Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f...

27
Chapter 2 Cartesian Tensors 1. Scalars and Vectors .............. 25 2. Rotation of Axes: Formal Definition of a Vector ....................... 26 3. Multiplication of Matrices ........ 29 4. Second-Order Tensor ............ 30 5. Contraction and Multiplication . . 32 6. Force on a Surface ............... 33 Example 2.1 .................... 35 7. Kronecker Delta and Alternating Tensor ........................... 36 8. Dot Product ..................... 37 9. Cross Product ................... 38 10. Operator : Gradient, Divergence, and Curl ........................ 38 11. Symmetric and Antisymmetric Tensors .......................... 40 12. Eigenvalues and Eigenvectors of a Symmetric Tensor ................ 41 Example 2.2 .................... 42 13. Gauss’ Theorem ................. 44 Example 2.3 .................... 45 14. Stokes’ Theorem ................. 47 Example 2.4 .................... 48 15. Comma Notation ................ 49 16. Boldface vs Indicial Notation ..... 49 Exercises ........................ 50 Literature Cited ................. 51 Supplemental Reading ........... 51 1. S c alars an d V ec to rs In fluid mechanics we need to deal with quantities of various complexities. Some of these are defined by only one component and are called scalars, some others are defined by three components and are called vectors, and certain other variables called tensors need as many as nine components for a complete description. We shall assume that the reader is familiar with a certain amount of algebra and calculus of vectors. The concept and manipulation of tensors is the subject of this chapter. A scalar is any quantity that is completely specified by a magnitude only, along with its unit. It is independent of the coordinate system. Examples of scalars are temperature and density of the fluid. A vector is any quantity that has a magnitude and a direction, and can be completely described by its components along three specified coordinate directions. A vector is usually denoted by a boldface symbol, for example, x for position and u for velocity. We can take a Cartesian coordinate system x 1 ,x 2 ,x 3 , with unit vectors a 1 , a 2 , and a 3 in the three mutually perpendicular directions (Figure 2.1). (In texts on vector analysis, the unit vectors are usually denoted by i, j, and k. We cannot use this simple notation here because we shall use ij k to 25 ©2010 Elsevier Inc. All rights reserved. DOI: 10.1016/B978-0-12-381399-2.50002-2

Transcript of Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f...

Page 1: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

Chapter 2

Cartesian Tensors

1. Scalars and Vectors . . . . . . . . . . . . . . 25

2. Rotation of Axes: Formal Definitionof a Vector . . . . . . . . . . . . . . . . . . . . . . . 26

3. Multiplication of Matrices . . . . . . . . 29

4. Second-Order Tensor . . . . . . . . . . . . 30

5. Contraction and Multiplication . . 32

6. Force on a Surface . . . . . . . . . . . . . . . 33

Example 2.1 . . . . . . . . . . . . . . . . . . . . 35

7. Kronecker Delta and AlternatingTensor. . . . . . . . . . . . . . . . . . . . . . . . . . . 36

8. Dot Product . . . . . . . . . . . . . . . . . . . . . 37

9. Cross Product . . . . . . . . . . . . . . . . . . . 3810. Operator ∇: Gradient, Divergence,

and Curl . . . . . . . . . . . . . . . . . . . . . . . . 38

11. Symmetric and AntisymmetricTensors . . . . . . . . . . . . . . . . . . . . . . . . . . 40

12. Eigenvalues and Eigenvectors of aSymmetric Tensor . . . . . . . . . . . . . . . . 41

Example 2.2 . . . . . . . . . . . . . . . . . . . . 42

13. Gauss’ Theorem . . . . . . . . . . . . . . . . . 44

Example 2.3 . . . . . . . . . . . . . . . . . . . . 45

14. Stokes’ Theorem . . . . . . . . . . . . . . . . . 47

Example 2.4 . . . . . . . . . . . . . . . . . . . . 48

15. Comma Notation . . . . . . . . . . . . . . . . 49

16. Boldface vs Indicial Notation. . . . . 49

Exercises . . . . . . . . . . . . . . . . . . . . . . . . 50

Literature Cited . . . . . . . . . . . . . . . . . 51Supplemental Reading . . . . . . . . . . . 51

1 . S c alars an d V ec to rs

In fluid mechanics we need to deal with quantities of various complexities. Some

of these are defined by only one component and are called scalars, some others are

defined by three components and are called vectors, and certain other variables called

tensors need as many as nine components for a complete description. We shall assume

that the reader is familiar with a certain amount of algebra and calculus of vectors.

The concept and manipulation of tensors is the subject of this chapter.

A scalar is any quantity that is completely specified by a magnitude only, along

with its unit. It is independent of the coordinate system. Examples of scalars are

temperature and density of the fluid. A vector is any quantity that has a magnitude

and a direction, and can be completely described by its components along three

specified coordinate directions. A vector is usually denoted by a boldface symbol,

for example, x for position and u for velocity. We can take a Cartesian coordinate

system x1, x2, x3, with unit vectors a1, a2, and a3 in the three mutually perpendicular

directions (Figure 2.1). (In texts on vector analysis, the unit vectors are usually denoted

by i, j, and k. We cannot use this simple notation here because we shall use ijk to

25©2010 Elsevier Inc. All rights reserved. DOI: 10.1016/B978-0-12-381399-2.50002-2

Page 2: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

26 Cartes ian T en s o rs

Fig ur e 2 .1 P osition vector O P and its three Cartesian components (x1, x2, x3). The three unit vectors

are a1, a2, and a3.

denote com p on en ts of a vector.) Then the position vector is written as

x = a1x1 + a2x2 + a3x3,

where (x1, x2, x3) are the components of x along the coordinate directions. (The

superscripts on the unit vectors a do n ot denote the components of a vector; the a’s

are vectors themselves.) Instead of writing all three components explicitly, we can

indicate the three Cartesian components of a vector by an index that takes all possible

values of 1, 2, and 3. For example, the components of the position vector can be

denoted by xi , where i takes all of its possible values, namely, 1, 2, and 3. To obey the

laws of algebra that we shall present, the components of a vector should be written

as a column. For example,

x =

x1

x2

x3

.

In matrix algebra, one defines the tran sp ose as the matrix obtained by interchanging

rows and columns. For example, the transpose of a column matrix x is the row matrix

xT = [x1 x2 x3].

