Scalars, Vectors, Tensors

Scalars, Vectors, Tensors Scalars, Vectors, Tensors

as.arizona.edu
from as.arizona.edu More from this publisher
20.01.2015 Views

Appendix B Scalars, Vectors, Tensors The fundamental requirement for the mathematical expression of each physical law is that it is written in a way that is independent of the particular coordinate system that is being used. For example, in Newtonian dynamics, the second law is expressed using vector notation as m d2 ⃗r dt = ⃗ F, (B.1) where m and ⃗r are the mass and position vector of a particle that is moving under the influence of a force ⃗ F .If,moreover,theforceispotential,i.e.,itcanbewritten as the gradient of a scalar potential function Φ, then Newton’s second law takes the form m d2 ⃗r d 2 t = −⃗ ∇Φ . (B.2) This is a symbolic representation of the vectorial form of Newton’s second law, Manifestly covariant forms but it is not particularly useful for computations. What we would like to have is aformofNewton’ssecondlawintermsofthe components of the various vectors, but written in a way that is invariant under coordinate transformations. We would call such an expression the manifestly covariant form of a physical law. In an orthonormal Cartesian coordinate system with unit vectors ˆx i (i=1,2,3), we can write the position vector in component form as ⃗r = x i ˆx i ,andthegradient operator as ⃗ ∇ =(∂/∂x i )ˆx i . In this case, Newton’s second law in Cartesian coordinates takes the form m d2 x i dt 2 = − ∂Φ ∂x i . (B.3) This is, however, not a manifestly covariant form of the physical law, as we can easily prove. Consider a transformation from the Cartesian coordinates x i to a set of general coordinates ξ i that may be neither orthogonal nor normal. We understand this transformation to imply the existence of one-to-one functions of the form x i = x i (ξ 1 ,ξ 2 ,ξ 3 ) (B.4) as well as of their inverse ξ i = ξ i (x 1 ,x 2 ,x 3 ) (B.5) that are well defined in all but a small number of spatial points in space that we 315

Appendix B<br />

<strong>Scalars</strong>, <strong>Vectors</strong>, <strong>Tensors</strong><br />

The fundamental requirement for the mathematical expression of each physical law<br />

is that it is written in a way that is independent of the particular coordinate system<br />

that is being used. For example, in Newtonian dynamics, the second law is expressed<br />

using vector notation as<br />

m d2 ⃗r<br />

dt = ⃗ F,<br />

(B.1)<br />

where m and ⃗r are the mass and position vector of a particle that is moving under<br />

the influence of a force ⃗ F .If,moreover,theforceispotential,i.e.,itcanbewritten<br />

as the gradient of a scalar potential function Φ, then Newton’s second law takes the<br />

form<br />

m d2 ⃗r<br />

d 2 t = −⃗ ∇Φ .<br />

(B.2)<br />

This is a symbolic representation of the vectorial form of Newton’s second law, Manifestly<br />

covariant forms<br />

but it is not particularly useful for computations. What we would like to have is<br />

aformofNewton’ssecondlawintermsofthe components of the various vectors,<br />

but written in a way that is invariant under coordinate transformations. We would<br />

call such an expression the manifestly covariant form of a physical law.<br />

In an orthonormal Cartesian coordinate system with unit vectors ˆx i (i=1,2,3),<br />

we can write the position vector in component form as ⃗r = x i ˆx i ,andthegradient<br />

operator as ⃗ ∇ =(∂/∂x i )ˆx i . In this case, Newton’s second law in Cartesian<br />

coordinates takes the form<br />

m d2 x i<br />

dt 2<br />

= − ∂Φ<br />

∂x i .<br />

(B.3)<br />

This is, however, not a manifestly covariant form of the physical law, as we can<br />

easily prove. Consider a transformation from the Cartesian coordinates x i to a set<br />

of general coordinates ξ i that may be neither orthogonal nor normal. We understand<br />

this transformation to imply the existence of one-to-one functions of the form<br />

x i = x i (ξ 1 ,ξ 2 ,ξ 3 )<br />

(B.4)<br />

as well as of their inverse<br />

ξ i = ξ i (x 1 ,x 2 ,x 3 )<br />

(B.5)<br />

that are well defined in all but a small number of spatial points in space that we<br />

