Fundamental Objects in the Euclidean Space

In the beginning, we studied the geometric elements of a Euclidean space in complete absence of coordinate systems. In recent chapters, however, in an effort to increase the analytical power of our framework, we introduced coordinates along with the most crucial elements of the tensor notation. We will now put all of this together and begin to study the key coordinate-dependent elements of a Euclidean space.
We have already encountered the position vector R\mathbf{R}, also known as the radius vector, which points from an arbitrary fixed origin OO to each point in the Euclidean space. The purpose of the position vector is, of course, to represent the position of the corresponding point by an object that is subject to analytical operations. Every one of our geometric analyses will start with the position vector. However, despite its fundamental role, it will rarely figure explicitly in advanced stages of most of our analyses, as we will be more interested in its derivatives rather than the position vector itself. This is part of the reason why the origin OO can be selected arbitrarily: its location has no bearing on the derivatives of R\mathbf{R}.
(9.1)
The position vector is a vector field, since to every point in space, there corresponds a particular value of R\mathbf{R}. However, as you see from the above figure, it is usually depicted in an unusual way compared to most vector fields. For most vector fields, such as the velocity distribution of a fluid, we usually imagine vectors emanating from the associated points in the Euclidean space, as in the following figure from Chapter 4.
(9.2)
The position vector, on the other hand, is best imagined emanating from the origin OO, as you see in the preceding figure. Because of this unconventional representation, we need to remind ourselves that the position vector is associated with the point to which it is pointing, rather than the origin.
Now, impose a coordinate system upon the Euclidean space. Assume that the Euclidean space is three-dimensional and, as usual, denote the coordinates by Z1Z^{1}, Z2Z^{2}, and Z3Z^{3} or, collectively, ZiZ^{i}. Importantly, the coordinate system is assumed to be completely arbitrary as long as it is sufficiently smooth in the sense described below. If the coordinate system has the concept of the coordinate origin -- as most of the standard coordinate systems do -- it need not coincide with the origin OO that anchors the position vector field R\mathbf{R}. We also remind the reader that the decision to use a superscript to enumerate the coordinates is completely arbitrary in its own right. However, once the decision to use a superscript to enumerate the coordinates is made, the placement of indices on all subsequent objects is uniquely prescribed by the rules of the tensor notation.
In the absence of a coordinate system, the position vector is described by the term field rather than function. A field is an association between the points of a Euclidean space and some quantity. As such, the position vector is an example of a vector field. Meanwhile, the temperature distribution in a room is an example of a scalar field. The term function, on the other hand, describes an association between sets of numbers and some quantity. Thus, with a coordinate system imposed upon the Euclidean space, we may begin to speak of the position vector function
R(Z1,Z2,Z3).(9.3)\mathbf{R}\left( Z^{1},Z^{2},Z^{3}\right) .\tag{9.3}
This function maps triplets of numbers Z1,Z2,Z3Z^{1},Z^{2},Z^{3} to the value of the position vector R\mathbf{R} at the point with coordinates Z1Z^{1}, Z2Z^{2}, and Z3Z^{3}.
By smoothness of the coordinate system we understand the differentiability characteristics of the function R(Z1,Z2,Z3)\mathbf{R}\left( Z^{1},Z^{2},Z^{3}\right) . For most analyses, existence of second derivatives is sufficient. We will often describe R(Z1,Z2,Z3)\mathbf{R}\left( Z^{1},Z^{2},Z^{3}\right) as sufficiently smooth, meaning that as many derivatives are available as required by the analysis at hand.
As we have similarly noted in a number of analogous contexts, the symbol R\mathbf{R} in the expression R(Z1,Z2,Z3)\mathbf{R}\left( Z^{1},Z^{2},Z^{3}\right) is being used in a new capacity compared to the symbol R\mathbf{R} in the absence of coordinates. In the absence of coordinates, R\mathbf{R} denotes a vector field, i.e. an association between points and vectors. In the expression R(Z1,Z2,Z3)\mathbf{R}\left( Z^{1},Z^{2},Z^{3}\right) , however, R\mathbf{R} denotes a vector-valued function of the three independent variables Z1Z^{1}, Z2Z^{2}, and Z3Z^{3}. This is another example of assigning the same symbol two closely related, albeit different, meanings.
From this point forward, we will collapse the functional arguments of all functions of coordinates into the symbol ZZ. For example, the function R(Z1,Z2,Z3)\mathbf{R}\left( Z^{1},Z^{2},Z^{3}\right) will be denoted by the symbol
R(Z).(9.4)\mathbf{R}\left( Z\right) .\tag{9.4}
In addition to its compactness, the symbol R(Z)\mathbf{R}\left( Z\right) has the advantage that it works in any number of dimensions. The alternative symbol R(Zi)\mathbf{R}\left( Z^{i}\right) cannot be used since it makes the index ii appear live and leaves it hanging, which violates the rules of the tensor notation.
The function R(Z)\mathbf{R}\left( Z\right) is the starting point for a crucial sequence of definitions. Before we turn our attention to those definitions, it is important to reiterate that the position vector R\mathbf{R} is a primary object that is to be understood and accepted on its own geometric terms. It is counterproductive to imagine some a priori basis with respect to which R\mathbf{R} could be represented by its components. There is simply no such basis available -- nor is one needed.

9.2.1A note on the overloaded use of the letter ZZ

The remainder of the objects introduced in this Chapter are denoted by symbols anchored by the letter ZZ. This includes the covariant basis Zi\mathbf{Z}_{i} , the covariant metric tensor ZijZ_{ij}, the volume element Z\sqrt{Z}, the contravariant metric tensor ZijZ^{ij}, and the contravariant basis Zi\mathbf{Z}^{i}. Despite the fact that the same letter is used in each of these symbols, they can be easily distinguished by their indicial signatures. Additionally, the symbol Zi\mathbf{Z}^{i} for the contravariant basis is distinguished from the symbol ZiZ^{i} denoting the coordinates by the fact that the letter ZZ is textbf{bold} in the former and plain in the latter.

