25
Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of elementary particles 1 that interact through mutual attraction or repulsion. In our everyday life, however, we do not see the elemental nature of matter, we see continuous regions of material such as the air in a room, the water in a glass or the wood of a desk. Even in a gas, the number of individual molecules in macroscopic volumes is enormous 10 19 cm 3 . Hence, when describing the behaviour of matter on scales of microns or above it is simply not practical to solve equations for each individual molecule. Instead, theoretical models are derived using average quantities that represent the state of the material. In general, these average quantities will vary with position in space, but an important concept to bear in mind is that the fundamental behaviour of matter should not depend on the particular coor- dinate system chosen to represent the position. The consequences of this almost trivial observation are far-reaching and we shall nd that it dictates the form of many of our governing equations. We shall always consider our space to be three-dimensional and Euclidean 2 and we describe position in space by a position vector, r, which runs from a specic point, the origin, to our chosen location. The exact coordinate system chosen will depend on the problem under consideration; ideally it should make the problem as “easy” as possible. 1.1 Vectors A vector is a geometric quantity that has a “magnitude and direction”. A more mathematically precise, but less intuitive, denition is that a vector is an element of a vector space. Many physical quantities are naturally described in terms of vectors, e.g. position, velocity and acceleration, force. The invariance of material behaviour under changes in coordinates means that if a vector represents a physical quantity then it must not vary if we change our coordinate system. Imagine drawing a line that connects two points in two-dimensional (Euclidean) plane, that line remains unchanged whether we describe it as “x units across and y units up from the origin” or “r units from the origin in the θ direction”. Thus, a vector is an object that exists independent of any coordinate system, but if we wish to describe it we must choose a specic coordinate system and its representation in that coordinate system (its components) will depend on the specic coordinates chosen. 1 The exact nature of the most basic unit is, of course, still debated, but the fundamental discrete nature of matter is not. 2 We won’t worry about relativistic eects at all. 7

Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

  • Upload
    others

  • View
    33

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

Chapter 1

Describing the Physical World:Vectors & Tensors

Lecture 1It is now well established that all matter consists of elementary particles1 that interact throughmutual attraction or repulsion. In our everyday life, however, we do not see the elemental natureof matter, we see continuous regions of material such as the air in a room, the water in a glass orthe wood of a desk. Even in a gas, the number of individual molecules in macroscopic volumes isenormous ≈ 1019 cm−3. Hence, when describing the behaviour of matter on scales of microns orabove it is simply not practical to solve equations for each individual molecule. Instead, theoreticalmodels are derived using average quantities that represent the state of the material.

In general, these average quantities will vary with position in space, but an important concept tobear in mind is that the fundamental behaviour of matter should not depend on the particular coor-dinate system chosen to represent the position. The consequences of this almost trivial observationare far-reaching and we shall find that it dictates the form of many of our governing equations. Weshall always consider our space to be three-dimensional and Euclidean2 and we describe position inspace by a position vector, r, which runs from a specific point, the origin, to our chosen location.The exact coordinate system chosen will depend on the problem under consideration; ideally itshould make the problem as “easy” as possible.

1.1 Vectors

A vector is a geometric quantity that has a “magnitude and direction”. A more mathematicallyprecise, but less intuitive, definition is that a vector is an element of a vector space. Many physicalquantities are naturally described in terms of vectors, e.g. position, velocity and acceleration, force.The invariance of material behaviour under changes in coordinates means that if a vector representsa physical quantity then it must not vary if we change our coordinate system. Imagine drawing aline that connects two points in two-dimensional (Euclidean) plane, that line remains unchangedwhether we describe it as “x units across and y units up from the origin” or “r units from the originin the θ direction”. Thus, a vector is an object that exists independent of any coordinate system,but if we wish to describe it we must choose a specific coordinate system and its representation inthat coordinate system (its components) will depend on the specific coordinates chosen.

1The exact nature of the most basic unit is, of course, still debated, but the fundamental discrete nature of matteris not.

2We won’t worry about relativistic effects at all.

7

Page 2: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

1.1.1 Cartesian components and summation convention

The fact that we have a Euclidean space means that we can always choose a Cartesian coordinatesystem with fixed orthonormal base vectors, e1 = i, e2 = j and e3 = k. For a compact notation, itis much more convenient to use the numbered subscripts rather than different symbols to distinguishthe base vectors. Any vector quantity a can be written as a sum of its components in the directionof the base vectors

a = a1e1 + a2e2 + a3e3; (1.1)

and the vector a can be represented via its components (a1, a2, a3); and so, e1 = (1, 0, 0), e2 =(0, 1, 0) and e3 = (0, 0, 1). We will often represent the components of vectors using an index, i.e.(a1, a2, a3) is equivalent to aI , where I ∈ {1, 2, 3}. In addition, we use the Einstein summationconvention in which any index that appears twice represents a sum over all values of that index

a = aJeJ =3X

J=1

aJeJ . (1.2)

Note that we can change the (dummy) summation index without affecting the result

aJeJ =3X

J=1

aJeJ =3X

K=1

aKeK = aKeK .

The summation is ambiguous if an index appears more than twice and such terms are not allowed.For clarity later, an upper case index is used for objects in a Cartesian (or, in fact, any orthonormal)coordinate system and, in general, we will insist that summation can only occur over a raised indexand a lowered index for reasons that will hopefully become clear shortly.

It is important to recognise that the components of a vector aI do not actually make sense unlesswe know the base vectors as well. In isolation the components give you distances but not direction,which is only half the story.

1.1.2 Curvilinear coordinate systems

For a complete theoretical development, we shall consider general coordinate systems3. Unfortu-nately the use of general coordinate systems introduces considerable complexity because the lines onwhich individual coordinates are constant are not necessarily straight lines nor are they necessarilyorthogonal to one another. A consequence is that the base vectors in general coordinate systemsare not orthonormal and vary as functions of position.

1.1.3 Tangent (covariant base) vectors

The position vector r = xKeK , is a function of the Cartesian coordinates, xK , where x1 = x, x2 = yand x3 = z. Note that the Cartesian base vectors can be recovered by differentiating the positionvector with respect to the appropriate coordinate

∂r

∂xK= eK . (1.3)

3For our purposes a coordinate system to be a set of independent scalar variables that can be used to describeany position in the entire Euclidean space.

Page 3: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

In other words the derivative of position with respect to a coordinate returns a vector tangent to thecoordinate direction; a statement that is true for any coordinate system. For a general coordinatesystem, ξi, we can write the position vector as

r(ξ1, ξ2, ξ3) = xK(ξi)eK , (1.4)

because the Cartesian base vectors are fixed. Here the notation xK(ξi) means that the Cartesiancoordinates can be written as functions of the general coordinates, e.g. in plane polars x(r, θ) =r cos θ, y(r, θ) = r sin θ, see Example 1.1. Note that equation (1.4) is the first time in which we useupper and lower case indices to distinguish between the Cartesian and general coordinate systems.A tangent vector in the ξ1 direction, t1, is the difference between two position vectors associatedwith a small (infinitesimal) change in the ξ1 coordinate

t1(ξi) = r(ξi + dξ1)− r(ξi), (1.5)

where dξ1 represents the small change in the ξ1 coordinate direction, see Figure 1.1.