315


316 APPENDIX B. SCALARS, VECTORS, TENSORS<br />

will call the poles. Then, we can use the chain rule of differentiation to write<br />

and<br />

d 2 x i 3∑<br />

dt 2 = ∂x i d 2 ξ j<br />

∂ξ j dt 2<br />

∂Φ<br />

∂x i =<br />

j=1<br />

3∑<br />

j=1<br />

∂Φ ∂ξ j<br />

∂ξ j ∂x i .<br />

Inserting these two expressions in equation (B.3) we obtain<br />

3∑<br />

j=1<br />

( ) ∂x<br />

i<br />

∂ξ j m d2 ξ j<br />

dt 2<br />

= − 3∑<br />

j=1<br />

( ∂ξ<br />

j<br />

∂x i ) ∂Φ<br />

∂ξ j .<br />

(B.6)<br />

(B.7)<br />

(B.8)<br />

where<br />

This last expression can be put in the form of equation (B.3) if and only if<br />

δ i j ≡<br />

∂x i<br />

∂ξ j = δi j ,<br />

{ 1, if i = j<br />

0, if i ≠ j<br />

(B.9)<br />

(B.10)<br />

is Kronecker’s delta. As a result, Newton’s second law in the coordinate form of<br />

equation (B.3) is not manifestly covariant.<br />

What went wrong in this derivation<br />

As we will see below, in writing the<br />

components of the two vectors as ⃗r = x i ˆx i and ⃗ ∇ =(∂/∂x i )ˆx i ,weactuallyusedtwo<br />

different basis vectors, even though we denoted them both by ˆx i . We, therefore,<br />

need to start the discussion from the beginning by defining properly coordinate<br />

systems and basis vectors.<br />

B.1 Coordinate and Dual Basis <strong>Vectors</strong><br />

We consider a flat space with N dimensions and define a coordinate system as an<br />

one-to-one map between an ordered set of N real numbers ξ 1 , ξ 2 ,...,ξ N and each<br />

individual point in space. The position of each point in space can, therefore, be<br />

written in the form<br />

⃗r = ⃗r(ξ 1 ,ξ 2 ,...,ξ N ) .<br />

(B.11)<br />

Coordinate Lines We define coordinate lines as the curves along which only one of the coordinates<br />

and Surfaces changes, whereas the other remain constant. This is illustrated in Figure B.1 for two<br />

sample coordinate systems. At the same time, we also define coordinate surfaces as<br />

the surfaces on which only one of the coordinates remains constant. It follows from<br />

their definition that, e.g., in a three dimensional space, the intersection between<br />

two coordinate surfaces is a coordinate line.<br />

Using this definition of coordinate lines and surfaces, we have an infinite number<br />

of options of defining basis vectors, three of which are particularly useful in<br />

describing physical laws.<br />

We define the coordinate basis vectors at each point in space as the ordered set<br />

of vectors ê i (i =1, 2,...,N), with the property that each of them is tangent to the<br />

corresponding coordinate line. Formally, we define them by relations of the form


B.1. COORDINATE AND DUAL BASIS VECTORS 317<br />

Figure B.1: Coordinate lines and coordinate basis vectors for two different coordinate systems.<br />

The left panel depicts a Cartesian orthonormal coordinate system. In both cases, a third coordinate<br />

is assumed to extend from each point, perpendicular to the plane of the paper.<br />