9.2.2The definition

The covariant basis Z1\mathbf{Z}_{1}, Z2\mathbf{Z}_{2}, Z3\mathbf{Z}_{3} -- or, collectively, Zi\mathbf{Z}_{i} -- is a generalization of the affine coordinate basis i\mathbf{i}, j\mathbf{j}, k\mathbf{k} to curvilinear coordinates. It is constructed from the position vector R(Z)\mathbf{R}\left( Z\right) by differentiation with respect to each of the coordinates, i.e.
Z1=R(Z1,Z2,Z3)Z1          (9.5)Z2=R(Z1,Z2,Z3)Z2          (9.6)Z3=R(Z1,Z2,Z3)Z3.          (9.7)\begin{aligned}\mathbf{Z}_{1} & =\frac{\partial\mathbf{R}\left( Z^{1},Z^{2},Z^{3}\right) }{\partial Z^{1}}\ \ \ \ \ \ \ \ \ \ \left(9.5\right)\\\mathbf{Z}_{2} & =\frac{\partial\mathbf{R}\left( Z^{1},Z^{2},Z^{3}\right) }{\partial Z^{2}}\ \ \ \ \ \ \ \ \ \ \left(9.6\right)\\\mathbf{Z}_{3} & =\frac{\partial\mathbf{R}\left( Z^{1},Z^{2},Z^{3}\right) }{\partial Z^{3}}.\ \ \ \ \ \ \ \ \ \ \left(9.7\right)\end{aligned}
Of course, the vectors Z1\mathbf{Z}_{1}, Z2\mathbf{Z}_{2}, and Z3\mathbf{Z} _{3} constitute a legitimate basis only when they form a linearly independent set. Points where this is not the case are called singular and almost always require special treatment.
As we have already discussed on several occasions, the term covariant describes the placement of the index as a subscript and, correspondingly, the manner in which the basis Zi\mathbf{Z}_{i} transforms under a coordinates transformation. For the time being, we will use the term covariant without exploring its deeper meaning. Also note that, in reference to Zi\mathbf{Z}_{i}, we will frequently drop the term covariant and refer to Zi\mathbf{Z}_{i} simply as the basis.
It is not surprising that the first new fundamental object in a Euclidean space is introduced by means of partial differentiation with respect to the coordinates ZiZ^{i}. After all, it is just about the only interesting operation that we have at our disposal that can be applied to R(Z)\mathbf{R} \left( Z\right) . Furthermore, it is remarkable that the generalization of the coordinate basis from affine to general coordinates is accomplished by such a simple operation. Just like that, we are able to decompose vectors, and thus conduct component analysis, in general coordinate systems. Keep in mind, however, that -- unlike the coordinate basis i\mathbf{i}, j\mathbf{j}, k\mathbf{k} in affine coordinates -- the covariant basis Zi\mathbf{Z}_{i} varies from one point to another and therefore the component space is specific to each point in space. As a result, we must imagine an independent linear space at each point and work out how each space interacts with the neighboring spaces. In the near future, we will have a lot to say about this.
Since we have decided to represent the function R(Z1,Z2,Z3)\mathbf{R}\left( Z^{1} ,Z^{2},Z^{3}\right) by the symbol R(Z)\mathbf{R}\left( Z\right) , the definition of the covariant basis can be expressed by the more compact equations
Z1=R(Z)Z1;    Z2=R(Z)Z2;    Z3=R(Z)Z3.(9.8)\mathbf{Z}_{1}=\frac{\partial\mathbf{R}\left( Z\right) }{\partial Z^{1} };\ \ \ \ \mathbf{Z}_{2}=\frac{\partial\mathbf{R}\left( Z\right) }{\partial Z^{2}};\ \ \ \ \mathbf{Z}_{3}=\frac{\partial\mathbf{R}\left( Z\right) }{\partial Z^{3}}.\tag{9.8}
With the help of the tensor notation, the above definitions can be combined into a single indicial equation, i.e.
Zi=R(Z)Zi.(9.9)\mathbf{Z}_{i}=\frac{\partial\mathbf{R}\left( Z\right) }{\partial Z^{i}}.\tag{9.9}
The fact that the covariant basis ends up with a subscript illustrates an important feature of the tensor notation which we have mentioned earlier on a number of occasions. Namely, that by strongly suggesting the proper placements of indices, the tensor notation tends to predict the manner in which objects transform under coordinate transformations. For the covariant basis Zi\mathbf{Z}_{i}, the subscript is a natural choice since, as we discussed in Chapter 7, in the expression
R(Z)Zi(9.10)\frac{\partial\mathbf{R}\left( Z\right) }{\partial Z^{i}}\tag{9.10}
the index ii appears as a superscript in the "denominator". Thus, the tensor notation predicts that the covariant basis transforms in a manner opposite of the coordinates. In Chapter 6, we have already observed that this is the case for transformations between affine coordinates. For general coordinate transformations, this will be confirmed in Chapter 14.
The covariant basis Zi\mathbf{Z}_{i} is used for the decomposition of vectors at a given point. When a vector U\mathbf{U} is decomposed with respect to Zi\mathbf{Z}_{i}, the resulting coefficients are naturally assigned superscripts, i.e.
U=U1Z1+U2Z2+U3Z3,(9.11)\mathbf{U}=U^{1}\mathbf{Z}_{1}+U^{2}\mathbf{Z}_{2}+U^{3}\mathbf{Z}_{3},\tag{9.11}
in order to utilize the summation conventions in the equation
U=UiZi.(9.12)\mathbf{U}=U^{i}\mathbf{Z}_{i}.\tag{9.12}
The components UiU^{i} of a vector U\mathbf{U} with respect to the covariant basis Zi\mathbf{Z}_{i} are referred to as the contravariant components. Of course, the broad goal of our analysis is to replace vectors with their components so that we are able to solve problems by analytical rather than geometric means. Thus, a detailed discussion of the components of vectors and their use will be postponed until the next Chapter devoted entirely to coordinate space analysis.
Finally, note that it is not advisable to decompose the elements of the basis Zi\mathbf{Z}_{i} itself with respect to some a priori supplementary basis. Given the prevalent use of Cartesian coordinates, the temptation may be great to think of each vector Zi\mathbf{Z}_{i} as a triplet of numbers with respect to some background Cartesian basis i\mathbf{i}, j\mathbf{j}, k\mathbf{k}. However, this would obscure the primary decompositional role of the covariant basis. The covariant basis Zi\mathbf{Z}_{i} is used for decomposing other vectors but itself need not be decomposed. The buck, so to say, stops with Zi\mathbf{Z}_{i}.

9.2.3Visualizing the covariant basis

The primary geometric characteristic of the vectors Zi\mathbf{Z}_{i} is that they are tangential to the corresponding coordinate lines.
(9.13)
To see why this is so for, say,textbf{ }Z1\mathbf{Z}_{1}, consider the function
R(γ)=R(γ,Z2,Z3).(9.14)\mathbf{R}\left( \gamma\right) =\mathbf{R}\left( \gamma,Z^{2},Z^{3}\right) .\tag{9.14}
(Note that we are now using the letter R\mathbf{R} to denote yet another function, this time of one variable γ\gamma.) The vectors R(γ)\mathbf{R}\left( \gamma\right) trace out the coordinate line corresponding to Z1Z^{1}. Therefore, as we studied in Chapter 5, the derivative
R(γ)(9.15)\mathbf{R}^{\prime}\left( \gamma\right)\tag{9.15}
is tangential to that coordinate line. Of course, since R(γ)\mathbf{R}^{\prime }\left( \gamma\right) is the same as
R(Z1,Z2,Z3)Z1,(9.16)\frac{\partial\mathbf{R}\left( Z^{1},Z^{2},Z^{3}\right) }{\partial Z^{1}},\tag{9.16}
we conclude that the vector Z1\mathbf{Z}_{1} is tangential to the coordinate line corresponding to Z1Z^{1}, as we set out to show.
The covariant basis is particularly easy to visualize when the coordinate lines are drawn at integer increments, as in the following two-dimensional figure.
(9.17)
In the context of such a representation, the approximate length of Zi\mathbf{Z}_{i} at a node is such that its tip is located in the vicinity of the neighboring node whose coordinate ZiZ^{i} exceeds that of its tail by h=1h=1. For example, the tip of the vector Z1\mathbf{Z}_{1} at the point with coordinates (Z1,Z2,Z3)\left( Z^{1},Z^{2},Z^{3}\right) falls near the point with coordinates (Z1+1,Z2,Z3)\left( Z^{1}+1,Z^{2},Z^{3}\right) . This is so because, in the limit
Z1=limh0R(Z1+h,Z2,Z3)R(Z1,Z2,Z3)h,(9.18)\mathbf{Z}_{1}=\lim_{h\rightarrow0}\frac{\mathbf{R}\left( Z^{1}+h,Z^{2} ,Z^{3}\right) -\mathbf{R}\left( Z^{1},Z^{2},Z^{3}\right) }{h},\tag{9.18}
the "intermediate" vector corresponding to the finite value h=1h=1 falls exactly at the point with coordinates (Z1+1,Z2,Z3)\left( Z^{1}+1,Z^{2} ,Z^{3}\right) and, in the limit as hh approaches 00, it ought to be approximately so.

9.2.4The covariant basis in various coordinate systems

9.2.5In affine coordinates

For general curvilinear coordinate systems, the covariant basis varies from one point to another. In an affine coordinate system, thanks to its regularity, the covariant basis is the same at all points and coincides with the coordinate basis i\mathbf{i}, j\mathbf{j}, k\mathbf{k}.
  (9.19)
In other words, the basis vector Zi\mathbf{Z}_{i} points from the point with coordinates (Z1,Z2,Z3)\left( Z^{1},Z^{2},Z^{3}\right) to the point whose ii-th coordinate is increased by 11. For example, Z2\mathbf{Z}_{2} points from the point with coordinates (Z1,Z2,Z3)\left( Z^{1},Z^{2},Z^{3}\right) to the point with coordinates (Z1,Z2+1,Z3)\left( Z^{1},Z^{2}+1,Z^{3}\right) . In curvilinear coordinates, the preceding statement is approximately true at points where the magnitude of the second derivatives of R(Z)\mathbf{R}\left( Z\right) is not too great.
Since Cartesian coordinates are a special case of affine coordinates, the Cartesian covariant basis is the same at all points. Cartesian coordinate lines form a regular orthogonal unit grid and therefore the covariant basis consists of orthogonal unit, i.e. orthonormal, vectors. The following figure shows the Cartesian covariant bases in two and three dimensions.
  (9.20)
In conclusion, the familiar affine basis i\mathbf{i}, j\mathbf{j}, k\mathbf{k} fits perfectly into the new framework where the coordinate basis is constructed by differentiating the position vector R\mathbf{R} with respect to each of the coordinates.

