0 be the density of reference configuration. Let F (X, t) be deformation gradient
and J be its determinant. Let T (X, t) and S(X, t) be first and second Piola-Kirchhoff
stress tensors. Let θ0 (X, t), ψ 0 (X, t), η 0 (X, t), Q0h (X, t) and q 0 (X, t) be material de-
scription of temperature, free-energy, entropy, heat generation per unit mass and heat
flux, respectively. Then we have the following relationships:

ρ0 = ρJ,
1 DF
τ :D=τ :L = T : ,
J Dt
q 0 = JF −1 q
∇X · q 0 = J∇x · q
∇X θ 0 = F T ∇x θ

Substituting all the terms in Eq. (13), we get the Clausius-Duhem inequality with respect
to reference configuration
Dψ 0 Dθ0 1 DF q 0 · ∇X θ 0
+ η0 − T : + ≤ 0. (15)
Dt Dt ρ0 Dt ρ0 θ 0
Since the total time derivative and material time derivative are equal, we can also write
∂ψ 0 ∂θ0 1 ∂F q 0 · ∇X θ 0
+ η0 − T : + ≤ 0. (16)
∂t ∂t ρ0 ∂t ρ0 θ 0

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Let E(X, t) be a Green strain tensor. Then we have the following Clausius-Duhem
inequality in Lagrangian description.
∂ψ 0 ∂θ0 1 ∂E q 0 · ∇X θ0
+ η0 − S: + ≤ 0. (17)
∂t ∂t ρ0 ∂t ρ0 θ0

The Clausius-Duhem inequality for control volume:


Let Ωc be a fixed control volume in space and Γc be its boundary. Let n be field of unit
outward normals to Γc . Considering the Clausius-Duhem inequality (see Eq. (8)), we can
write
Dη Qh q
Z Z Z  
ρ ∂Ω ≥ ρ ∂Ω − ∇x · ∂Ω.
Ωc Dt Ωc θ Ωc θ
Using divergence theorem to second term on the right hand side, we get
Z
D(ρη) Dρ Z
Qh Z
1
− η ∂Ω ≥ ρ ∂Ω − q · n ∂Γ.
Ωc Dt Dt Ωc θ Γc θ


Applying the conservation of mass, i.e., = −ρ(∇x · v), we get
Dt
Z
D(ρη) Z
Qh Z
1
+ ρ(∇x · v)η ∂Ω ≥ ρ ∂Ω − q · n ∂Γ.
Ωc Dt Ωc θ Γc θ

D(ρη) ∂(ρη)
Since = + ∇x (ρη) · v, we can write
Dt ∂t
Z
∂(ρη) Z
Qh Z
1
+ ∇x (ρη) · v + ρ(∇x · v)η ∂Ω ≥ ρ ∂Ω − q · n ∂Γ.
Ωc ∂t Ωc θ Γc θ

Substituting the relation ∇x · (ρηv) = ∇x (ρη) · v + ρ(∇x · v)η yields


Z
∂(ρη) Z Z
Qh Z
1
∂Ω + ∇x · (ρηv) ∂Ω ≥ ρ ∂Ω − q · n ∂Γ.
Ωc ∂t Ωc Ωc θ Γc θ

Using divergence theorem, we obtain the following Clausius-Duhem inequality for control
volume: Z
∂(ρη) Z Z
Qh Z
1
∂Ω + ρηv · n ∂Γ ≥ ρ ∂Ω − q · n ∂Γ. (18)
Ωc ∂t Γc Ωc θ Γc θ

Since Ωc is fixed, we can also write


d Z Z Z
Qh Z
1
ρη ∂Ω + ρηv · n ∂Γ ≥ ρ ∂Ω − q · n ∂Γ. (19)
dt Ωc Γc Ωc θ Γc θ

We use the inequality statement of second law of thermodynamics to impose some restric-
tions on constitutive relations.

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

4. M. E. Gurtin, E. Fried and L. Anand, The Mechanics and Thermodynamics of


Continua, 2010, Cambridge University Press, New York.

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-4: Balance Laws


Lecture-31: The Frame of Reference and
the Transformation of Kinematical Quantities

We need a frame of reference to describe any physical quantity. For example, the earth
velocity is measured with respect to the sun. Unfortunately, the frame of reference is not
necessarily a fixed object (i.e., object at rest). For example, the frame of reference can
be attached either to the earth or to the moon. Clearly, the frames of reference can have
a relative motion with respect to each other. The description of motion of one frame
with respect to other is known as Euclidean transformation. This transformation resem-
bles rigid body motion. Furthermore, the physical quantities that are observed in two
frames of reference are related by the Euclidean transformation. In particular, kinematical
quantities as well as governing equations can be transformed from one frame of reference
to other by the Euclidean transformation. In this lecture, we discuss transformation of
kinematical quantities.

Frame of reference:
A frame of reference is fixed to the observer1 who makes observations in terms of position
and time. We now consider two observers who are in relative motion with each other and
start observing the same event using their rulers and clocks. Since both observers posses
their individual clocks and rulers, they can make different measurement of distances and
times of the same event. However, we assume that both observers are measuring length
and time in same units. In other words, both the observers can have different time
settings in their clocks although they have same units of measurements. We now derive
the relationship between observed kinematical quantities.
Let B0 be a reference configuration and B be a deformed or observed configuration.
Let F ∗ and F be two reference frames which are in relative motion. Without loss of
generality, let a fixed-star in the universe be a frame of reference F ∗ and F be a moving
frame with respect to fixed star. Let O∗ be a origin of F ∗ and O be a origin of F . Let
{e∗1 , e∗2 , e∗3 } and {e1 , e2 , e3 } be orthonormal basis to the frame F ∗ and F , respectively,
as shown in Fig. (1). Since both clocks of observers can set to different initial time, Let
t∗ and t be time setting in F ∗ and F , respectively. Let s be difference in time settings.
Then the position of a point in B and times are related by

x∗ = Q(t)x + c(t),
t∗ = t − s, (1)

where Q(t) represents rotation of frame F with respect to F ∗ and c(t) represents position
1
The observer could be any object, i.e., human, animal, instrument etc.

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

of O with respect to O∗ . As noted, the position and time relationship shown in Eq. (1)
is known as Euclidean transformation.

Figure 1: Fixed star and moving frame of references

The change of frame of reference should not be confused with a change of basis. In
any given frame, we can choose infinitely many orthonormal basis. Transformation from
one basis to another basis can be done using a rotation tensor which is time independent.
Whereas, the rotation tensor appearing in change of frame is a function of time. Since the
rotation tensor in frame change is function of time, we can have a skewsymmetric tensor
DQ T
Wf = Q . (2)
Dt
The skewsymmetric tensor W f is known as frame-spin and it represents the rate of spin
of frame F with respect to fixed frame F ∗ .

For simplicity, we set Q(0) to be identity and c(0) to be zero vector. One can draw
all the conclusions that we make in this discussion, without making any assumption on
Q(0) and c(0). Clearly, the change of frame affects the quantities that are pertaining to
the deformed configuration but it does not affect reference configuration.

• Material points and material vectors are invariant, i.e., points and vector in refer-
ence configuration are not affected by the change of frame.

Frame-indifferent fields:
Let φ(x, t), v(x, t) and T (x, t) be scalar, vector and tensor fields over the deformed
configuration B by the frame F and φ∗ (x∗ , t∗ ), v ∗ (x∗ , t∗ ) and T ∗ (x∗ , t∗ ) be same scalar,
tensor and vector fields observed by the frame F ∗ . If these fields follow the relation

φ∗ (x∗ , t∗ ) = φ(x, t),


v ∗ (x∗ , t∗ ) = Q(t)v(x, t), (3)
∗ ∗ ∗ T
T (x , t ) = Q(t)T (x, t)Q(t) ,

then they are said to be frame-indifferent or objective. We now show that some kine-
matical quantities are frame-indifferent while some others are not. If the quantity is
frame-indifferent then the magnitude of quantity observed by observers in both frames
is same. In addition, the frame-indifferent second-order tensors have same eigenvalues in
both the frames of reference.

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Transformation rules for kinematical quantities:


Suppose that the motion of continuum body is described in the frame F by x = χ(X, t)
and also the same event described in the frame F ∗ by the relation x∗ = χ∗ (X ∗ , t∗ ). Since
material points and vector are invariant, i.e., X ∗ = X, we have

χ∗ (X, t∗ ) = Q(t)χ(X, t) + c(t), (4)

where Q is rotation tensor, i.e., relative rotation between frames, and c is relative trans-
lation of frames. Of course t∗ = t − a as noted in Eq. (1).
We now determine the relation between various kinematical quantities using Eq. (4). If
the quantity is a function of time alone then we indicate time derivative with superposed
dQ
dot, e.g., Q̇ ≡ . We note that time derivative of any quantity with respect to t∗ and
dt
t are same as t∗ = t − a.

Deformation gradient:
Let F ∗ and F be deformation gradients observed in frame F ∗ and F . Taking material
gradient on both sides of Eq. (1), we get

F ∗ = ∇χ∗ = Q∇χ = QF . (5)

Thus, the deformation gradient depends on frame. However, it can be observed that
det(F ∗ ) = det(F ) as det(Q) = 1. Hence, the transformation of volume element is invariant
under the change of frame.
Right and left Cauchy-Green strain tensors:
Let C ∗ and B ∗ be right and left Cauchy-Green strain tensors observed in frame F ∗ ; C
and B be right and left Cauchy-Green strain tensors observed in frame F . Then, using
Eq. (5) and definitions, we have

C ∗ = (F ∗ )T F ∗ = (QF )T QF = F T QT QF = F T F = C,
B ∗ = F ∗ (F ∗ )T = QF (QF )T = QF F T QT = QBQT .

Thus, the right Cauchy-Green strain tensor is invariant and the left Cauchy-Green strain
tensor is frame-indifferent.
Right and left stretch tensors:
Let U ∗ and U be right stretch tensors observed from the F ∗ and F frames; V ∗ and V
be left stretch tensors observed from the F ∗ and F frames. Then, using deformation
gradient in F ∗ and F , we get

F ∗ = R∗ U ∗ = V ∗ R∗ , F = RU = V R,

where R∗ and R are rotation tensors observed from F ∗ and F .


Using the relation between right stretch tensor and right Cauchy-Green strain (see
Lecture-18), we can get √ √
U ∗ = C ∗ = C = U ∗. (6)
Substituting this relation in above polar decomposition, we obtain

F ∗ = R∗ U

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Using relation between F ∗ and F , i.e., F ∗ = QF , we can have

R∗ = QR, V ∗ R∗ = QV R.

Since R∗ and Q are orthogonal tensors, above relation yields

V ∗ = QV QT . (7)

Thus, the right stretch tensor is invariant and the left stretch tensor is frame-indifferent.
Green strain and Almansi strain tensors:
Recall the definition of Green strain E and Almansi strain Ẽ from Lecture-18. Using
right and left Cauchy-Green strain in F ∗ and F frames, we get

E ∗ = E, (8)
∗ T
Ẽ = QẼQ . (9)

Similar to right and left Cauchy-Green strain tensors, the Green strain tensor is invariant
whereas Almansi strain tensor is frame-indifferent.

Velocity field:
Let v ∗ and v be velocity fields observed from F ∗ and F . Then taking material time
derivative on both sides of Eq. (4), we get

Dχ∗
v∗ =
Dt
Dχ DQ Dc
= Q + χ+
Dt Dt Dt
= Qv + Q̇x + ċ. (10)

Clearly, the velocity is neither invariant nor frame-indifferent. Physically, it can be seen
that velocity is not frame-indifferent as different observers can record different velocities.
For example consider a small ball moving radially outward in a straight groove on spinning
disc. Let F ∗ be fixed to the center of spinning disc and F be fixed to the ball. Then
velocity observed in frame F is zero whereas non-zero velocity is observed in frame F ∗ .
Thus, velocity is not frame-indifferent, i.e., velocity in one frame cannot be obtained by
simple orthogonal transformation of velocity in another frame.
Acceleration field:
Let a∗ and a be acceleration field observed from F ∗ and F ∗ . Then taking the material
derivative on both sides of Eq. (10), we get the following relation between accelerations:

a∗ = Qa + 2Q̇v + Q̈x + c̈. (11)

Similar to velocity, the acceleration field is also neither invariant nor frame-indifferent.

Velocity gradient:
The velocity gradient is defined as spatial gradient of velocity. Let L∗ and L be velocity
gradient observed from F ∗ and F . Then using relation between deformation gradient

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

and velocity gradient we can write


DF ∗ ∗ −1
L∗ = (F )
Dt
D(QF )
= (QF )−1 (Using Eq. (5))
Dt !
DQ DF
= F +Q F −1 QT .
Dt Dt

DF
Substituting the relation = LF , we get
Dt
L∗ = QLQT + Q̇QT . (12)

Thus, the velocity gradient is neither invariant nor frame-indifferent. Since the tensor
Q̇QT is skewsymmetric, the trace of velocity gradient tensor, i.e., spatial divergence of
velocity, is frame-indifferent.

∇x∗ · v ∗ = tr L∗ = tr L = ∇x · v. (13)

Rate of deformation and spin tensors:


The symmetric part and skewsymmetric parts of velocity gradient are defined as rate of
deformation and spin tensors. Let D ∗ and W ∗ be rate of deformation and spin tensors
observed in F ∗ and D and W be rate of deformation and spin tensors observed in F .
Then, using skewsymmetric nature of Q̇QT , we have
1 ∗ 
D∗ = L + (L∗ )T = QDQT , (14)
2
1 ∗ 
W∗ = L − (L∗ )T = QW QT + Q̇QT . (15)
2
Thus, rate of deformation is frame-indifferent, whereas spin tensor is not.
Material time derivative of Green and Almansi strain tensors:
Taking material time derivative of Green strain tensor, we get
DE ∗
= (F ∗ )T D ∗ F ∗
Dt
= (QF )T QDQT QF
= F T DF
DE
= .
Dt
Using the relation between material time derivative of Almansi strain, rate of deformation
and velocity gradient (see Lecture-21), we get

DẼ ∗
= D ∗ − (L∗ )T Ẽ ∗ − Ẽ ∗ L∗
Dt
= QDQT − (QLQT + Q̇Q)T (QẼQT ) − (QẼQT )(QLQT + Q̇Q)
= Q(D − LT Ẽ − ẼL)QT − QQ̇T QẼQT − QẼQT Q̇QT
DẼ T
= Q Q + Q̇ẼQT + QẼ Q̇T (Since Q̇QT = −QQ̇T ).
Dt

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Thus, material time derivative of Green strain is invariant and the material time derivative
of Almansi strain is neither invariant nor frame-indifferent.

In summary, we have the following transformation laws for the kinematical quantities:

F ∗ = QF ,
C ∗ = C,
B ∗ = QBQT ,
U∗ = U,
V ∗ = QV QT ,
E ∗ = E,
Ẽ ∗ = QẼQT ,
v ∗ = Qv + Q̇x + c,
a∗ = Qa + 2Q̇v + Q̈x + c̈,
L∗ = QLQT + Q̇QT ,
D ∗ = QDQT ,
W ∗ = QW QT + Q̇QT ,
DE ∗ DE
= ,
Dt Dt
DẼ ∗ DẼ T
= Q Q + Q̇ẼQT + QẼ Q̇T .
Dt Dt
These quantities can be grouped into following three categories:

• The tensors B, V , Ẽ and D and the scalars J (i.e., det F ) and tr(L) (i.e., ∇x · v),
are frame-indifferent.
DE
• The tensors C, U , E and are invariant.
Dt
DẼ
• The tensors F , L, W and and the vectors v and a are neither frame-indifferent
Dt
nor invariant.

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-4: Balance Laws


Lecture-32: The Principle of Frame-Indifference

The principle of frame-indifference states that the form of governing equations must be
independent of frame of reference. In other words, there is no preferred frame such as
‘inertial’ frame in classical mechanics setup. In this lecture, we shall see the assumptions
that are required to make mass balance and laws of dynamics independent of frame of
reference.

Law of mass conservation:


Let ρ∗ , F ∗ and v ∗ be density, deformation gradient and velocity fields observed by observer
in frame F ∗ and ρ, F and v be density, deformation gradient and velocity fields observed
by observer in frame F . As it was pointed out in previous lecture, the position vectors
and times in frames F ∗ and F are related by the following Euclidean transformation:

x∗ = Q(t)x + c(t),
t∗ = t − s,

where Q(t) represents relative rotation, c(t) represents relative translation and s repre-
sents the time difference in clocks.

Let J ∗ and J be determinant of F ∗ and F . Then, as we proved in previous lecture, the


relation between the determinant of deformation gradients and divergence of velocities
can be written as
J ∗ = J and ∇x∗ · v ∗ = ∇x · v. (1)
Let ρ0 (X) be density of reference configuration and which is not affected by frame of
reference. Then the mass balance in Lagrangian framework can be written as

ρ0 = ρ∗ J ∗ = ρJ. (2)

Substituting Eq. (1)1 in Eq. (2), we get

ρ∗ = ρ. (3)

Thus, the density is frame-indifferent. Substituting Eqs. (3) and (1)2 in conservation of
mass, we get
Dρ∗ Dρ
+ ρ∗ ∇x∗ · v ∗ = 0 =⇒ + ρ∇x · v = 0. (4)
Dt Dt
Thus, conservation of mass has same form in all frames.
Conservation of linear momentum:
We now postulate that traction vector is frame-indifferent. Consequently, the Cauchy

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

stress tensor becomes frame-indifferent. We emphasize that the frame-indifference nature


of stress tensor cannot be derived from kinematics without using the postulate.

Let Ω be a material volume in deformed configuration B and Γ be its boundary. Let


n and n be normal field observed by the frames of reference F ∗ and F , respectively.