ê i ≡ ∂⃗r<br />

∂ξ i (B.12) Coordinate<br />

Basis <strong>Vectors</strong><br />

and use a subscript notation to denote the coordinate along which this vector is<br />

tangent. Figure B.1 shows the coordinate basis vectors for two sample coordinate<br />

systems.<br />

We can also define the dual basis vectors at each point in space in terms of the<br />

coordinate surfaces. Given than each coordinate surface can be represented by an<br />

equation of the form ξ i (⃗r) =constant,wedefinethedualbasisvectorsas<br />

ê i ≡ ∇ξ ⃗ i . (B.13) Dual Basis<br />

<strong>Vectors</strong><br />

Note that we use a subscript notation for the dual basis vectors, in order to distinguish<br />

them from the coordinate basis vectors. Figure B.2 shows the dual basis<br />

vectors for two sample coordinate systems. Note that for an orthonormal Cartesian<br />

coordinate system<br />

ê ∗ i =ê∗i ,<br />

where we have used the star to denote a Cartesian system.<br />

(B.14)<br />

In general, the coordinate and dual basis vectors depend on position in space.<br />

However, they always satisfy<br />

ê i ê j = δ j i .<br />

(B.15)<br />

In order to prove this property, we will use an orthonormal Cartesian coordinate<br />

system (x 1 ,x 2 ,...,x N )andmakeacoordinatetransformationtosomeunspecified<br />

coordinate system (ξ 1 ,ξ 2 ,...,ξ N ). The position vector of any point in space can be<br />

written in terms of the Cartesian basis vectors as<br />

N∑<br />

⃗r = x k ê ∗ k .<br />

k=1<br />

(B.16)


318 APPENDIX B. SCALARS, VECTORS, TENSORS<br />

Figure B.2: Coordinates lines and dual basis vectors for the coordinate systems shown in Figure<br />

B.1.<br />

Figure B.3: Aexamplesetofcoordinatebasisvectors(ê 1 , ê 2 )andofthecorrespondingdual<br />

basis vectors (ê 1 , ê 2 ). Equation (B.15) requires that ê 1 ⊥ ê 2 and ê 2 ⊥ ê 1 but not necessarily that<br />

ê 1 ↑↑ ê 1 or ê 2 ↑↑ ê 2 .<br />

We use this together with the definition of the coordinate basis vectors to obtain<br />

ê i = ∂⃗r<br />

∂ξ i = N ∑<br />

k=1<br />

∂x k<br />

∂ξ i ê∗ k<br />

We also write explicitly the definition of the dual basis vectors as<br />

ê j = ⃗ ∇ξ j =<br />

N∑<br />

k=1<br />

∂ξ j<br />

∂x k ê∗ k .<br />

(B.17)<br />

(B.18)<br />

(Don’t worry for the moment about the apparent asymmetry in the k−index in the<br />

last expression; this is a Cartesian system for which ê ∗ k =ê∗k .) Taking the product<br />

of the two vectors we finally obtain<br />

ê i ê j =<br />

N∑<br />

k=1<br />

∂ξ j ∂x k<br />

∂x k ∂ξ i<br />

= ∂ξj<br />

∂ξ i = δj i .<br />

(B.19)<br />

Relation (B.15) leads to a number of important results regarding the two sets<br />

of basis vectors. In general, it implies that each coordinate basis vector ê i is perpendicular<br />

to all the dual basis vectors ê j with j ≠ i, butmaynotbeparallelto


B.2. COVARIANT AND CONTRAVARIANT COMPONENTS 319<br />

the dual basis vector ê i . The same is also true for each dual basis vector ê i : it is<br />

perpendicular to all the coordinate basis vectors ê j with j ≠ i, butmaynotbe<br />

parallel to the coordinate basis vector ê i .ThisisillustratedinFigureB.3.<br />

In an orthogonal coordinate system, i.e., one in which the coordinate lines are<br />

intersecting at right angles, the coordinate basis vectors and the dual basis vectors<br />

are mutually orthogonal and, hence, in this case<br />

ê i ↑↑ ê i . (B.20) Orthogonal Coordinate<br />

System<br />

B.2 Covariant and Contravariant Components<br />

Having defined different basis vectors, we can express the components of any vector<br />