9.2.6In polar coordinates

In polar coordinates r,θr,\theta, the coordinate Z1Z^{1} corresponds to rr and Z2Z^{2} corresponds to θ\theta. The covariant basis vector Z1\mathbf{Z}_{1}, given by the equation
Z1=R(r,θ)r,(9.21)\mathbf{Z}_{1}=\frac{\partial\mathbf{R}\left( r,\theta\right) }{\partial r},\tag{9.21}
is a unit vector that points in the radial direction away from the origin OO. Demonstrating this fact is left as an exercise. The vector Z2\mathbf{Z}_{2} is given by
Z2=R(r,θ)θ.(9.22)\mathbf{Z}_{2}=\frac{\partial\mathbf{R}\left( r,\theta\right) } {\partial\theta}.\tag{9.22}
Recall from Section 4.3 that the derivative R(γ)\mathbf{R}^{\prime}\left( \gamma\right) of the position vector R(γ)\mathbf{R}\left( \gamma\right) tracing out a circle of radius rr is a vector tangential to the circle also of length rr.
(4.24)
Of course, this configuration corresponds exactly to the calculation of Z2\mathbf{Z}_{2}. Thus, Z2\mathbf{Z}_{2} is a vector of length rr that points in the counterclockwise tangential direction to the coordinate circle. These findings regarding the covariant basis Zi\mathbf{Z}_{i} for polar coordinates are illustrated in the following figure.
(9.23)
Observe that the covariant basis in polar coordinates is orthogonal. Interestingly, in some textbooks, the vector Z2\mathbf{Z}_{2} is scaled to unit length in order to produce an orthonormal basis at every point. This is highly inadvisable due to the numerous disadvantages in exchange for little gain.
Polar coordinates present us with our first example of a coordinate basis that varies from one point to another. This variability is of enormous importance. Some of its most crucial implications are described below in Section 9.2.10 and will be discussed further later in our narrative.

9.2.7In cylindrical coordinates

Cylindrical coordinates augment polar coordinates r,θr,\theta with the coordinate zz whose coordinate lines are straight lines orthogonal to the coordinate plane. Within each plane parallel to the coordinate plane, i.e. within each coordinate surface corresponding to a constant zz, the vectors
Z1=R(r,θ,z)r and Z2=R(r,θ,z)θ(9.24)\mathbf{Z}_{1}=\frac{\partial\mathbf{R}\left( r,\theta,z\right) }{\partial r}\text{ and }\mathbf{Z}_{2}=\frac{\partial\mathbf{R}\left( r,\theta ,z\right) }{\partial\theta}\tag{9.24}
coincide with the covariant basis in polar coordinates. The additional vector Z3\mathbf{Z}_{3}, given by
Z3=R(r,θ,z)z,(9.25)\mathbf{Z}_{3}=\frac{\partial\mathbf{R}\left( r,\theta,z\right) }{\partial z},\tag{9.25}
is a constant unit vector orthogonal to the coordinate plane that points in the "upward" direction -- in other words, the set Z1\mathbf{Z}_{1}, Z2\mathbf{Z}_{2}, Z3\mathbf{Z}_{3} is positively oriented.
(9.26)
Note that even with the addition of Z3\mathbf{Z}_{3}, the covariant basis remains orthogonal.

9.2.8In spherical coordinates

In order to describe the covariant basis Zi\mathbf{Z}_{i} in spherical coordinates r,θ,φr,\theta,\varphi, we will use a combination of words and figures. The first vector
Z1=R(r,θ,φ)r(9.27)\mathbf{Z}_{1}=\frac{\partial\mathbf{R}\left( r,\theta,\varphi\right) }{\partial r}\tag{9.27}
is a unit vector that points in the radial direction away from the origin OO. The second vector
Z2=R(r,θ,φ)θ(9.28)\mathbf{Z}_{2}=\frac{\partial\mathbf{R}\left( r,\theta,\varphi\right) }{\partial\theta}\tag{9.28}
corresponds to the rate of change in the position vector R\mathbf{R} along a meridian. Thus, Z2\mathbf{Z}_{2} is a vector of length rr that points in the direction tangential to the meridian away from the north pole, i.e. the point with θ=0\theta=0. The third vector
Z3=R(r,θ,φ)φ(9.29)\mathbf{Z}_{3}=\frac{\partial\mathbf{R}\left( r,\theta,\varphi\right) }{\partial\varphi}\tag{9.29}
corresponds to the rate of change in the position vector R\mathbf{R} along a parallel and is thus tangential to that parallel. Since a parallel at colatitude θ\theta has radius rsinθr\sin\theta, the length of Z3\mathbf{Z}_{3} is also rsinθr\sin\theta. The following figure shows the basis vectors Z2\mathbf{Z}_{2} and Z3\mathbf{Z}_{3} on the coordinate sphere of radius rr.
(9.30)
As was the case with cylindrical coordinates, the basis Z1,Z2,Z3\mathbf{Z} _{1},\mathbf{Z}_{2},\mathbf{Z}_{3} is positively oriented. Furthermore, it is orthogonal but not orthonormal.

9.2.9Orthogonal coordinate systems

As we have pointed out, the covariant bases for Cartesian, polar, cylindrical, and spherical coordinates are orthogonal. Coordinate systems that have this property are themselves known as orthogonal. Since the covariant basis vectors are tangential to the coordinate lines, orthogonal systems can be equivalently characterized by orthogonal coordinate lines. There is no particular significance to orthogonal coordinate systems, except that some calculations are simplified. The following figure shows the coordinate lines for a generic orthogonal system.
(9.31)
You may be wondering whether there exist orthonormal coordinate systems other than Cartesian. The surprising answer is no: as we discuss below in Section 9.3.6, all orthonormal coordinates are necessarily Cartesian.

9.2.10On the spatial variability of the covariant basis

If you are coming from an exclusively affine background, the implications of the spatial variability of the covariant basis in curvilinear coordinates may require some getting used to. In some ways, the difference is dramatic. For example, the components of one and the same vector calculated at different points are likely to be distinct. Conversely, two vectors with identical components are likely to be distinct. Furthermore, vectors at different points cannot be added together by adding their components. This insight has a number implications, including for integration which, in essence, is a form of addition. Suppose that gi(Z)g^{i}\left( Z\right) represent the components of the gravitational force field per unit mass acting upon a body with density ρ\rho that occupies the domain Ω\Omega. If the coordinate system as affine, then the components FiF^{i} of the total force F\mathbf{F} are given by the integral
Fi=ΩρgidΩ.(9.32)F^{i}=\int_{\Omega}\rho g^{i}d\Omega.\tag{9.32}
In curvilinear coordinates, on the other hand, the above integral is meaningless since it requires the addition of components of vectors calculated with respect to (an infinity of) different bases.
These difficulties, however, should not in any way dissuade us from utilizing curvilinear coordinates whose use is essential in most situations. Furthermore, in many geometric spaces, such as the surface of a sphere and most other curved surfaces, curvilinear coordinates are the only available option. And even when affine coordinates are feasible, curvilinear coordinates may nevertheless be a natural choice and experience shows that overcoming difficulties that stem from natural choices is always a worthwhile endeavor. Indeed, our future experience will demonstrate that allowing curvilinear coordinates is an unequivocal improvement over the exclusive use of affine coordinates. The additional complexity will prove to be not an obstacle but an impetus for deeper insights that would not have been come form affine analysis.