O

r(ξi)r(ξi + dξ1)

t1(ξi) ξ2 constant

Figure 1.1: Sketch illustrating the tangent vector t1(ξ) corresponding to a small change dξ1 in thecoordinate ξ1. The tangent lies along a line of constant ξ2 in two dimensions, or a plane of constantξ2 and ξ3 in three dimensions.

Assuming that r is differentiable and Taylor expanding the first term in (1.5) demonstrates that

t1 = r(ξi) +∂r

∂ξ1dξ1 − r(ξi) +O((dξi)2),

which yields

t1 =∂r

∂ξ1dξ1,

if we neglect the (small) quadratic and higher-order terms. Note that exactly the same argumentcan be applied to increments in the ξ2 and ξ3 directions and because dξi are scalar lengths, it followsthat

gi =∂r

∂ξi

is also a tangent vector in the ξi direction, as claimed. Hence, using equation (1.4), we can computetangent vectors in the general coordinate directions via

gi =∂r

∂ξi=

∂xK

∂ξieK . (1.6)

Page 4: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

We can interpret equation (1.6) as defining a local linear transformation between the Cartesianbase vectors and our new tangent vectors gi. The transformation is linear because gi is a linearcombination of the vectors eK , which should not be a surprise because we explicitly neglected thequadratic and higher terms in the Taylor expansion. The transformation is local because, in general,the coefficients will change with position. The coefficients of the transformation can be written asentries in a matrix, M, in which case equation (1.6) becomes

g1

g2

g3

=

Mz }| {

∂x1

∂ξ1∂x2

∂ξ1∂x3

∂ξ1

∂x1

∂ξ2∂x2

∂ξ2∂x3

∂ξ2

∂x1

∂ξ3∂x2

∂ξ3∂x3

∂ξ3

e1

e2

e3

.

(1.7)

Provided that the transformation is non-singular (the determinant of the matrix M is non-zero)the tangent vectors will also be a basis of the space and they are called covariant base vectorsbecause the transformation preserves the tangency of the vectors to the coordinates. In general,the covariant base vectors are neither orthogonal nor of unit length. It is also important to notethat the covariant base vectors will usually be functions of position.

Example 1.1. Finding the covariant base vectors for plane polar coordinatesA plane polar coordinate system is defined by the two coordinates ξ1 = r, ξ2 = θ such that

x = x1 = r cos θ and y = x2 = r sin θ.

Find the covariant base vectors.

Solution 1.1. The position vector is given by

r = x1e1 + x2e2 = r cos θe1 + r sin θe2 = ξ1 cos ξ2e1 + ξ1 sin ξ2e2,

and using the definition (1.6) gives

g1 =∂r

∂ξ1= cos ξ2e1 + sin ξ2e2, and g2 =

∂r

∂ξ2= −ξ1 sin ξ2e1 + ξ1 cos ξ2e2.

Note that g1 is a unit vector,

|g1| =√g1·g1 =

qcos2 ξ2 + sin2 ξ2 = 1,

but g2 is not, |g2| = ξ1. The vectors are orthogonal g1·g2 = 0 and can be are related to the standardorthonormal polar base vectors via g1 = er and g2 = reθ.

1.1.4 Contravariant base vectorsLecture 2

The fact that the covariant basis is not necessarily orthonormal makes life somewhat awkward. Fororthonormal systems we are used to the fact that when a = aKeK , then unique components can beobtained via a dot product4.

a·eI = aKeK·eI = aI , (1.8)

4The dot or scalar product is an operation on two vectors that returns a unique scalar: the product of the lengthsof the two vectors and the cosine of the angle between them. In the present context we only need to know thatfor orthogonal vectors the dot product is zero and the dot product of a unit vector with itself is one, see §1.3.1 forfurther discussion.

Page 5: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

where the last equality is a consequence of the orthonormality. In our covariant basis, we wouldwrite a = akgk, so that

a·gi = akgk·gi, (1.9)

but no further simplification can be made. Equations (1.9) are a linear system of simultaneousequations that must be solved in order to find the values of ak, which is considerably more effortthan using an explicit formula such as (1.8). The explicit formula (1.8) arises because the matrixeK·eI is diagonal, which means that the equations decouple and are no longer simultaneous.

We can, however, recover most of the nice properties of an orthonormal coordinate system if wedefine another set of base vectors that are each orthogonal to two of the covariant base vectors andhave unit length when projected in the direction of the remaining covariant base vector. In otherwords, the new vectors gi are defined such that

gi·gj = δij ≡

(1, if i = j,

0, otherwise,(1.10)

where the object δij is known as the Kronecker delta. In orthonormal coordinate systems the twosets of base vectors coincide; for example, in our global Cartesian coordinates eI ≡ eI .

We can decompose gi into its components in the Cartesian basis gi = giKeK , where we have

used the raised index on the base vectors for consistency with our summation convention. Notethat giK is thus defined to be the K-th Cartesian component of the i-th contravariant base vector.From the definition (1.10) and (1.6)

�giKe

K�·

�∂xL

∂ξjeL

�= giK

∂xL

∂ξjeK

·eL = giK∂xL

∂ξjδKL = giL

∂xL

∂ξj= δij. (1.11)

Note that we have used the “index-switching” property of the Kronecker delta to write giKδKL = giL,

which can be verified by writing out all terms explicitly.Multiplying both side of equation (1.11) by ∂ξj

∂xK yields

giL∂xL

∂ξj∂ξj

∂xK= δij

∂ξj

∂xK=

∂ξi

∂xK;

and from the chain rule∂xL

∂ξj∂ξj

∂xK=

∂xL

∂xK= δLK ,

because the Cartesian coordinates are independent. Hence,

giLδLK = giK =

∂ξi

∂xK,

and so the new set of base vectors are

gi =∂ξi

∂xKeK . (1.12)

The equation (1.12) defines a local linear transformation between the Cartesian base vectors andthe vectors gi. In a matrix representation, equation (1.12) is

g1

g2

g3

=

M−T

z }| {

∂ξ1

∂x1

∂ξ1

∂x2

∂ξ1

∂x3

∂ξ2

∂x1

∂ξ2

∂x2

∂ξ2

∂x3

∂ξ3

∂x1

∂ξ3

∂x2

∂ξ3

∂x3

e1

e2

e3

,

(1.13)

Page 6: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

and we see that the new transformation is the inverse transpose5 of the linear transformation thatdefines the covariant base vectors (1.6). For this reason, the vectors gi are called contravariant basevectors.

Example 1.2. Finding the contravariant base vectors for plane polar coordinatesFor the plane polar coordinate system defined in Example 1.1, find the contravariant base vectors.

Solution 1.2. The contravariant base vectors are defined by equation (1.12) and in order to use thatequation directly, we must express our polar coordinates as functions of the Cartesian coordinates

r = ξ1 =√x1x1 + x2x2, and tan θ = tan ξ2 =

x2

x1,

and then we can compute

∂ξ1

∂x1= cos ξ2,

∂ξ1

∂x2= sin ξ2,

∂ξ2

∂x1= −sin ξ2

ξ1and

∂ξ2

∂x1=

cos ξ2

ξ1

Thus, using the transformation (1.13), we have

g1 =∂ξ1

∂x1e1 +

∂ξ1

∂x2e2 = cos ξ2e1 + sin ξ2e2 = g1,

where we have used the fact that eI = eI , and also

g2 =∂ξ2

∂x1e1 +

∂ξ2

∂x2e2 = −sin ξ2

ξ1e1 +

cos ξ2

ξ1e2 =

1

(ξ1)2g2.