⃗A with respect to either of them. When we use the coordinate basis vectors, i.e.,<br />

⃗A = A i ê i (B.21) Contravariant<br />

Components<br />

we use a superscript (or upstairs) notation for the components, which we call the<br />

contravariant components of the vector. On the other hand, when we use the dual<br />

basis vector, i.e.,<br />

⃗A = A i ê i (B.22) Covariant<br />

Components<br />

we use a subscript (or downstairs) notation for the components, which we call the<br />

covariant components of the vector ⃗ A.<br />

The contravariant and covariant components of a vector transform in different<br />

ways from one coordinate system to another. In order to study their transformations,<br />

we will consider two coordinate systems ξ i and ξ ′j and use unprimed and<br />

primed quantities, respectively, to denote the various vector components in the two<br />

systems.<br />

We will start by inserting the definition of the coordinate basis (B.12) into<br />

equation (B.21), i.e.,<br />

⃗A = A i ê i = A i ∂⃗r<br />

∂ξ i<br />

(B.23)<br />

We will then perform a change of coordinates in the derivatives using the chain rule,<br />

⃗A = A i ∂⃗r ∂ξ ′j<br />

∂ξ ′j ∂ξ i<br />

= A i ∂ξ′j<br />

∂ξ i ê′i .<br />

(B.24)<br />

In the last equation we have used again the definition (B.12) but for the primed<br />

coordinate frame. Comparing the last term with the definition (B.21) we finally<br />

find<br />

A ′j = A i ∂ξ′j<br />

∂ξ i . (B.25) Transformation of<br />

Contravariant Components<br />

In order to derive the transformation rule for the covariant components of a<br />

vector, we well use an auxiliary Cartesian coordinate system x i with basis vectors


320 APPENDIX B. SCALARS, VECTORS, TENSORS<br />

ê ∗i . We will start again by inserting the definition of the dual basis (B.13) into<br />

equation (B.22) and use the chain rule to perform a change of coordinates in the<br />

derivatives. In detail<br />

⃗A = A i ê i = A i<br />

∂ξ i<br />

∂ξ i ∂ξ ′k ∂ξ i<br />

∂x j ê∗j = A i<br />

∂ξ ′k ∂x j ê∗j = A i<br />

∂ξ ′k ê′k .<br />

(B.26)<br />

Comparing the last term with the definition (B.22) and simply changing the dummy<br />

index for k to j, weobtain<br />

Transformation of<br />

Covariant Components<br />

A ′ j = A i<br />

∂ξ i<br />

∂ξ ′j .<br />

(B.27)<br />

Note in these transformations how useful the notation of subscripts and superscripts<br />

has been.<br />

Until this point, we have assumed that the space on which we have defined coordinates,<br />

basis vectors, and vector components is flat. Extending these definitions<br />

to a general curved space will require a different understanding of the various quantities<br />

involved. For example, in a curved space,directedlinesegmentscanonly<br />

be infinitesimally short and, therefore, the position vector ⃗r that we used in the<br />

definition of the coordinate basis is not well defined. Even though there is a way<br />

of extending all these geometric interpretations in curved spaces, it is sufficient for<br />

the purposes of this class to take a somewhat backward approach.<br />

We will define as the contravariant components of a vector in an N-dimensional<br />

space, and denote them with superscript notation, an ordered set of N physical<br />

quantities (i.e., components of velocities, momenta, fields) that transform between<br />

coordinate systems according to equation (B.25).<br />

Similarly, we will define as the covariant components of a vector in an N-<br />

dimensional space, and denote them with subscript notation, an ordered set of N<br />

physical quantities (i.e., components of velocities, momenta, fields) that transform<br />