2. R o tatio n o f A x es : F o rm al D efi n itio n o f a V ec to r

A vector can be formally defined as any quantity whose components change similarly

to the components of a position vector under the rotation of the coordinate system.

Page 3: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

2. R o tatio n o f A x es : F o rm al D efi n itio n o f a V ec to r 27

Fig ur e 2 .2 Rotation of coordinate system O 1 2 3 to O 1′ 2′ 3′.

L et x1 x2 x3 be the original axes, and x′1 x′2 x′3 be the rotated system (Figure 2.2). The

components of the position vector x in the original and rotated systems are denoted

by xi and x′i , respectively. The cosine of the angle between the old i and new j axes

is represented by Cij . H ere, the fi rst index of the C matrix refers to the old axes,

and the second index of C refers to the new axes. It is apparent that Cij 6= Cji . A

little geometry shows that the components in the rotated system are related to the

components in the original system by

x′j = x1C1j + x2C2j + x3C3j =3

i=1

xiCij . (2.1)

For simplicity, we shall verify the validity of equation (2.1) in two dimensions only.

Referring to Figure 2.3, let αij be the angle between old i and new j axes, so that

Cij = cos αij . Then

x′1 = O D = O C+ AB = x1 cos α11 + x2 sin α11. (2.2)

As α11 = 90◦−α21, we have sin α11 = cos α21 = C21. Equation (2.2) then becomes

x′1 = x1C11 + x2C21 =2

i=1

xiCi1. (2.3)

In a similar manner

x′2 = P D = P B − D B = x2 cos α11 − x1 sin α11.

Page 4: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

28 Cartes ian T en s o rs

Fig ur e 2 .3 Rotation of a coordinate system in two dimensions.

As α11 = α22 = α12 − 90◦ (Figure 2.3), this becomes

x′2 = x2 cos α22 + x1 cos α12 =2

i=1

xiCi2. (2.4)

In two dimensions, equation (2.1) reduces to equation (2.3) for j = 1, and to equa-

tion (2.4) for j = 2. This completes our verification of equation (2.1).

N ote that the index i appears twice in the same term on the right-hand side of

equation (2.1), and a summation is carried out over all values of this repeated index.

This type of summation over repeated indices appears frequently in tensor notation.

A convention is therefore adopted that, w h en ever an in d ex occu rs tw ice in a term , a

su m m ation over th e rep eated in d ex is im p lied , alth ou g h n o su m m ation sig n is ex p licitly

w ritten . This is frequently called the E in stein su m m ation con ven tion . Equation (2.1)

is then simply written as

x′j = xiCij , (2.5)

where a summation over i is understood on the right-hand side.

The free index on both sides of equation (2.5) is j , and i is the repeated or dummy

index. O bviously any letter (other than j ) can be used as the dummy index without

changing the meaning of this equation. For example, equation (2.5) can be written

equivalently as

xiCij = xkCkj = xmCmj = ··· ,

because they all mean x′j = C1jx1 + C2jx2 + C3jx3. L ikewise, any letter can also

be used for the free index, as long as the same free index is used on b oth sides of

the equation. For example, denoting the free index by i and the summed index by k,

equation (2.5) can be written as

x′i = xkCki . (2.6)

Page 5: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

3 . M u ltiplic atio n o f M atric es 29

This is because the set of three equations represented by equation (2.5) corresponding

to all values of j is the same set of equations represented by equation (2.6) for all

values of i.

It is easy to show that the components of x in the old coordinate system are

related to those in the rotated system by

xj = Cjix′i . (2.7)

N ote that the indicial positions on the right-hand side of this relation are different

from those in equation (2.5), because the first index of C is summed in equation (2.5),

whereas the second index of C is summed in equation (2.7).

We can now formally define a Cartesian vector as any quantity that transforms like

a position vector under the rotation of the coordinate system. Therefore, by analogy

with equation (2.5), u is a vector if its components transform as

u′j = uiCij . (2.8)

3 . M u ltiplic atio n o f M atric es

In this chapter we shall generally follow the convention that 3× 3 matrices are repre-

sented by uppercase letters, and column vectors are represented by lowercase letters.

(An exception will be the use of lowercase τ for the stress matrix.) L et A and B be

two 3×3 matrices. The product of A and B is defined as the matrix P whose elements

are related to those of A and B by

Pij =3

k=1

AikBkj ,

or, using the summation convention

Pij = AikBkj . (2.9)

Symbolically, this is written as

P = A • B. (2.10)

A single dot between A and B is included in equation (2.10) to signify that a single

index is summed on the right-hand side of equation (2.9). The important thing to note

in equation (2.9) is that the elements are summed over the inner or ad jacen t index k.

It is sometimes useful to write equation (2.9) as

Pij = AikBkj = (A • B)ij ,

where the last term is to be read as the “ ij -element of the product of matrices A

and B.”

Page 6: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

30 Cartes ian T en s o rs

In explicit form, equation (2.9) is written as

P11 P12 P13

P21 P22 P23

P31 P32 P33

=

A11 A12 A13

A21 A22 A23

A31 A32 A33

B11 B12 B13

B21 B22 B23

B31 B32 B33

(2.11)

N ote that equation (2.9) signifies that the ij -element of P is determined by multiplying

the elements in the i-row of A and the j -column of B, and summing. For example,

P12 = A11B12 + A12B22 + A13B32.

This is indicated by the dotted lines in equation (2.11). It is clear that we can define

the product A • B only if the number of columns of A equals the number of rows of B.

Equation (2.9) can be used to determine the product of a 3 × 3 matrix and a

vector, if the vector is written as a column. For example, equation (2.6) can be written

as x′i = CTikxk , which is now of the form of equation (2.9) because the summed index

k is adjacent. In matrix form equation (2.6) can therefore be written as

x′1x′2x′3

=

C11 C12 C13

C21 C22 C23

C31 C32 C33

T

x1

x2

x3

.

Symbolically, the preceding is

x′ = CT • x,

whereas equation (2.7) is

x = C • x′.

4 . S ec o n d - O rd er T en s o r

We have seen that scalars can be represented by a single number, and a Cartesian

vector can be represented by three numbers. There are other quantities, however, that

need more than three components for a complete description. For example, the stress

(equal to force per unit area) at a point in a material needs nine components for a

complete specification because tw o directions (and, therefore, tw o free indices) are

involved in its description. O ne direction specifies the orientation of the su rface on

which the stress is being sought, and the other specifies the direction of the force on

that surface. For example, the j -component of the force on a surface whose outward

normal points in the i-direction is denoted by τij . (H ere, we are departing from the

convention followed in the rest of the chapter, namely, that tensors are represented by

uppercase letters. It is customary to denote the stress tensor by the lowercase τ .) The

first index of τij denotes the direction of the normal, and the second index denotes

the direction in which the force is being projected.