We can now easily verify that gi·gj = δij.An alternative (and often easier) approach is to find the contravariant base vectors by finding

the inverse transpose of the matrix M that defines the covariant base vectors and using equation(1.13).

1.1.5 Components of vectors in covariant and contravariant bases

We can find the components of a vector a in the covariant basis by taking the dot product with theappropriate contravariant base vectors

a = akgk, where ai = a·gi�= akgk·g

i = akδik = ai.�

(1.14)

Similarly components of the vector a in the contravariant basis are given by taking the dot productwith the appropriate covariant base vectors

a = akgk, where ai = a·gi

�= akg

k·gi = akδ

ki = ai.

�(1.15)

5That the inverse matrix is given by

M−1 =

∂ξ1

∂x1

∂ξ2

∂x1

∂ξ3

∂x1

∂ξ1

∂x2

∂ξ2

∂x2

∂ξ3

∂x2

∂ξ1

∂x3

∂ξ2

∂x3

∂ξ3

∂x3

,

can be confirmed by checking that MM−1 = M−1M = I, the identity matrix. Alternatively, the relationship followsdirectly from equation (1.11) written in matrix form.

Page 7: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

In fact, we can obtain the components of a general vector in either the covariant or contravariantbasis directly from the Cartesian coordinates. If a = aKeK = aKe

K , then the components in thecovariant basis associated with the curvilinear coordinates ξi are

ai = a·gi = aKeK·gi = aK

∂ξi

∂xJeJ

·eK = aK∂ξi

∂xJδJK = aK

∂ξi

∂xK,

a contravariant transform. Similarly, the components of the vector in the contravariant basis maybe obtained by covariant transform from the Cartesian components and so

ai =∂xK

∂ξiaK (1.16a)

and

ai =∂ξi

∂xKaK . (1.16b)

1.1.6 Invariance of vectors:(significance of index position)

Having established the need for two different types of transformations in curvilinear coordinatesystems, we are now in a position to consider the significance of the raised and lowered indicesin our summation convention. We shall insist that for an index to be lowered the object musttransform covariantly under a change in coordinates and for an index to be raised the object musttransform contravariantly under a change in coordinates6. An important exception to this ruleare the coordinates themselves: ξi represents the three scalar coordinates, e.g. in spherical polarcoordinates ξ1 = r, ξ2 = θ and ξ3 = φ; ξi are not the components of a vector and do not obeycontravariant transformation rules. Equation (1.16a) demonstrates that the components of a vectorin the contravariant basis are indeed covariant, justifying the lowered index, and equation (1.16b)provides similar justification for the contravariance of components in the covariant basis. We shallnow demonstrate that these transformation properties also follow directly from the requirementthat a physical vector should be independent of the coordinate system.

Consider a vector a, which can be written in the covariant or contravariant basis

a = aigi = aigi. (1.17)

We now consider a change in coordinates from ξi to another general coordinate system χi. It willbe of vital importance later on to know which index corresponds to which coordinate system sowe have chosen to add an overbar to the index to distinguish components associated with the twocoordinate systems, ξi and χi. The covariant base vectors associated with χi are then

gi ≡∂r

∂χi=

∂xJ

∂χieJ =

∂ξk

∂χi

∂xJ

∂ξkeJ =

∂ξk

∂χigk; (1.18)

and the transformation between gi and gk is of the same (covariant) type as that between gi and eK

in equation (1.6). The transformation is covariant because the “new” coordinate is the independent

6The logic for the choice of index location is the position of the generalised coordinate in the partial derivativedefining the transformation:

gi =∂xK

∂ξieK (lowered index), gi =

∂ξi

∂xKeK (raised index).

Page 8: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

variable in the partial derivative (it appears in the denominator). In our new basis, the vectora = ai gi and because a must remain invariant

a = ai gi = aigi.

Using the transformation of the base vectors (1.18) to replace gi gives

ai∂ξk

∂χigk = aigi = akgk ⇒ ak =

∂ξk

∂χiai. (1.19)

Hence, the components of the vector must transform contravariantly because multiplying both sidesof equation (1.19) by the inverse transpose transformation ∂χj/∂ξk gives

ai =∂χi

∂ξkak. (1.20)

This transformation is contravariant because the “new” coordinate is the dependent variable in thepartial derivative (it appears in the numerator).

A similar approach can be used to show that the components in the contravariant basis musttransform covariantly in order to ensure that the vector remains invariant. Thus, the use of oursummation convention ensures that the summed quantities remain invariant under coordinate trans-formations, which will be essential when deriving coordinate-independent physical laws.

Interpretation

The fact that base vectors and vector components must transform differently for the vector toremain invariant is actually quite obvious. Consider a one-dimensional Euclidean space in whicha = a1g1. If the base vector is rescaled7 by a factor λ so that g1 = λg1 then to compensate thecomponent must be rescaled by the factor 1/λ: a1 = 1

λa1. Note that for a 1 × 1 transformation

matrix with entry λ, the inverse transpose is 1/λ.

1.1.7 Orthonormal coordinates

If the coordinates are orthonormal then, by construction, there is no distinction between the co-variant and contravariant basis, gi = gi. Using equations (1.6) and (1.12), we see that

gi =∂xK

∂ξieK = gi =

∂ξi

∂xKeK ,

and so∂xK

∂ξi=

∂ξi

∂xK. (1.21)

Hence, the covariant and contravariant transformations are identical in orthonormal coordinatesystems, which means that there is no need to distinguish between raised and lowered indices. Thissimplification is adopted in many textbooks and the convention is to use only lowered indices. Whenworking with orthonormal coordinates we will also adopt this convention for simplicity, but we mustalways make sure that we know when the coordinate system is orthonormal. It is for this reasonthat we have adopted the convention that upper case indices are used for orthonormal coordinates.

7In one dimension all we can do is rescale the length, although the scaling can vary with position.

Page 9: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

If the coordinate system is not known to be orthonormal, we will use lower case indices and mustdistinguish between the covariant and contravariant transformations.

Condition (1.21) implies that

∂xK

∂ξj∂xK

∂ξi=

∂xK

∂ξj∂ξi

∂xK= δij. (1.22)

In the matrix representation, equation (1.22) is

MMT = I ⇒ MTM = I,

where I is the identity matrix. In other words the components of the transformation form anorthogonal matrix. It follows that (all) orthonormal coordinates can only be generated by anorthogonal transformation from the reference Cartesians. This should not be a big surprise: anyother transform will change the angles between the base vectors or their relative lengths whichdestroys orthonormality. The argument is entirely reversible: if either the covariant or contravarianttransform is orthogonal then the two transforms are identical and the new coordinate system isorthonormal.

An aside

Further intuition for the reason why the covariant and contravariant transformations are identicalwhen the coordinate transform is orthogonal can be obtained as follows. Imagine that we have ageneral linear transformation represented as a matrix M the acts on vectors such that componentsin the fixed Cartesian coordinate system p = pKeK transform as follows

pK = MKJ pJ .