between coordinate systems according to equation (B.27).<br />

In general, we will define as a tensor and denote by<br />

T ijkl... αβγδ...<br />

(B.28)<br />

an ordered set of physical quantities, some of which transform according to the rules<br />

for contravariant components (and we will use superscript notation) and some of<br />

which transform according to the rule for covariant components, i.e.,<br />

Transformation of<br />

Tensor Components<br />

T ′ijk... αβγ... = ∂ξ′i ∂ξ ′j ∂ξA ∂ξ B<br />

∂ξ I ...<br />

∂ξJ ∂ξ ′α ∂ξ ′β ...T IJK... ABΓ... .<br />

(B.29)<br />

The total number of indices is called the rank of the tensor. A vector is a tensor of<br />

rank one. A scalar is a tensor of rank zero.<br />

Example: The inner product of two vectors<br />

In this example, we will derive some useful expressions for the inner product<br />

between two vectors and show that it is a scalar quantity, i.e., that it is invariant<br />

between coordinate transformations.


B.3. THE METRIC TENSOR 321<br />

We will start with two vectors,<br />

⃗A = A i ê i = A i ê i (B.30)<br />

⃗B = B j ê j = B j ê j (B.31)<br />

and calculate their product as<br />

⃗A · ⃗B =(A i ê i ) · (B j ê j )=A i B j (ê i · ê j )=A i B j δ i j = A i B i .<br />

(B.32)<br />

Here we used the fact that ê i · ê j = δ i j . We can follow the exact same procedure<br />

using the covariant components of vector ⃗ A and the contravariant components of<br />

vector ⃗ B.Thefinalsetofexpressionsfortheinnerproductoftwovectorsis<br />

⃗A · ⃗B = A i B i = A i B i (B.33) Inner product<br />

of <strong>Vectors</strong><br />

We will now consider a change of coordinates from a system ξ i to another system<br />

ξ ′i and evaluate the inner product of the two vectors in that system:<br />

⃗A· ⃗B = A ′i B i ′ = ∂ξ′i ∂ξk<br />

Aj<br />

∂ξj ∂ξ ′i B k =( ∂ξk ∂ξ ′i<br />

∂ξ ′i ∂ξ j )Aj B k = ∂ξk<br />

∂ξ j Aj B k = δ k jA j B k = A j B j .<br />

(B.34)<br />

This last expression proves that the inner product of two vectors is a scalar quantity.<br />

Example: Projecting onto basis vectors<br />

In an orthonormal coordinate system, e.g., in a Cartesian system (ê x , ê y , ê z ), we<br />

can calculate the component of a vector A ⃗ along one of the coordinate lines using<br />

inner products of the form A x = A ⃗ · ê x .Thisisnot,ofcourse,thecaseifthesystem<br />

is non-orthogonal. However, the definitions of the coordinate and dual basis provide<br />

us with a very useful tool in calculating contravariant and covariant components of<br />

vectors, independent of whether the coordinate system is orthogonal or not.<br />

Starting with the definition of the contravariant components of a vector A ⃗ and<br />

multiplying both sides of the equation with a dual basis vector ê i we obtain<br />

⃗A = A j ê j ⇒ ⃗ A · ê i = A j (ê j · ê i )=A j δ j<br />

i<br />

(B.35)<br />

and, therefore,<br />

A i = ⃗ A · ê i<br />

(B.36)<br />

Similarly, we can also prove that<br />

A i = ⃗ A · ê i<br />

(B.37)<br />

B.3 The Metric Tensor<br />

We will use the dot product of two vectors, and in particular of two basis vectors,<br />

in order to specify the geometry of a general curved space. Starting from<br />

⃗A · ⃗B =(A i ê i ) · (B j ê j )=A i B j (ê i · ê j ) ,<br />

(B.38)