Let Ω0 be a part of reference configuration B0 which is corresponding to Ω. Let Γ0 be


boundary of Ω0 and n0 be field of unit normal to Γ0 . Then
∗ (cof F ∗ )n0
n =
|(cof F ∗ )n0 |
(cof(QF ))n0
=
|(cof(QF ))n0 |
(cof(Q)cof(F ))n0
=
|(cof(Q)cof(F ))n0 |
Q(cof(F )n0 )
=
|(Q(cof(F )n0 )|
(cof F )n0
= Q
|(cof F )n0 |
= Qn. (5)
Let t∗ (x∗ , t∗ , n∗ ) and t(x, t, n) be traction vectors observed by frames F ∗ and F , respec-
tively. Then the postulate of traction frame-indifferent can be written as
t∗ = Qt. (6)
Let τ ∗ and τ be Cauchy stress field in F ∗ and F . Then we obtain
t∗ = τ ∗ n∗ = τ ∗ Qn (Using Eq. (5))
Qt = τ ∗ Qn (Using Eq. (6))
=⇒ t = QT τ ∗ Qn
=⇒ τ n = QT τ ∗ Qn
=⇒ (τ − QT τ ∗ Q)n = 0.
Since n is arbitrary, we get
τ ∗ = Qτ QT . (7)
Thus, the Cauchy stress tensor is frame-indifferent.
We now see transformation for the term ∇x · τ that appear in first Cauchy’s equation
of motion.
∂τij∗ ∗
∇ x∗ · τ ∗ = e
∂x∗j i
∂τmn ∗
= Qim Qjn e
∂x∗j i
∂τmn
= Qim Qjn Qjk e∗i
∂xk
∂τmn ∗
= Qim δnk e
∂xk i
∂τmn ∗
= Qim e
∂xn i
= Q∇x · τ .

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Therefore, the vector ∇x · τ is frame-indifferent. Let a and b be acceleration and body


force per unit mass fields, respectively, in frame F ; a∗ and b∗ be acceleration and body
force per unit mass fields, respectively, in frame F ∗ . Then we have

a∗ = Qa + 2Q̇v + Q̈x + c̈.

We note that the first Cauchy’s equation of motion in F ∗ and F can be written as

∇x∗ · τ ∗ = ρ∗ (a∗ − b∗ ), ∇x · τ = ρ(a − b).

In order to satisfy principle of frame-indifference for the linear momentum balance, we


must have
b∗ = Qb + 2Q̇v + Q̈x + c̈. (8)
We can observe that both acceleration and body force per unit mass are not frame-
indifferent but the difference of these two, i.e., the quantity ρ(a − b), is frame-indifferent.
The body force per unit mass observed by moving frame F can be written as

b = QT b∗ − 2QT Q̇v − QT Q̈x − QT c̈.

The frame-spin W f = Q̇QT is skewsymmetric tensor. Consequently, QT W f Q and


QT Ẇ f Q are also skewsymmetric tensors. The vector field b can be represented in terms
of W f as   2 
b = QT b∗ − 2QT W f Qv − QT Ẇ f + W f Qx − QT c̈.

Using the relation Q̇ = W f Q, we can obtain


d
(QT W f Q) = QT Ẇ f Q.
dt
Let ω be an axial vector of QT W f Q. Then the time derivative of axial vector, i.e., ω̇, is
the axial vector of QT Ẇ f Q. Thus, we have

b = QT b∗ − 2ω × v − ω̇ × x − ω × (ω × x) − QT c̈.

The terms other than QT b∗ are arising due to relative motion between reference frames.
These terms are usually known as

• 2ω × v Coriolis force,

• ω̇ × x Euler force,

• ω × (ω × x) centrifugal force,

• QT c̈ inertial force of relative translation.

Galilean transformation:
The Galilean transformation is a special case of Euclidean transformation where Q and
ċ are constants. Physically, in Galilean transformation the frames move at constant lin-
ear velocity with respect to each other. It can be noted that Coriolis force, Euler force,

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

centrifugal force and inertial force of relative translation becomes zero in Galilean trans-
formation. Hence, the acceleration and body force per unit mass are frame-indifferent,
i.e,
b∗ = Qb and a∗ = Qa.
Thus, the linear momentum balance is naturally frame-indifferent in Galilean transfor-
mation. In other words, the linear momentum balance has same form in all frames that
are related by Galilean transformation.
Conservation of angular momentum:
We have shown in Lecture-27 that the conservation of angular momentum is equivalent
to symmetry of Cauchy stress tensor. Let τ ∗ and τ be Cauchy stress fields observed by
the frames F ∗ and F , respectively. Using frame-indifference nature of Cauchy stress and
symmetry property in frame F , we have
 T
(τ ∗ )T = Qτ QT = Qτ T QT = Qτ QT = τ ∗ .

Thus, angular momentum has same form in all frames.


We note that the first law of thermodynamics along with principle of frame-indifference
give rise to all other balance laws. This topic will be discussed in the next lecture. We
now see the material time derivative of stress with change of frame.
Frame-indifferent stress rates:
Even though the Cauchy stress tensor is frame-indifferent, we now show that the ma-
terial time derivative of Cauchy stress tensor is not frame-indifferent. Let τ ∗ and τ be
Cauchy stress fields observed from frames F ∗ and F , respectively. Then taking material
derivative of τ ∗ , we get
Dτ ∗ D   Dτ T
= Qτ QT = Q̇τ QT + Q Q + Qτ Q̇T .
Dt Dt Dt
Clearly, the stress rate is objective if and only if

Q̇τ QT + Qτ Q̇T = 0, ∀Q ∈ Orth+ .

Since W f is skewsymmetric tensor and Q̇ = W f Q, we get

W f Qτ QT = Qτ QT W f .

The skewsymmetric tensor W f is arbitrary as Q is arbitrary rotation tensor. This implies

Qτ QT = λI,

where λ is some constant. Hence, Cauchy stress tensor must have the form τ = λI if
its material time derivative is frame-indifferent. However, in general the Cauchy stress
tensor need not have the form of λI. Thus, in general, the material time derivative of
Cauchy stress is not frame-indifferent.
The rate of stress is required to represent the constitutive relations of rate-dependent
materials. These constitutive relations demand the frame-indifferent stress rates as they
are necessarily frame-indifferent. Clearly, as shown above, the rate of Cauchy stress is not

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

frame-indifferent. Thus, we now construct a stress rate that is frame-indifferent using the
invariance nature of second Piola-Kirchhoff stress tensor.
S ∗ = J (F )−1 τ ∗ (F )−T
= JF −1 Q−1 Qτ QT Q−T F −T
= JF −1 τ F −T
= S
Since the Piola-Kirchhoff stress is invariant, the material time derivative is also invariant,
i.e.,
DS ∗ DS
= .
Dt Dt
Taking material derivative of relation S = JF −1 τ F −T , we get
DS D  −1 
= JF τ F −T
Dt Dt
DJ −1 DF −1 Dτ −T DF −T
= F τ F −T + J τ F −T + JF −1 F + JF −1 τ
Dt Dt Dt Dt

= (tr L)F −1 τ F −T + JF −1 Lτ F −T + JF −1 F −T + JF −1 τ LT F −T
Dt

 
= JF −1 − Lτ − τ LT + (tr L)τ F −T ,
Dt
where

τ◦ = − Lτ − τ LT + (tr L)τ . (9)
Dt
is known as Truesdell stress rate. It is easy to see that the Truesdell stress rate τ ◦ is
frame-indifferent. Using material rate of second Piola-Kirchhoff stress tensor, we can
write
1 DS T
τ◦ = F F .
J Dt
Let (τ ◦ )∗ and τ ◦ be Truesdell stress rates which are observed by frames F ∗ and F . Then
1 ∗ DS ∗
(τ ◦ )∗ = ∗
F (F ∗ )T
J Dt
1 DS T T
= QF F Q
J Dt
= Qτ ◦ QT .
We can obtain alternative frame-indifferent stress rates using Lie derivative. The Lie
derivative of spatial tensor field G with respect to another tensor field P is defined by
D  −1
  
LP (G) = P P GP −T P T .
Dt
Many frame-indifferent stress rates can be obtained using the Lie derivative by appropri-
ately choosing G and P in above definition. We note that the Truesdell stress rate can
also be defined using Lie derivative.
Truesdell stress rate:
Let LF (Jτ ) be the Lie derivative of Jτ with respect to F . Then the Truesdell stress rate
τ ◦ is defined by LF (Jτ )/J. Hence, we can write
1 D  −1
  
τ◦ = F F Jτ F −T F T .
J Dt

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

DF −1
Substituting = −F −1 L, we get Eq. (9).
Dt
Oldroyd stress rate:
The Oldroyd stress rate τ is defined as Lie derivative of τ with respect to F , i.e.,
D  −1
  

τ =F F τ F −T F T .
Dt
DF −1
Using the relation = −F −1 L, we have
Dt

τ = − Lτ − τ LT . (10)
Dt

Convective stress rate:


The convective stress τ  is defined as the Lie derivative of τ with respect to F −T , i.e.,
D  T
 
 −T
τ =F F τ F F −1 .
Dt
DF
Using the relation = LF , we have
Dt

τ = + LT τ + τ L. (11)
Dt

Green-Naghdi stress rate:


Let R be a rotation tensor that appears in polar decomposition of F , i.e., F = RU =
V R, where U and V are right and left stretch tensors. Then the Green-Naghdi stress
rate τ M is defined as the Lie derivative of τ with respect to R.
D  T  
τ =RM
R τ R RT .
Dt
DR T
Since R is skewsymmetric tensor, we have
Dt
Dτ DR T DR T
τM = − R τ +τ R . (12)
Dt Dt Dt

Jaumann stress rate:


We define Jaumann stress rate τ ∇ as average of Oldroyd and convective stress rates. Since
sum of frame-indifferent quantities is frame-indifferent, Jaumann stress rate is also frame
1
indifferent. Let W be skewsymmetric part of L, i.e., W = (L − LT ). Then Jaumann
2
stress rate can be written as

τ∇ = − Wτ + τW. (13)
Dt
It can be verified that τ , τ  , τ M and τ ∇ are frame-indifferent. Observing all the
stress rates that are discussed above, we can generalize the frame-indifferent material
time derivative of tensor fields.

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

Frame-indifferent material time derivative of tensor fields:


Let G be a frame-indifferent field, i.e., G∗ and G in frames F ∗ and F are related by

G∗ = QGQT .

Then the material time derivative of G is not frame-indifferent.


We now choose a tensor field P such that

P ∗ = QP ,

where P ∗ and P are observed by F ∗ and F . Then the frame-indifferent rate can be
DG DP
obtained in terms of G, P , and using the Lie derivative
Dt Dt
D  −1
  
LP (G) = P P GP −T P T . (14)
Dt
Thus, various frame-indifferent rates of both stresses and strains can be obtained.

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. J. Bonet and R. D. Wood, Nonlinear Continuum Mechanics for Finite Element


Analysis, 1997, Cambridge University Press, Cambridge.

5. I-S. Liu, Continuum Mechanics, 2002, Springer, Berlin.

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-4: Balance Laws


Lecture-33: Application of the Principle of
Frame-Indifference to
the First Law of Thermodynamics

Here, we show that the principle of frame-indifference plus the first law of thermodynamics
is equivalent to all other conservation laws (i.e., conservation of mass, linear momentum
and angular momentum). As discussed in Lecture-29, the first law of thermodynamics
states that rate of change of total energy (i.e., the kinetic and internal energy) is equal to
the rate of work done by external forces plus rate of change of heat energy, i.e.,
d Z 1
 Z  Z
ρ e + v · v ∂Ω = (t · v + γ) ∂Γ + ρ (b · v + Qh ) ∂Ω. (1)
dt Ω 2 Γ Ω

• The principle of frame-indifference states that the first law of thermodynamics as-
sumes same form in all frames of reference.

Let density (ρ), internal energy (e), heat flux across the surface (γ), heat supply per unit
mass (Qh ) and traction vector (t) be frame-indifferent. Let b be a body force that follows
transformation rule stated in principle of frame-indifference for linear momentum balance
(see Lecture-32). Then we have the following result.
The result of frame-indifference applied to first law of thermodynamics:
The principle of frame-indifference applied to the first law of thermodynamics implies:

i. there exist Cauchy stress tensor τ (x, t) such that t = τ n,

ii. there exist a heat flux vector q(x, t) such that γ = q · n,

iii. conservation of mass,

iv. conservation of linear momentum,

v. conservation of angular momentum.

Conversely, frame-indifference of conservation of mass, linear momentum and angular


momentum along with the existence of Cauchy stress tensor τ (x, t) imply the frame-
indifference of energy balance.
Proof: Let Ω be a material volume in B and Γ be its boundary. Let ρ∗ (x∗ , t∗ ), t∗ (x∗ , t∗ , n∗ ),
a∗ (x∗ , t∗ ), b∗ (x∗ , t∗ ), e∗ (x∗ , t∗ ), Q∗h (x∗ , t∗ ) and γ ∗ (x∗ , t∗ , n∗ ) be density, traction, acceler-
ation, body force per unit mass, specific internal energy, heat generation per unit mass

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

and heat flux across normal to the Γ in the frame F ∗ , respectively. Let ρ(x, t), t(x, t, n),
a(x, t), b(x, t), e(x, t), Qh (x, t) and γ(x, t, n) be density, traction, acceleration, body
force per unit mass, specific internal energy, heat generation per unit mass and heat flux
across normal to the Γ in the frame F . We note that γ(x, t, n) is analogous to traction
t(x, t, n).
As mentioned in previous lecture, the Euclidean transformation between the frames
F and F is given by

x∗ = Q(t)x + c(t), (2)


t∗ = t − s,

where s is difference in clocks setting.

The assumption we considered for the density, internal energy, heat flux, heat supply,
traction and body force can be written as

ρ∗ (x∗ , t∗ ) = ρ(x, t), e∗ (x∗ , t∗ ) = e(x, t), γ ∗ (x∗ , t∗ , n∗ ) = γ(x, t, n),


Q∗h (x∗ , t∗ ) = Qh (x, t), t∗ (x∗ , t∗ , n∗ ) = Qt(x, t, n), (a∗ − b∗ ) = Q(a − b).

Let us defined the energy density (total energy per unit volume) of continuum body f
defined by
v·v
 
f =ρ e+ .
2
Since Ω is a part of deformed configuration B, the balance of energy for Ω observed by
the frame F
d Z Z Z
f ∂Ω = (t · v + γ) ∂Γ + ρ (b · v + Qh ) ∂Ω.
dt Ω Γ Ω
Using first transport theorem, we get
!
Z
Df Z Z
+ f ∇x · v ∂Ω = (t · v + γ) ∂Γ + ρ (b · v + Qh ) ∂Ω. (3)
Ω Dt Γ Ω

Similarly, the energy balance observed by the frame F ∗ can be written as


Df ∗
Z ! Z Z
+ f ∗ ∇ x∗ · v ∗ ∂Ω = ∗ ∗ ∗
(t · v + γ ) ∂Γ + ρ∗ (b∗ · v ∗ + Q∗h ) ∂Ω.
Ω Dt Γ Ω

where f ∗ = ρ∗ (e∗ + v ∗ · v ∗ /2). Since ρ∗ = ρ, Q∗h = Qh and γ ∗ = γ, ∇x∗ · v ∗ = ∇x · v, we


get
Df ∗
Z ! Z Z
+ f ∇x · v ∂Ω = (t · v + γ) ∂Γ + ρ (b∗ · v ∗ + Qh ) ∂Ω.
∗ ∗ ∗
(4)
Ω Dt Γ Ω

Subtracting Eq. (3) from Eq. (4), we obtain


Z 
D ∗ Z  Z
(f − f ) + (f ∗ − f ) ∇x · v ∂Ω = (t∗ · v ∗ − t · v) ∂Γ+ ρ (b∗ · v ∗ − b · v) ∂Ω.
Ω Dt Γ Ω

Since e∗ = e, we have f ∗ − f = ρ(v ∗ · v ∗ − v · v)/2. Substituting f ∗ − f in the above


equation, we get
Z  ∗
v · v∗ − v · v Dρ
 Z 
+ ρ∇x · v ∂Ω + ρ(a∗ · v ∗ − a · v) ∂Ω

Z 2 Dt Z Ω

= (t∗ · v ∗ − t · v) ∂Γ + ρ (b∗ · v ∗ − b · v) ∂Ω.


Γ Ω

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Rearrangement of terms yields


Z  ∗
v · v∗ − v · v Dρ
 
Z
+ ρ∇x · v ∂Ω + ρ ((a∗ − b∗ ) · v ∗ − (a − b) · v) ∂Ω

Z 2 Dt Ω

= (t∗ · v ∗ − t · v) ∂Γ. (5)


Γ

Existence of Cauchy stress tensor:


Consider the Euclidean transformation, i.e., Eq. (2), where Q(t) and c(t) are arbitrary
rotation tensor and translation vectors. We now choose Q(t) = I and ċ = λû, where λ
is an arbitrary constant and û is an arbitrary constant unit vector. Substituting Q and
ċ in Eq. (2) and differentiating with respect to time, we get

v ∗ = v + ċ, and a∗ = a.

Using the transformation law for the field of body force per unit mass, we have

b∗ = b.

Since traction is frame-indifferent, substituting the above relations in Eq. (5), we get
!
λ2 Dρ
Z 
Z Z
λv · û + + ρ∇x · v ∂Ω + λρ(a − b) · û ∂Ω = λt · û ∂Γ. (6)
Ω 2 Dt Ω Γ

The argument similar to the Cauchy’s theorem (see Lectures 26 and 27) yields the exis-
tence of Cauchy stress tensor τ (x, t) such that t = τ n.