Note that the index K does not have an overbar because pK is a component in the fixed Cartesiancoordinate system, eK . The transformation can, of course, also be applied to the base vectors ofthe fixed Cartesian coordinate system eI ,

[eI ]K = MK

J [eI ]J ,

where [ ]K indicates the K-th component of the base vector. Now, [eI ]J = δJI and it follows that

[eI ]K = MK

I ,

which allows us to define the operation of the matrix components on the base vectors directlybecause

eI = [eI ]KeK = MK

I eK . (1.23a)

Thus the operation of the transformation on the components is the transpose of its operation onthe base vectors8. We could write the new base vectors as eI = MK

IeK to be consistent with our

previous notation, but this will probably lead to more confusion in the current exposition.Now consider a vector a that must remain invariant under our transformation. Let the vector

a′ be the vector with the same numerical values of its components as a but with transformed basevectors, i.e. a′ = aK eK . Thus, the vector a′ will be a transformed version of a. In order to ensurethat the vector remains unchanged under transformation we must apply the appropriate inverse

8This statement also applies to general bases.

Page 10: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

transformation to a′ relative to the new base vectors, eI . In other words, the transformation of thecoordinates must be

aK = [M−1]KJ a′J = [M−1]KJ aJ , (1.23b)

where we have used the fact that a′J = aJ by definition. Using the two transformation equations(1.23a,b) we see that

aK eK = [M−1]KJ aJMLKeL = [M−1]KJ ML

KaJeL = δLJ a

JeL = aJeJ ,

as required.Thus, we have the two results: (i) a general property of linear transformations is that the matrix

representation of the transformation of vector components is the transpose of the matrix represen-tation of the transformation of base vectors; (ii) in order to remain invariant the transform of thecomponents of the vector must actually undergo the inverse of the coordinate transformation. Thus,the transformations of the base vectors and the coordinates coincide when the inverse transform isequal to its transpose, i.e. when the transform is orthogonal.

If that all seems a bit abstract, then hopefully the following specific example will help make theideas a little more concrete.

Example 1.3. Equivalence of covariant and contravariant transformations under affinetransformationsConsider a two-dimensional Cartesian coordinate system with base vectors eI . A new coordinatesystem with base vectors eI is obtained by rotation through an angle θ in the anticlockwise directionabout the origin. Derive the transformations for the base vectors and components of a general vectorand show that they are the same.

Solution 1.3. The original and rotated bases are shown in Figure 1.2(a) from which we determinethat the new base vectors are given by

e1 = cos θe1 + sin θe2 and e2 = − sin θe1 + cos θe2.

e1

e2

e1

e2

θ

θ

(a) e1

e2

e1

e2

θp

p′

(b)

Figure 1.2: (a) The base vectors eI are the Cartesian base vectors eI rotated through an angle θabout the origin. (b) If the coordinates of the position vector p are unchanged it is also rotated byθ to p′.

Page 11: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

Consider a position vector p = pI eI in the original basis. If we leave the coordinates unchangedthen the new vector p′ = pI eI is the original vector rotated by θ, see Figure 1.2(b).

We must therefore rotate the position vector p′ through an angle −θ relative to the fixed basiseI , but this is actually equivalent to a positive rotation of the base vectors. Hence the transformsfor the components of vector and the base vectors are the same.

p1 = cos θ p1 + sin θ p2 and p2 = − sin θ p1 + cos θ p2.

1.2 TensorsLecture 3

Tensors are geometric objects that have magnitude and zero, one or many associated directions,but are linear in character. A more mathematically precise definition is to say that a tensor ismultilinear map or alternatively an element of a tensor product of vector spaces, which is somewhattautological and really not helpful at this point. The order (or degree or rank) of a tensor isthe number of associated directions and so a scalar is a tensor of order zero and a vector is atensor of order one. Many quantities in continuum mechanics such as strain, stress, diffusivity andconductivity are naturally expressed as tensors of order two. We have already seen an exampleof a tensor in our discussion of vectors: linear transformations from one set of vectors to another,e.g. the transformation from Cartesian to covariant base vectors, are second-order tensors. If thevectors represent physical objects, then they must not depend on the coordinate representationchosen. Hence, the linear transformation must also be independent of coordinates because thesame vectors must always transform in the same way. We can write our linear transformation in acoordinate-independent manner as

a = M(b), (1.24)

and the transformation M is a tensor of order two. In order to describe M precisely we mustpick a specific coordinate system for each vector in equation (1.24). In the global Cartesian basis,equation (1.24) becomes

aIeI = M(bJeJ) = bJM(eJ), (1.25)

because it is a linear transformation. We now take the dot product with eK to obtain

aIeK·eI = bJeK·M(eJ) ⇒ aK = bJeK·M(eJ),

where the dot product is written on the left to indicate that we are taking the dot product afterthe linear transformation has operated on the base vector eJ . Hence, we can write the operation ofthe transformation on the components in the form

aI = MIJbJ , (1.26)

where MIJ = eI·M(eJ). Equation (1.26) can be written in a matrix form to aid calculation

a1a2a3

=

M11 M12 M13

M21 M22 M23

M31 M32 M33

b1b2b3

.

The quantity MIJ represents the component of the transformed vectors in the I-th Cartesian di-rection if the original vector is of unit length in the J-th direction. Hence, the quantity MIJ ismeaningless without knowing the coordinate system associated with both I and J .

Page 12: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

In fact, there is no need to choose the same coordinate system for I and J . If we write thevector a in the covariant basis, equation (1.25) becomes

aigi = bJM(eJ).

Taking the dot product with the appropriate contravariant base vector gives

ak = bJgk·M(eJ) = bJ

∂ξk

∂xKeK·M(eJ),

which means that

ak = MkJ bJ =

∂ξk

∂xKMKJbJ ⇒ Mk

J =∂ξk

∂xKMKJ .

In other words the components of each (column) vector corresponding to a fixed second index ina coordinate representation of M must obey a contravariant transformation if the associated basisundergoes a covariant transform, i.e. the behaviour is exactly the same as for the components of avector.

If we now also represent the vector b in the covariant basis, equation (1.25) becomes

aigi = bjM(gj).

Taking the dot product with the appropriate contravariant base vector gives

ak = bjgk·M(gj) = bj

∂ξk

∂xKeK·M

�∂xJ

∂ξjeJ

�= bj

∂ξk

∂xK

∂xJ

∂ξjeK·M(eJ),

on using the linearity of the transformation. Hence,

ak = Mkj b

j =∂ξk

∂xKMKJ

∂xJ

∂ξjbj ⇒ Mk

j =∂ξk

∂xKMKJ

∂xJ

∂ξj,

and the components of each (row) vector associated with a fixed first index in a coordinate repre-sentation of M undergo a covariant transformation when the associated basis undergoes a covarianttransform, i.e. the “opposite behaviour” to the components of a vector. The difference in behaviourbetween the two indices of the components of the linear transformation arises because one indexcorresponds to the basis of the “input” vector, whereas the other corresponds to the basis of the“output” vector. There is a sum over the second (input) index and the components of the vector band in order for this sum to remain invariant the transform associated with the second index mustbe the opposite to the components of the vector b, in other words the same as the transformationof the base vectors of that vector.