This is shown in Figure 2.4, which gives the normal and shear stresses on an

infinitesimal cube whose surfaces are parallel to the coordinate planes. The stresses

Page 7: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

4 . S ec o n d - O rd er T en s o r 31

Fig ur e 2 .4 Stress field at a point. P ositive normal and shear stresses are shown. For clarity, the stresses

on faces FB CG and CD H G are not labeled.

are positive if they are directed as in this figure. The sign convention is that, on a

surface whose outward normal points in the positive direction of a coordinate axis,

the normal and shear stresses are positive if they point in the positive direction of

the axes. For example, on the surface AB CD , whose outward normal points in the

positive x2 direction, the positive stresses τ21, τ22, and τ23 point toward the x1, x2

and x3 directions, respectively. (Clearly, the normal stresses are positive if they are

tensile and negative if they are compressive.) O n the opposite face EFG H the stress

components have the same value as on AB CD , but their directions are reversed. This

is because Figure 2.4 shows the stresses at a p oin t. The cube shown is supposed to be

of “ zero” size, so that the faces AB CD and EFG H are just opposite faces of a plane

perpendicular to the x2-axis. That is why the stresses on the opposite faces are equal

and opposite.

Recall that a vector u can be completely specified by the three components ui

(where i = 1, 2, 3). We say “ completely specified” because the components of u in

any direction other than the original axes can be found from equation (2.8). Similarly,

the state of stress at a point can be completely specified by the nine components τij

(where i, j = 1, 2, 3), which can be written as the matrix

τ =

τ11 τ12 τ13

τ21 τ22 τ23

τ31 τ32 τ33

.

The specification of the preceding nine components of the stress on surfaces parallel

to the coordinate axes completely determines the state of stress at a point, because

Page 8: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

32 Cartes ian T en s o rs

the stresses on any arbitrary plane can then be determined. To find the stresses on any

arbitrary surface, we shall consider a rotated coordinate system x′1 x′2 x′3 one of whose

axes is perpendicular to the given surface. It can be shown by a force balance on a

tetrahedron element (see, e.g., Sommerfeld (1964), page 59) that the components of

τ in the rotated coordinate system are

τ ′mn = CimCjnτij . (2.12)

N ote the similarity between the transformation rule equation (2.8) for a vector, and the

rule equation (2.12). In equation (2.8) the first index of C is summed, while its second

index is free. The rule equation (2.12) is identical, except that this happens twice. A

quantity that obeys the transformation rule equation (2.12) is called a secon d -ord er

ten sor.

The transformation rule equation (2.12) can be expressed as a matrix product.

Rewrite equation (2.12) as

τ ′mn = CTmiτijCjn,

which, with adjacent dummy indices, represents the matrix product

τ′ = CT • τ • C.

This says that the tensor τ in the rotated frame is found by multiplying C by τ and

then multiplying the product by CT.

The concepts of tensor and matrix are not quite the same. A matrix is any arran g e-

m en t of elements, written as an array. The elements of a matrix represent the compo-

nents of a tensor only if they obey the transformation rule equation (2.12).

Tensors can be of any order. In fact, a scalar can be considered a tensor of zero

order, and a vector can be regarded as a tensor of first order. The number of free

indices correspond to the order of the tensor. For example, A is a fourth-order tensor

if it has four free indices, and the associated 81 components change under the rotation

of the coordinate system according to

A′mnpq = CimCjnCkpClqAijkl . (2.13)

Tensors of various orders arise in fluid mechanics. Some of the most frequently

used are the stress tensor τij and the velocity gradient tensor ∂ui/∂xj . It can be shown

that the nine products uivj formed from the components of the two vectors u and

v also transform according to equation (2.12), and therefore form a second-order

tensor. In addition, certain “ isotropic” tensors are also frequently used; these will be

discussed in Section 7.

5 . Co n trac tio n an d M u ltiplic atio n

When the two indices of a tensor are equated, and a summation is performed over

this repeated index, the process is called con traction . An example is

Page 9: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

6 . F o rc e o n a S u rfac e 33

Ajj = A11 + A22 + A33,

which is the sum of the diagonal terms. Clearly, Ajj is a scalar and therefore inde-

pendent of the coordinate system. In other words, Ajj is an in varian t. (There are

three independent invariants of a second-order tensor, and Ajj is one of them; see

Exercise 5.)

H igher-order tensors can be formed by multiplying lower tensors. If u and v are

vectors, then the nine components uivj form a second-order tensor. Similarly, if A

and B are two second-order tensors, then the 81 numbers defined by Pijkl ≡ AijBkl

transform according to equation (2.13), and therefore form a fourth-order tensor.

L ower-order tensors can be obtained by performing contraction on these multi-

plied forms. The four contractions of AijBkl are

AijBki = BkiAij = (B • A)kj ,

AijBik = ATjiBik = (AT • B)jk,

AijBkj = AijBTjk = (A • BT)ik,

AijBjk = (A • B)ik.

(2.14)

All four products in the preceding are second-order tensors. N ote in equation (2.14)

how the terms have been rearranged until the summed index is adjacent, at which

point they can be written as a product of matrices.

The contracted product of a second-order tensor A and a vector u is a vector. The

two possibilities are

Aijuj = (A • u)i,

Aijui = ATjiui = (AT • u)j .

The doubly contracted product of two second-order tensors A and B is a scalar. The

two possibilities are AijBji (which can be written as A :B in boldface notation) and

AijBij (which can be written as A :BT).

6 . F o rc e o n a S u rfac e

A surface area has a magnitude and an orientation, and therefore should be treated as

a vector. The orientation of the surface is conveniently specified by the direction of

a unit vector normal to the surface. If dA is the magnitude of an element of surface

and n is the unit vector normal to the surface, then the surface area can be written as

the vector

dA = n dA.

Suppose the nine components of the stress tensor with respect to a given set of

Cartesian coordinates are given, and we want to find the force per unit area on a

surface of given orientation n (Figure 2.5). O ne way of determining this is to take

a rotated coordinate system, and use equation (2.12) to find the normal and shear

stresses on the given surface. An alternative method is described in what follows.