9.3.1The definition

We have arrived at one of the central objects in our subject: the covariant metric tensor. By definition, the elements of the covariant metric tensor ZijZ_{ij} are the pairwise dot products of the covariant basis vectors Zi\mathbf{Z}_{i}, i.e.
Zij=ZiZj.(9.33)Z_{ij}=\mathbf{Z}_{i}\cdot\mathbf{Z}_{j}.\tag{9.33}
Since the dot product is commutative, the metric tensor is symmetric, i.e.
Zij=Zji,(9.34)Z_{ij}=Z_{ji},\tag{9.34}
and can be organized into a symmetric 3×33\times3 matrix
 [Z1Z1Z1Z2Z1Z3Z2Z1Z2Z2Z2Z3Z3Z1Z3Z2Z3Z3].(9.35)\text{ }\left[ \begin{array} {ccc} \mathbf{Z}_{1}\cdot\mathbf{Z}_{1} & \mathbf{Z}_{1}\cdot\mathbf{Z}_{2} & \mathbf{Z}_{1}\cdot\mathbf{Z}_{3}\\ \mathbf{Z}_{2}\cdot\mathbf{Z}_{1} & \mathbf{Z}_{2}\cdot\mathbf{Z}_{2} & \mathbf{Z}_{2}\cdot\mathbf{Z}_{3}\\ \mathbf{Z}_{3}\cdot\mathbf{Z}_{1} & \mathbf{Z}_{3}\cdot\mathbf{Z}_{2} & \mathbf{Z}_{3}\cdot\mathbf{Z}_{3} \end{array} \right] .\tag{9.35}
This matrix is entirely analogous to the dot product matrix
M=[b1b1b1b2b1b3b2b1b2b2b2b3b3b1b3b2b3b3](2.53)M=\left[ \begin{array} {ccc} \mathbf{b}_{1}\cdot\mathbf{b}_{1} & \mathbf{b}_{1}\cdot\mathbf{b}_{2} & \mathbf{b}_{1}\cdot\mathbf{b}_{3}\\ \mathbf{b}_{2}\cdot\mathbf{b}_{1} & \mathbf{b}_{2}\cdot\mathbf{b}_{2} & \mathbf{b}_{2}\cdot\mathbf{b}_{3}\\ \mathbf{b}_{3}\cdot\mathbf{b}_{1} & \mathbf{b}_{3}\cdot\mathbf{b}_{2} & \mathbf{b}_{3}\cdot\mathbf{b}_{3} \end{array} \right] \tag{2.53}
from Chapter 2.
The central role of the covariant metric tensor is to facilitate the calculation of dot products -- and thus of lengths and angles -- in the coordinate space. In the next Chapter, we will show that the dot product UV\mathbf{U}\cdot\mathbf{V} of vectors with components UiU^{i} and ViV^{i} is given by
UV=ZijUiVj.(10.25)\mathbf{U}\cdot\mathbf{V}=Z_{ij}U^{i}V^{j}. \tag{10.25}
Thus, the term metric refers to covariant metric tensor's role in representing geometric measurements.
A few sections below, we will introduce the contravariant metric tensor ZijZ^{ij} as the matrix inverse of the covariant metric tensor ZijZ_{ij}. Because it will almost always be clear from the context which of the two objects we are referring to, the adjectives covariant and contravariant are often dropped and the shorter term metric tensor is used to describe both tensors. The terms fundamental tensor and fundamental form can also be used to describe both metric tensors.
When we write the covariant basis directly in terms of the position vector R\mathbf{R}, i.e.
Zij=R(Z)ZiR(Z)Zj,(9.36)Z_{ij}=\frac{\partial\mathbf{R}\left( Z\right) }{\partial Z^{i}}\cdot \frac{\partial\mathbf{R}\left( Z\right) }{\partial Z^{j}},\tag{9.36}
we see the entire scope of its construction from the position vector which we chose as the starting point of our analysis. The recipe requires the ability to measure distances and angles (which go into the dot product), as well as the ability to differentiate vector-valued functions with respect to the coordinates. However, once the metric tensor ZijZ_{ij} is calculated, it can be used to calculate the dot product, and therefore distances and angles, in the component space by analytical means. Furthermore, as we will soon discover, the metric tensor plays a crucial role in the component space differentiation of vector-valued functions, again by analytical means. These abservations suggest the possibility of an entirely new approach to the subject of Geometry where the metric tensor -- rather than lengths and angles -- serves as the starting point. The fundamentals of this approach, known as Riemannian Geometry, will be outlined in Chapter 20.

9.3.2In affine and Cartesian coordinates

In affine coordinates, the covariant basis Zi\mathbf{Z}_{i} is the same at all points and, therefore, so is the metric tensor ZijZ_{ij}. Using the symbols i\mathbf{i}, j\mathbf{j}, and k\mathbf{k} to denote the elements of the covariant basis, the metric tensor ZijZ_{ij} correspond to the constant matrix
[iiijikjijjjkkikjkk].(9.37)\left[ \begin{array} {ccc} \mathbf{i\cdot i} & \mathbf{i\cdot j} & \mathbf{i\cdot k}\\ \mathbf{j\cdot i} & \mathbf{j\cdot j} & \mathbf{j\cdot k}\\ \mathbf{k\cdot i} & \mathbf{k\cdot j} & \mathbf{k\cdot k} \end{array} \right] .\tag{9.37}
For a specific two-dimensional example, consider the affine coordinate system illustrated in the following figure, where the length of i\mathbf{i} is 11, the length of j\mathbf{j} is 22, and the angle between i\mathbf{i} and j\mathbf{j} is π/3\pi/3.
(9.38)
Then
Zij corresponds to [1×1×cos01×2×cosπ32×1×cosπ32×2×cos0]=[1114].(9.39)Z_{ij}\text{ corresponds to }\left[ \begin{array} {ll} 1\times1\times\cos0 & 1\times2\times\cos\frac{\pi}{3}\\ 2\times1\times\cos\frac{\pi}{3} & 2\times2\times\cos0 \end{array} \right] =\left[ \begin{array} {cc} 1 & 1\\ 1 & 4 \end{array} \right] .\tag{9.39}
In Cartesian coordinates, the covariant basis i,j,k\mathbf{i,j,k} is orthonormal and, therefore, ZijZ_{ij} corresponds to the identity matrix
 [100010001].(9.40)\text{ }\left[ \begin{array} {ccc} 1 & 0 & 0\\ 0 & 1 & 0\\ 0 & 0 & 1 \end{array} \right] \text{.}\tag{9.40}
Generally speaking, in an nn-dimensional Euclidean space, the elements of the Cartesian metric tensor can be captured by the equation
Zij={1, if i=j0, if ij.(9.41)Z_{ij}=\left\{ \begin{array} {l} 1\text{, if }i=j\\ 0\text{, if }i\neq j. \end{array} \right.\tag{9.41}
Note that we are careful not to use the Kronecker delta symbol δji\delta_{j}^{i} on the right since its indicial signature does not match that of ZijZ_{ij} and therefore the equation Zij=δjiZ_{ij}=\delta_{j}^{i} would have been invalid from the tensor notation point of view. That said, once we introduce index juggling in Chapter 11, we will establish a surprising equivalence between the metric tensor and the Kronecker delta symbol.

9.3.3In polar coordinates

Recall that the covariant basis in polar coordinates consists of two orthogonal vectors Z1\mathbf{Z}_{1} and Z2\mathbf{Z}_{2} of lengths 11 and rr. Therefore, the covariant metric tensor ZijZ_{ij} has two nonzero elements
Z11=1 and Z22=r2.(9.42)Z_{11}=1\text{ and }Z_{22}=r^{2}.\tag{9.42}
Thus,
Zij corresponds to [100r2].(9.43)Z_{ij}\text{ corresponds to }\left[ \begin{array} {cc} 1 & 0\\ 0 & r^{2} \end{array} \right] .\tag{9.43}
The diagonal property of the resulting matrix reflects the orthogonality of the coordinate system.
It is interesting to note that the elements of ZijZ_{ij} have varying "units" of length. If meter is used as the unit of length in the Euclidean space, then Z11Z_{11} is dimensionless while Z22Z_{22} is measured in square meters. The units of the off-diagonal zero entries are meters. While the issue of units is quite subtle, we will be able to side step it completely throughout our narrative.