The obvious relationships between components can easily be deduced when we represent ourvectors in the contravariant basis,

ai = M ijbj, ai = M ji bj, ai = Mijb

j. (1.27)

Many books term M ij a contravariant second-order tensor; Mik a covariant second-order tensorand M j

i a mixed second-order tensor, but they are simply representations of the same coordinate-independent object in different bases. Another more modern notation is to say that M ij is a type(2, 0) tensor, MiJ is type (0, 2) and M i

j is a type (1, 1) tensor, which allows the distinction betweenmixed tensors or orders greater than two.

Page 13: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

1.2.1 Invariance of second-order tensors

Let us now consider a general change of coordinates from ξi to χi. Given that

ai = M ijb

j, (1.28a)

we wish to find an expression for M ijsuch that9

ai = M ijbj. (1.28b)

Using the transformation rules for the components of vectors (1.19) it follows that (1.28a)becomes

∂ξi

∂χnan = M i

j

∂ξj

∂χnbn.

We now multiply both sides by ∂χm/∂ξi to obtain

∂χm

∂ξi∂ξi

∂χnan = δmn a

n = am =∂χm

∂ξiM i

j

∂ξj

∂χnbn.

Comparing this expression to equation (1.28b) it follows that

M ij=

∂χi

∂ξnMn

m

∂ξm

∂χj,

and thus we see that covariant components must transform covariantly and contravariant compo-nents must transform contravariantly in order for the invariance properties to hold. Similarly, itcan be shown that

M i j =∂χi

∂ξn∂χj

∂ξmMnm, and Mi j =

∂ξn

∂χi

∂ξm

∂χjMnm. (1.29)

An alternative definition of tensors is to require that they are sets of index quantities (multi-dimensional arrays) that obey these transformation laws under a change of coordinates.

The transformations can be expressed in matrix form, but we must distinguish between thecovariant and contravariant cases. We shall write M♭ to indicate a matrix where all componentstransform covariantly and M♯ for the contravariant case. We define the transformation matrix F tohave the components

F =

∂χ1

∂ξ1∂χ1

∂ξ2∂χ1

∂ξ3

∂χ2

∂ξ1∂χ2

∂ξ2∂χ2

∂ξ3

∂χ3

∂ξ1∂χ3

∂ξ2∂χ3

∂ξ3

, or F i

j =∂χi

∂ξj;

and then from the chain rule and independence of coordinates

F−1 =

∂ξ1

∂χ1

∂ξ1

∂χ2

∂ξ1

∂χ3

∂ξ2

∂χ1

∂ξ2

∂χ2

∂ξ2

∂χ3

∂ξ3

∂χ1

∂ξ3

∂χ2

∂ξ3

∂χ3

, or [F−1]i

j=

∂ξi

∂χj.

IfM♯andM

♭represent the matrices of transformed components then the transformation laws (1.29)

becomeM

♯= FM♯FT , and M

♭= F−TM♭F−1. (1.30)

9This is a place where the use of overbars makes the notation look cluttered, but clarifies precisely which coordinatesystem is associated with each index. This notation also allows the representation of components in the two differentcoordinate systems, so-called two-point tensors, e. g. M i

j, which will be useful.

Page 14: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

1.2.2 Cartesian tensors

If we restrict attention to orthonormal coordinate systems, then the transformation between co-ordinate systems must be orthogonal10 and we do not need to distinguish between covariant andcontravariant behaviour. Consider the transformation from our Cartesian basis eI to another or-thonormal basis eI . The transformation rules for components of a tensor of order two become

MI J =∂xN

∂xI

∂xM

∂xJMNM .

The transformation between components of two vectors in the different bases are given by

aI =∂xI

∂xKaK =

∂xK

∂xIaK ,

which can be written the form

aI = QIKaK , where QIK =∂xI

∂xK=

∂xK

∂xI,

and the components QIK form an orthogonal matrix. Hence the transformation property of a(Cartesian) tensor of order two can be written as

MI J = QINMNMQJM

or in matrix formM = QMQT . (1.31)

In many textbooks, equation (1.31) is defined to be the transformation rule satisfied by a (Cartesian)tensor of order two.

1.2.3 Tensors vs matrices

There is a natural relationship between tensors and matrices because, as we have seen, we can writethe components of a second-order tensor in a particular coordinate system as a matrix. It is oftenhelpful to think of a tensor as a matrix when working with it, but the two concepts are distinct.

A summary of all the above is that a tensor is an geometric object that does not depend on anyparticular coordinate system and expresses a linear relationship between other geometric objects.

1.3 Products of vectors: scalar, vector and tensorRequiredReading(not cov-ered inlectures)

1.3.1 Scalar product

We have already used the scalar or dot product of two vectors, and the discussion here is includedonly for completeness. The scalar product is the product of two vectors that returns a unique scalar:the product of the lengths of the vectors and the cosine of the angle between them. Thus far, wehave only used the dot product to define orthonormal sets of vectors.

If we represent two vectors a and b in the co- and contravariant bases, then

a · b = (aigi)·(bjgj),

10Although the required orthogonal transformation may vary with position, as is the case in plane polar coordinates.

Page 15: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

and soa · b = aib

jgi·gj = aib

jδij = aibi.

An alternative decomposition demonstrates that

a · b = aibi,

and we note that the scalar product is invariant under coordinate transformation, as expected. Inorthonormal coordinate systems, there is no distinction between co and contravariant bases and soa · b = aKbK .

1.3.2 Vector product

The vector or cross product is a product of two vectors that returns a unique vector that is orthogonalto both vectors. In orthonormal coordinate systems, the vector product is defined by

a× b =

a1a2a3

×

b1b2b3

=

a2b3 − b2a3a3b1 − a1b3a1b2 − a2b1

.

In order to represent the vector product with index notation it is convenient to define a quantityknown as the alternating, or Levi-Civita, symbol eIJK . In orthonormal coordinate systems thecomponents of eIJK are defined by

eIJK = eIJK =

0 when any two indices are equal;

+1 when I, J,K is an even permutation of 1, 2, 3;

−1 when I, J,K is an odd permutation of 1,2,3.

(1.32)

e.g.

e112 = e122 = 0, e123 = e312 = e231 = 1, e213 = e132 = e321 = −1.

Strictly speaking eIJK thus defined is not a tensor because if the handedness of the coordinate systemchanges then the sign of the entries in eIJK should change in order for it to respect the appropriateinvariance properties; such objects are sometimes called pseudo-tensors. We could ensure that eIJK

is a tensor by restricting our definition to right-handed (or left-handed) orthonormal systems, whichwill be the approach taken in later chapters.

The vector product of two vectors a and b in orthonormal coordinates is

[a× b]I = eIJKaJbK , (1.33)

which can be confirmed by writing out all the components. In addition, the relationship betweenthe Cartesian base vectors eI can be expressed as a vector product using the alternating tensor

eI × eJ = eIJKeK . (1.34)

Let us now consider the case of general coordinates: the cross product between covariant basevectors is given by

gi × gj =∂xI

∂ξieI ×

∂xJ

∂ξjeJ =

∂xI

∂ξi∂xJ

∂ξjeI × eJ =

∂xI

∂ξi∂xJ

∂ξjeIJKeK .