Page 10: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

34 Cartes ian T en s o rs

Fig ur e 2 .5 Force f per unit area on a surface element whose outward normal is n.

Fig ur e 2 .6 (a) Stresses on surfaces of a two-dimensional element; (b) balance of forces on element AB C.

For simplicity, consider a two-dimensional case, for which the known stress

components with respect to a coordinate system x1 x2 are shown in Figure 2.6a. We

want to find the force on the face AC, whose outward normal n is known (Figure 2.6b).

Consider the balance of forces on a triangular element AB C, with sides AB = dx2,

B C = dx1, and AC = ds ; the thickness of the element in the x3 direction is unity. If

F is the force on the face AC, then a balance of forces in the x1 direction gives the

component of F in that direction as

F1 = τ11 dx2 + τ21 dx1.

Page 11: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

6 . F o rc e o n a S u rfac e 35

D ividing by ds , and denoting the force per unit area as f = F/ds , we obtain

f1 =F1

ds= τ11

dx2

ds+ τ21

dx1

ds= τ11 cos θ1 + τ21 cos θ2 = τ11n1 + τ21n2,

where n1 = cos θ1 and n2 = cos θ2 because the magnitude of n is unity (Figure 2.6b).

U sing the summation convention, the foregoing can be written as f1 = τj1nj , where

j is summed over 1 and 2. A similar balance of forces in the x2 direction gives

f2 = τj2nj . G eneralizing to three dimensions, it is clear that

fi = τjinj .

B ecause the stress tensor is symmetric (which will be proved in the next chapter),

that is, τij = τji , the foregoing relation can be written in boldface notation as

f = n • τ. (2.15)

Therefore, the contracted or “ inner” product of the stress tensor τ and the unit outward

vector n gives the force per unit area on a surface. Equation (2.15) is analogous to

un = u • n, where un is the component of the vector u along unit normal n; however,

whereas un is a scalar, f in equation (2.15) is a vector.

Exam p le 2 .1 . Consider a two-dimensional parallel flow through a channel. Take

x1, x2 as the coordinate system, with x1 parallel to the flow. The viscous stress tensor

at a point in the flow has the form

τ =[

0 a

a 0

]

,

where the constant a is positive in one half of the channel, and negative in the other

half. Find the magnitude and direction of force per unit area on an element whose

outward normal points at 30◦ to the direction of flow.

S olu tion b y u sin g eq u ation (2.15): B ecause the magnitude of n is 1 and it points

at 30◦ to the x1 axis (Figure 2.7), we have

n =[ √

3/2

1/2

]

.

The force per unit area is therefore

f = τ • n =[

0 a

a 0

] [ √3/2

1/2

]

=[

a/2√3 a/2

]

=[

f1

f2

]

.

The magnitude of f is

f = (f 21 + f 2

2 )1/2 = |a|.

Page 12: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

36 Cartes ian T en s o rs

Fig ur e 2 .7 D etermination of force on an area element (Example 2.1).

If θ is the angle of f with the x1 axis, then

sin θ =f2

f=√

3

2

a

|a|and cos θ =

f1

f=

1

2

a

|a|.

Thus θ = 60◦ if a is positive (in which case both sin θ and cos θ are positive), and

θ = 240◦ if a is negative (in which case both sin θ and cos θ are negative).

S olu tion b y u sin g eq u ation (2.12): Take a rotated coordinate system x′1, x′2,

with x′1 axis coinciding with n (Figure 2.7). U sing equation (2.12), the components

of the stress tensor in the rotated frame are

τ ′11 = C11C21τ12 + C21C11τ21 =√

32

12a + 1

2

√3

2a =

√3

2a,

τ ′12 = C11C22τ12 + C21C12τ21 =√

32

√3

2a − 1

212a = 1

2a.

The normal stress is therefore√

3 a/2, and the shear stress is a/2. This gives a

magnitude a and a direction 60◦ or 240◦ depending on the sign of a.

7 . K ro n ec k er D elta an d A ltern atin g T en s o r

The K ron eck er d elta is defined as

δij ={

1 if i = j

0 if i 6= j, (2.16)

which is written in the matrix form as

δ =

1 0 0

0 1 0

0 0 1

.

Page 13: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

8 . D o t P ro d u c t 37

The most common use of the K ronecker delta is in the following operation: If we

have a term in which one of the indices of δij is repeated, then it simply replaces the

dummy index by the other index of δij . Consider

δijuj = δi1u1 + δi2u2 + δi3u3.

The right-hand side is u1 when i = 1, u2 when i = 2, and u3 when i = 3. Therefore

δijuj = ui . (2.17)

From its definition it is clear that δij is an isotrop ic ten sor in the sense that its

components are unchanged by a rotation of the frame of reference, that is, δ′ij = δij .

Isotropic tensors can be of various orders. There is no isotropic tensor of first order,

and δij is the only isotropic tensor of second order. There is also only one isotropic

tensor of third order. It is called the altern atin g ten sor or p erm u tation sy m b ol, and is

defined as

εijk =

1 if ijk = 123, 231, or 312 (cyclic order),

0 if any two indices are equal,

−1 if ijk = 321, 213, or 132 (anticyclic order).

(2.18)

From the definition, it is clear that an in d ex on εijk can b e m oved tw o p laces (eith er

to th e rig h t or to th e left) w ith ou t ch an g in g its valu e. For example, εijk = εjki where

i has been moved two places to the right, and εijk = εkij where k has been moved

two places to the left. For a movement of one place, however, the sign is reversed.

For example, εijk = −εikj where j has been moved one place to the right.

A very frequently used relation is the ep silon d elta relation

εijkεklm = δilδjm − δimδj l . (2.19)

The reader can verify the validity of this relationship by taking some values for ij lm.

Equation (2.19) is easy to remember by noting the following two points: (1) The

adjacent index k is summed; and (2) the first two indices on the right-hand side,

namely, i and l, are the first index of εijk and the first free index of εklm. The remaining

indices on the right-hand side then follow immediately.

8 . D o t P ro d u c t

The dot product of two vectors u and v is defined as the scalar

u • v = v • u = u1v1 + u2v2 + u3v3 = uivi .