Existence of heat flux vector:


Substituting the relation between Cauchy stress and traction, i.e., t = τ n in Eq. (3), we
obtain
!
Z
Df Z Z Z
+ f ∇x · v ∂Ω = τ n · v ∂Γ + γ ∂Γ + ρ (b · v + Qh ) ∂Ω.
Ω Dt Γ Γ Ω

Using the relation τ n · v = τ T v · n and divergence theorem, we get


!
Z
Df Z Z Z
+ f ∇x · v ∂Ω − ρ (b · v + Qh ) ∂Ω − ∇x · (τ T v) ∂Ω = γ ∂Γ.
Ω Dt Ω Ω Γ

Using the similar arguments that were used in Cauchy’s theorem, we get the existence of
heat flux vector q(x, t) such that γ = q · n.
Conservation of mass:
Substituting the relation t = τ n in Eq. (6), we get
!
λ2 Dρ
Z 
Z Z
λ(v · û) + + ρ∇x · v ∂Ω + λρ(a − b) · û ∂Ω = λτ n · û ∂Γ.
Ω 2 Dt Ω Γ

Using the relation τ n · û = n · τ T û, the divergence theorem and ∇x · (τ T û) = (∇x · τ ) · û,
we can write
!
λ2 Dρ
Z  Z
λ(v · û) + + ρ∇x · v ∂Ω = λ (∇x · τ + ρ(b − a)) · û ∂Ω. (7)
Ω 2 Dt Ω

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Differentiating twice with respect to λ, we get


Z 


+ ρ∇x · v ∂Ω = 0.
Ω Dt
Arbitrariness of Ω implies the mass balance in Eulerian description, i.e.,

+ ρ∇x · v = 0. (8)
Dt

Conservation of linear momentum:


Substituting the mass balance in Eq. (7), we get
Z
(∇x · τ + ρ(b − a)) · û ∂Ω = 0.

Arbitrariness of Ω and û imply the linear momentum balance:

∇x · τ + ρ(b − a) = 0. (9)

Conservation of angular momentum:


Substituting Eq. (8) in Eq. (5), we get
Z Z
∗ ∗ ∗
ρ ((a − b ) · v − (a − b) · v) ∂Ω = (t∗ · v ∗ − t · v) ∂Γ. (10)
Ω Γ

Using the transformation law between frames, we have (a∗ − b∗ ) = Q(a − b) and t∗ = Qt.
Substituting these relations in above equation, we can write
Z Z

(ρQ(a − b) · v − ρ(a − b) · v) ∂Ω = (Qt · v ∗ − t · v) ∂Γ.
Ω Γ

Using the property of dot product, we get


Z Z

T
ρ(a − b) · (Q v − v) ∂Ω = t · (QT v ∗ − v) ∂Γ.
Ω Γ

Substituting the relation t = τ n on the right hand side of equation, we obtain


Z Z

T
ρ(a − b) · (Q v − v) ∂Ω = τ n · (QT v ∗ − v) ∂Γ.
Ω Γ

Using the property of dot product and the divergence theorem, we can write
Z Z  

T
ρ(a − b) · (Q v − v) ∂Ω = ∇x · τ T (QT v ∗ − v) ∂Ω
Ω ZΩ  
= (∇x · τ ) · (QT v ∗ − v) + τ : ∇x (QT v ∗ − v) ∂Ω.

Substituting Eq. (9), we get


Z
τ : ∇x (QT v ∗ − v) ∂Ω.

Taking time differentiation of Euclidean transformation, we obtain v ∗ = Qv + Q̇x + ċ.


Spatial gradient of resulting relation yields ∇x (QT v ∗ − v) = QT Q̇. Substituting the
relation in above equation, we get
Z
τ : (QT Q̇) ∂Ω.

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

It is easy to see that QT Q̇ is skewsymmetric tensor as QT Q = I. Furthermore, the


arbitrariness of Q implies the arbitrariness of QT Q̇. The arbitrariness of Ω and QT Q̇
imply the symmetry of τ (x, t). Clearly, the symmetry of stress tensor is equivalent to the
balance of angular momentum (see Lecture-27). Thus, the first law of thermodynamics
and the principle of frame-indifference imply the conservation of angular momentum.
The converse is equivalent to showing Eq. (5) as it is difference of energy balance in
both frames. We get Eq. (10), by reversing steps in the proof of conservation of angular
momentum with help the relation t = τ n and the linear momentum balance. Finally, the
conservation of mass and Eq. (10) imply Eq. (5). Thus, Converse is proved.

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-5: Constitutive Relations


Lecture-34: Necessity of Constitutive Equations and
the Principle of Material Frame-Indifference

Here, we discuss the need of constitutive relations and also the restrictions imposed by
the principle of material frame-indifference.
Necessity of constitutive relations:
We observe in physical world that different materials respond differently for the same
forces. For example, steel and rubber with same geometry undergoes different deforma-
tions for the same amount of forces. Similarly, response of water or gas is different from
the solids under the action of forces. Though all materials follow the same basic principles
but they do exhibit different behaviors. It shows that the basic principles that we dis-
cussed are not sufficient to characterize the behavior of materials. In fact, we now show
the same argument mathematically.
We have the following basic laws, i.e., conservation of mass, conservation of momenta,
conservation of energy and second law of thermodynamics:

+ ρ∇x · v = 0, (1)
Dt
Dv
ρ = ∇x · τ + ρb, and τ T = τ , (2)
Dt
De
ρ = τ : D − ∇x · q + ρQh , (3)
Dt
Dη ρQh 1 q
ρ ≥ − ∇x · q + 2 · ∇x θ. (4)
Dt θ θ θ
In summary, we have five independent equations as second law of thermodynamics is
inequality. On the other hand, we have following 16 unknowns in the equations,

• mass density ρ, (scalar: one unknown)

• velocity v, (vector: three unknowns)

• Cauchy stress tensor τ , (symmetric second-order tensor: six unknowns)

• internal energy e, (scalar: one unknown)

• heat flux vector q, (vector: three unknowns)

• temperature θ, (scalar: one unknown)

• entropy η (scalar: one unknown)

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

These unknowns (ρ, v, τ , e, q, θ and η) are also known a state variables as they represent
the state of continuum at every point.

Assuming that the continuum body interaction with external environment is known,
i.e., body forces per unit mass b and heat sources distribution Qh are known1 . Clearly,
the unknowns are more than the equations and hence it is not possible to determine
the unknowns uniquely. Thus, the problem is ill-posed with above governing equations
except for some trivial cases of rigid body motion. In order to make problem well-posed,
we need to specify 11 more equations known as constitutive relations. These constitutive
relations are specified on the basis of mechanical and thermal response of material, i.e.,
these equations are dependent on the internal constitution of the material.
Constitutive relations cannot be chosen arbitrarily and they must follow a set of prin-
ciples and rules imposed by physical laws and material symmetry. Though there is great
reduction in the constitutive relation due to these principles and rules but there are con-
stants in the constitutive relations that cannot be evaluated alone by these principles. So,
finally experiments are needed to obtain the constants in the constitutive relations for a
given material. We now discuss the formulation of constitutive relations.
Constitutive relations and the principle of determinism:
Let B0 be a reference configuration and B be a deformed configuration at an instant of
time t. Let x = χ(X, t) be a map from B0 to B. The constitutive relations can be stated
as eleven implicit functionals of 16 unknown fields in the governing equations, i.e.,

R [ρ(Y , t − p), χ(Y , t − p), τ (Y , t − p), e(Y , t − p), q(Y , t − p),
p=0
∀Y ∈B0
θ(Y , t − p), η(Y , t − p), X, t] = 0. (5)

where p represents the time history and t represent present time. The implicit constitutive
relation indicating that R at material a point X ∈ B0 may also depend on other martial
points Y ∈ B0 . We note that the functional R should be function of velocity, but the
mapping function χ is introduced in the arguments of R as the velocity is material (total)
time derivative of χ.
Notation used in above constitutive relations:

• R represents the eleven implicit relations in terms of sixteen unknown fields at an


instant of time t and at a material coordinate X.

• The letter p = 0 to ∞ indicates total time history

• The symbol ∀Y ∈ B0 shows that the constitutive response at material coordinate


X depends on every point Y in the domain B0 .

The principle of determinism states that the past history of motion and temperature at
all points of continuum body can determine the present state of the body. Furthermore, it
excludes the dependence of behavior on outside of body and future events. Therefore, we
1
The assumption of known body forces and heat source distribution is criticized by L. C. Woods in “The
bogus axioms of continuum mechanics”, Bulletin of the Institute of Mathematics and its Applications, 17,
pp. 98-102 (1981). However, the assumption is widely accepted. The framework of continuum mechanics
avoiding this assumption can be seen in I-Shih Liu, “Continuum Mechanics”, Springer, 2002.

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

use explicit functional equations such that they depend on history of motion, temperature
and density. This argument is also consistent with physical perception as density, motion
and temperature can be perceived by our physical senses. Thus, we choose density, motion
and temperature as independent variables whereas the Cauchy stress tensor, heat flux
vector, internal energy and entropy as dependent variables. Eleven constitutive relations
in explicit form can be written as

τ (x, t) = τ̂ (ρ(Y , t − p), χ(Y , t − p), θ(Y , t − p), X, t),
p=0
∀Y ∈B0

q(x, t) = q̂ (ρ(Y , t − p), χ(Y , t − p), θ(Y , t − p), X, t),
p=0
∀Y ∈B0

e(x, t) = ê (ρ(Y , t − p), χ(Y , t − p), θ(Y , t − p), X, t),
p=0
∀Y ∈B0

η(x, t) = η̂ (ρ(Y , t − p), χ(Y , t − p), θ(Y , t − p), X, t),
p=0
∀Y ∈B0

where x = χ(X, t). We note that all constitutive relations τ̂ , q̂, ê and η̂ are dependent
on same set of independent variables. The rule of considering same set of independent
variables in the explicit constitutive relations is known as principle of equipresence.
The constitutive equations need to satisfy following three requirements:

• principle of material frame-indifference,

• material symmetry,

• second law of thermodynamics.

These three requirements impose severe restrictions on the constitutive relations. There-
fore, it is useful for practical problems as constitutive relations can be obtained with few
experiments. Before imposing these requirements, we also assume that the material is
simple. We now define simple material or principle of local action.
Simple materials:
In general, the constitutive relations (relation between state variables such as stress, defor-
mation, density, temperature, heat flux, internal energy and entropy) at a point depend
on thermomechanical history of every point in the body. This is known as non-local
property of constitutive relationship. The non-local nature of constitutive relations poses
difficulty in obtaining solution to the thermomechanical problems. However, it is rarely
relevant to many practical problems in the mechanics. Therefore, we assume that the
thermomechanical history of neighbourhood can affect the response at a point but not
the furthest points. This assumption is known as simple material or principle of local
action. Furthermore, we consider first-order approximation of dependent variable to ac-
count for local action. Let X be a material point and Y be a point in the neighbourhood.
Let dX = Y − X. Then, using Taylor’s series approximation, we get
ρ(Y , t − p) = ρ(X, t − p) + ∇X ρ(X, t − p) · dX + o(dX),
χ(Y , t − p) = χ(X, t − p) + F (X, t − p)dX + o(dX), (6)
θ(Y , t − p) = θ(X, t − p) + ∇X (X, t − p) · dX + o(dX),

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

where F (X, t − p) is the deformation gradient.


Recall the relation ρ0 (X) = det(F (X, t − p))ρ(X, t − p), where ρ0 (X) is density of
reference configuration, from Lecture-24. Therefore, the history of density and its gradient
can be related to the history of deformation gradient and its derivatives and X as density
of reference configuration is constant. Since the derivative of deformation gradient is not
considered (see Eq.(6)2 ), the ρ(X, t − p) and ∇X ρ(X, t − p) can be directly related to
F (X, t − p) and X. Thus, we drop the argument of density and its derivative from
the constitutive relations. Let us denote gradient of temperature ∇X θ with g. Then,
substituting first-order approximation of independent variables in the general form of
explicit constitutive relations, we get the following response functions for simple material.

τ (X, t) = τ̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X, t),
p=0

q(X, t) = q̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X, t),
p=0

e(X, t) = ê (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X, t),
p=0

η(X, t) = η̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X, t).
p=0

In addition to simple material, we also assume that constitutive relations are not depen-
dent explicitly on time. In other words, we are assuming that material do not experience
aging2 .
The constitutive relations for the simple material with the assumption of no aging can
be written as

τ (X, t) = τ̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X),
p=0

q(X, t) = q̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X),
p=0

e(X, t) = ê (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X),
p=0

η(X, t) = η̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X).
p=0

Despite the assumptions of principle of local action (simple material) and no aging, the
present model includes the most of constitutive relations which are in practice. For
example Hooke’s law for elastic solids and Newton’s law of viscosity for fluids. As noted
earlier, we impose principle of material frame-indifference on constitutive relations.
Principle of material frame-indifference (MFI):
The principle of material frame-indifference postulates that the constitutive relations have
same form in all frames of reference. This is similar to the principle of frame-indifference
for the governing equations. However, the principle of frame-indifference for the gov-
erning equations has been universally accepted but not the principle of material frame-
indifference. There are some works that show the constitutive relations do not have
2
The assumption of no aging can be deduced from the principle of material frame-indifference, see
I-Shih Liu, “Continuum mechanics”, Springer, 2002.

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

same form in all frames (see for example Müller, 1972)3 . There are series of discussions
on the aspects of this principle and it has been shown that there are two assumptions
involved in this principle4 . The two assumptions involved are observer-invariance and
form-invariance.
Let τ̂ ∗ and τ̂ be constitutive relations for the quantities τ ∗ and τ , respectively. We
note that the quantities τ ∗ and τ are description of same physical variable from the
frames F ∗ and F , respectively. As stated in previous lectures, the frames of reference
are related by Euclidean transformation. Let l∗ and l be list of independent variables,
that are present in the constitutive relations, observed by F ∗ and F , respectively. Then
the observer-invariance can be written as

τ̂ ∗ (l∗ ) = Qτ̂ (l)QT . (7)

We note that τ̂ ∗ and τ̂ (l) could be different functions. In addition to observer-invariance,


we assume following form-invariance, i.e.,

τ̂ ∗ (l∗ ) = τ̂ (l∗ ). (8)

Substituting Eq. (8) in Eq. (7), we get

τ̂ (l∗ ) = Qτ̂ (l)QT . (9)

This condition (Eq. 9) is known as the principle of material frame-indifference or invari-


ance of constitutive relation under superposed rigid-body motion. It has also been shown
that any two conditions among observer-invariance, form-invariance and invariance of con-
stitutive relation under superposed rigid-body motion (Eqs. (7), (8) and (9)) can imply
other one.

Without debating much on controversy of this principle, in the absence of any experi-
mental evidence, we assume the principle of material frame-indifference as an axiom.
We present a simple example to understand the consequence of principle of material
frame-indifference. In the next lecture, we discuss the consequence of principle of material
frame-indifference on general constitutive relations of simple materials.
Example 1. Let τ be a stress response. Let ρ, v, L and D be density, velocity, velocity
gradient and rate of deformation, respectively. Let the constitutive relation of material be

τ = τ̂ (ρ, v, L).

If the constitutive relation follows the principle of material frame-indifference then the
constitutive relation reduces to
τ = τ̂ (ρ, D). (10)

Proof. Let F ∗ and F be fixed and moving frames. Then they are related by the Euclidean
transformation
χ∗ (X, t∗ ) = Q(t)χ(X, t) + c(t), (11)
3
Müller, I., “On the frame dependence of stress and heat flux”, Archive for Rational Mechanics and
Analysis, 45, pp. 241-250, 1972
4
Jog, C. S., “Foundations and applications of mechanics: continuum mechanics”, Narosa Publishing
House, New Delhi, 2007.

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

where t∗ = t − s, Q(t) represents relative rotation between frames and c(t) represents
translation.

Let us assume ρ, v and L be density, velocity, velocity gradient, respectively, in frame


of reference F ; ρ∗ , v ∗ and L∗ be density, velocity, velocity gradient, respectively, in frame
of reference F ∗ .
We also have the relations ρ∗ = ρ, L∗ = QLQT + Q̇QT and v ∗ = Qv + Q̇x + ċ (see
Lectures 31 and 32).
By applying the principle of material frame-indifference to the given constitutive rela-
tion, we get
τ̂ (ρ∗ , v ∗ , L∗ ) = Qτ̂ (ρ, v, L)QT .
Substituting ρ∗ = ρ, L∗ = QLQT + Q̇QT and v ∗ = Qv + Q̇x + ċ in the above equation,
we can write

τ̂ (ρ, Qv + Q̇x + ċ, QLQT + Q̇QT ) = Qτ̂ (ρ, v, L)QT . (12)

Since Q(t) is arbitrary, choosing Q = I, we get

τ̂ (ρ, v + ċ, L) = τ̂ (ρ, v, L).

This relation shows that the stress response cannot be a function of velocity as c(t)
is arbitrary. Let D and W be rate of deformation and spin tensors. Then we have
L = D + W . Of course D and W are symmetric and skewsymmetric part of velocity
gradient. Substitution of relation L = D + W in Eq. (12) yields

τ̂ (ρ, QDQT + QW QT + Q̇QT ) = Qτ̂ (ρ, L)QT . (13)

Note that the velocity v in the argument is removed as we have shown it is independent
of velocity. Since the orthogonal tensor Q(t) is arbitrary, we choose Q(t0 ) = I and
Q̇(t0 ) = −W . Evaluating the above equation at t = t0 , we get

τ̂ (ρ, D) = τ̂ (ρ, L).

Thus, the principle of material frame-indifference reduces the constitutive relation τ =


τ̂ (ρ, v, L) to τ = τ̂ (ρ, D).

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. Z. Martinec, Lecture Notes on Continuum Mechanics


(Link: http://geo.mff.cuni.cz/vyuka/Martinec-ContinuumMechanics.pdf)

3. I-S. Liu, Continuum Mechanics, 2002, Springer, Berlin.

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-5: Constitutive Relations


Lecture-35: Implications of
the Principle of Material Frame-Indifference and
the Material Symmetry

Here, we discuss the restrictions imposed by the principle of material frame-indifference


and material symmetry on simple materials.

Principle of material frame-indifference:


Recall the principle of material frame-indifference from previous lecture. Let F ∗ and F
be two frames of reference related by the Euclidean transformation

χ∗ (X, t∗ ) = Q(t)χ(X, t) + c(t), (1)

where t∗ = t − s, Q(t) represents relative rotation between frames and c(t) represents
translation.
Let G, g and φ be tensor, vector and scalar valued constitutive relations of continuum
body B observed by the frame F . Let G∗ , g ∗ and φ∗ be corresponding constitutive
relations of same body B observed by the frame of reference F ∗ . Then the principle
of material frame-indifference provides the following relation between these constitutive
relations.

G(l∗ ) = QG(l)QT ,
g(l∗ ) = Qg(l), (2)

φ(l ) = φ(l),

where l indicates the list of arguments in the response function.