Page 16: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

The expression on the right-hand side corresponds to the first two indices of the alternating tensorundergoing a covariant transformation so that

gi × gj = eijKeK .

If we now transform the third index covariantly we must transform the base vector contravariantlyso that

gi × gj = ǫijk∂ξk

∂xKeK = ǫijkg

k; (1.35)

where ǫijk ≡ ∂xI

∂ξi∂xJ

∂ξj∂xK

∂ξkeIJK . A similar argument shows that

gi × gj = ǫijkgk,

where ǫijk =≡ ∂ξi

∂xI

∂ξj

∂xJ

∂ξk

∂xK eIJK .If we decompose the vectors a and b into the contravariant basis we have

a× b = (aigi)× (bjg

j) = aibjgi × gj = aibjǫ

ijkgk.

Thus, if we decompose the vector product into the covariant basis we have the following expressionfor the components

[a× b]k = ǫijkaibj, or [a× b]i = ǫijkajbk.

1.3.3 Tensor product

The tensor product is a product of two vectors that returns a second-order tensor. It can bemotivated by the following discussion. Recall that equation (1.24) can be written in the form

ai = Mijbj, where Mij = gi·M(gj).

The components Mij correspond to the representation of tensor with respect to a basis, but whichbasis? We shall define the basis to be that formed from the tensor product of pairs of base vectors:gi⊗gj, where the symbol ⊗ is used to denote the tensor product. Hence, we can represent a tensorin the different forms

M = Mij gi ⊗ gj = MIJ eI ⊗ ej = M i

j gi ⊗ gj · · · ,

which is analogous to representing vectors in the different forms

a = aIeI = ai gi = ai gi.

Returning to equation (1.24), we have

a = M(b) ⇒ a = (Mijgi ⊗ gj)(b),

and because Mij are just coefficients it follows that gi⊗gj are themselves tensors of second order11.Decomposing a and b into the contravariant and covariant bases respectively gives

aigi = (Mijg

i ⊗ gj)(bngn) = Mijbn(gi ⊗ gj)(gn), (1.36)

11You should think carefully to convince yourself that this is true.

Page 17: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

but from equation (1.27)

ai = Mij bj ⇒ aig

i = Mij bjgi = Mij b

n δjn gi. (1.37)

Equating (1.36) and (1.37), it follows that the tensor product acting on the base vectors is

(gi ⊗ gj)(gn) = δjn gi = (gj

·gn) gi. (1.38)

Equation (1.38) motivates an alternative definition of the tensor product, which in the case of theproduct of two vectors is also called the dyadic or outer product. The dyadic product of two vectorsa and b is a second-order tensor that is defined through its action on an arbitrary vector v

(a⊗ b)(v) = (b · v)a. (1.39)

Consider the tensor T = a⊗ b; the components of that tensor are given by

Tij = gi·T (gj) = gi·(b·gj)a = (gi·a)(b·gj) = aibj,

which gives a third (equivalent) definition of the tensor product. Alternatively, we can write

T ij = aibj or T ij = aibj or T j

i = aibj. (1.40)

Note that we can write

T = a⊗ b = aigi ⊗ bjg

j = aibjgi ⊗ gj = Tijg

i ⊗ gj,

again demonstrating that Tij are the components of the tensor T in the basis gi ⊗ gj.The tensor product can easily be extended to tensors of arbitrary order via its component-wise

definition. e.g.T imn

jk = AimnB

jk

1.4 The metric tensorLecture 4

We can reinterpret the Kronecker delta as a tensor with components in the global Cartesian coor-dinate system, so that

δIJ = δIJ = δIJ ≡(1, if I = J,

0, otherwise.

If we transform the Kronecker delta into the general coordinate system ξi, then we obtain theso-called metric tensors

gij ≡ ∂xI

∂ξi∂xJ

∂ξjδIJ =

∂xI

∂ξi∂xI

∂ξj= gi·gj, (1.41a)

gij ≡ ∂ξi

∂xI

∂ξj

∂xJδIJ =

∂ξi

∂xI

∂ξj

∂xI= gi

·gj, (1.41b)

gij ≡ ∂ξi

∂xI

∂xJ

∂ξjδIJ =

∂ξi

∂xI

∂xI

∂ξj= δij. (1.41c)

We note that all the tensors are symmetric gij = gji and that the mixed metric tensor gij is invariantand equal to δij in all coordinate systems. From equations (1.41a,b) we deduce that

gijgjk =∂ξi

∂xI

∂ξj

∂xI

∂xJ

∂ξj∂xJ

∂ξk=

∂ξi

∂xIδJI

∂xJ

∂ξk=

∂ξi

∂xI

∂xI

∂ξk= δik, (1.42)

Page 18: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

in other words if we represent the components gij as a matrix, we can find the components gij bytaking the inverse of that matrix, i.e.

gij =Dij

g, (1.43)

where Dij is the matrix of cofactors of the matrix of coefficients gij and g is the determinant of thematrix gij

g ≡ |gij| =����∂xI

∂ξj

����2

and |gij| =����∂ξi

∂xJ

����2

=1

g, (1.44)

which demonstrates that the determinants are positive.

1.4.1 Properties of the metric tensor

Line elements

We have already seen that an infinitesimal line element corresponding to a change dξi in the ξi

coordinate is given by

ti = dsi =∂r

∂ξidξi = gidξ

i (not summed).

Hence, a general infinitesimal line element corresponding to changes in all coordinates is given by

dr = gidξi, (summed over i),

and its length squared isds2 = dr·dr = gidξ

i·gjdξ

j = gijdξidξj.

Thus, the infinitesimal length change due to increments dξi in general coordinates is given bypgijdξidξj, which explains the use of the term metric to describe the tensor gij.

Surface elements

Any surface ξ1 is constant is spanned by the covariant base vectors g2 and g3 and so an infinitesimalarea within that surface is given by

dS(1) = |ds2 × ds3| = |g2 × g3|dξ2dξ3,

and

|g2 × g3| = |ǫ231g1| = |ǫ231||g1| =����∂xI

∂ξ2∂xJ

∂ξ3∂xK

∂ξ1eIJK

����p

g1·g1. (1.45)

The term between the modulus signs is the determinant12 of the matrix with components ∂xI/∂ξj ,so from the definition of the determinant of the metric tensor (1.44),

dS(1) =pgg11 dξ2dξ3.

Hence, an element of area in the surface ξi is constant is

dS(i) =p

ggii dξjdξk, (i 6= j 6= k and i not summed ).

12It is the scalar triple product of the three rows of the matrix.

Page 19: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

Volume elements

An infinitesimal parallelepiped spanned by the vectors dsi has volume given by

dV = |ds1·(ds2 × ds3)| = |g1·(g2 × g3)|dξ1dξ2dξ3 = |g1·ǫ231g1|dξ1dξ2dξ3.

Now|g1·ǫ231g

1| = |ǫ231| =√g,

using the same argument as in equation (1.45). Thus, the volume element is given by

dV =√gdξ1dξ2dξ3, (1.46)

which is, of course, equivalent to the standard expression for change of coordinates in multipleintegrals

dV =

����∂(x1, x2, x3)

∂(ξ1, ξ2, ξ3)

���� dξ1dξ2dξ3.