It is easy to show that u • v = uv cos θ , where u and v are the magnitudes and θ is the

angle between the vectors. The dot product is therefore the magnitude of one vector

times the component of the other in the direction of the first. Clearly, the dot product

u • v is equal to the sum of the diagonal terms of the tensor uivj .

Page 14: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

38 Cartes ian T en s o rs

9 . Cro s s P ro d u c t

The cross product between two vectors u and v is defined as the vector w whose

magnitude is uv sin θ , where θ is the angle between u and v, and whose direction is

perpendicular to the plane of u and v such that u, v, and w form a right-handed system.

Clearly, u × v = −v × u, and the unit vectors obey the cyclic rule a1× a2 = a3. It

is easy to show that

u× v = (u2v3 − u3v2)a1 + (u3v1 − u1v3)a

2 + (u1v2 − u2v1)a3, (2.20)

which can be written as the symbolic determinant

u× v =

a1 a2 a3

u1 u2 u3

v1 v2 v3

.

In indicial notation, the k-component of u× v can be written as

(u× v)k = εijkuivj = εkijuivj . (2.21)

As a check, for k = 1 the nonzero terms in the double sum in equation (2.21) result

from i = 2, j = 3, and from i = 3, j = 2. This follows from the definition

equation (2.18) that the permutation symbol is zero if any two indices are equal. Then

equation (2.21) gives

(u× v)1 = εij1uivj = ε231u2v3 + ε321u3v2 = u2v3 − u3v2,

which agrees with equation (2.20). N ote that the second form of equation (2.21) is

obtained from the first by moving the index k two places to the left; see the remark

below equation (2.18).

1 0 . O perato r ∇: G rad ien t, D iv erg en c e, an d Cu rl

The vector operator “ del” 1 is defined symbolically by

∇ ≡ a1 ∂

∂x1+ a2 ∂

∂x2+ a3 ∂

∂x3= ai ∂

∂xi

. (2.22)

When operating on a scalar function of position φ, it generates the vector

∇φ = ai ∂φ

∂xi

,

1The inverted G reek delta is called a “ nabla” (ν αβ λα). The origin of the word is from the H ebrew

(pronounced navel), which means lyre, an ancient harp-like stringed instrument. It was on this

instrument that the boy, D avid, entertained K ing Saul (Samuel II) and it is mentioned repeatedly

in P salms as a musical instrument to use in the praise of G od.

Page 15: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

1 0 . O perato r ∇: G rad ien t, D iv erg en c e, an d Cu rl 39

whose i-component is

(∇φ)i =∂φ

∂xi

.

The vector ∇φ is called the g rad ien t of φ. It is clear that ∇φ is perpendicular to the

φ = constant lines and gives the magnitude and direction of the m ax im u m spatial rate

of change of φ (Figure 2.8). The rate of change in any other direction n is given by

∂φ

∂n= (∇φ) • n.

The d iverg en ce of a vector field u is defined as the scalar

∇ • u ≡∂ui

∂xi

=∂u1

∂x1+

∂u2

∂x2+

∂u3

∂x3. (2.23)

So far, we have defined the operations of the gradient of a scalar and the diver-

gence of a vector. We can, however, generalize these operations. For example, we can

define the divergence of a second-order tensor τ as the vector whose i-component is

(∇ • τ)i =∂τij

∂xj

.

It is evident that the divergence operation d ecreases the order of the tensor by one.

In contrast, the gradient operation in creases the order of a tensor by one, changing

Fig ur e 2 .8 L ines of constant φ and the gradient vector ∇φ.

Page 16: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

40 Cartes ian T en s o rs

a zero-order tensor to a first-order tensor, and a first-order tensor to a second-order

tensor.

The cu rl of a vector field u is defined as the vector ∇ × u, whose i-component

can be written as (using equations (2.21) and (2.22))

(∇ × u)i = εijk

∂uk

∂xj

. (2.24)

The three components of the vector ∇ × u can easily be found from the right-hand

side of equation (2.24). For the i = 1 component, the nonzero terms in the double

sum in equation (2.24) result from j = 2, k = 3, and from j = 3, k = 2. The three

components of ∇ × u are finally found as

(

∂u3

∂x2−

∂u2

∂x3

)

,

(

∂u1

∂x3−

∂u3

∂x1

)

, and

(

∂u2

∂x1−

∂u1

∂x2

)

. (2.25)

A vector field u is called solen oid al if ∇ • u = 0, and irrotation al if ∇ × u = 0. The

word “ solenoidal” refers to the fact that the magnetic induction B always satisfies

∇ • B = 0. This is because of the absence of magnetic monopoles. The reason for the

word “ irrotational” will be clear in the next chapter.

1 1 . S y m m etric an d A n tis y m m etric T en s o rs

A tensor B is called sy m m etric in the indices i and j if the components do not change

when i and j are interchanged, that is, if Bij = Bji . The matrix of a second-order

tensor is therefore symmetric about the diagonal and made up of only six distinct

components. O n the other hand, a tensor is called an tisy m m etric if Bij = −Bji . An

antisymmetric tensor must have zero diagonal terms, and the off-diagonal terms must

be mirror images; it is therefore made up of only three distinct components. Any

tensor can be represented as the sum of a symmetric part and an antisymmetric part.

For if we write

Bij = 12(Bij + Bji)+ 1

2(Bij − Bji)

then the operation of interchanging i and j does not change the first term, but changes

the sign of the second term. Therefore, (Bij +Bji)/2 is called the symmetric part of

Bij , and (Bij − Bji)/2 is called the antisymmetric part of Bij .

Every vector can be associated with an antisymmetric tensor, and vice versa. For

example, we can associate the vector

ω =

ω1

ω2

ω3

,

with an antisymmetric tensor defined by

Page 17: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

1 2. E ig en v alu es an d E ig en v ec to rs o f a S y m m etric T en s o r 41

R ≡

0 −ω3 ω2

ω3 0 −ω1

−ω2 ω1 0

, (2.26)

where the two are related as

Rij= −εijkωk

ωk =− 12εijkRij .

(2.27)

As a check, equation (2.27) gives R11 = 0 and R12 = −ε123ω3 = −ω3, which is in

agreement with equation (2.26). (In Chapter 3 we shall call R the “ rotation” tensor

corresponding to the “ vorticity” vector ω.)