Constitutive relations for simple material:
Recall the constitutive relations for simple material from previous lecture.

τ (X, t) = τ̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X),
p=0

q(X, t) = q̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X),
p=0

e(X, t) = ê (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X),
p=0

η(X, t) = η̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X).
p=0

We now apply the principle of material frame-indifference to the simple material.

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Implication of principle of material frame-indifference:


Applying the principle of material frame-indifference (i.e., Eq. (2)) to the constitutive
relations of simple material, we get

τ̂ (χ∗ (X, t∗ − p), F ∗ (X, t∗ − p), θ∗ (X, t∗ − p), g ∗ (X, t∗ − p), X)
p=0
 ∞ 
= Q(t) τ̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X) QT (t),
p=0

q̂ (χ∗ (X, t∗ − p), F ∗ (X, t∗ − p), θ∗ (X, t∗ − p), g ∗ (X, t∗ − p), X)
p=0
!

= Q(t) q̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X) ,
p=0

ê (χ∗ (X, t∗ − p), F ∗ (X, t∗ − p), θ∗ (X, t∗ − p), g ∗ (X, t∗ − p), X)
p=0

= ê (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X),
p=0

η̂ (χ∗ (X, t∗ − p), F ∗ (X ∗ , t − p), θ∗ (X, t∗ − p), g ∗ (X, t∗ − p), X)
p=0

= η̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X).
p=0

Using Eq. (1), we can write

χ∗ (X, t∗ ) = Q(t∗ + s)χ(X, t∗ + s) + c(t∗ + s)

Replacing the variable t∗ with t∗ − p, we obtain following relation

χ∗ (X, t∗ − p) = Q(t∗ − p + s)χ(X, t∗ − p + s) + c(t∗ − p + s)


= Q(t − p)χ(X, t − p) + c(t − p).

Taking material gradient on both sides of Eq. (1), yields

F (X, t∗ ) = Q(t)F (X, t)


= Q(t∗ + s)F (X, t∗ + s).

Consequently,

F (X, t∗ − p) = Q(t∗ + s − p)F (X, t∗ + s − p)


= Q(t − p)F (X, t − p). (3)

Since the temperature field θ is assumed to be frame-indifferent, i.e., θ∗ = θ, we get g ∗ = g


as it is material gradient. Substituting all arguments in above constitutive relations, we
can have

τ̂ (Q(t − p)χ(X, t − p) + c(t − p), Q(t − p)F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0
 ∞ 
= Q(t) τ̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X) QT (t),
p=0

q̂ (Q(t − p)χ(X, t − p) + c(t − p), Q(t − p)F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0
!

= Q(t) q̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X) ,
p=0

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics


ê (Q(t − p)χ(X, t − p) + c(t − p), Q(t − p)F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= ê (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X),
p=0

η̂ (Q(t − p)χ(X, t − p) + c(t − p), Q(t − p)F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= η̂ (χ(X, t − p), F (X, t − p), θ(X, t − p), g(X, t − p), X).
p=0

We now choose Q(t) = I ∀t ∈ [0, ∞) in Euclidean transformation and substitute in above


relations. Then arbitrariness of c(t) implies that constitutive relations must be inde-
pendent of motion. Therefore, the constitutive relations under the principle of material
frame-indifference reduces to

τ̂ (Q(t − p)F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0
 ∞ 
= Q(t) τ̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X) QT (t), (4)
p=0

q̂ (Q(t − p)F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0
!

= Q(t) q̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X) , (5)
p=0

ê (Q(t − p)F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= ê (F (X, t − p), θ(X, t − p), g(X, t − p), X), (6)
p=0

η̂ (Q(t − p)F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= η̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X). (7)
p=0

Reduction of constitutive relations using polar decomposition theorem:


We now recall the polar decomposition of deformation gradient from Lecture-18. Using
the polar decomposition, we can write F (X, t) = R(X, t)U (X, t), where R(X, t) is the
rotation tensor and U (X, t) is right stretch tensor. We choose Q(t − s) = RT (X, t − s)
as X is fixed then Eq. (4) yields

τ̂ (U (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0
 ∞ 
T
= R (X, t) τ̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X) R(X, t).
p=0

This relation can be written as



τ̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0
 ∞ 
= R(X, t) τ̂ (U (X, t − p), θ(X, t − p), g(X, t − p), X) RT (X, t). (8)
p=0

Since the stretch tensor, U (X, t) is invertible tensor, we can have R(X, t) = F (X, t)U −1 (X, t).
Substituting this relation in Eq. (8), we get

τ̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X) =
p=0
 ∞ 
F (X, t)U −1 (X, t) τ̂ (U (X, t − p), θ(X, t − p), g(X, t − p), X) U −1 (X, t)F T (X, t). (9)
p=0

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

We now recall the relation between the second Piola-Kirchhoff stress tensor and the
Cauchy stress tensor from Lecture-28, i.e., S = det(F )F −1 τ F −T . Hence, we now de-
fine the constitutive response function for second Piola-Kirchhoff stress

Ŝ (F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= det(F (X, t))F −1 (X, t) τ̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X)F −T (X, t). (10)
p=0

Substituting Eq. (9) and the relation det(F (X, t)) = det(U (X, t)) in Eq. (10), we get

Ŝ (F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0
∞ 
= det(U (X, t))U −1 (X, t) τ̂ (U (X, t − p), θ(X, t − p), g(X, t − p), X) U −1 (X, t), (11)
p=0
q
Since the right stretch tensor U (X, t) = F T (X, t)F (X, t) and the right Cauchy-Green
strain C(X, t) = F T (X, t)F (X, t), we can define a new function
∞ ∞
S̃ (C(X, t−p), θ(X, t−p), g(X, t−p), X) = Ŝ (F (X, t−p), θ(X, t−p), g(X, t−p), X)
p=0 p=0

so that it is a function of C. Using Eq. (11), the new function can explicitly be written
as

S̃ (C(X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= det(U (X, t))U −1 (X, t) τ̂ (U (X, t − p), θ(X, t − p), g(X, t − p), X) U −1 (X, t), (12)
p=0
q
where U (X, t) = C(X, t).

Similar to previous procedure, using polar decomposition and the principle of material
frame-indifference, heat flux vector shown in Eq. (5) can be written as

q̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= R(X, t) q̂ (U (X, t − p), θ(X, t − p), g(X, t − p), X), (13)
p=0

Since R(X, t) = F (X, t)U −1 (X, t), we can have



q̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= F (X, t)U −1 (X, t) q̂ (U (X, t − p), θ(X, t − p), g(X, t − p), X), (14)
p=0

Recall transformation of heat flux into Lagrangian description (see Lecture-29), i.e.,
q 0 (X, t) = det(F )F −1 q(x, t). Therefore, we define a new vector valued function

q̃ 0 (C(X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= det(U (X, t))U −1 (X, t) q̂ (U (X, t − p), θ(X, t − p), g(X, t − p), X), (15)
p=0

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

q
where U (X, t) = C(X, t).
Applying similar steps to internal energy and entropy response function i.e., to Eqs.
(6) and (7), we get
∞ ∞
ê (F (X, t−p), θ(X, t−p), g(X, t−p), X) = ê (U (X, t−p), θ(X, t−p), g(X, t−p), X),
p=0 p=0
(16)
∞ ∞
η̂ (F (X, t−p), θ(X, t−p), g(X, t−p), X) = η̂ (U (X, t−p), θ(X, t−p), g(X, t−p), X).
p=0 p=0
(17)
In conclusion all constitutive relations can be expressed as right Cauchy-Green strain. In
particular, stress response is related to strain tensor and its history. Hence, the consti-
tutive relation in solid mechanics is known as stress-strain relation. We now discuss the
restrictions imposed by the material symmetry on constitutive relation.
Material symmetry:
Most materials exhibits some material symmetries, i.e., material properties about some
planes are symmetric. The constitutive relations must reflect the symmetry possessed by
the material. The symmetries can be two types: (i) rotational symmetry (ii) reflectional
symmetry. We discuss these symmetries through examples.
Rotational symmetry:
Rotation of reference configuration leaves the indistinguishable deformed configuration
for some angle of rotations. We now present two examples to understand the rotational
symmetry. The angle of rotation corresponding to symmetry transformation is π in first
π
example while in second example.
2
Example-1 (square plate under bi-axial stretch):
Consider a square plate of homogeneous, fiber reinforce, composite material with a size
of a × a as shown in Fig. (1a). Let the fibers are aligned along AB and we assume that
the stiffness is more along the direction of fiber (i.e., along AB) than that of orthogonal
direction (i.e., along BC) as shown in Fig. (1d). We apply a uniform traction tu on
edges AD and BC as shown in Fig. (1b) and let the elongation along the load be ∆ax1 .
Similarly, we apply the uniform traction tu on edges AB and CD and let the elongation
along loading direction be ∆ay2 (see Fig. (1c)). Clearly, the elongation ∆ay2 is more than
∆ax1 as the stiffness along BC is less than AB.
We now consider another loading of bi-axial traction on the reference configuration.
Let the traction in both the directions be equal as shown in Fig. (2). Since the stiffness
is more along the fibers (see distribution of stiffness in Fig. (1d)), the plate deforms less
along fiber direction. If we mark a circle in the center of plate, it deforms into an elliptical
shape as shown in Fig. (2). It can be noted that the rotation of reference configuration by
an angle of π also leaves the same deformed configuration under same loading conditions
as shown in Fig. (2). In other words, if we do not mark A, B, C and D for the corners
then it is not possible to recognize the reference configuration by looking at deformed
configuration. Thus, the angle of symmetry for this problem is π.
Example-2 (circular plate under hydrostatic pressure):
Consider a circular plate of homogeneous material with stiffness distribution that is shown

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

B A B A

C D C D
(a) (b)

A
B

C
D
(d)
(c)

Figure 1: Square plate of homogeneous, fiber reinforce, composite material: (a) reference
configuration (b) loading along fiber direction (b) loading along orthogonal to fibers (d)
stiffness distribution of plate

B A

C D

Rotation of
reference
configuration

D C
Deformed
configuration

A B

Figure 2: Material symmetry for an angle rotation π

in Fig. (3). The variation in stiffness is responsible for the deformed shape depicted
in Fig. (3) under hydrostatic pressure. It can be observed that rotation of reference
configuration by an angle of π/2 then applying the hydrostatic pressure would also assume
same deformed shape as there is symmetry in stiffness distribution. Similar to earlier
example, if there are no marking on the reference configuration then it is not possible to
distinguish the orientation of reference configuration by looking at deformed configuration.

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

Clearly, the angle of symmetries are integer multiples of π/2.

A
A
B H B H
C G C G
D F
Rotation of E F
D
reference E
configuration
Stiffness distribution

G
H F
A E
B D
C
Deformed configuration

Figure 3: Material symmetry for an angle of rotation π/2

Reflectional symmetry:
We present an example to explain the reflectional symmetry. Let us consider a circular
plate with distribution of stiffness shown in Fig. (4). It can be observed that the reference
configuration can be deformed into a star like shape as shown in Fig. (4) under a hydro-
static pressure. Clearly, the stiffness distribution is symmetric about x2 -axis. Therefore,
if the reference configuration were to make reflection about the axis then the deformation
is same under the hydrostatic pressure as shown in Fig. (4).

A A
B H
B H
C G
C G
D F
Reflection of E
D E F
reference
configuration Stiffness distribution

A
H B
G C
F D
E
Deformed configuration

Figure 4: Example for a reflectional material symmetry

We note that both rotational and reflectional transformations belong to orthogonal

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

tensors. Therefore, we deal with group of orthogonal transformations in the material


symmetry.

The notion of a group:


A set G with a binary operation ‘*’ defined over every pair of elements is said to be a
group if it obeys the following axioms.

• Closed under binary operation: For all Q1 , Q2 ∈ G, the element generated by the
operation should also be in G, i.e., Q1 ∗ Q2 ∈ G.

• Associativity: For all Q1 , Q2 , Q3 ∈ G,

(Q1 ∗ Q2 ) ∗ Q3 = Q1 ∗ (Q2 ∗ Q3 ).

• Existence of identity element: There exist a unique identity element I ∈ G such


that
Q ∗ I = I ∗ Q = Q, ∀Q ∈ G.

• Existence of inverse element: For each element Q ∈ G, there exist Q−1 ∈ G such
that
Q−1 ∗ Q = Q ∗ Q−1 = I.

Orthogonal group:
Collect all orthogonal tensors and denote with Orth. In other words,

Orth = {Q | Q is orthogonal tensor}.

We now define usual product between tensors as binary operation between tensors. Then,
clearly, the product of any two orthogonal tensors is again an orthogonal tensor and also
usual product is associative. We also note that the identity tensor also belongs to the
set. In addition, by definition of orthogonal tensor, the transpose operator is the inverse
and also orthogonal tensor. Thus, the set of all orthogonal tensor follows all the axioms
of group and it is known as orthogonal group.
Material symmetry group:
A transformation defined over a reference configuration is called as material symmetry
transformation if it does not alter the deformed configuration. The collection of these
material symmetry transformations forms a group and this group is known as material
symmetry group. For example the material symmetry group is a subset of orthogonal
group for solids. This fact can be observed from the previous examples. Materials are
classified based on these symmetry groups. We now define the fluid and solid based on
these symmetry groups.
The fluid and the solid:
Fluid:
Take some fluid (water) into one test-tube and let it be a reference configuration. We now
pour the fluid from one test-tube into another test-tube of same size. We can see that
the fluid undergoes large deformation, still the fluid in second test-tube starts behaving
exactly similar to the previous one. The reason for this behavior is that the fluid does not

Joint initiative of IITs and IISc – Funded by MHRD 8


NPTEL – Mechanical Engineering – Continuum Mechanics

have any preferred reference configuration. The only similarity in both configuration is
that the density is same at similar locations. In other words, the transformation is density
preserving. Thus, the material symmetry group for fluid is unimoduluar transformations,
i.e., the collection of tensors those have unit determinant. Clearly, the fluid is an isotropic
material.
Solid:
Let us consider a bar and apply loads to deform the bar. We now consider the deformed
configuration as new reference configuration and apply similar loads as that of previous
case. But then the deformation response would be different from the previous case. Thus,
the solids posses a preferred reference configurations unlike fluids. The symmetry group
of solids is subgroup of orthogonal group. Furthermore, if the solid is isotropic then the
symmetry group is proper orthogonal group, i.e., group of rotational tensors.
We now define the material based on degree of symmetry.
Isotropic material:
A material whose symmetry group is orthogonal group is called as isotropic material.
Clearly, the orthogonal group includes all rotations and reflections. A material that
consists of rotation tensors and not reflections as symmetry group is known as hemitropic
material. In our discussion, we don’t distinguish the difference between isotropic and
hemitropic material. We refer these two materials as isotropic material.

Anisotropic material:
A material that has fewer symmetries than isotropic is called anisotropic material.
Transversely isotropic material:
A material that has all rotation tensors about a fixed axis as symmetry group is known
as transversely isotropic material. Of course this is a special case of anisotropic material.
Orthotropic material:
A material that has reflectional symmetry about three mutually orthogonal planes is
known as orthotropic material. Of course this is another special case of anisotropic ma-
terial.
Special class of materials:
In many practical problems, we make the following assumption to simplify the problem.

Homogeneous material:
The continuum body is said to be homogeneous if the constitutive relations are not ex-
plicitly dependent on material point, i.e, material properties are same at all points. In
fact this assumption has been made in previous examples to explain material symmetry.

Elastic material:
The material is said to be elastic if the constitutive relation is independent of the history
of deformation and temperature.
We now see that the material symmetry imposes certain restrictions on constitutive
relations along with principle of material frame-indifference.

Joint initiative of IITs and IISc – Funded by MHRD 9


NPTEL – Mechanical Engineering – Continuum Mechanics

Consequence of material symmetry on constitutive relations:


Let H be an orthogonal tensor that belongs to material symmetry group. Then the
following constitutive response relations are defined to account for material symmetry.

τ̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= τ̂ (F (X, t − p)H, θ(X, t − p), H T g(X, t − p), X),
p=0

q̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= q̂ (F (X, t − p)H, θ(X, t − p), H T g(X, t − p), X),
p=0

ê (F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= ê (F (X, t − p)H, θ(X, t − p), H T g(X, t − p), X),
p=0


η̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= η̂ (F (X, t − p)H, θ(X, t − p), H T g(X, t − p), X).
p=0

The determinant of tensor QF QT is positive as det(F ) > 0 and Q is orthogonal


tensor.. Therefore, replacing F with QF QT and also g with Qg in above equations, we
get

τ̂ (QF (X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0

= τ̂ (QF (X, t − p)QT H, θ(X, t − p), H T Qg(X, t − p), X),
p=0

q̂ (QF (X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0

= q̂ (QF (X, t − p)QT H, θ(X, t − p), H T Qg(X, t − p), X),
p=0

ê (QF (X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0

= ê (QF (X, t − p)QT H, θ(X, t − p), H T Qg(X, t − p), X),
p=0

η̂ (QF (X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0

= η̂ (QF (X, t − p)QT H, θ(X, t − p), H T Qg(X, t − p), X).
p=0

Let us choose the orthogonal tensor Q such that Q = H, i.e., Q is in symmetry group.

Joint initiative of IITs and IISc – Funded by MHRD 10


NPTEL – Mechanical Engineering – Continuum Mechanics

Since QT = Q−1 , preceding equations transformed to



τ̂ (QF (X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0

= τ̂ (QF (X, t − p), θ(X, t − p), g(X, t − p), X),
p=0

q̂ (QF (X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0

= q̂ (QF (X, t − p), θ(X, t − p), g(X, t − p), X),
p=0

ê (QF (X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0

= ê (QF (X, t − p), θ(X, t − p), g(X, t − p), X),
p=0

η̂ (QF (X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0

= η̂ (QF (X, t − p), θ(X, t − p), g(X, t − p), X).
p=0

Applying the principle of material frame-indifference, i.e., Eqs. (4), (5) (6) and (7),
to the right hand side of above equations, we get the following final form of constitutive
equations.