9.3.4In cylindrical coordinates

Since the cylindrical covariant basis consists of orthogonal vectors Z1\mathbf{Z}_{1}, Z2\mathbf{Z}_{2} and Z3\mathbf{Z}_{3} of lengths 11, rr, and 11,
Zij corresponds to [1000r20001].(9.44)Z_{ij}\text{ corresponds to }\left[ \begin{array} {ccc} 1 & 0 & 0\\ 0 & r^{2} & 0\\ 0 & 0 & 1 \end{array} \right] .\tag{9.44}

9.3.5In spherical coordinates

Based on the above description of the covariant basis Zi\mathbf{Z}_{i} in spherical coordinates, it is an entirely straightforward matter to show that in spherical coordinates,
Zij corresponds to [1000r2000r2sin2θ].(9.45)Z_{ij}\text{ corresponds to }\left[ \begin{array} {ccc} 1 & 0 & 0\\ 0 & r^{2} & 0\\ 0 & 0 & r^{2}\sin^{2}\theta \end{array} \right] .\tag{9.45}

9.3.6Restoring the coordinate system from the metric tensor

Two different coordinate systems related by a rigid transformation, i.e. a combination of rotation, reflection, and translation, produce the same covariant metric tensor function Zij(Z)Z_{ij}\left( Z\right) . For example, the following figure shows a rotation of a coordinate system by an angle α\alpha followed by translation by a vector d\mathbf{d}.
(9.46)
It is clear that Zij(Z)Z_{ij}\left( Z\right) is unchanged under this transformation since the relative arrangement of the covariant basis vectors is exactly the same at corresponding points.
Interestingly, the converse is also true: if two coordinate systems are characterized by the same covariant basis function Zij(Z)Z_{ij}\left( Z\right) then they are necessarily related by a rigid transformation. This statement is the object of Problem 9.1 at the end of this Chapter. This observation shows that the covariant metric tensor ZijZ_{ij} retains a great deal of information about the coordinate system. In fact, it retains all the information, except its absolute location and orientation (in the sense of rotation and reflection).
In particular, if the elements of the metric tensor ZijZ_{ij} form the identity matrix, then the coordinate system is necessarily Cartesian. In other words, all orthonormal coordinates are Cartesian. In other words, there cannot exist a coordinate system, such as the one "illustrated" in the following figure, where the covariant basis Zi\mathbf{Z}_{i} is orthonormal at every point, while the coordinate lines are not straight.
(9.47)

9.4.1The definition

Denote by ZZ the determinant of the matrix corresponding to covariant metric tensor ZijZ_{ij}, i.e.
Z=Z1Z1Z1Z2Z1Z3Z2Z1Z2Z2Z2Z3Z3Z1Z3Z2Z3Z3.(9.48)Z=\left\vert \begin{array} {ccc} \mathbf{Z}_{1}\cdot\mathbf{Z}_{1} & \mathbf{Z}_{1}\cdot\mathbf{Z}_{2} & \mathbf{Z}_{1}\cdot\mathbf{Z}_{3}\\ \mathbf{Z}_{2}\cdot\mathbf{Z}_{1} & \mathbf{Z}_{2}\cdot\mathbf{Z}_{2} & \mathbf{Z}_{2}\cdot\mathbf{Z}_{3}\\ \mathbf{Z}_{3}\cdot\mathbf{Z}_{1} & \mathbf{Z}_{3}\cdot\mathbf{Z}_{2} & \mathbf{Z}_{3}\cdot\mathbf{Z}_{3} \end{array} \right\vert .\tag{9.48}
According to our discussion in Chapter 3, ZZ equals the square of the volume of the parallelepiped formed by the vectors Z1\mathbf{Z}_{1}, Z2\mathbf{Z}_{2}, and Z3\mathbf{Z}_{3}. Therefore, its square root
Z(9.49)\sqrt{Z}\tag{9.49}
equals the conventional (unsigned) volume of the same parallelepiped. As a result, Z\sqrt{Z} is known as the volume element. In a two-dimensional space, Z\sqrt{Z} may be referred to as the area element while in a one-dimensional space, it may be referred to as the line element. When the dimension of the space is not specified, the term volume element is preferred.
You will recall that the symbol ZZ is also used to denote the coordinate arguments of a function in compressed fashion, as in the symbol R(Z)\mathbf{R} \left( Z\right) which represents the function R(Z1,Z2,Z3)\mathbf{R}\left( Z^{1},Z^{2},Z^{3}\right) . Of course, whether ZZ is used in the sense of the determinant of ZijZ_{ij} or the coordinates ZiZ^{i} as the arguments of a function is always clear from the context. In fact, since the determinant of ZijZ_{ij} can be treated as a function of the coordinates ZiZ^{i}, the two symbols can be combined into a single expression Z(Z)Z\left( Z\right) that represents the function Z(Z1,Z2,Z3)Z\left( Z^{1},Z^{2},Z^{3}\right) , i.e. the determinant of ZijZ_{ij} as a function of the coordinates. In Chapter 16, we will consider the derivative Z(Z)/Zi\partial Z\left( Z\right) /\partial Z^{i} of this function.

9.4.2The role of Z\sqrt{Z} in integration

The volume element Z\sqrt{Z} plays an important role in many aspects of Tensor Calculus. Due to the fact that it equals volume of the parallelepiped formed by the elements of the covariant basis, it is not surprising that it is featured prominently in integration. While the topic of integration will be discussed thoroughly in a future book, the purpose of this Section is to illustrate, on an intuitive level, why Z\sqrt{Z} appears as a factor in the coordinate representation of integrals.
Suppose that a two-dimensional domain Ω\Omega is referred to a curvilinear coordinate system ZiZ^{i}. Consider the covariant basis Z1,Z2\mathbf{Z} _{1},\mathbf{Z}_{2} at a point with coordinates (Z1,Z2)\left( Z^{1},Z^{2}\right) .
(9.50)
It is apparent that the area of the parallelogram formed by Z1\mathbf{Z}_{1} and Z2\mathbf{Z}_{2} is approximately equal to the coordinate parallelogram formed by the curved segments of the coordinate lines connecting the points with coordinates (Z1,Z2)\left( Z^{1},Z^{2}\right) , (Z1+1,Z2)\left( Z^{1}+1,Z^{2}\right) , (Z1+1,Z2+1)\left( Z^{1}+1,Z^{2}+1\right) , and (Z1,Z2+1)\left( Z^{1},Z^{2}+1\right) . Of course, the difference between the two areas is relatively significant since h=1h=1 is a relatively large increment. For a smaller increment ΔZ1=ΔZ2=h\Delta Z^{1}=\Delta Z^{2}=h, the difference between the scaled value Zh2\sqrt{Z}h^{2} and the area of the coordinate parallelogram with vertices at the points (Z1,Z2)\left( Z^{1} ,Z^{2}\right) , (Z1+h,Z2)\left( Z^{1}+h,Z^{2}\right) , (Z1+h,Z2+h)\left( Z^{1} +h,Z^{2}+h\right) , and (Z1,Z2+h)\left( Z^{1},Z^{2}+h\right) is smaller. The following figure shows the coordinate parallelogram for h=1/3h=1/3 as well as the parallelogram formed by Z1/3\mathbf{Z}_{1}/3 and Z2/3\mathbf{Z}_{2}/3. It is apparent that the discrepancy between the two areas is dramatically smaller.
(9.51)
As the increment hh approaches zero, the discrepancy between the areas approaches zero faster than h2h^{2} and therefore the sum of the growing number of terms Zh2\sqrt{Z}h^{2} approaches the area ΩdΩ\int_{\Omega}d\Omega of the domain, i.e.
ZΔZ1ΔZ2ΩdΩ.(9.52)\sum\sqrt{Z}\Delta Z^{1}\Delta Z^{2}\rightarrow\int_{\Omega}d\Omega.\tag{9.52}
Meanwhile, the sum
ZΔZ1ΔZ2(9.53)\sum\sqrt{Z}\Delta Z^{1}\Delta Z^{2}\tag{9.53}
approaches the repeated integral
ZdZ1dZ2,(9.54)\int\int\sqrt{Z}dZ^{1}dZ^{2},\tag{9.54}
where the limits of integration on the right are chosen as to describe the region Ω\Omega. Thus,
ΩdΩ=ZdZ1dZ2.(9.55)\int_{\Omega}d\Omega=\int\int\sqrt{Z}dZ^{1}dZ^{2}.\tag{9.55}
Similarly, the geometric integral ΩFdΩ\int_{\Omega}Fd\Omega for a field FF can be converted to the repeated integral
ΩFdΩ=F(Z)ZdZ1dZ2.(9.56)\int_{\Omega}Fd\Omega=\int\int F\left( Z\right) \sqrt{Z}dZ^{1}dZ^{2}.\tag{9.56}
We will refer to the repeated integrals on the right as arithmetic integrals since they are divorced from the geometric problem from which they arose and can be evaluated by the techniques of Calculus or by a computational method.
Our experience with geometric integrals such as ΩFdΩ\int_{\Omega}Fd\Omega reflects our broader experience with geometric objects. That is, while they provide us with great geometric insight, they do not give us the ability to make any specific calculations for specific problems. In order to perform a specific calculation, we must convert the geometric integral to an arithmetic integral which requires the use of the volume element Z\sqrt{Z}. The foregoing discussion was our first example of translating a geometric analysis into the coordinate space.