A very frequently occurring operation is the doubly contracted product of a

sy m m etric tensor τ and any tensor B. The doubly contracted product is defined as

P ≡ τijBij = τij (Sij + Aij ),

where S and A are the symmetric and antisymmetric parts of B, given by

Sij ≡1

2(Bij + Bji) and Aij ≡

1

2(Bij − Bji).

Then

P = τijSij + τijAij (2.28)

= τijSji − τijAji because Sij = Sji and Aij = −Aji,

= τjiSji − τjiAji because τij = τji,

= τijSij − τijAij interchanging dummy indices. (2.29)

Comparing the two forms of equations (2.28) and (2.29), we see that τijAij = 0, so

that

τijBij =1

2τij (Bij + Bji).

The important rule we have proved is that the d ou b ly con tracted p rod u ct of a sy m m etric

ten sor τ w ith an y ten sor B eq u als τ tim es th e sy m m etric p art of B. In the process,

we have also shown that the doubly contracted product of a symmetric tensor and an

antisymmetric tensor is zero. This is analogous to the result that the definite integral

over an even (symmetric) interval of the product of a symmetric and an antisymmetric

function is zero.

1 2. E ig en v alu es an d E ig en v ec to rs o f a S y m m etric T en s o r

The reader is assumed to be familiar with the concepts of eigenvalues and eigenvectors

of a matrix, and only a brief review of the main results is given here. Suppose τ is a

Page 18: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

42 Cartes ian T en s o rs

symmetric tensor with real elements, for example, the stress tensor. Then the following

facts can be proved:

(1) There are three real eigenvalues λk (k = 1, 2, 3), which may or may not be all

distinct. (The superscripted λk does not denote the k-component of a vector.)

The eigenvalues satisfy the third-degree equation

det |τij − λδij | = 0,

which can be solved for λ1, λ2, and λ3.

(2) The three eigenvectors bk corresponding to distinct eigenvalues λk are mutu-

ally orthogonal. These are frequently called the p rin cip al ax es of τ. Each b is

found by solving a set of three equations

(τij − λδij ) bj = 0,

where the superscript k on λ and b has been omitted.

(3) If the coordinate system is rotated so as to coincide with the eigenvectors, then

τ has a diagonal form with elements λk . That is,

τ′ =

λ1 0 0

0 λ2 0

0 0 λ3

in the coordinate system of the eigenvectors.

(4) The elements τij change as the coordinate system is rotated, but they cannot be

larger than the largest λ or smaller than the smallest λ. That is, the eigenvalues

are the extremum values of τij .

Exam p le 2 .2 . The strain rate tensor E is related to the velocity vector u by

Eij =1

2

(

∂ui

∂xj

+∂uj

∂xi

)

.

For a two-dimensional parallel flow

u =[

u1(x2)

0

]

,

show how E is diagonalized in the frame of reference coinciding with the principal

axes.

S olu tion : For the given velocity profile u1(x2), it is evident that E11 = E22 = 0,

and E12 = E21 = 12(du1/dx2) = 0. The strain rate tensor in the unrotated coordinate

system is therefore

E =[

0 0

0 0

]

.

Page 19: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

1 2. E ig en v alu es an d E ig en v ec to rs o f a S y m m etric T en s o r 43

Fig ur e 2 .9 O riginal coordinate system O x1 x2 and rotated coordinate system O x′1x′

2coinciding with

the eigenvectors (Example 2.2).

The eigenvalues are given by

det |Eij − λδij | =∣

−λ 0

0 −λ

= 0,

whose solutions are λ1 = 0 and λ2 = −0. The first eigenvector b1 is given by

[

0 0

0 0

] [

b11

b12

]

= λ1

[

b11

b12

]

,

whose solution is b11 = b1

2 = 1/√

2, thus normalizing the magnitude to unity. The

first eigenvector is therefore b1 = [1/√

2, 1/√

2], writing it in a row. The second

eigenvector is similarly found as b2 = [−1/√

2, 1/√

2]. The eigenvectors are shown

in Figure 2.9. The direction cosine matrix of the original and the rotated coordinate

system is therefore

C =

1√2− 1√

2

1√2

1√2

,

which represents rotation of the coordinate system by 45◦. U sing the transformation

rule (2.12), the components of E in the rotated system are found as follows:

E′12 = Ci1Cj2Eij = C11C22E12 + C21C12E21

=1√

2

1√

20 −

1√

2

1√

20 = 0

Page 20: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

44 Cartes ian T en s o rs

E′21 = 0

E′11 = Ci1Cj1Eij = C11C21E12 + C21C11E21 = 0

E′22 = Ci2Cj2Eij = C12C22E12 + C22C12E21 = −0

(Instead of using equation (2.12), all the components of E in the rotated system can be

found by carrying out the matrix product CT • E • C.) The matrix of E in the rotated

frame is therefore

E′ =[

0 0

0 −0

]

.

The foregoing matrix contains only diagonal terms. It will be shown in the next

chapter that it represents a linear stretching at a rate 0 along one principal axis, and a

linear compression at a rate −0 along the other; there are no shear strains along the

principal axes.

1 3 . G au s s ’ T heo rem

This very useful theorem relates a volume integral to a surface integral. L et V be a

volume bounded by a closed surface A. Consider an infinitesimal surface element

dA, whose outward unit normal is n (Figure 2.10). The vector n dA has a magnitude

dA and direction n, and we shall write dA to mean the same thing. L et Q(x) be a

scalar, vector, or tensor field of any order. G auss’ theorem states that

V

∂Q

∂xi

dV =∫

A

dAi Q. (2.30)

Fig ur e 2 .1 0 Illustration of G auss’ theorem.

Page 21: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

1 3 . G au s s ’ T heo rem 45

The most common form of G auss’ theorem is when Q is a vector, in which case the

theorem is

V

∂Qi

∂xi

dV =∫

A

dAi Qi,

which is called the d iverg en ce th eorem . In vector notation, the divergence theorem is

V

∇ • Q dV =∫

A

dA • Q.

P hysically, it states that the volume integral of the divergence of Q is equal to the

surface integral of the outflux of Q. Alternatively, equation (2.30), when considered

in its limiting form for an infintesmal volume, can define a generalized field derivative

of Q by the expression

DQ = limV→0

1

V

A

dAiQ. (2.31)

This includes the gradient, divergence, and curl of any scalar, vector, or tensor Q.