τ̂ (QF (X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0
∞ 
=Q τ̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X) QT , (18)
p=0

q̂ (QF (X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0

= Q q̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X), (19)
p=0

ê (QF (X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0

= ê (F (X, t − p), θ(X, t − p), g(X, t − p), X), (20)
p=0

η̂ (QF (X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0

= η̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X). (21)
p=0

Clearly, the orthogonal tensor Q belongs to the material symmetry group. Till now
there are no assumptions made over the material symmetry. We now make assumption
of isotropic material and show that the constitutive relations indeed reduces to isotropic
functions.
Constitutive relations for isotropic materials:
As noted earlier, full orthogonal group is symmetry group for isotropic materials. There-
fore, Q in above equations (i.e., Eqs. (18), (19), (20), and (21)) is arbitrary for isotropic
material.
We now recall that the constitutive equations can be represented as function of right
stretch tensor and thereby the right Green strain tensor as shown in Eqs .(9), (14), (16)

Joint initiative of IITs and IISc – Funded by MHRD 11


NPTEL – Mechanical Engineering – Continuum Mechanics

and (17). In other words, there exist constitutive response functions such that

τ̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0
∞ 
= F (X, t) τ̃ (U (X, t − p), θ(X, t − p), g(X, t − p), X) F T (X, t), (22)
p=0

q̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= F (X, t) q̃ (C(X, t − p), θ(X, t − p), g(X, t − p), X), (23)
p=0

ê (F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= ẽ (C(X, t − p), θ(X, t − p), g(X, t − p), X), (24)
p=0

η̂ (F (X, t − p), θ(X, t − p), g(X, t − p), X)
p=0

= η̃ (C(X, t − p), θ(X, t − p), g(X, t − p), X), (25)
p=0

as U = C. Applying these relations to above constitutive responses shown in Eqs. (18)
and (19), we get
∞ 
QF QT τ̃ (QC(X, t − p)QT , θ(X, t − p), Qg(X, t − p), X) QF T QT
p=0
∞ 
= QF τ̃ (C(X, t − p), θ(X, t − p), g(X, t − p), X) F T QT ,
p=0
!

T T
QF Q q̃ (QC(X, t − p)Q , θ(X, t − p), Qg(X, t − p), X)
p=0
!

= QF q̃ (C(X, t − p), θ(X, t − p), g(X, t − p), X) .
p=0

Since Q and F are invertible tensors, we get the following form of stress and heat flux
constitutive responses for an isotropic material.

τ̃ (QC(X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0
∞ 
=Q τ̃ (C(X, t − p), θ(X, t − p), g(X, t − p), X) QT , (26)
p=0

q̃ (QC(X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0

= Q q̃ (C(X, t − p), θ(X, t − p), g(X, t − p), X). (27)
p=0

Applying Eqs. (24) and (25) to the constitutive responses (20) and (21), we get the
following form of constitutive responses for internal energy and entropy.

ẽ (QC(X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0

= ẽ (C(X, t − p), θ(X, t − p), g(X, t − p), X), (28)
p=0

η̃ (QC(X, t − p)QT , θ(X, t − p), Qg(X, t − p), X)
p=0

= η̃ (C(X, t − p), θ(X, t − p), g(X, t − p), X). (29)
p=0

Joint initiative of IITs and IISc – Funded by MHRD 12


NPTEL – Mechanical Engineering – Continuum Mechanics

Recall isotropic functions from Lecture-12. Clearly, the constitutive relations for isotropic
material are isotropic functions. More detailed account of material symmetry and its
rigorous treatment can be found in Jog (2007) or in Liu (2002).

In forthcoming lectures, we specialize the constitutive relations to isotropic elastic solids


and classical fluids. We also discuss the implications of second law of thermodynamics.

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. Z. Martinec, Lecture Notes on Continuum Mechanics


(Link: http://geo.mff.cuni.cz/vyuka/Martinec-ContinuumMechanics.pdf)

3. E. B. Tadmor, R. E. Miller and R. S. Elliott, Continuum Mechanics and Thermo-


dynamics from Fundamental Concepts to Governing Equations, 2012, Cambridge
University Press, UK.

4. I-S. Liu, Continuum Mechanics, 2002, Springer, Berlin.

5. M. E. Gurtin, E. Fried and L. Anand, The Mechanics and Thermodynamics of


Continua, 2010, Cambridge University Press, New York.

Joint initiative of IITs and IISc – Funded by MHRD 13


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-5: Constitutive Relations


Lecture-36: Thermomechanics of Elastic Solids

In this lecture, we discuss the restrictions imposed by the second law of thermodynamics
on constitutive relations of elastic solids. As mentioned in Lecture-35, the constitutive
response of elastic material does not depend on thermomechanical history.
Constitutive relations in Lagrangian framework:
It is convenient to work in Lagrangian framework for the solid mechanics. Therefore, we
consider the constitutive relations that are defined in material description, rather than
the spatial description. The constitutive equations for the first Piola-Kirchhoff stress,
internal energy and entropy are given by

T = T̂ (X, F , g 0 , θ0 ), e0 = ê0 (X, F , g 0 , θ0 ), and η 0 = η̂ 0 (X, F , g 0 , θ0 ), (1)

where X is material particle, F is deformation gradient, θ0 is material description of


temperature and g 0 = ∇X θ0 . Since the material is elastic, the thermomechanical history
is not considered in constitutive relations. We now show that the constitutive relations
are indeed independent of temperature gradient.

Application of the second law of thermodynamics (Clausius-Duhem inequality) to con-


stitutive relations yields
∂e0 ∂η 0 1 q0 · g0
− θ0 − T : Ḟ + ≤ 0, (2)
∂t ∂t ρ0 ρ0 θ 0
where ρ0 is density of reference configuration and Ḟ is the material time derivative of
deformation gradient. Substituting constitutive relations, we get
! ! !
∂e0 0 ∂η
0
1 ∂e0 0 ∂η
0
0 ∂e0 0 ∂η
0
0 q0 · g0
−θ − T : Ḟ + −θ · ġ + −θ θ̇ + ≤ 0. (3)
∂F ∂F ρ0 ∂g 0 ∂g 0 ∂θ0 ∂θ0 ρ0 θ0

Since Ḟ , ġ 0 , and θ̇0 are arbitrary, we have


∂e0 0 ∂η
0
1
−θ − T = O (4)
∂F ∂F ρ0
0 0
∂e 0 ∂η
−θ = 0 (5)
∂g 0 ∂g 0
∂e0 0 ∂η
0
−θ = 0. (6)
∂θ0 ∂θ0
Taking derivative of Eq. (5) with respect to θ0 , we get
∂ 2 e0 2 0
0 ∂ η ∂η 0
−θ − = 0. (7)
∂θ0 ∂g 0 ∂θ0 ∂g 0 ∂g 0

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Taking differentiation of Eq. (6) with respect to g 0 , we get

∂ 2 e0 2 0
0 ∂ η
− θ = 0. (8)
∂g 0 ∂θ0 ∂g 0 ∂θ0

Substituting Eq. (8) in Eq. (7), we obtain

∂η 0
= 0. (9)
∂g 0

Thus, the entropy η 0 is independent of gradient of temperature g 0 .


Substituting Eq. (9) in Eq. (5), we get

∂e0
= 0. (10)
∂g 0

Thus, the internal energy e0 is also independent of gradient of temperature.

We can write Eq. (4) as !


∂e0 ∂η 0
T = ρ0 − θ0 . (11)
∂F ∂F
The first Piola-Kirchhoff stress T is also independent of g 0 as it is function of e0 and η 0 .

We now define Helmholtz free energy ψ 0 by

ψ 0 = e0 − η 0 θ 0 . (12)

Substituting this relation in above equation, we get

∂ψ 0
T = ρ0 . (13)
∂F
Taking the derivative of free energy ψ 0 with respect to temperature θ0 and substituting
in Eq. (6), yields
∂ψ 0
η0 = − 0 . (14)
∂θ
Substituting Eqs. (4), (5) and (6) in Clausius-Duhem inequality, i.e., in Eq. (3), we get

q0 · g0 ≤ 0 (15)

This shows that heat flux flows in the negative gradient direction, or, alternatively, heat
flows from hot to cold regions. This is a classical statement of second law of thermody-
namics.
Summary of the relations:
In summary, we have the following three observations:

• The constitutive relations of first Piola-Kirchhoff stress, internal energy and entropy
are independent of the gradient of temperature

T = T̂ (X, F , θ0 ), e0 = ê0 (X, F , θ0 ), and η 0 = η̂ 0 (X, F , θ0 ). (16)

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

• There is a free energy ψ 0 = e0 − η 0 θ0 such that the first Piola-Kirchhoff stress and
entropy are given by
∂ψ 0 ∂ψ 0
T = ρ0 , and η 0 = − 0 .
∂F ∂θ
• Heat flows from hot region to cold region, i.e.,

q 0 · g 0 ≤ 0.

We now discuss two special cases of the general constitutive relations: (i) The heat con-
duction in solids (ii) The deformation of hyperelastic solids.
Heat conduction in rigid solids:
Let the heat flux vector q 0 = q̂ 0 (X, F , θ0 , g 0 ). Then the heat conductivity tensor is
defined by
0 ∂ q̂i0 0

0
[K(F , θ )]ij = − 0 (X, F , θ , g ) . (17)
∂gj 0
g =0

Using Taylor’s expansion of q̂ 0 about g 0 = 0, we get

q̂ 0 (g 0 ) = q̂ 0 (0) − Kg 0 + o(|g 0 |). (18)

If g 0 → 0 then o(|g 0 |) can be neglected. Substitution of Eq. (18) in Eq. (15) yields

q̂ 0 (0) · g 0 − g 0 · Kg 0 + o(|g 0 |2 ) ≤ 0. (19)

This inequality holds if and only if q̂ 0 (0) = 0 and g 0 · Kg 0 ≥ 0 for every g 0 .


In conclusion, we have q̂ 0 (F , θ0 , 0) = 0 and the conductivity tensor K(F , θ0 ) is positive
semi-definite. We also note that the conductivity tensor K is not necessarily symmetric.
The result of first approximation, we have

q 0 (X, F , θ0 , g 0 ) = −K(X, F , θ0 )g 0 . (20)

This relation is known as Fourier law of heat conduction.

We now derive the governing equation of heat conduction for rigid solids, i.e., for the
solids with negligible deformation.
Taking the material time derivative of Helmholtz free energy function ψ 0 (X, F , θ0 )
(see Eq. (12)), we get
∂ψ 0 ∂ψ 0 ∂ψ 0
= : Ḟ + 0 θ̇0 . (21)
∂t ∂F ∂θ
Substituting Eqs. (13) and (14), we obtain

∂ψ 0 1
= T : Ḟ − η 0 θ̇0 . (22)
∂t ρ0

Using Eq. (12), we can write the material time derivative of free energy

∂ψ 0 ∂e0 ∂η 0 0 ∂θ0
= − θ − η0 . (23)
∂t ∂t ∂t ∂t

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

We can have the following relation by equating total time derivative of free energy in Eqs.
(22) and (23).
∂e0 ∂η 0
ρ0 = T : Ḟ + ρ0 θ0 . (24)
∂t ∂t
Upon substitution of Eq. (24) in the first law of thermodynamics (i.e., in Eq. (11) of
Lecture-29), we get
∂η 0
ρ0 θ 0 = ρ0 Q0h − ∇X · q 0 . (25)
∂t
We now define a specific heat capacity c by
∂ 0
c(X, F , θ0 ) = 0
e (X, F , θ0 ). (26)
∂θ
Combining Eqs. (12) and (14), we can have the following expression for specific heat
capacity.
∂η 0
c = θ0 0 . (27)
∂θ
If the deformations are negligible (i.e., solid is rigid) then Eq. (16)3 give rise to the relation
∂η 0 ∂η 0 ∂θ0
= 0 . (28)
∂t ∂θ ∂t
Substitution of Eqs. (28) and (20) in Eq. (25), we can obtain the following heat conduction
equation.
∂θ0  
ρ0 c = ∇X · Kg 0 + ρ0 Q0h . (29)
∂t
Though the conductivity tensor K is not necessarily symmetric, in practice we assume
that the conductivity tensor is symmetric. This assumption is known as Onsager reciprocal
relations. If the medium (continuum body) is isotropic and homogeneous then the heat
conduction equation can be written as
∂θ0
ρ0 c = κ∇X · g 0 + ρ0 Q0h , (30)
∂t
where κ is conductivity constant, i.e., K = κI.
Before discussing hyperelastic solids, we discuss the relation between second Piola-
Kirchhoff stress and Helmholtz free energy.

Second Piola-Kirchhoff stress:


Equations (9) and (10) shows that both entropy (η 0 ) and internal energy (e0 ) are not
function of temperature gradient. Hence, Constitutive response of internal energy and
entropy are function of deformation gradient and temperature as shown in Eq. (16).
Since free energy ψ 0 = e0 − η 0 θ0 , the constitutive response can be written as
ψ 0 = ψ̂ 0 (X, F , θ0 ). (31)
The constitutive relation for free energy can be expressed as function of right Cauchy-
Green strain C instead of deformation gradient F as it follows the principle of material
1
frame-indifference. Since Green strain tensor E = (C − I), we can have a function ψ̃ 0
2
such that the Helmholtz free energy
ψ 0 = ψ̃ 0 (X, E, θ0 ), (32)
where E is Green strain tensor.

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Problem 1. If the first Piola-Kirchhoff stress tensor T is defined by the relation

∂ ψ̂ 0
T = ρ0 , (33)
∂F
then the second Piola-Kirchhoff stress S can be obtained by

∂ ψ̃ 0
S = ρ0 , (34)
∂E

Proof. Consider the direction derivative of ψ̂ 0 along U , i.e.,

∂ ψ̂ 0
: U = DF ψ̂ 0 (F , θ0 )[U ]
∂F
= DF ψ̃ 0 (E, θ0 )[U ]
= DE ψ̃ 0 (E, θ0 )[DF (E)[U ]] (Using chain rule)
1
= DE ψ̃ 0 (E, θ0 )[F T U + U T F ]
2
1 ∂ ψ̃ 0  T 
= : F U + UTF
2 ∂E
 " #T 
F  ∂ ψ̃ 0 ∂ ψ̃ 0 
= + :U (Since A : (B T C) = (CAT ) : B)
2 ∂E ∂E
! !
∂ ψ̃ 0 ∂ ψ̃ 0
= F :U Since is symmetric
∂E ∂E
Therefore, the first Piola-Kirchhoff stress

∂ ψ̂ 0 ∂ ψ̃ 0
T = ρ0 = ρ0 F (35)
∂F ∂E
Recall the following relation between first and second Piola-Kirchhoff stress tensors from
Lecture-28.
S = F −1 T . (36)
Substitution of Eq. (35) in Eq. (36) give rise to Eq. (34). This completes the proof.

Since ψ̂ 0 and ψ̃ 0 are having same dependency on θ0 , Eq. (14) implies the following
relation for entropy.
∂ ψ̃ 0
η 0 = η̃ 0 (X, E, θ0 ) = − 0 . (37)
∂θ
In summary, we have the following relations for the second Piola-Kirchhoff stress and
entropy
∂ ψ̃ 0 ∂ ψ̃ 0
S = ρ0 and η 0 = − 0 .
∂E ∂θ
We now specialize these relations to hyperelastic materials.

Hyperelastic solids:
The hyperelastic solid is defined as a elastic solid with a stored energy function called
strain energy density function W (X, E) such that
∂W
S(X, E) = . (38)
∂E

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

We now show that thermoelastic solid is hyperelastic solid in two special cases. These
special case are (i) isothermal deformation (ii) isentropic deformation. In former case the
temperature is constant during the process of deformation and in later case the entropy
is constant during deformation.
Isothermal case:
Since θ0 is constant in isothermal deformation, the thermoelastic solid is clearly a hy-
perelastic solid with W (X, E) = ρ0 ψ̃ 0 (X, E, θ0 ). Using Eq. (34), we can obtain second
Piola-Kirchhoff stress
∂ ψ̃ 0 ∂W
S = ρ0 = . (39)
∂E ∂E
Isentropic case:
In isentropic deformation η 0 is constant. Taking derivative of Helmholtz free energy with
respect to temperature θ0 (see Eq. (12)), yields

∂ψ 0 ∂e0
= − η0. (40)
∂θ0 ∂θ0
Combining Eqs.(37) and (40), we get

∂e0
= 0. (41)
∂θ0
Clearly, the internal energy e0 is independent of temperature. Thus, the stored energy
function or strain energy density function W (X, E) = ρ0 e0 for isentropic process.
Using the principle of material frame-indifference, we can write the internal energy

e0 = ê0 (X, F , θ0 ) = ẽ0 (X, E, θ0 ). (42)

The second Piola-Kirchhoff stress for isentropic case can be written as

∂ ψ̃ 0 ∂ẽ0 ∂W
S = ρ0 = ρ0 = (43)
∂E ∂E ∂E
In conclusion, either isothermal or isentropic cases the thermoelastic solid can be consid-
ered as hyperelastic solid.
Isotropic hyperelastic solid:
From the previous discussion the stress response (second Piola-Kirchhoff stress) for the
hyperelastic solid can be written as

S = S̃(X, E) (44)

For the isotropic solid, using Rivlin-Ericksen representation theorem (see Lecture-12), we
can have
S = α(X, IE )I + γ(X, IE )E + β(X, IE )E 2 , (45)
where IE indicates three principal invariants of E. This is the most general stress-strain
relation for the isotropic hyperelastic solids. Linearized version of this model is known as
St Venant-Kirchhoff model.

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

St Venant-Kirchhoff model:
The linear relation between second Piola-Kirchhoff stress and Green strain is known as
St Venant-Kirchhoff model. The linearization of Eq. (45) yields

S = λ(tr E)I + 2µE, (46)

where λ and µ are Lamé constants. The stored energy or strain energy density function
for St Venant-Kirchhoff model can be written as
λ
W (E) = (tr E)2 + µ(tr E 2 ). (47)
2
Limitation of the model: We expect the strain energy density W (X, E) → ∞ when the
material is compressed to a point, i.e., det(F ) → 0. However, this is not true with the St
Venant-Kirchhoff model. Thus, this model is good approximation when strains are small.
We now discuss few other common models for elastic materials such that the shortcoming
in St Venant-Kirchhoff model is accounted.
Other common models:
Neo-Hookean model:
The strain energy density
q 
W (E) = a tr(I + 2E) + Γ det(I + 2E) , (48)
q
where Γ(s) = c s2 − d ln(s) and a, c, d > 0. We note that det(F ) = det(F T F ) =
q
det(I + 2E).
Mooney-Rivlin model:
The strain energy density
q    q 
−1
W (E) = a tr(I + 2E) + b det(I + 2E) tr (I + 2E) +Γ det(I + 2E) , (49)

where Γ(s) = c s2 − d ln(s) and a, b, c, d > 0.