Index raising and lowering

We conclude this brief section by noting one of the most useful properties of the metric tensor. Fora vector a = ajg

j, we can find the components in the covariant basis by taking the dot product

ai = a·gi = ajgj·gi = ajg

ji.

Thus the contravariant metric tensor can be used to raise indices and similarly the covariant metrictensor can be used to lower indices

ai = ajgji.

1.5 Vector and tensor calculus

1.5.1 Covariant differentiation

We have already established that the derivative of the position vector r with respect to the coor-dinate ξi gives a tangent vector gi. If we wish to express the rate of change of a vector field in thedirection of a coordinate then we also need to be able to calculate

∂a

∂ξi≡ a,i,

where we shall use a comma to indicate differentiation with respect to the i-th general coordinate.Under a change of coordinates from ξi to χi, the derivative becomes

∂a

∂χi=

∂ξj

∂χi

∂a

∂ξj,

so it transforms covariantly. If we decompose the vector into the covariant basis we have that

a,j =�aigi

�,j= ai,j gi + aigi,j , (1.47)

by the product rule.

Page 20: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

Vector equations are often written by just writing out the components (in an assumed basis)

v = u ⇒ vIeI = uIeI ⇒ vI = uI ,

where formally the last equation is obtained via dot product with a suitable base vector. In Carte-sian components we can simply take derivatives of the component equations with respect to thecoordinates because the base vectors do not depend on the coordinates

uI = vI leads to uI,J = vI,J because (uIeI),J = uI,J eI .

In other words, the components of the differentiated vector equation are simply the derivatives ofthe components of the original equation.

In a general coordinate system, the base vectors are not independent of the coordinates and sothe second term in equation (1.47) cannot be neglected. The vector equation

a = b ⇒ aigi = bigi ⇒ ai = bi,

where the last equation is now obtained by taking the dot product with the contravariant basevectors. However, in general coordinates,

ai = bi does not (directly) lead to ai,j = bi,j.

Although the final equation is (obviously) true from direct differentiation of each component thestatement is not coordinate-independent because it does not obey the correct transformation prop-erties for a mixed second-order tensor.

In fact, the derivatives of the base vectors are given by

gi,j =∂2r

∂ξj∂ξi= r,ij = r,ji,

assuming symmetry of partial differentiation. If we decompose the position vector into the fixedCartesian coordinates, we have

gi,j =∂2xK

∂ξj∂ξieK ,

which can be written in the contravariant basis as

gi,j = Γijkgk,

where

Γijk = gi,j·gk =∂2r

∂ξj∂ξi·gk =

∂2xK

∂ξi∂ξj∂xK

∂ξk. (1.48)

The symbol Γijk is called the Christoffel symbol of the first kind. Thus, we can write

a,j = ai,jgi + aiΓijkgk = ai,jgi + aiΓijkg

klgl =�al,j + Γl

ijai�gl,

whereΓlij = gklΓijk = gklgk·gi,j = gl

·gi,j , (1.49)

and Γlij is called the Christoffel symbol of the second kind.

Finally, we obtaina,j = ak|jgk, where ak|j = ak,j + Γk

ijai,

Page 21: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

and the expression ak|j is the covariant derivative of the component ak. In many books the covariantderivative is represented by a subscript semicolon ak;j and in some it is denoted by a comma, justto confuse things. The fact that a,j is a covariant vector and that gk is a covariant vector meansthat the covariant derivative ak|j is a (mixed) second-order tensor13. Thus when differentiatingequations expressed in components in a general coordinate system we should write

ai = bi ⇒ ai|j = bi|j.

If we had decomposed the vector a into the contravariant basis then

a,j = ai,jgi + aig

i,j.

and from equation (1.12),

gk =∂ξk

∂xKeK ⇒ gk ∂x

I

∂ξk=

∂ξk

∂xK

∂xI

∂ξkeK = δIKe

K = eI . (1.50)

We differentiate equation (1.50) with respect to ξj, remembering that eI is a constant base vector,to obtain

gk,j

∂xI

∂ξk+ gk ∂2xI

∂ξj∂ξk= 0,

⇒ gk,j

∂xI

∂ξk∂ξi

∂xI= −gk ∂ξ

i

∂xI

∂2xI

∂ξj∂ξk⇒ gi

,j = −gk,j·gigk = −Γi

kj gk.

Thus the derivative of the vector a when decomposed into the contravariant basis is

a,j = ai,jgi − Γi

jk ai gk =

�ak,j − Γi

jk ai�gk;

and finally, we obtaina,j = ai|jgi, where ai|j = ai,j − Γk

jiak,

which gives the covariant derivative of the covariant components of a vector.In Cartesian coordinate systems, the base vectors are independent of the coordinates, so ΓIJK ≡

0 for all I, J,K and the covariant derivative reduces to the partial derivative:

aI |K = aI |K = aI,K .

Therefore, in most cases, when generalising equations derived in Cartesian coordinates to othercoordinate systems we simply need to replace the partial derivatives with covariant derivatives.

The partial derivative of a scalar already exhibits covariant transformation properties because

∂φ

∂χi=

∂ξj

∂χi

∂φ

∂ξj;

and the partial derivative of a scalar coincides with the covariant derivative

φ,i = φ|i.

The covariant derivative of higher-order tensors can be constructed by considering the covariantderivative of invariants and insisting that the covariant derivative obeys the product rule.

13Thus, the covariant derivative exhibits tensor transformation properties, which the partial derivative does notbecause

∂ai

∂χj=

∂ξk

∂χj

∂ξk

∂χi

∂ξlal

!=

∂ξk

∂χj

∂χi

∂ξl∂al

∂ξk+

∂ξk

∂χj

∂2χi

∂ξk∂ξlal,

and the presence of the second term violates the tensor transformation rule.

Page 22: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

1.5.2 Vector differential operators

Gradient

The gradient of a scalar field f(x) is denoted by ∇f , or gradf , and is defined to be a vector fieldsuch that

f(x+ dx) = f(x) +∇f ·dx+ o(|dx|); (1.51)

i.e. the gradient is the vector that expresses the change in the scalar field with position.Letting dx = th, dividing by t and taking the limit as t → 0 gives the alternative definition

∇f ·h = limt→0

f(x+ th)− f(x)

t=

d

dtf(x+ th)|t=0. (1.52)

Here, the derivative is the directional derivative in the direction h.If we decompose the vectors into components in the global Cartesian basis equation (1.52)

becomes

[∇f ]K hK =∂f

∂xK

d

dt(xK + thK)|t=0 =

∂f

∂xK

hK ,

and so because h is arbitrary, we can define

∇f = [∇f ]K eK =∂f

∂xK

eK , (1.53)

which should be a familiar expression for the gradient. Relative to the general coordinate systemξi equation (1.53) becomes

∇f =∂f

∂ξi∂ξi

∂xK

eK =∂f

∂ξigi = f,ig

i, (1.54)

so we can write the vector differential operator ∇ = gi ∂∂ξi

. Note that because the derivativetransforms covariantly and the base vector transform contravariantly the gradient is invariant undercoordinate transform.

Example 1.4. Gradient in standard coordinate systemsFind the gradient in a plane polar coordinate system.

Solution 1.4. The gradient of a scalar field f is simply

∇f = gi ∂f

∂ξi.