M oreover, by regarding equation (2.31) as a definition, the recipes for the computation

of the vector field derivatives may be obtained in any coordinate system. For a tensor

Q of any order, equation (2.31) as written defines the gradient. For a tensor of order

one (vector) or higher, the divergence is defined by using a dot (scalar) product under

the integral

div Q = limV→0

1

V

A

dA • Q, (2.32)

and the curl is defined by using a cross (vector) product under the integral

curl Q = limV→0

1

V

A

dA×Q. (2.33)

In equations (2.31), (2.32), and (2.33), A is the closed surface bounding the volume V.

Exam p le 2 .3 . O btain the recipe for the divergence of a vector Q(x) in cylindri-

cal polar coordinates from the integral definition equation (2.32). Compare with

Appendix B .1.

S olu tion : Consider an elemental volume bounded by the surfaces R −1R/2,

R + 1R/2, θ − 1θ/2, θ + 1θ/2, x − 1x/2 and x + 1x/2. The volume enclosed

1V is R1θ1R1x. We wish to calculate div Q = lim1V→01

1V

AdA • Q at the

central point R, θ , x by integrating the net outward flux through the bounding surface

A of 1V:

Q = iRQR(R, θ, x)+ iθ Qθ (R, θ, x)+ ixQx(R, θ, x).

Page 22: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

46 Cartes ian T en s o rs

In evaluating the surface integrals, we can show that in the limit taken, each of the

six surface integrals may be approximated by the product of the value at the center

of the surface and the surface area. This is shown by Taylor expanding each of the

scalar products in the two variables of each surface, carrying out the integrations, and

applying the limits. The result is

div Q = lim1R→01θ→01x→0

{

1

R1θ1R1x

[

QR

(

R +1R

2, θ, x

)(

R +1R

2

)

1θ1x

−QR

(

R −1R

2, θ, x

)(

R −1R

2

)

1θ1x

+Qx

(

R, θ, x+1x

2

)

R1θ1R −Qx

(

R, θ, x−1x

2

)

R1θ1R

+Q

(

R, θ +1θ

2, x

)

(

iθ − iR1θ

2

)

1R1x

+Q

(

R, θ −1θ

2, x

)

(

− iθ − iR1θ

2

)

1R1x

]}

,

where an additional complication arises because the normals to the two planes θ ±1θ/2 are not antiparallel:

Q

(

R, θ ±1θ

2, x

)

= QR

(

R, θ ±1θ

2, x

)

iR

(

R, θ ±1θ

2, x

)

+Qθ

(

R, θ ±1θ

2, x

)

(

R, θ ±1θ

2, x

)

+Qx

(

R, θ ±1θ

2, x

)

ix.

N ow we can show that

iR

(

θ ±1θ

2

)

= iR(θ)±1θ

2iθ (θ), iθ

(

θ ±1θ

2

)

= iθ (θ)∓1θ

2iR(θ).

Evaluating the last pair of surface integrals explicitly,

div Q = lim1R→01θ→01x→0

{

1

R1θ1R1x

[

QR

(

R +1R

2, θ, x

) (

R +1R

2

)

1θ1x

−QR

(

R −1R

2, θ, x

) (

R −1R

2

)

1θ1x

Page 23: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

1 4 . S to k es ’ T heo rem 47

+(

Qx

(

R, θ, x+1x

2

)

−Qx

(

R, θ, x−1x

2

) )

R1θ1R

+(

QR

(

R, θ +1θ

2, x

)

2−QR

(

R, θ +1θ

2, x

)

2

)

1R1x

+(

(

R, θ +1θ

2, x

)

−Qθ

(

R, θ −1θ

2, x

)

1R1x

−(

QR

(

R, θ −1θ

2, x

)

2−QR

(

R, θ −1θ

2, x

)

2

)

1R1x

]}

,

where terms of second order in the increments have been neglected as they will vanish

in the limits. Carrying out the limits, we obtain

div Q =1

R

∂R(RQR)+

1

R

∂Qθ

∂θ+

∂Qx

∂x.

H ere, the physical interpretation of the divergence as the net outward flux of a vector

field per unit volume has been made apparent by its evaluation through the integral

definition.

This level of detail is required to obtain the gradient correctly in these coordinates.

1 4 . S to k es ’ T heo rem

Stokes’ theorem relates a surface integral over an open surface to a line integral

around the boundary curve. Consider an open surface A whose bounding curve is C

(Figure 2.11). Choose one side of the surface to be the outside. L et ds be an element of

the bounding curve whose magnitude is the length of the element and whose direction

is that of the tangent. The positive sense of the tangent is such that, when seen from

the “ outside” of the surface in the direction of the tangent, the interior is on the left.

Then the theorem states that

A

(∇ × u) • dA =∫

C

u • ds, (2.34)

which signifies that the surface integral of the curl of a vector field u is equal to the

line integral of u along the bounding curve.

The line integral of a vector u around a closed curve C (as in Figure 2.11)

is called the “ circulation of u about C.” This can be used to define the curl of a

vector through the limit of the circulation integral bounding an infinitesmal surface

as follows:

n • curl u = limA→0

1

A

C

u • ds, (2.35)

Page 24: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

48 Cartes ian T en s o rs

Fig ur e 2 .1 1 Illustration of Stokes’ theorem.

where n is a unit vector normal to the local tangent plane of A. The advantage of the

integral definitions of the field derivatives is that they may be applied regardless of

the coordinate system.

Exam p le 2 .4 . O btain the recipe for the curl of a vector u(x) in Cartesian coordinates

from the integral definition given by equation (2.35).

S olu tion : This is obtained by considering rectangular contours in three perpen-

dicular planes intersecting at the point (x, y, z). First, consider the elemental rectangle

in the x = const. plane. The central point in this plane has coordinates (x, y, z) and

the area is 1y 1z . It may be shown by careful integration of a Taylor expansion of

the integrand that the integral along each line segment may be represented by the

product of the integrand at the center of the segment and the length of the segment

with attention paid to the direction of integration ds. Thus we obtain

(curl u)x = lim1y→01z→0

{

1

1y1z

[

uz

(

x, y+1y

2, z

)

− uz

(

x, y−1y

2, z

) ]

1z

+1

1y1z

[

uy

(

x, y, z−1z

2

)

− uy

(

x, y, z+1z

2

) ]

1y

}

.

Taking the limits,

(curl u)x =∂uz

∂y−

∂uy

∂z.