322 APPENDIX B. SCALARS, VECTORS, TENSORS<br />

we only need to specify the elements of the rank-2 covariant tensor<br />

Metric<br />

Tensor<br />

Line<br />

Element<br />

g ij =ê i · ê j<br />

(B.39)<br />

which we will call the metric tensor.<br />

If, instead of the product of two vectors, we calculate the product of an infinitesimal<br />

translational vector to itself, i.e.,<br />

ds 2 = d⃗x · d⃗x = g ij dx i dx j<br />

(B.40)<br />

we call the result the line element of the space.<br />

We can also define the contravariant components of the metric tensor as<br />

and the mixed components as<br />

g ij =ê i · ê j<br />

g i j =ê i · ê j .<br />

(B.41)<br />

(B.42)<br />

Because of the orthogonality of the coordinate and dual basis vectors, g i j = δ i j.<br />

The metric tensor has very many uses in problem solving, one of which is to help<br />

us transform the components of a tensor between the coordinate and dual basis (i.e.,<br />

to raise on lower indices). For example, we show earlier that the inner product of<br />

two vectors is equal to<br />

⃗A · ⃗B = A i B i .<br />

(B.43)<br />

However, we can write the same inner product using the definition of the metric<br />

tensor as<br />

⃗A · ⃗B = A i B j g ij .<br />

(B.44)<br />

Comparing these two equations we obtain<br />

Lowering<br />

an Index<br />

Similarly, we can prove that<br />

B i = g ij B j .<br />

(B.45)<br />

Raising<br />

an Index<br />

B i = g ij B j .<br />

(B.46)<br />

Finally, we can use this last property of the metric tensor to prove that g ij is<br />

the inverse of g ij .Wewillstartfromthedotproductofacoordinatetoadualbasis<br />

vector,<br />

ê i · ê k = δ i k ⇒ (g ij ê j ) · ê k = δ i k ⇒ g ij (ê j · ê k )=δ i k (B.47)<br />

from which we obtain<br />

g ij g jk = δ i k .<br />

(B.48)


B.3. THE METRIC TENSOR 323<br />

Useful Expressions<br />

Coordinate Basis <strong>Vectors</strong><br />

Dual Basis <strong>Vectors</strong><br />

Orthogonality of Basis <strong>Vectors</strong><br />

ê i ≡ ∂⃗r<br />

∂ξ i<br />

ê i ≡ ⃗ ∇ξ i .<br />

ê i ê j = δ j i<br />

(B.49)<br />

(B.50)<br />

(B.51)<br />

Contravariant Components of Vector<br />

⃗A = A i ê i<br />

(B.52)<br />

Covariant Components of Vector<br />

⃗A = A i ê i<br />

(B.53)<br />

Transformation of Contravariant Components<br />

Transformation of Covariant Components<br />

A ′j = A i ∂ξ′j<br />

∂ξ i . (B.54)<br />

A ′ j = A ∂ξ i<br />

i<br />

∂ξ ′j .<br />

(B.55)<br />

Transformation of General Tensor Components<br />

T ′ijk... αβγ... = ∂ξ′i ∂ξ ′j ∂ξA ∂ξ B<br />

∂ξ I ...<br />

∂ξJ ∂ξ ′α ∂ξ ′β ...T IJK... ABΓ... .<br />

(B.56)<br />

Inner Product of <strong>Vectors</strong><br />

Vector Components<br />

The Metric Tensor<br />

⃗A · ⃗B = A i B i = A i B i<br />

A i = A ⃗ · ê i<br />

A i = A ⃗ · ê i<br />

g ij =ê i · ê j<br />

(B.57)<br />

(B.58)<br />

(B.59)<br />

(B.60)<br />

Other components of the metric tensor<br />

g ij g jk = δ i<br />

k<br />

g i j = δ i j<br />

(B.61)<br />

(B.62)<br />

Lowering and Raising an Index<br />

B i = g ij B j B i = g ij B j (B.63)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!