In plane polar coordinates,

g1 = cos ξ2 e1 + sin ξ2 e2, g2 = −sin ξ2

ξ1e1 +

cos ξ2

ξ1e2,

so

∇f =�cos ξ2 e1 + sin ξ2 e2

� ∂f

∂ξ1+

�−sin ξ2

ξ1e1 +

cos ξ2

ξ1e2

�∂f

∂ξ2,

which can be written in the (hopefully) familiar form

∇f =∂f

∂rer +

1

r

∂f

∂θeθ.

Page 23: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

Gradient of a Vector Field

The gradient of a vector field F (x) is a second-order tensor ∇F (x) also often written as ∇ ⊗ F

that arises when the vector differential operator is applied to a vector

∇⊗ F = gi ⊗ F ,i = F k|i gi ⊗ gk = Fk|i gi ⊗ gk.

Note that we have used the covariant derivative because we are taking the derivative of a vectordecomposed into the covariant or contravariant basis.

Divergence

The divergence of a vector field is the trace or contraction of the gradient of the vector field. It isformed by taking the dot product of the vector differential operator ∇ with the vector field:

divF = ∇·F = gi·F ,i = gi

·∂

∂ξi�F jgj

�= gi

·F j|igj = δijFj|i = F i|i,

so thatdivF = F i|i = F i

,i + ΓijiF

j = F i,i + Γi

ijFj. (1.55)

Curl

The curl of a vector field is the cross product of our vector differential operator ∇ with the vectorfield

curlF = ∇× F = gi × ∂F

∂ξi= gi × ∂ (Fjg

j)

∂ξi= gi × Fj|igj = ǫijkFj|igk.

Laplacian

The Laplacian of a scalar field f is defined by

∇2f = Δf = ∇·∇f.

Thus in general coordinates

∇2f = gj·∂

∂ξj

�gi ∂f

∂ξi

�= gj

·f,i|j gi = gijf,i|j = gijf |ij,

because the partial and covariant derivatives are the same for scalar quantities. Now gij|j is a tensor(covariant first order), which means that we know how it transforms under change in coordinates,but in Cartesian coordinates gij|j becomes δIJ,J = 0, so gij|j = 0 in all coordinate systems14. Thus,

∇2f = (gijf,i)|j =1√g

∂ξj

�√g gij

∂f

∂ξi

�.

The last equality is not obvious and is proved in the next section, equation (1.58)

14This is a special case of Ricci’s Lemma.

Page 24: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

1.6 Divergence theorem

The divergence theorem is a generalisation of the fundamental theorem of calculus into higherdimensions. Consider the volume given by a ≤ ξ1 ≤ b, c ≤ ξ2 ≤ d and e ≤ ξ3 ≤ f . If we want tocalculate the outward flux of a vector field F (x) from this volume then we must calculate

ZZF · n dS,

where n is the outer unit normal to the surface and dS is an infinitesimal element of the surfacearea.

From section 1.4.1, we know that on the faces ξ1 = a, b,

dS =p

gg11dξ2dξ3,

and a normal to the surface is given by the contravariant base vector g1, with length |g1| =pg1·g1 =

pg11. On the face ξ1 = a the vector g1 is directed into the volume (it is oriented along

the line of increasing coordinate) and on ξ1 = b, g1 is directed out of the volume, so the net outwardflux through these faces is

Z d

c

Z f

e

F (a, ξ2, ξ3)·(−g1/p

g11)pgg11dξ2dξ3 +

Z d

c

Z f

e

F (b, ξ2, ξ3)·(+g1/pg11)

pgg11dξ2dξ3

=

Z d

c

Z f

e

hF 1(b, ξ2, ξ3)

pg(b, ξ2, ξ3)− F 1(a, ξ2, ξ3)

pg(a, ξ2, ξ3)

idξ2dξ3

From the fundamental theorem of calculus, the integral becomes

Z d

c

Z f

e

Z b

a

(F 1√g),1 dξ1dξ2dξ3,

and using similar arguments, the total outward flux from the volume is given by

ZZF · n dS =

Z b

a

Z d

c

Z f

e

(F i√g),idξ1dξ2dξ3.

From section 1.4.1 we also know that an infinitesimal volume element is given by dV =√g dξ1dξ2dξ3,

so ZZF · n dS =

ZZZ1√g(F i√g),i

√g dξ1dξ2dξ3 =

ZZZ1√g(F i√g),i dV.

The volume integrand is1√g(F i√g),i = F i

,i +1√gF i(

√g),i,

but1√g(√g),i =

1

2g

∂g

∂ξi=

1

2g

∂g

∂gjk

∂gjk∂ξi

,

by the chain rule. From equations (1.42) and (1.44)

gjkgkl = gjk

Dkl

g= δlj ⇒ gjkD

kl = g δlj ⇒ ∂g

∂gjkδlj =

∂g

∂glk= Dkl = ggkl = gglk, (1.56)

Page 25: Chapter 1 Describing the Physical World: Vectors & …...Chapter 1 Describing the Physical World: Vectors & Tensors Lecture 1 It is now well established that all matter consists of

which means that

1√g(√g),i =

1

2gggjk

∂gjk∂ξi

=1

2gjk

�gj,i·gk + gj·gk,i

�=

1

2

�gj,i·g

j + gk,i·gk�=

1

2

�Γjji + Γk

ki

�= Γj

ji,

(1.57)from the definition of the Christoffel symbol (1.49). Thus the volume integrand becomes

1√g

�F i√g

�,i= F i

,i + F i Γjji = F i|i, (1.58)

and so, ZZF · n dS =

ZZZF i,i + F iΓj

ji dV =

ZZZF i|i dV =

ZZZdivF dV,

using the definition of the divergence (1.55). In fact, one should argue that the definition of thedivergence is chosen so that it is the quantity that appears within the volume integral in thedivergence theorem. The form of the theorem that we shall use most often is

ZZF ini dS =

ZZFin

i dS =

ZZZF i|i dV =

ZZZ1√g(F i√g),i dV. (1.59)

You might argue that the proof above is only valid for simple parallelepipeds in our general coordi-nate system, but we can extend the result to more general volumes by breaking the volume up intoregions, each of which can be described by a parallelepiped in a general coordinate systems.

1.7 The Stokes Theorem

For completeness, we also state, but do not prove the Stokes theorem, which states that for anyvector v, the integral of its tangential component around a closed path is equal to the flux of thecurl of the vector through any surface bounded by the path:

I

C

v·dr =

ZZ

A

(curl v)·dS. (1.60)

In general tensor form, equation (1.60) becomes

I

C

v · gi dξi =

Ivi dξ

i =

ZZ

A

ǫijkvj|igk·dS =

ZZ

A

ǫijkvj|ink dS, (1.61)

where the vector area is dS = n dS, and n is a unit vector normal to the surface. Again we canmake the argument that the definition of the curl is chosen so that this relationship is satisfied.

You may be interested to know that an entire theory (differential forms) can be constructed inwhich the Stokes theorem is a fundamental result

Z

∂C

ω =

Z

C

dω,

where ω is an n-form and dω is its differential. The theory encompasses all known integral theoremsbecause the appropriate n-dimensional differential forms coincide with the divergence and curl andthe theory provides a rational extension to higher dimensions.