9.4.3The volume element Z\sqrt{Z} in various coordinates

In affine coordinates, the volume element is a constant given by the equation
Z=iiijikjijjjkkikjkk.(9.57)\sqrt{Z}=\sqrt{\left\vert \begin{array} {ccc} \mathbf{i\cdot i} & \mathbf{i\cdot j} & \mathbf{i\cdot k}\\ \mathbf{j\cdot i} & \mathbf{j\cdot j} & \mathbf{j\cdot k}\\ \mathbf{k\cdot i} & \mathbf{k\cdot j} & \mathbf{k\cdot k} \end{array} \right\vert }.\tag{9.57}
A more specific expression can be given only if further details about the coordinate system are available. For the two-dimensional affine coordinates considered earlier, we have
Z=1114=3.(9.58)\sqrt{Z}=\sqrt{\left\vert \begin{array} {cc} 1 & 1\\ 1 & 4 \end{array} \right\vert }=\sqrt{3}.\tag{9.58}
In Cartesian coordinates, the volume element has a constant value of 11, i.e.
Z=1.(9.59)\sqrt{Z}=1.\tag{9.59}
In polar coordinates, we have
Z=100r2=r.(9.60)\sqrt{Z}=\sqrt{\left\vert \begin{array} {cc} 1 & 0\\ 0 & r^{2} \end{array} \right\vert }=r.\tag{9.60}
In cylindrical coordinates, we similarly have
Z=1000r20001=r.(9.61)\sqrt{Z}=\sqrt{\left\vert \begin{array} {ccc} 1 & 0 & 0\\ 0 & r^{2} & 0\\ 0 & 0 & 1 \end{array} \right\vert }=r.\tag{9.61}
Finally, in spherical coordinates, we have
Z=1000r2000r2sin2θ=r2sinθ.(9.62)\sqrt{Z}=\sqrt{\left\vert \begin{array} {ccc} 1 & 0 & 0\\ 0 & r^{2} & 0\\ 0 & 0 & r^{2}\sin^{2}\theta \end{array} \right\vert }=r^{2}\sin\theta.\tag{9.62}
As a demonstration of the utility of Z\sqrt{Z} in integration, let us calculate the area AA of a circle of radius RR in polar coordinates. We have
A=ZdZ1dZ2.(9.63)A=\int\int\sqrt{Z}dZ^{1}dZ^{2}.\tag{9.63}
In other words,
A=02π0Rrdrdθ,(9.64)A=\int\limits_{0}^{2\pi}\int\limits_{0}^{R}rdrd\theta,\tag{9.64}
which is easily evaluate to produce
A=πR2.(9.65)A=\pi R^{2}.\tag{9.65}
Similarly, the volume VV of a sphere of radius RR is given by
V=02π0π0Rr2sinθdrdθdφ=43πR3.(9.66)V=\int\limits_{0}^{2\pi}\int\limits_{0}^{\pi}\int\limits_{0}^{R}r^{2} \sin\theta drd\theta d\varphi=\frac{4}{3}\pi R^{3}.\tag{9.66}
These are classical examples of the advantage of using the appropriate coordinate system for every problem.

9.5.1The definition

The contravariant metric tensor ZijZ^{ij} is defined as the "matrix inverse" of the covariant metric tensor ZijZ_{ij}, i.e.
Zij corresponds to [Z1Z1Z1Z2Z1Z3Z2Z1Z2Z2Z2Z3Z3Z1Z3Z2Z3Z3]1.(9.67)Z^{ij}\text{ corresponds to }\left[ \begin{array} {ccc} \mathbf{Z}_{1}\cdot\mathbf{Z}_{1} & \mathbf{Z}_{1}\cdot\mathbf{Z}_{2} & \mathbf{Z}_{1}\cdot\mathbf{Z}_{3}\\ \mathbf{Z}_{2}\cdot\mathbf{Z}_{1} & \mathbf{Z}_{2}\cdot\mathbf{Z}_{2} & \mathbf{Z}_{2}\cdot\mathbf{Z}_{3}\\ \mathbf{Z}_{3}\cdot\mathbf{Z}_{1} & \mathbf{Z}_{3}\cdot\mathbf{Z}_{2} & \mathbf{Z}_{3}\cdot\mathbf{Z}_{3} \end{array} \right] ^{{\LARGE -1}}.\tag{9.67}
Since the inverse of a symmetric positive definite matrix is itself symmetric and positive definite, the contravariant metric tensor also has this important property.
The definition of the contravariant metric tensor can be captured in the tensor notation, albeit implicitly, by the equation
ZijZjk=δki.(9.68)Z^{ij}Z_{jk}=\delta_{k}^{i}.\tag{9.68}
Thus, the tensor notation strongly suggests the use of superscripts for the new object and the corresponding use of the term contravariant. As usual, this placement will prove to be the correct predictor of the manner in which this object transforms under coordinate transformations.
A well-known fact from Linear Algebra tells us that a matrix commutes with its inverse. Therefore, the contraction ZikZijZ_{ik}Z^{ij}, which we can also write as ZijZkiZ^{ij}Z_{ki}, yields the Kronecker delta symbol, as well, i.e.
ZijZki=δkj.(9.69)Z^{ij}Z_{ki}=\delta_{k}^{j}.\tag{9.69}
Of course, in the case of the metric tensors, this relationship also follows from their symmetric property. However the argument based on the commutativity of a matrix with its inverse is more general.
Looking ahead, the contravariant metric tensor will provide our framework with much needed "contravariance" and thus help balance the numbers of superscripts and subscripts. As we discussed in Chapter 14, such balance is essential to achieving invariance.

9.5.2The operation of inverting the metric tensor

Suppose that the system PiP^{i} is obtained from QjQ_{j} by contraction with the contravariant metric tensor ZijZ^{ij}, i.e.
Pi=ZijQj.(9.70)P^{i}=Z^{ij}Q_{j}.\tag{9.70}
From this relationship, it can be shown -- thanks to the matrix inverse relationship between the two metric tensors -- that QjQ_{j} can be obtained from PiP^{i} by contraction with the covariant metric tensor, i.e.
Qj=ZjiPi.(9.71)Q_{j}=Z_{ji}P^{i}.\tag{9.71}
Of course, the connection between the above two identities is analogous to that between the Linear Algebra identities
y=Ax(9.72)y=Ax\tag{9.72}
and
x=A1y,(9.73)x=A^{-1}y,\tag{9.73}
where AA is an invertible n×nn\times n matrix, while xx and yy are, typically, n×1n\times1 matrices or, more generally, n×mn\times m matrices.
In elementary algebra, we show that x=15yx=\frac{1}{5}y follows from y=5xy=5x by dividing both sides of y=5xy=5x by 55. In Linear Algebra, the analogous tactic is stated in terms of multiplication by A1A^{-1}. Multiplying both sides of
y=Ax(9.72)y=Ax \tag{9.72}
by A1A^{-1}, we have
A1y=A1Ax.(9.74)A^{-1}y=A^{-1}Ax.\tag{9.74}
Since A1A=IA^{-1}A=I and Ix=xIx=x, we find that
A1y=x(9.75)A^{-1}y=x\tag{9.75}
or
x=A1y,(9.73)x=A^{-1}y, \tag{9.73}
as we set out to show.
Let us now use the tensor notation to apply the same logic to the identity
Pi=ZijQj.(9.70)P^{i}=Z^{ij}Q_{j}. \tag{9.70}
Contracting both sides with ZkiZ_{ki}, we have
ZkiPi=ZkiZijQj.(9.76)Z_{ki}P^{i}=Z_{ki}Z^{ij}Q_{j}.\tag{9.76}
Since ZkiZij=δkjZ_{ki}Z^{ij}=\delta_{k}^{j} and δkjQj=Qk\delta_{k}^{j}Q_{j}=Q_{k}, we find
ZkiPi=Qk(9.77)Z_{ki}P^{i}=Q_{k}\tag{9.77}
or
Qk=ZkiPi.(9.78)Q_{k}=Z_{ki}P^{i}.\tag{9.78}
Renaming kk into jj, we arrive at the desired identity
Qj=ZjiPi.(9.71)Q_{j}=Z_{ji}P^{i}. \tag{9.71}
In summary,
Pi=ZijQj     implies     Qj=ZjiPi.(9.79)P^{i}=Z^{ij}Q_{j}\ \ \ \ \ \text{implies }\ \ \ \ Q_{j}=Z_{ji}P^{i}.\tag{9.79}
By analogy with Linear Algebra, we will refer to the tactic that converts the former identity into the latter as inverting the metric tensor. Note that inverting the metric tensor works just as well in the reverse direction, i.e.
Qj=ZjiPi     implies     Pi=ZijQj.(9.80)Q_{j}=Z_{ji}P^{i}\ \ \ \ \ \text{implies }\ \ \ \ P^{i}=Z^{ij}Q_{j}.\tag{9.80}
Furthermore, it also works for systems with indicial signatures of arbitrary complexity. For example,
Pklirst=ZijQjklrst     implies     Qiklrst=ZijPkljrst.(9.81)P_{kl}^{irst}=Z^{ij}Q_{jkl}^{rst}\ \ \ \ \ \text{implies }\ \ \ \ Q_{ikl} ^{rst}=Z_{ij}P_{kl}^{jrst}.\tag{9.81}