Page 25: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

1 6 . B o ld fac e v s I n d ic ial N o tatio n 49

Similarly, integrating around the elemental rectangles in the other two planes

(curl u)y =∂ux

∂z−

∂uz

∂x,

(curl u)z =∂uy

∂x−

∂ux

∂y.

1 5 . Co m m a N o tatio n

Sometimes it is convenient to introduce the notation

A,i ≡∂A

∂xi

, (2.36)

where A is a tensor of any order. In this notation, therefore, the comma denotes a

spatial derivative. For example, the divergence and curl of a vector u can be written,

respectively, as

∇ • u =∂ui

∂xi

= ui,i,

(∇ × u)i = εijk

∂uk

∂xj

= εijkuk,j .

This notation has the advantages of economy and that all subscripts are written on

one line. Another advantage is that variables such as ui,j “ look like” tensors, which

they are, in fact. Its disadvantage is that it takes a while to get used to it, and that

the comma has to be written clearly in order to avoid confusion with other indices

in a term. The comma notation has been used in the book only in two sections, in

instances where otherwise the algebra became cumbersome.

1 6 . B o ld fac e v s I n d ic ial N o tatio n

The reader will have noticed that we have been using both boldface and indicial nota-

tions. Sometimes the boldface notation is loosely called “ vector” or d y ad ic notation,

while the indicial notation is called “ tensor” notation. (Although there is no reason

why vectors cannot be written in indicial notation!). The advantage of the boldface

form is that the physical meaning of the terms is generally clearer, and there are no

cumbersome subscripts. Its disadvantages are that algebraic manipulations are dif-

ficult, the ordering of terms becomes important because A • B is not the same as

B • A, and one has to remember formulas for triple products such as u× (v×w) and

u • (v × w). In addition, there are other problems, for example, the order or rank of

a tensor is not clear if one simply calls it A, and sometimes confusion may arise in

products such as A • B where it is not immediately clear w h ich index is summed. To

add to the confusion, the singly contracted product A • B is frequently written as AB

in books on matrix algebra, whereas in several other fields AB usually stands for the

uncontracted fourth-order tensor with elements Aij Bkl .

Page 26: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

50 Cartes ian T en s o rs

The indicial notation avoids all the problems mentioned in the preceding. The

algebraic manipulations are especially simple. The ordering of terms is unneces-

sary because Aij Bkl means the same thing as BklAij . In this notation we deal with

com p on en ts only, which are scalars. Another major advantage is that one does not

have to remember formulas except for the product εijkεklm, which is given by equa-

tion (2.19). The disadvantage of the indicial notation is that the physical meaning of a

term becomes clear only after an examination of the indices. A second disadvantage

is that the cross product involves the introduction of the cumbersome εijk . This, how-

ever, can frequently be avoided by writing the i-component of the vector product of u

and v as (u× v)i using a mixture of boldface and indicial notations. In this book we

shall use boldface, indicial and mixed notations in order to take advantage of each. As

the reader might have guessed, the algebraic manipulations will be performed mostly

in the indicial notation, sometimes using the comma notation.

E x erc is es

1. U sing indicial notation, show that

a × (b× c) = (a • c)b− (a • b)c.

[H in t: Call d ≡ b × c. Then (a × d)m = εpqmapdq = εpqmapεijqbicj . U sing

equation (2.19), show that (a × d)m = (a • c)bm − (a • b)cm.]

2. Show that the condition for the vectors a, b, and c to be coplanar is

εijkaibj ck = 0.

3. P rove the following relationships:

δij δij = 3

εpqr εpqr = 6

εpqiεpqj = 2δij .

4. Show that

C • CT = CT • C = δ,

where C is the direction cosine matrix and δ is the matrix of the K ronecker delta.

Any matrix obeying such a relationship is called an orth og on al m atrix because it

represents transformation of one set of orthogonal axes into another.

5. Show that for a second-order tensor A, the following three quantities are

invariant under the rotation of axes:

I1 = Aii

I2 =∣

A11 A12

A21 A22

+∣

A22 A23

A32 A33

+∣

A11 A13

A31 A33

I3 = det(Aij ).

Page 27: Fluid Mechanics Kundu Cohen · 2020-01-13 · 2. R o tatio n o f A x es: F o rm al D e n itio n o f a V ecto r 27 Fig ure 2 .2 Rotation of coordinate system O 1 2 3 to O 102030. L

S u pplem en tal R ead in g 51

[H in t: U se the result of Exercise 4 and the transformation rule (2.12) to show that

I ′1 = A′ii = Aii = I1. Then show that AijAji and AijAjkAki are also invariants. In

fact, all contracted scalars of the form AijAjk · · ·Ami are invariants. Finally, verify

that

I2 = 12[I 2

1 − AijAji]I3 = AijAjkAki − I1AijAji + I2Aii .

B ecause the right-hand sides are invariant, so are I2 and I3.]

6. If u and v are vectors, show that the products uivj obey the transformation

rule (2.12), and therefore represent a second-order tensor.

7. Show that δij is an isotropic tensor. That is, show that δ′ij = δij under rotation

of the coordinate system. [H in t: U se the transformation rule (2.12) and the results of

Exercise 4.]

8. O btain the recipe for the gradient of a scalar function in cylindrical polar

coordinates from the integral definition.

9. O btain the recipe for the divergence of a vector in spherical polar coordinates

from the integral definition.

10. P rove that div(curl u) = 0 for any vector u regardless of the coordinate

system. [H int: use the vector integral theorems.]

11. P rove that curl(grad φ) = 0 for any single-valued scalar φ regardless of the

coordinate system. [H int: use Stokes’ theorem.]

L iteratu re Cited

Sommerfeld, A. (1964). M ech an ics of D eform ab le B od ies, N ew Y ork: Academic P ress. (Chapter 1 contains

brief but useful coverage of Cartesian tensors.)

S u pplem en tal R ead in g

Aris, R. (1962). V ectors, T en sors, an d th e B asic E q u ation s of F lu id M ech an ics, Englewood Cliffs, N J :

P rentice-H all. (This book gives a clear and easy treatment of tensors in Cartesian and non-Cartesian

coordinates, with applications to fluid mechanics.)

P rager, W. (1961). I n trod u ction to M ech an ics of C on tin u a, N ew Y ork: D over P ublications. (Chapters 1

and 2 contain brief but useful coverage of Cartesian tensors.)