Ogden model:
The strain energy density
m   n q βi /2  
αi /2
tr (I + 2E)−βi /2
X X
W (E) = ai tr (I + 2E) + bj det(I + 2E)
i=1 j=1
q 
+Γ det(I + 2E) , (50)

where Γ(s) = c s2 − d ln(s), c, d > 0, ai , bj > 0, αi , βj ≥ 1, and m, n ≥ 1

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. Z. Martinec, Lecture Notes on Continuum Mechanics


(Link: http://geo.mff.cuni.cz/vyuka/Martinec-ContinuumMechanics.pdf)

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-5: Constitutive Relations


Lecture-37: Linear Elasticity

In this lecture, we discuss the theory of linear elasticity. Though there are strong assump-
tions in this theory, due to simplicity, it is widely used for predicting the behaviour of
solids.
Kinematics:
We want to have a strain measure that is linear function of displacements. In order to
obtain such a linear strain, we assume that the displacement gradient ∇X u is small.
Therefore, the quadratic term that appears in the Green strain tensor (see Eq. (6) in
Lecture-20) can be neglected. Consequently, we have a linear strain measure
1 
= ∇X u + (∇X u)T .
2
This linear strain is also known as small strain tensor. In addition to small displacement
gradients, if the displacements are assumed to be small then the linearization of Almansi
strain tensor also coincides with the small strain tensor. In other words, if displacement
gradients and displacements are small then we have the linear strain
1  1 
= ∇X u + (∇X u)T ≈ ∇x u + (∇x u)T
2 2
We note that there is no difference between material gradient and spatial gradient in
linear elasticity. Therefore, we can write the small strain tensor
1 
= ∇u + (∇u)T . (1)
2

Characterization of rigid motion under small strain measure:


The displacement field u represents (infinitesimal) rigid motion (i.e., under the assumption
of displacement gradients are small) if and only if
u = W X + c, (2)
where W is constant skewsymmetric tensor and c is constant vector. An alternative form
of rigid displacements can be represented by
u = ω × X + c, (3)
where ω is the axial vector of W .

We now show that Eq. (3) represents the rigid motion with the assumptions of small
displacements and displacement gradients. Recall the following equation for rigid motion
from Lecture-18.
x = Q(t)X + c(t). (4)

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

We note that Q(0) = I. Using Taylor’s expansion of Q(t) about t = 0, we get

Q(t) = Q(0) + Q̇|t=0 t + o(t).

It is easy to see Q̇ is skewsymmetric tensor as Q̇QT is skewsymmetric and Q(0) = I. Let


us denote Ŵ ≡ Q̇|t=0 . Then Q(t) can be written as

Q(t) = I + Ŵ t + o(t). (5)

Let W = Ŵ t. Then for small t we have the following relation from Eqs. (4) and (5).

x − X = W X + c. (6)

Since displacement field u = x − X, we get the result stated in Eq. (2). It is easy to see
that W and c can be function of X but they are not function of time t.
As defined in Eq. (1), six components of strain tensor depends on three components of
displacements. Hence, a unique strain field can be obtained for a give displacement field.
But for a given strain components, the displacement field is not necessarily unique and
compatible. In order to ensure a single valued displacement field, the given strain field
should follow compatibility conditions discussed below.
Compatibility conditions:
The strain field  should satisfy the following condition known as compatibility equations
to ensure single valued displacement field.

∇ × (∇ × ) = O. (7)

In indicial notation
∂ 2 nq
(∇ × (∇ × ))ij = imn jpq = 0. (8)
∂Xm ∂Xp
Clearly, there are six equations as the strain tensor is symmetric. We now show that every
displacement field satisfies the compatibility conditions.

Let u be a displacement field. Then the strain field is defined by


1 
= ∇u + (∇u)T .
2
Taking curl of curl of strain field, we get
1   
∇ × (∇ × ) = ∇ × ∇ × ∇u + (∇u)T
2
1 1  
= ∇ × (∇ × (∇u)) + ∇ × ∇ × (∇u)T
2 2
1  
= ∇ × ∇ × (∇u)T (Since ∇ × (∇u) = O)
2
1  
= ∇ × (∇(∇ × u)) Since ∇ × (∇u)T = ∇(∇ × u)
2
= O (Since ∇ × (∇v) = O, ∀v ∈ V)

Thus, for a given displacement field the compatibility conditions are automatically satis-
fied. On the other hand, for a give strain field the compatibility ensures the single valued
displacement field.

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

We now state two important properties of curl of curl of strain tensor which will be
used for simplifying the compatibility conditions.

Properties of curl of curl of strain field:

• The tensor ∇ × (∇ × ) is symmetric as the strain field is symmetric, i.e.,

(∇ × (∇ × ))T = ∇ × (∇ × ). (9)

• It follows from previous proof that the trace of tensor ∇ × (∇ × ) is zero, i.e.,

tr (∇ × (∇ × )) = 0, (10)

or, alternatively,
(∇ × (∇ × ))ii = 0. (11)
Using  − δ identity (see Lecture-3) and Eq. (8), we get

∇2 (tr ) − ∇ · (∇ · ) = 0. (12)

Problem 1. The compatibility stated in Eq. (7), i.e., ∇×(∇×) = O, can be alternatively
expressed as
∇2  + ∇ (∇(tr )) − ∇(∇ · ) − (∇(∇ · ))T = O. (13)
or, alternatively in indicial notation,

ij,kk + kk,ij − ik,kj − jk,ki = 0. (14)

It can be noted that comma represents partial derivative.

Proof. We use indicial notation to prove the result. Consider the ij th component of curl
of curl of strain field
∂ 2 nq
(∇ × (∇ × ))ij = imn jpq (15)
∂Xm ∂Xp
Recall the following general  − δ relation from Lecture-3.
 
δij δip δiq
 
imn jpq = det  δmj δmp δmq 

δnj δnp δnq


= δij (δmp δnq − δmq δnp ) − δip (δmj δnq − δmq δnj ) + δiq (δmj δnp − δmp δnj ).

Upon substituting the general  − δ relation in Eq. (15), yields


! !
∂ 2 nn ∂ 2 mn ∂ 2 nn ∂ 2 jm
(∇ × (∇ × ))ij = − δij − −
∂Xm ∂Xm ∂Xm ∂Xn ∂Xi ∂Xj ∂Xi ∂Xm
!
2 2
∂ ni ∂ ji
+ −
∂Xj ∂Xn ∂Xm ∂Xm
 
= ∇2 (tr ) − ∇ · (∇ · ) δij
h i
+ −∇(∇(tr )) + ∇(∇ · ) + (∇(∇ · ))T − ∇2  . (16)
ij

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Substitution of Eq. (12) in above equation give rise to

∇ × (∇ × ) = −∇2  − ∇(∇(tr )) + ∇(∇ · ) + (∇(∇ · ))T (17)

Since ∇ × (∇ × ) = O, we get

∇2  + ∇(∇(tr )) − ∇(∇ · ) − (∇(∇ · ))T = O.

The component form can be written as

∂ 2 ij ∂ 2 kk ∂ 2 ik ∂ 2 jk


+ − − = 0. (18)
∂Xk ∂Xk ∂Xj ∂Xj ∂Xj ∂Xk ∂Xi ∂Xk

Alternatively, it can be represented by

ij,kk + kk,ij − ik,kj − jk,ki = 0.

This is a convenient component form of compatibility.

The stress measure:


In linear elasticity, we assume that the displacement gradient is small and displacements
are small. Therefore, all the stress measures, i.e., Cauchy stress τ , first Piola-Kirchhoff
stress T and second Piola-Kirchhoff stress S, are approximately equal. Recall the follow-
ing relations between Cauchy stress and Piola-Kirchhoff stresses from Lecture-28.

T = Jτ F −1 and S = JF −1 τ F −T , (19)

where F and J are deformation gradient and its determinant, respectively. Since F =
I + ∇u (see Lecture-17) and the displacement gradient ∇u is small, F ≈ I. Hence, the
relations shown in Eq. (19) imply

T ≈ τ and S ≈ τ . (20)

Thus, there is no difference between various stress measures in linear elasticity.

The governing equations:


From the above discussion, the Cauchy stress, first Piola-Kirchhoff stress, and second
Piola-Kirchhoff stress are identical. Thus, we have the following governing equation for
linear elasticity.
∂ 2u
ρ0 2 = ∇ · τ + ρ0 b0 . (21)
∂t
Constitutive relation:
We now use the linear constitutive relation that is similar to St venant-Kirchhoff hypere-
lastic model. Of course the material is also assumed to be isotropic. The only difference
is that the small strain measure  is used in linear elasticity whereas Green strain is used
as strain measure in St venant-Kirchhoff model. Therefore, the constitutive relation is
given by
τ = λ(tr )I + 2µ, (22)
where λ and µ are Lamé constants.

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

We will show the relation between engineering constants (i.e., Young’s modulus E and
Poisson’s ratio ν) and Lamé constants using thought experiments. For isotropic material,
the strains can also be represented in terms of stress tensor
1+ν ν
= τ − tr(τ )I. (23)
E E
We now present relation between engineering constants (Young’s modulus E and Poisson’s
ratio ν) and Lamé constants along with bounds on these constants.

Bounds on Lamé constants:


We conduct three thought experiments to find bounds on Lamé constants and correspond-
ing Young’s modulus and Poisson’s ratio. We assume that the reference configuration is
stress free (i.e., in natural state) and material is isotropic. Since there is no difference
between Cauchy stress and second Piola-Kirchhoff stress in linear elasticity, we use the
Cauchy stress as stress measure.
Experiment of simple shear:
The following equation of motion (or mapping) represents the simple shear (see Fig. 1)

x1 = X1 + γX2 , x2 = X2 , and x3 = X3 .

Evaluating  and substituting in Eq. (22), we get


1 unit

1 unit

Figure 1: Simple shear

τ12 = µγ.

Since τ12 > 0 for γ > 0, we have


µ > 0. (24)

Experiment of uniform compression of a sphere:


The equation of motion for the uniform compression (hydrostatic compression) is given
by
x = (1 − )X,
where the constant  ∈ (0, 1).
The Cauchy stress tensor
τ = −(3λ + 2µ)I.

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Since the Cauchy stress is hydrostatic pressure, i.e., τ = −pI, and  and p are positive
constants, we have
3λ + 2µ > 0. (25)

Experiment of uniform extension of a bar:


The following motion describes uniform extension (accounting Poisson’s contraction).

x1 = (1 + )X1 , x2 = (1 − ν)X2 , and x3 = (1 − ν)X3 .

Substitution of computed strain tensor  in Eq. (22) yields

τ22 = τ33 =  [λ(1 − 2ν) − 2µν] .

Since there are no lateral stresses, i.e, τ22 = τ33 = 0, we get


λ
ν= (26)
2(λ + µ)
The stress along e1 direction on the e1 -plane is given by

τ11 =  [λ(1 − 2ν) + 2µ] .

Since the axial stress τ11 = E, we get


µ(3λ + 2µ)
E= . (27)
(λ + µ)
The engineering constants E and ν are known as Young’s modulus and Poisson’s ratio,
respectively. In summary, we have
µ(3λ + 2µ) λ
E= , ν= ,
(λ + µ) 2(λ + µ)
Eν E
λ= , µ= . (28)
(1 + ν)(1 − 2ν) 2(1 + ν)

The inequalities shown in Eq. (24) and (25) are equivalent to E > 0 and −1 < ν < 0.5,
i.e.,
µ > 0 and 3λ + 2µ > 0 ⇐⇒ E > 0 and − 1 < ν < 0.5.

Approach for solving linear elasticity problems:


In linear elasticity, our aim is to find the stress and displacement fields for a given traction
and displacement boundary conditions. In order to find the stress and displacement fields,
there are two distinct approaches called displacement approach and stress approach.

In displacement approach, the governing equation shown in Eq. (21) is expressed in


terms of displacements known as Navier equation of elasticity. The solution is obtained
by solving the Navier equation subjected to the given boundary conditions.

In stress approach, the compatibility conditions shown in Eq. (7) is expressed in terms
of stress components known as Beltrami-Michell compatibility conditions. The stress field
is guessed such that it satisfies the Beltrami-Michell compatibility conditions and the
boundary conditions.

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

We now present both the Navier equation of elasticity and Beltrami-Michell compati-
bility conditions.

Navier equation of elasticity:


Substitution of the strain displacement relation, i.e, Eq. (1) in constitutive relation (i.e.,
in Eq. (22)), we get
τ = λ(∇ · u)I + µ(∇u + ∇uT ), (29)
Upon substituting Eq. (29) in Eq. (21), we get the following Navier equation of elasticity.
∂ 2u
ρ0 = (λ + µ)∇(∇ · u) + µ∇2 u + ρ0 b, (30)
∂t2
where ρ0 is density, λ and µ are Lamé constants.
Beltrami-Michell compatibility conditions:
In theory of elasticity problems, it is easy to guess stresses rather than solving for displace-
ments. Therefore, it is necessary to verify the compatibility conditions. We now derive
the compatibility conditions in terms of stress components known as Beltrami-Michell
compatibility conditions.
We consider the static problem in linear elasticity. Hence, the governing equation Eq.
(21) for static case reduces to
∇ · τ + ρ0 b = 0. (31)
Substituting the constitutive relation shown in Eq. (23) in compatibility conditions (i.e.,
in Eq. (13)), we get
1+ν 2 ν 2 1 − 2ν 1+ν ν
     
∇ τ − ∇ (tr τ )I + ∇ (∇(tr τ )) − ∇(∇ · τ ) − ∇ (∇(tr τ ))
E E E E E
1+ν ν
 
− [∇(∇ · τ )]T − ∇(∇(tr τ )) = 0.
E E
Rearrangement of terms yields
 
(1 + ν)∇2 τ + ∇ (∇(tr τ )) = ν∇2 (tr τ )I + (1 + ν) ∇(∇ · τ ) + [∇(∇ · τ )]T .
Substituting the ∇ · τ = −ρ0 b from equilibrium equation (see Eq. (31)), we get
1 ν  
∇2 τ + ∇ (∇(tr τ )) = ∇2 (tr τ )I − ρ0 ∇b + (∇b)T .
1+ν 1+ν
Taking trace on both sides of equation, we obtain
1+ν
∇2 (tr τ ) = − ρ0 (∇ · b). (32)
1−ν
Substituting this relation in above equation we get the following Beltrami-Michell com-
patibility equation.
1 ν
∇2 τ + ∇ (∇(tr τ )) = − ρ0 (∇ · b)I − ρ0 ∇b − ρ0 (∇b)T . (33)
1+ν 1−ν
In any linear elasticity problem, the stress field should be guessed such that it satisfies
the Beltrami-Michell compatibility condition, i.e., Eq. (33).

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-5: Constitutive Relations


Lecture-38: Thermomechanics of Fluids

In this lecture, we discuss the implications of second law of thermodynamics on simple


fluids (i.e., the fluid with assumption of simple material). The short-range time history is
accounted along with the assumption of simple fluid while ignoring the long-range time
history. In particular, we consider the constitutive relations such that they are functions
of rate of deformation gradient Ḟ and not higher time derivatives of deformation gradi-
ent, i.e., F̈ , · · · . Furthermore, the Eulerian framework is convenient for fluid mechanics.
Hence, constitutive relations for Cauchy stress, heat flux, internal energy and entropy are
given by

τ = τ̂ (X, F , θ, g, Ḟ ),
q = q̂(X, F , θ, g, Ḟ ),
e = ê(X, F , θ, g, Ḟ ),
η = η̂(X, F , θ, g, Ḟ ),

where g is temperature gradient, i.e., g = ∇X θ and X represents the material point or


particle. We use the terminology ‘viscous fluid’ to indicate the simple fluid without long-
range memory effects. As mentioned in Lecture-35, fluids known to exhibit all possible
symmetries. Hence, we have

τ̂ (X, F , θ, g, Ḟ ) = τ̂ (X, F H, θ, g, Ḟ H),


q̂(X, F , θ, g, Ḟ ) = q̂(X, F H, θ, g, Ḟ H),
ê(X, F , θ, g, Ḟ ) = ê(X, F H, θ, g, Ḟ H),
η̂(X, F , θ, g, Ḟ ) = η̂(X, F H, θ, g, Ḟ H),

for every unimodular tensor H, i.e., for every H such that det(H) = 1. Let L and ρ be
velocity gradient and density field, respectively. Then, choosing H = J 1/3 F −1 , we get

τ̂ (X, F , θ, g, Ḟ ) = τ̂ (X, J 1/3 I, θ, g, J 1/3 L) = τ̃ (X, ρ, θ, g, L)


q̂(X, F , θ, g, Ḟ ) = q̂(X, J 1/3 I, θ, g, J 1/3 L) = q̃(X, ρ, θ, g, L)
ê(X, F , θ, g, Ḟ ) = ê(X, J 1/3 I, θ, g, J 1/3 L) = ẽ(X, ρ, θ, g, L)
η̂(X, F , θ, g, Ḟ ) = η̂(X, J 1/3 I, θ, g, J 1/3 L) = η̃(X, ρ, θ, g, L), (1)

since ρ0 = ρJ.

Problem 1. The constitutive relation τ = τ̃ (X, ρ, θ, g, L) is isotropic and follows the


principle of material frame-indifference if and only if τ̃ (X, ρ, θ, g, L) = τ̃ (X, ρ, θ, g, D),
where D is symmetric part of L.