9.5.3The contravariant metric tensor in various coordinate systems

In a general affine coordinate system,
Zij corresponds to [iiijikjijjjkkikjkk]1.(9.82)Z^{ij}\text{ corresponds to }\left[ \begin{array} {ccc} \mathbf{i\cdot i} & \mathbf{i\cdot j} & \mathbf{i\cdot k}\\ \mathbf{j\cdot i} & \mathbf{j\cdot j} & \mathbf{j\cdot k}\\ \mathbf{k\cdot i} & \mathbf{k\cdot j} & \mathbf{k\cdot k} \end{array} \right] ^{-1}.\tag{9.82}
In particular, for the specific affine coordinates considered above, where
Zij corresponds to [1114],(9.83)Z_{ij}\text{ corresponds to }\left[ \begin{array} {cc} 1 & 1\\ 1 & 4 \end{array} \right] ,\tag{9.83}
we find, by inverting this matrix, that
Zij corresponds to [43131313].(9.84)Z^{ij}\text{ corresponds to }\left[ \begin{array} {rr} \frac{4}{3} & -\frac{1}{3}\\ -\frac{1}{3} & \frac{1}{3} \end{array} \right] .\tag{9.84}
In Cartesian coordinates, where the covariant metric tensor ZijZ_{ij} corresponds to the identity matrix, so does the contravariant metric tensor ZijZ^{ij}, i.e.
Zij corresponds to [100010001].(9.85)Z^{ij}\text{ corresponds to }\left[ \begin{array} {ccc} 1 & 0 & 0\\ 0 & 1 & 0\\ 0 & 0 & 1 \end{array} \right] .\tag{9.85}
In polar coordinates,
Zij corresponds to [1001r2].(9.86)Z^{ij}\text{ corresponds to }\left[ \begin{array} {cc} 1 & 0\\ 0 & \frac{1}{r^{2}} \end{array} \right] .\tag{9.86}
In cylindrical coordinates,
Zij corresponds to [10001r20001].(9.87)Z^{ij}\ \text{corresponds to }\left[ \begin{array} {ccc} 1 & 0 & 0\\ 0 & \frac{1}{r^{2}} & 0\\ 0 & 0 & 1 \end{array} \right] .\tag{9.87}
In spherical coordinates,
Zij corresponds to [10001r20001r2sin2θ].(9.88)Z^{ij}\text{ corresponds to }\left[ \begin{array} {ccc} 1 & 0 & 0\\ 0 & \frac{1}{r^{2}} & 0\\ 0 & 0 & \frac{1}{r^{2}\sin^{2}\theta} \end{array} \right] .\tag{9.88}
As a preview of the important role of the contravariant metric tensor, consider the formula
iiU=1r2r(r2Ur)+1r2sinθθ(sinθUθ)+1r2sin2θ2Uφ2(18.62)\nabla_{i}\nabla^{i}U=\frac{1}{r^{2}}\frac{\partial}{\partial r}\left( r^{2}\frac{\partial U}{\partial r}\right) +\frac{1}{r^{2}\sin\theta} \frac{\partial}{\partial\theta}\left( \sin\theta\frac{\partial U} {\partial\theta}\right) +\frac{1}{r^{2}\sin^{2}\theta}\frac{\partial^{2} U}{\partial\varphi^{2}} \tag{18.62}
for the Laplacian of a function U(r,θ,φ)U\left( r,\theta,\varphi\right) in spherical coordinates (derived in Chapter 18), where you can see, especially in the last two terms, the distinctive influence of the contravariant metric tensor.

9.5.4The vital role of the contravariant metric tensor

The role of the contravariant metric tensor in Tensor Calculus is indeed indispensable. As an object with superscripts, it provides the necessary counterbalance for the objects with subscripts, of which there is no shortage since differentiation naturally leads to objects with subscripts. Once again, the significance of such a balance is straightforward: invariants -- the ultimate objects of our investigations -- are formed by contracting away all indices, which naturally requires a balance of superscripts and subscripts.
Another insight into the importance of the contravariant metric tensor comes from Section 2.4 where we discussed linear decomposition by means of the dot product. Operating without superscripts, we showed that the components U1U_{1}, U2U_{2}, and U3U_{3} of a vector U\mathbf{U} with respect to a basis b1,b2,b3\mathbf{b}_{1},\mathbf{b}_{2},\mathbf{b}_{3}, not necessarily orthogonal, are given by the equation
[U1U2U3]=[b1b1b1b2b1b3b2b1b2b2b2b3b3b1b3b2b3b3]1[b1Ub2Ub3U].(2.54)\left[ \begin{array} {c} U_{1}\\ U_{2}\\ U_{3} \end{array} \right] =\left[ \begin{array} {ccc} \mathbf{b}_{1}\cdot\mathbf{b}_{1} & \mathbf{b}_{1}\cdot\mathbf{b}_{2} & \mathbf{b}_{1}\cdot\mathbf{b}_{3}\\ \mathbf{b}_{2}\cdot\mathbf{b}_{1} & \mathbf{b}_{2}\cdot\mathbf{b}_{2} & \mathbf{b}_{2}\cdot\mathbf{b}_{3}\\ \mathbf{b}_{3}\cdot\mathbf{b}_{1} & \mathbf{b}_{3}\cdot\mathbf{b}_{2} & \mathbf{b}_{3}\cdot\mathbf{b}_{3} \end{array} \right] ^{-1}\left[ \begin{array} {c} \mathbf{b}_{1}\cdot\mathbf{U}\\ \mathbf{b}_{2}\cdot\mathbf{U}\\ \mathbf{b}_{3}\cdot\mathbf{U} \end{array} \right] . \tag{2.54}
Since the matrix that is being inverted corresponds to the covariant metric tensor ZijZ_{ij}, the inverse corresponds to the contravariant metric tensor ZijZ^{ij}. Thus, we can anticipate that the contravariant metric tensor plays a central role in the determination of the components of a vector. This is, indeed, the case as we will shortly discover in Chapter 10.

9.6.1The definition

The contravariant basis Zi\mathbf{Z}^{i} is obtained by contracting the covariant basis Zj\mathbf{Z}_{j} and the contravariant metric tensor ZijZ^{ij}, i.e.
Zi=ZijZj.(9.89)\mathbf{Z}^{i}=Z^{ij}\mathbf{Z}_{j}.\tag{9.89}
Since ZijZ^{ij} is symmetric, it makes no difference whether the first or the second index is used in the contraction. In Chapter 11, we will learn that the contraction on the right is an example of index juggling.
By inverting the metric tensor, as described in Section 9.5.2, we note the inverse relationship
Zi=ZijZj.(9.90)\mathbf{Z}_{i}=Z_{ij}\mathbf{Z}^{j}.\tag{9.90}
In other words, the covariant basis Zi\mathbf{Z}_{i} can be obtained by contracting the contravariant basis with the covariant metric tensor. Thus, we are able to easily move back and forth between the covariant and contravariant bases by contraction with the appropriate metric tensor.