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Proof. The given constitutive relation τ = τ̃ (X, ρ, θ, g, L). Let F ∗ and F be fixed and
moving frames as discussed in the principle of material frame-indifference (see Lecture-34).
Then
τ̃ (X ∗ , ρ∗ , θ∗ , g ∗ , L∗ ) = Qτ̃ (X, ρ, θ, g, L)QT ,
where Q(t) is relative rotation between frames. Since X ∗ = X, ρ∗ = ρ, θ∗ = θ, g ∗ = g and
L∗ = QLQT + Q̇QT (see Lectures 31, 32 and 35) and the velocity gradient L = D + W
where D is symmetric part and W is skew symmetric part, above equation can be written
as
τ̃ (X, ρ, θ, g, QDQT + QW QT + Q̇QT ) = Qτ̃ (X, ρ, θ, g, L)QT , (2)
where Q is arbitrary rotation tensor. For any given skewsymmetric tensor W , we can
choose an orthogonal tensor such that
Q(0) = I and Q̇(0) = −W .
Substituting these relations in Eq. (2), we get
τ̃ (X, ρ, θ, g, L) = τ̃ (X, ρ, θ, g, D).
Thus, the principle of material frame-indifference implies above relation. Since the given
function is isotropic, the converse is trivial.

Applying similar steps that are discussed in Problem-1 to other constitutive relations,
we can show that they are also function of rate of deformation. In summary, we have
τ = τ̃ (X, ρ, θ, g, D)
q = q̃(X, ρ, θ, g, D)
e = ẽ(X, ρ, θ, g, D)
η = η̃(X, ρ, θ, g, D), (3)
The following Clausius-Duhem inequality (second law of thermodynamics) in terms of
internal energy e and entropy η can be obtained from Eqs. (10) and (13) of Lecture-30.
De Dη 1 q·g
−θ − τ :D+ ≤ 0. (4)
Dt Dt ρ ρθ
Since the constitutive relations are not direct function of time, we have
De ∂e ∂e ∂e ∂e Dη ∂η ∂η ∂η ∂η
= ρ̇ + θ̇ + · ġ + : Ḋ and = ρ̇ + θ̇ + · ġ + : Ḋ.
Dt ∂ρ ∂θ ∂g ∂D Dt ∂ρ ∂θ ∂g ∂D
Substituting these relations in above Clausius-Duhem inequality, we get
! ! !
∂e ∂η ∂e ∂η ∂e ∂η
−θ ρ̇ + −θ θ̇ + −θ · ġ
∂ρ ∂ρ ∂θ ∂θ ∂g ∂g
!
∂e ∂η 1 q·g
+ −θ : Ḋ − τ :D+ ≤ 0.
∂D ∂D ρ ρθ
Let v be a spatial velocity field. Then the conservation of mass can be written as ρ̇ =
−ρ ∇x · v = −ρ tr(L) = −ρ I : L. Substitution of this relation yields
! ! !
∂e ∂η ∂e ∂η ∂e ∂η
−θ θ̇ + −θ · ġ + −θ : Ḋ
∂θ ∂θ ∂g ∂g ∂D ∂D
! !
∂e ∂η 1 q·g
+ − ρ − ρθ I− τ :D+ ≤ 0. (5)
∂ρ ∂ρ ρ ρθ

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Clearly, the constitutive relations are not function of θ̇, ġ and Ḋ as shown in Eq. (3).
Since θ̇, ġ and Ḋ are arbitrary, we get
∂e ∂η ∂e ∂η ∂e ∂η
−θ = 0, −θ = 0, and −θ =O (6)
∂θ ∂θ ∂g ∂g ∂D ∂D
! !
∂e ∂η 1 q·g
ρ − ρθ I+ τ :D− ≥0 (7)
∂ρ ∂ρ ρ ρθ

Differentiating Eq. (6)1 with g and Eq. (6)2 with θ, we can write

∂ 2e ∂ 2η ∂ 2e ∂ 2η ∂η
−θ = 0 and −θ − = 0.
∂g∂θ ∂g∂θ ∂θ∂g ∂θ∂g ∂g
Combining these two equations yield
∂η
= 0. (8)
∂g

Substitution of this relation in Eq. (6)2 give rise to

∂e
= 0. (9)
∂g
Thus, the internal energy and entropy are independent of temperature gradient g. Similar
steps applied to Eq. (6)1 and Eq. (6)3 , we get

∂η ∂e
= O and =O (10)
∂D ∂D
In conclusion, constitutive relation to internal energy and entropy reduces to

e = ẽ(X, ρ, θ) and η = η̃(X, ρ, θ). (11)

We now consider the dissipation inequality (see Eq. (7))

δ(ρ, θ, g, D) ≥ 0 (12)

where the dissipation function


! !
∂e ∂η q·g
δ(ρ, θ, g, D) = ρ2 −θ I +τ :D− (13)
∂ρ ∂ρ θ

The thermodynamic equilibrium is defined as the process that is independent of time and
also uniform. Physically, the thermodynamic equilibrium refers to mechanical equilibrium
(force balance), thermal equilibrium (no heat transfer between system and surroundings)
and chemical equilibrium (no change in reactants and products with time). For the
thermodynamic equilibrium, we have the following velocity and temperature fields.

v(x, t) = constant and θ(x, t) = constant, ∀x = Ω. (14)

The velocity and temperature field imply

D = O and g = 0. (15)

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Therefore, at thermodynamic equilibrium, Eq. (13) shows zero dissipation, i.e,

δ(ρ, θ, 0, O) = 0. (16)

It is clear from Eqs. (12) and (16) that the dissipation function δ is minimum at thermody-
namic equilibrium. In order to have minimum dissipation at thermodynamic equilibrium,
we need to have the necessary condition
∂δ ∂δ
= O and =0 (17)
∂D ∂g
and positive definite Hessian (second derivative of dissipation) for sufficiency condition.
The Hessian of dissipation function is given by
∂ 2δ ∂ 2δ
 

 ∂D∂D ∂D∂g 
 
H=  
. (18)
 ∂ 2δ ∂ 2δ 
∂D∂g ∂g∂g
Taking derivatives to the dissipation function (see Eq. (13)), we get
!
∂δ ∂e ∂η ∂τ 1 ∂q
= τ + ρ2 −θ I+ :D− · g, (19)
∂D ∂ρ ∂ρ ∂D θ ∂D
∂δ ∂τ q 1 ∂q
= :D− − · g, (20)
∂g ∂g θ θ ∂g
where
! !
∂τ ∂τkl ∂q ∂qk
:D = Dkl , ·g = gk ,
∂D ij
∂Dij ∂D ij
∂Dij
! !
∂τ ∂τjk ∂q ∂qj
:D = Djk , and ·g = gj .
∂g i
∂gi ∂g i
∂gi

The derivatives at thermodynamic equilibrium, i.e., the evaluation of derivatives at D =


O and g = 0, yield
!
∂δ ∂e ∂η ∂δ q
= τ + ρ2 −θ I, and =− . (21)
∂D T .E. ∂ρ ∂ρ ∂g T .E. θ

In above relations the subscript T .E. represents thermodynamic equilibrium. Since the
derivatives should vanish at thermodynamic equilibrium (see Eq. (17)), we have


τ = −pI and q = 0, (22)


T .E. T .E.

where thermodynamic pressure


!
∂e ∂η
p = ρ2 −θ . (23)
∂ρ ∂ρ

Since the free energy ψ = e − ηθ (see Lecture-30), the thermodynamic pressure can be
written as
∂ψ
p = ρ2 . (24)
∂ρ

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

In conclusion, at thermodynamic equilibrium the state of stress is isotropic and it is


determined by the free energy. In addition, the heat flux vanishes at the thermodynamic
equilibrium.
In general (including for non-equilibrium processes) the constitutive relations for stress
and heat flux are given by

τ = τ̃ (X, ρ, θ, g, D) and q = q̃(X, ρ, θ, g, D). (25)

We now define the viscous stress

σ = σ̃(X, ρ, θ, g, D) = τ̃ (X, ρ, θ, g, D) − τ̃ (X, ρ, θ, 0, O). (26)

In other words, the stress tensor can be decomposed into two parts: (i) thermodynamic
equilibrium stress (ii) viscous stress. Therefore, the stress tensor

τ = −pI + σ. (27)

Substituting this relation in dissipation inequality (i.e., in Eq. (12)), we get


1
σ : D − q · g ≥ 0. (28)
θ

Summary of the relations:


In summary, we have the following relations

• The constitutive relations for internal energy and entropy are given by

e = ẽ(X, ρ, θ) and η = η̃(X, ρ, θ).

• The thermodynamic equilibrium stress




τ = τ̃ (X, ρ, θ, 0, O) = −pI

T .E

where !
2 ∂e ∂η
p=ρ −θ .
∂ρ ∂ρ
• Let σ = σ̃(X, ρ, θ, g, D) be viscous stress then the total stress is given by

τ = −pI + σ.

• The dissipation inequality


1
σ : D − q · g ≥ 0.
θ

The momentum balance and energy balance:


The following momentum and energy balance equations can be obtained with the substi-
tution of the stress decomposition, i.e., decomposition of viscous stress and equilibrium
stress, into previously stated momentum and energy balances.
!
∂v
ρ + (∇v)v = −∇p + ∇ · σ + ρb, (29)
∂t
∂e
ρ = −p∇ · v + σ : D − ∇ · q + ρQh , (30)
∂t

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

where v is velocity field. The term σ : D is known as mechanical dissipation as it


appears in the dissipation function. Physically the term σ : D represents the conversion
of mechanical energy into thermal energy due to viscous effects.
Viscous heat-conducting fluids:
It is reasonable to assume the isotropic constitutive relations for fluids as they exhibit all
possible material symmetries. In particular, the constitutive relations for Cauchy stress,
heat flux, internal energy and entropy are isotropic. The isotropy of constitutive relation
for Cauchy stress implies the isotropy of viscous stress as the thermodynamic pressure is
an isotropic tensor. In result, we have the following isotropic functions.

σ = σ̃(ρ, θ, g, D),
q = q̃(ρ, θ, g, D),
e = ẽ(ρ, θ),
η = η̃(ρ, θ).

The isotropic functions of viscous stress and heat flux have the following representation1

σ = γ0 I + γ1 D + γ2 D 2 + γ3 g ⊗ g + γ4 (Dg ⊗ g + g ⊗ Dg) + γ5 (Dg ⊗ Dg),


q = (β0 I + β1 D + β2 D 2 )g, (31)

where the scalars γi are isotropic functions of ρ, θ, g and D, i.e.,

γi = γ̃i (ρ, θ, tr(D), tr(D 2 ), tr(D 3 ), g · g, g · Dg, g · D 2 g).

It is clear that both the viscous stress σ and also the heat flux q are dependent on rate
of deformation D and temperature gradient g.
Uncoupled viscous heat-conducting fluids:
In case of uncoupled system, we assume that viscous stress is independent of temperature
gradient and the heat flux is independent of rate of deformation. The constitutive relations
for uncoupled system can be written as

σ = γ0 (ρ, θ, 0, ID )I + γ1 (ρ, θ, 0, ID )D + γ2 (ρ, θ, 0, ID )D 2 ,


q = β0 (ρ, θ, g · g, O)g,

where ID represents set of three invariants of D. We can have two special cases: (i) rate
of deformation is zero (D = O) and temperature gradient is non-zero (g 6= 0) (ii) rate
of deformation is non-zero (D 6= O) and temperature gradient is zero (g = 0). We note
that later case is also known as Reiner-Rivlin fluid.
Heat conducting fluid:
Since the rate of deformation is zero (D = O), we only have heat conduction inequality

q · g ≤ 0 =⇒ β0 (ρ, θ, g · g, O) g · g ≤ 0. (32)

Since g · g ≥ 0, we have β0 ≤ 0. We now define the conductivity

k = −β0 . (33)
1
Liu, I-S., Continuum Mechanics, Springer-Verlag, Berlin, 2002.

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

Therefore, the heat flux relation can be written as

q = −kg. (34)

This relation is known as Fourier law of heat conduction for fluids. Furthermore, the
Fourier law is similar to both solids and also for fluids. Through this law is important for
fluids, there are many instances where convective heat transfer is more than conduction.
Thus, in many practical problems involving fluids the conduction can be ignored. There
are few problems where the conduction is important. For example if the fluid does not have
sufficient space (thin fluid layer between walls) to convect then conduction is important.

Reiner-Rivlin fluids:
In this case, we assume that the temperature gradient is zero (g = 0). The constitutive
relation for the viscous stress can be written as

σ = γ0 (ρ, θ, 0, ID )I + γ1 (ρ, θ, 0, ID )D + γ2 (ρ, θ, 0, ID )D 2 , (35)

subjected to the (Clausius-Duhem) inequality

σ : D ≥ 0. (36)

References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. Z. Martinec, Lecture Notes on Continuum Mechanics


(Link: http://geo.mff.cuni.cz/vyuka/Martinec-ContinuumMechanics.pdf)

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-5: Constitutive Relations


Lecture-39: Linear Viscous Heat-Conducting Fluids

Recall the general constitutive relations for heat conducting fluid from Lecture-38. In
this lecture, we specialize these general constitutive relations to the linear, viscous heat-
conducting fluid. The linear, viscous heat-conducting fluid is also known as classical
heat-conducting fluid. The classical heat-conducting fluid without temperature gradient
is known as Newtonian fluid.

We recall from Lecture-38 that the Reiner-Rivlin fluid is the general heat-conducting
fluid without temperature gradient. Therefore, the Newtonian fluid model can also be
thought of the linearized Reiner-Rivlin fluid model. In conclusion, either ignoring temper-
ature gradient in classical heat-conducting fluid model or linearizing Reiner-Rivlin fluid
model, we get Newtonian fluid model.
From above discussion, clearly, the Newtonian fluid model is a special case of Reiner-
Rivlin fluid model. However, there are no practical fluids that can be model with the
Reiner-Rivlin fluid model but the Newtonian fluid model. On the other hand, the non-
Newtonian fluids like polymer solutions, blood, etc., need to be accounted for long-range
time history through higher time derivatives of deformation gradient. Since the Reiner-
Rivlin fluid model based short-range time history, it is not suitable for non-Newtonian
fluids. Thus, the Newtonian fluid model is an important model in comparison with the
Reiner-Rivlin fluid model. Since linear models are important, in this lecture, the consti-
tutive relations for the classical heat-conducting fluid model are discussed.
Linear (classical) viscous heat-conducing fluid:
We now consider a special class of viscous heat-conducting fluids obeying linear consti-
tutive relationship with respect to rate of deformation D and temperature gradient g.
The following linear constitutive relations for viscous stress and heat flux can be obtained
from the general relations (see Eq. (31) in Lecture-38).
σ = λ (tr D)I + 2µ D, (1)
q = −k g, (2)
where λ = λ̃(ρ, θ) is known as dilatational viscosity, µ = µ̃(ρ, θ) is known as shear viscosity
or dynamic viscosity, and k = k̃(ρ, θ) is known as coefficient of thermal conductivity.
Recall the following dissipation function from Lecture-38.
q·g
δ(ρ, θ, g, D) = σ : D − , (3)
θ
where σ is viscous stress, q is heat flux, D is rate of deformation, θ is temperature and
g is temperature gradient.

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

Utilizing the positive definiteness of Hessian of dissipation function (see Lecture-38),


∂ 2δ ∂ 2δ
 

 ∂D∂D ∂D∂g 
 
H=  
, (4)
 ∂ 2δ ∂ 2δ 
∂D∂g ∂g∂g
we derive the bounds on λ, µ, and k.
Since the viscous stress σ is independent of g and the heat flux q is independent of D
(see Eqs. (1) and (2)), it is easy to see that the cross derivative terms involved in Hessian
H are zero, i.e., in Eq. (4)
∂ 2δ ∂ 2δ
= 0 and = 0.
∂D∂g ∂g∂D
∂ 2δ ∂ 2δ
Therefore, it is sufficient to have positive definiteness of and to have semi-
∂D∂D ∂g∂g
positive definiteness of Hessian H. The positive definiteness implies
! !
∂ 2δ ∂ 2δ
Sij Skl > 0 and ui uj > 0. (5)
∂Dij ∂Dkl ∂gi ∂gj
for every non-zero arbitrary symmetric second-order
! tensor S and for every non-zero
∂ 2δ
arbitrary vector u. Of course Sij Skl = 0 if and only if S = O and
!
∂Dij ∂Dkl
∂ 2δ
ui uj = 0 if and only if u = 0.
∂gi ∂gj
Since the viscous stress tensor σ is not a function of g, using Eqs. (2) and (3), we get
∂ 2δ
= kI. (6)
∂g∂g
Substituting the relation in Eq. (5)2 , we obtain

k > 0. (7)

In conclusion the coefficient of thermal conductivity is a positive scalar.

Since q is not a function of D, taking derivative of dissipation function δ (see Eq. (3))
with respect to D, we get
!
∂δ ∂δ ∂σpq
= = σij + Dpq . (8)
∂D ij
∂Dij ∂Dij

Again taking differentiation with respect to D yields


!
∂ 2δ ∂ 2δ ∂σij ∂σpq ∂Dpq
= = +
∂D∂D ijkl
∂Dij ∂Dkl ∂Dkl ∂Dij ∂Dkl
∂σij ∂σkl
= + . (9)
∂Dkl ∂Dij
Using Eq. (1), we get
∂σkl
= λδij δpq + µ(δip δjq + δiq δjp ). (10)
∂Dij

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

Upon substituting Eq. (10) in Eq. (9), we get


!
∂ 2δ
= 2λδij δkl + 2µ(δik δlj + δkj δil ). (11)
∂D∂D ijkl

Substitution of this relation in Eq. (5)1 yields


!
∂ 2δ
Sij Skl = 2λ(tr S)2 + 4µS : S > 0,
∂D∂D ijkl

where S is a non-zero arbitrary symmetric tensor. This inequality can be written as

λ(tr S)2 + 2µS : S > 0. (12)

The symmetric tensor S can be decomposed into spherical and deviatoric parts. Let
deviatoric part of S be S̃. Then
tr(S)
S̃ = S − I. (13)
3
Substituting this relation in above inequality and using property I : S̃ = 0, we get
2
 
λ + µ (tr S)2 + 2µS̃ : S̃ > 0. (14)
3
Choosing S = I, we get
2
λ + µ > 0, (15)
3
and choosing symmetric tensor S such that tr(S) = 0, we get

µ > 0. (16)

Starting with Eq. (12), we can have an alternative proof for bounds on dilatational and
shear viscosities which is presented in the following problem.

Problem 1. Let S be a non-zero arbitrary symmetric tensor. Then

λ(tr S)2 + 2µS : S > 0.


2
if and only if λ + µ > 0 and µ > 0.
3

Proof. Given inequality can explicitly be written as

λ(S11 + S22 + S33 )2 + 2µ(S11


2 2
+ S22 2
+ S33 2
+ 2S12 2
+ 2S13 2
+ 2S23 ) > 0, (17)

as S is a non-zero symmetric tensor.


Let us define a vector s and matrix M by
   


 S11 

 λ + 2µ λ λ 0 0 0
   
S22 λ λ + 2µ λ 0 0 0

 
  

 
  

 
  
S33 λ λ λ + 2µ 0 0 0

 
  
s= , M= 





 S12 



 0 0 0 2µ 0 0 


 
  



 S13 




 0 0 0 0 2µ 0 

 
S23 0 0 0 0 0 2µ

 

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

The inequality shown in Eq. (17) can be written as

s · M s > 0. (18)

Of course s · M s = 0 if and only if s = 0.


Let α be an eigenvalue of M . Then we can have the characteristic equation
 
−α3 + 3(λ + 2µ)α2 − 12µ(λ + µ)α + 4µ2 (3λ + 2µ) (2µ − α)3 = 0.

This equation can also be written as

(α − (3λ + 2µ))(α − 2µ)5 = 0.

A known result from linear algebra (see Lecture-11) that above positive definiteness of
matrix M , i.e., Eq. (18), implies all eigenvalues are positive. Furthermore, converse is
also true. Thus, we have
3λ + 2µ > 0 and µ > 0.
This completes the proof.

Relation between thermodynamic pressure and mean pressure:


Let us consider the constitutive relation for stress in case of Newtonian fluid, i.e.,

τ = −pI + λtr(D) + 2µD.

Let p̂ be a mean pressure. Then


1
p̂ = − tr(τ )
3
2
= p − λtr(D) − µtr(D)
3
2
 
= p − λ + µ tr(D)
3 
2

= p − λ + µ ∇ · v. (19)
3
where v is velocity field. The factor λ + 2µ/3 is known as bulk viscosity. Clearly, the
thermodynamic pressure p and mean pressure p̂ are equal, i.e., p = p̂, if the fluid is
incompressible as ∇ · v = 0.

The Navier-Stokes equation:


We get the following famous Navier-Stokes equation with substitution of linear stress
and rate of deformation relation (Newtonian fluid model) in momentum balance, i.e., the
Cauchy stress τ = −pI ++λtr(D)+2µD, where D is symmetric part of velocity gradient,
in momentum equation.
!
∂v
ρ + (∇v)v = −∇p + (λ + µ)∇(∇ · v) + µ∇ · (∇v) + ρb. (20)
∂t
The Navier-Stokes equation in indicial notation can be written as
!
∂vi ∂vi ∂p ∂ 2 vk ∂ 2 vi
ρ + vj =− + (λ + µ) +µ + ρbi . (21)
∂t ∂xj ∂xi ∂xi ∂xk ∂xj ∂xj

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

The term (∇v)v contributes to the nonlinearity with respect to velocity in the Navier-
Stokes equation. Furthermore, the greatest difficulty involved in getting analytical solu-
tion is due to this nonlinear term.
Specific volume and specific heat capacity:
We now define two quantities known as specific volume and specific heat at constant
volume. The specific volume v is defined as the reciprocal of density, i.e.,
1
v= . (22)
ρ
It is clear that constant density is also equivalent to constant specific volume. The specific
heat capacity at constant volume is defined by
!
∂e
cv = . (23)
∂θ v

We note that subscript v represents the thermodynamic process with constant volume.
Physically, the specific heat at constant volume indicates the increase of internal energy
required to raise one degree of temperature in Kelvin scale.
Clearly, the specific heat can be function of both temperature and density as internal
energy e = ẽ(ρ, θ). In practice, there is a change in specific heat with temperature but
the change is small. So, usually, in engineering problems the specific heat is assumed to
be constant.
Energy balance:
We now derive the energy balance in terms of specific heat, pressure, temperature, viscous
stress, heat flux and heat generation. Consider the first law of thermodynamics (see Eq.
(10) in Lecture-29)
De
ρ = τ : D − ∇ · q + ρQh . (24)
Dt
Substitution of the decomposition of Cauchy stress, i.e., τ = −pI + σ, in energy balance
yields
De
ρ = −pI : D + σ : D − ∇ · q + ρQh .
Dt
Since I : D = tr(D) = ∇ · v, we get
De
ρ + p∇ · v = σ : D − ∇ · q + ρQh . (25)
Dt
We note that the boldface letter v is velocity and lightface letter v is specific volume.
Since the internal energy e = ẽ(ρ, θ), the material time derivative can be written as
De ∂e ∂e
= θ̇ + ρ̇.
Dt ∂θ ∂ρ
Using Eq. (23) and mass conservation ρ̇ = −ρ∇ · v, we get
De ∂e
= cv θ̇ − ρ (∇ · v). (26)
Dt ∂ρ
As defined in Lecture-38, the thermodynamic pressure
!
2 ∂e ∂η
p=ρ −θ .
∂ρ ∂ρ

Joint initiative of IITs and IISc – Funded by MHRD 5


NPTEL – Mechanical Engineering – Continuum Mechanics

Taking partial derivative with temperature, we get


!
∂p ∂ 2e ∂η ∂η
= ρ2 −θ −
∂θ ∂θ∂ρ ∂θ∂ρ ∂ρ
!
∂η ∂e ∂η
= −ρ2 Since =θ (see (6)1 in Lecture-38)
∂ρ ∂θ ∂θ
Substituting this relation in above thermodynamic pressure, we get
∂e ∂p
p = ρ2 +θ .
∂ρ ∂θ
Rearrangement of terms yields
∂e ∂p
ρ2 =p−θ . (27)
∂ρ ∂θ
Combining Eqs. (27) and (26), yields
!
De ∂p
ρ = ρcv θ̇ − p∇ · v + θ ∇ · v. (28)
Dt ∂θ

Substituting Eq. (28) in Eq. (25), we get


!
Dθ ∂p
ρcv = −θ (∇ · v) + σ : D − ∇ · q + ρQh . (29)
Dt ∂θ v

Substituting constitutive relations for classical viscous heat conducting fluid, i.e., σ =
λtr(D)I + 2µD and q = −k∇θ, we get
!
Dθ ∂p
ρcv = −θ (∇ · v) + λ(∇ · v)2 + 2µD : D + k∇2 θ + ρQh . (30)
Dt ∂θ v

Summary of governing equations:


In summary, we have the following governing equation for classical heat-conducting fluid.
Continuity equation:


+ ρ∇ · v = 0,
Dt
or, alternatively,
∂ρ
+ ∇ · (ρv) = 0.
∂t
The Navier-Stokes equations:
!
∂v
ρ + (∇v)v = −∇p + (λ + µ)∇(∇ · v) + µ∇ · (∇v) + ρb.
∂t
Energy equation:
!
Dθ ∂p
ρcv = −θ (∇ · v) + λ(∇ · v)2 + 2µD : D + k∇2 θ + ρQh ,
Dt ∂θ v
Equation of state:

f (p, ρ, θ) = 0. (31)

Joint initiative of IITs and IISc – Funded by MHRD 6


NPTEL – Mechanical Engineering – Continuum Mechanics

We have six equations and six unknown fields, i.e., the velocity field v, pressure field p,
temperature field θ and density field ρ. So we can solve the problem with appropriate
boundary conditions. Furthermore, the following differential equation can be used for
evaluating entropy production.

ρθ = λ(∇ · v)2 + 2µD : D + k∇2 θ + ρQh . (32)
Dt
References

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

2. Z. Martinec, Lecture Notes on Continuum Mechanics


(Link: http://geo.mff.cuni.cz/vyuka/Martinec-ContinuumMechanics.pdf)

Joint initiative of IITs and IISc – Funded by MHRD 7


NPTEL – Mechanical Engineering – Continuum Mechanics

Module-5: Constitutive Relations


Lecture-40: Thermomechanics of
Incompressible Fluids

In this lecture, we discuss the governing equations for incompressible Newtonian fluid.
The density remains constant throughout the motion, i.e., ρ = ρ0 , for incompressible
fluid. In addition, if the density is same at every point then it is called homogeneous
incompressible fluid. In the following discussion, we assume that the fluid is homogeneous
incompressible.
Kinematic constraint:
The incompressibility is a kinematic constraint on a fluid which can be written as

∇ · v = 0, or, alternatively, tr(D) = 0, (1)

where v is the velocity field and D is rate of deformation tensor. The hydrostatic pressure
p developed in the fluid as reaction to incompressible constraint. However, the hydrostatic
pressure cannot be determined by the constitutive law as density is invariant under the
stress. Furthermore, the pressure can only be determined by the boundary conditions and
equations of motion.
Constitutive relations:
Recall the following constitutive relations for Cauchy stress, heat flux, internal energy,
and entropy from Lecture-38.

τ = τ̃ (X, θ, g, D)
q = q̃(X, θ, g, D)
e = ẽ(X, θ, g, D)
η = η̃(X, θ, g, D), (2)

Since the density is constant for in compressible fluid, the density does not appear in
the argument of constitutive relations. Recall from Eqs. (8) and (9) in Lecture-38 that
the internal energy and the entropy for viscous heat-conducting fluid are independent of
temperature gradient. Now following same procedure that was discussed in Lecture-38
for incompressible fluid, we can also show that the constitutive relations for the internal
energy and the entropy are independent of temperature gradient, i.e.,

e = ẽ(X, θ) and η = η̃(X, θ). (3)

The dissipation function (see for discussion on dissipation in Lecture-38)


q·g
δ(θ, g, D) = σ : D − , (4)
θ

Joint initiative of IITs and IISc – Funded by MHRD 1


NPTEL – Mechanical Engineering – Continuum Mechanics

and
δ(θ, g, D) ≥ 0. (5)
tr(τ )
Let τ be Cauchy stress. Then the hydrostatic pressure p = − and the deviatoric
3
stress σ = τ + pI. Then the stress power

σ : D = (pI + τ ) : D
= −p (tr D) + τ : D
= τ :D (Since tr(D) = 0) . (6)

The Cauchy stress tensor


τ = −pI + σ. (7)
The deviatoric stress σ is also known as viscous stress.
Governing equations:
The governing equations shown for compressible in Lecture-39 are also applicable for
incompressible fluids. Of course the kinematic constraint, i.e., constant volume, need to
be accounted. With substitution of the constant volume constraint, i.e., Eq. (1), in the
governing equations of linear viscous heat conducting fluid (i.e., in Eqs. (19), (29) and
(31) of Lecture-39), we get the following reduced governing equations for incompressible
fluids.
Continuity equation:

∇ · v = 0.
Navier-Stokes equation:
!
∂v
ρ0 + (∇v)v = −∇p + µ∇ · (∇v) + ρ0 b.
∂t
Energy equation:


= 2µD : D + k∇2 θ + ρ0 Qh ,
ρ0 cv
Dt
The entropy production can be evaluated using

ρ0 θ = 2µD : D + k∇2 θ + ρ0 Qh . (8)
Dt
Clearly, the governing equations for incompressible fluids are simple when they compared
with the general governing equations that are presented in Lecture-39. Since there are five
unknowns (i.e., velocity, pressure and temperature) and five governing equations (i.e., one
continuity equation, three Navier-Stokes equations, and one energy balance), the problem
is solvable with appropriate boundary conditions. In general, the shear viscosity µ is
assumed to be independent of temperature so that the energy equation is decoupled from
continuity and Navier-Stokes equations.
Bernoulli’s theorem:
The Bernoulli’s theorem (or the Bernoulli’s principle) is a special case of conservation

Joint initiative of IITs and IISc – Funded by MHRD 2


NPTEL – Mechanical Engineering – Continuum Mechanics

of energy. This principle can also be derived from momentum equation for incompress-
ible fluids. However, we derive the Bernoulli’s equation from energy equation so that it
holds for both compressible and incompressible fluids. We need to have the following
assumptions to derive the result.
!
∂α
• The flow is steady state, i.e, any flow quantity α is independent of time =0 .
∂t
• Fluid is inviscid, i.e., viscous stress σ = O.

• The fluid has negligible thermal conductivity (q = 0).

• Body forces are conservative, i.e., there exist a scalar function φ such that the body
force b = ∇φ.

• There are no heat sources, i.e., Qh = 0.

• There is no shaft work (there is no work interaction with surroundings).

Consider the general mass conservation, i.e,



+ ρ∇ · v = 0.
Dt
Dρ ∂ρ
Substituting the material time derivative = + v · ∇ρ, we get
Dt ∂t
∂ρ
+ v · ∇ρ + ρ∇ · v = 0.
∂t
∂ρ
Since the flow is steady state, we have = 0. Substitution of this relation in above
∂t
equation yields
v · ∇ρ + ρ∇ · v = 0
∇·v v · ∇ρ
=⇒ = − 2
ρ ρ
!
1
=⇒ ∇ · v = ρv · ∇ . (9)
ρ
Consider the following conservation of linear momentum
Dv
∇·τ = ρ − ρb
Dt !
∂v
= ρ + (∇v)v − ρb.
∂t
∂v
Since the stress tensor τ = −pI and flow is steady state, i.e., = 0, we get
∂t
−∇p = ρ(∇v)v − ρb.
As stated in assumptions, the body forces are conservative, i.e., b = ∇φ where φ is scalar
field. Therefore, we can write the above equation
−∇p = ρ(∇v)v − ρ∇φ
v·v
 
= ρ∇ − ρ∇φ. (10)
2

Joint initiative of IITs and IISc – Funded by MHRD 3


NPTEL – Mechanical Engineering – Continuum Mechanics

Let us consider the conservation of energy (see Eq. (10) in Lecture-29), i.e.,

De
ρ = τ : D − ∇ · q + ρQh .
Dt
Since q = 0 and Qh = 0, above equation reduces to
De
ρ = τ : D. (11)
Dt
Using the relation between total time derivative and spatial time derivative (see Lecture-
De ∂e ∂e
20), we have = + v · ∇e. Since the flow is steady state, i.e., = 0, and τ = −pI,
Dt ∂t ∂t
Eq. (11) can be written as
ρv · ∇e = −p(∇ · v).
Subtracting v · ∇p on both sides of equation, we get

ρv · ∇e − v · ∇p = −p(∇ · v) − v · ∇p
!
1
= −pρv · ∇ − v · ∇p (using Eq. (9))
ρ
!
p
= −ρv · ∇
ρ
!
v·v p
 
=⇒ ρv · ∇e + ρv · ∇ − ρv · ∇φ = −ρv · ∇ (using Eq. (10))
2 ρ

Combining all the terms, we get


!
p v·v
ρv · ∇ + − φ + e = 0. (12)
ρ 2

We note that the velocity vector is a tangent to the streamline at any point in fluid. Let
s be the arc length of streamline. Then the above equation can be written as
!
p v·v
0 = ρv · ∇ + −φ+e
ρ 2
!
vi ∂ p v·v
= ρ|v| + −φ+e
|v| ∂xi ρ 2
!
dxi ∂ p v·v
= ρ|v| + −φ+e
ds ∂xi ρ 2
!
d p v·v
= ρ|v| + −φ+e .
ds ρ 2

This implies
p v·v
+ − φ + e = constant along streamline.
ρ 2
We have not made any assumption about the compressibility of fluid and hence it is valid
for both compressible and incompressible fluids. Under gravitational body force field,
φ = −gz, above equation reduces to the famous Bernoulli’s equation.
p v·v
+ + gz + e = constant along streamline. (13)
ρ 2

Joint initiative of IITs and IISc – Funded by MHRD 4


NPTEL – Mechanical Engineering – Continuum Mechanics

Bernoulli’s equation for compressible fluids:


The quantity enthalpy for compressible fluids, denoted by h, and defined by the relation

h = e + pv, (14)
1
where v is specific volume. Recall from Lecture-39 that the specific volume, v = .
ρ
Substituting the enthalpy in Eq. (13), we get the following form of Bernoulli’s equation
for compressible fluids.
v·v
h+ + gz = constant along streamline. (15)
2
Applying the Bernoulli’s equation for two points on a streamline, we get
v1 · v1 v2 · v2
h1 + + gz1 = h2 + + gz2 . (16)
2 2
Bernoulli’s equation for incompressible fluids:
If the fluid is incompressible then we have
Dρ ∂ρ
= 0 =⇒ + v · ∇ρ = 0.
Dt ∂t
In addition to incompressible if the flow is steady state then we get v · ∇ρ = 0. Thus, the
density is constant along streamline.

We have the kinematic constraint ∇ · v = 0 (see Eq. (1)) if the fluid is incompressible.
The conservation of energy (with all assumptions stated in Bernoulli’s theorem) shown in
Eq. (11) can be written as
De
= −p∇ · v (Since τ = −pI). (17)
Dt
De
The incompressible condition (∇ · v = 0) implies = 0. The relation between total
Dt
De ∂e
time derivative and spatial time derivative = + v · ∇e along steady state condition
Dt ∂t
∂e
= 0, we get v · ∇e = 0. This relation shows that the internal energy is constant along
∂t
each streamline. If the internal energy is constant then Eq. (15) reduces to the following
Bernoulli’s equation for incompressible fluid.
p v·v
+ + gz = constant along streamline. (18)
ρ 2
Applying Eq. (18) at two points on a streamline for incompressible fluid, we get the
following form of Bernoulli’s equation.
p1 v1 · v1 p2 v2 · v2
+ + gz1 = + + gz2 . (19)
ρ 2 ρ 2
This equation is usually referred as Bernoulli’s equation.

Reference

1. C. S. Jog, Foundations and Applications of Mechanics: Continuum Mechanics,


Volume-I, 2007, Narosa Publishing House Pvt. Ltd., New Delhi.

Joint initiative of IITs and IISc – Funded by MHRD 5

Vous aimerez peut-être aussi