9.6.2The identity Zij=ZiZjZ^{ij}=\mathbf{Z}^{i}\cdot\mathbf{Z}^{j}

Recall that, by definition, the pairwise dot products of the covariant basis vectors Zi\mathbf{Z}_{i} produce the elements of the covariant metric tensor ZijZ_{ij}, i.e.
Zij=ZiZj.(14.5)Z_{ij}=\mathbf{Z}_{i}\cdot\mathbf{Z}_{j}. \tag{14.5}
Similarly, the pairwise dot products of the elements of contravariant basis Zi\mathbf{Z}^{i} produce the contravariant metric tensor ZijZ^{ij}, i.e.
Zij=ZiZj.(9.91)Z^{ij}=\mathbf{Z}^{i}\cdot\mathbf{Z}^{j}.\tag{9.91}
This relationship is demonstrated by the chain of identities
ZiZj=ZikZkZjlZl=ZikZjlZkl=Zikδkj=Zij.(9.92)\mathbf{Z}^{i}\cdot\mathbf{Z}^{j}=Z^{ik}\mathbf{Z}_{k}\cdot Z^{jl} \mathbf{Z}_{l}=Z^{ik}Z^{jl}Z_{kl}=Z^{ik}\delta_{k}^{j}=Z^{ij}.\tag{9.92}
Justifying each step in the above derivation is left as an exercise.
When viewed side by side, the relationships
Zij=ZiZj(14.5)Z_{ij}=\mathbf{Z}_{i}\cdot\mathbf{Z}_{j} \tag{14.5}
and
Zij=ZiZj(9.91)Z^{ij}=\mathbf{Z}^{i}\cdot\mathbf{Z}^{j} \tag{9.91}
show the close parallel between the covariant and contravariant types of objects, which will become more general with the introduction of index juggling in Chapter 11.

9.6.3The identity ZiZj=δji\mathbf{Z}^{i}\cdot\mathbf{Z}_{j}=\delta_{j}^{i}

Finally, we will now demonstrate the identity
ZiZj=δji(9.93)\mathbf{Z}^{i}\cdot\mathbf{Z}_{j}=\delta_{j}^{i}\tag{9.93}
which shows that the contravariant and covariant bases are mutually orthonormal. In other words, each element of the contravariant basis is orthogonal to every differently-numbered element of the covariant basis and vice versa. Additionally, the dot product of each element of one basis with the same-numbered element of the other basis is 11. This relationship can also be proved by the simple chain of identities
ZiZj=ZikZkZj=ZikZkj=δji.(9.94)\mathbf{Z}^{i}\cdot\mathbf{Z}_{j}=Z^{ik}\mathbf{Z}_{k}\cdot\mathbf{Z} _{j}=Z^{ik}Z_{kj}=\delta_{j}^{i}.\tag{9.94}
As we concluded in the very last sentence of Section 2.4, a vector is uniquely determined by its dot products with the elements of a basis. Thus, the identity
ZiZj=δji(9.93)\mathbf{Z}^{i}\cdot\mathbf{Z}_{j}=\delta_{j}^{i} \tag{9.93}
can actually be taken as the definition of the contravariant basis Zi\mathbf{Z}^{i} as it specifies the values of the dot products of each vector Zi\mathbf{Z}^{i} with every element of the covariant basis Zj\mathbf{Z}_{j}. Indeed, the equation
ZiZj=δji(9.93)\mathbf{Z}^{i}\cdot\mathbf{Z}_{j}=\delta_{j}^{i} \tag{9.93}
gives us an actionable geometric recipe for constructing the contravariant basis. For example, in order to construct Z1\mathbf{Z}^{1} note that it is orthogonal to Z2\mathbf{Z}_{2} and Z3\mathbf{Z}_{3} which uniquely determines the straight line containing Z1\mathbf{Z}^{1}. The fact that the dot product Z1Z1\mathbf{Z}^{1}\cdot\mathbf{Z}_{1} is 11, which is positive, implies that Z1\mathbf{Z}^{1} lies in the same half-space as Z1\mathbf{Z}_{1} -- otherwise the angle between Z1\mathbf{Z}^{1} and Z1\mathbf{Z}_{1} would be greater than π/2\pi/2 resulting in a negative dot product. Therefore, the only remaining unknown is the length of Z1\mathbf{Z}^{1} which is easily determined from the equation
Z1Z1=1,(9.95)\mathbf{Z}^{1}\cdot\mathbf{Z}_{1}=1,\tag{9.95}
i.e.
lenZ1lenZ1cosγ=1,(9.96)\operatorname{len}\mathbf{Z}^{1}\operatorname{len}\mathbf{Z}_{1}\cos\gamma=1,\tag{9.96}
where cosγ\cos\gamma is the already-determined angle between Z1\mathbf{Z}^{1} and Z1\mathbf{Z}_{1}. Thus,
lenZ1=1lenZ1cosγ(9.97)\operatorname{len}\mathbf{Z}^{1}=\frac{1}{\operatorname{len}\mathbf{Z}_{1} \cos\gamma}\tag{9.97}
which completes the construction of Z1\mathbf{Z}^{1}.
The reader may want to think through the two-dimensional case on their own. The figure below shows a covariant basis Zi\mathbf{Z}_{i} and the corresponding contravariant basis Zi\mathbf{Z}^{i} from the example that follows.
(9.98)

9.6.4In various coordinate systems

In affine coordinates, the contravariant basis is given by the matrix identity
[Z1Z2Z3]=[iiijikjijjjkkikjkk]1[ijk],(9.99)\left[ \begin{array} {c} \mathbf{Z}^{1}\\ \mathbf{Z}^{2}\\ \mathbf{Z}^{3} \end{array} \right] =\left[ \begin{array} {ccc} \mathbf{i\cdot i} & \mathbf{i\cdot j} & \mathbf{i\cdot k}\\ \mathbf{j\cdot i} & \mathbf{j\cdot j} & \mathbf{j\cdot k}\\ \mathbf{k\cdot i} & \mathbf{k\cdot j} & \mathbf{k\cdot k} \end{array} \right] ^{-1}\left[ \begin{array} {c} \mathbf{i}\\ \mathbf{j}\\ \mathbf{k} \end{array} \right] ,\tag{9.99}
and nothing more specific can be said in general. For the two-dimensional example considered previously, we have
[Z1Z2]=[43131313][ij],(9.100)\left[ \begin{array} {c} \mathbf{Z}^{1}\\ \mathbf{Z}^{2} \end{array} \right] =\left[ \begin{array} {rr} \frac{4}{3} & -\frac{1}{3}\\ -\frac{1}{3} & \frac{1}{3} \end{array} \right] \left[ \begin{array} {c} \mathbf{i}\\ \mathbf{j} \end{array} \right] ,\tag{9.100}
i.e.
Z1=43i13j          (9.101)Z2=13i+13j.          (9.102)\begin{aligned}\mathbf{Z}^{1} & = \phantom{-} \frac{4}{3}\mathbf{i}-\frac{1}{3}\mathbf{j}\ \ \ \ \ \ \ \ \ \ \left(9.101\right)\\\mathbf{Z}^{2} & =-\frac{1}{3}\mathbf{i}+\frac{1}{3}\mathbf{j.}\ \ \ \ \ \ \ \ \ \ \left(9.102\right)\end{aligned}
The resulting vectors Z1\mathbf{Z}^{1} and Z2\mathbf{Z}^{2} are illustrated in the following figure.
(9.103)
As an exercise, confirm by evaluating dot products, that Z1\mathbf{Z}^{1} is orthogonal to j\mathbf{j} and Z2\mathbf{Z}^{2} is orthogonal to i\mathbf{i}.
In Cartesian coordinates, since the contravariant metric tensor corresponds to the identity matrix, the contravariant and the covariant bases are the same, i.e.
Z1=i          (9.104)Z2=j          (9.105)Z3=k.          (9.106)\begin{aligned}\mathbf{Z}^{1} & =\mathbf{i}\ \ \ \ \ \ \ \ \ \ \left(9.104\right)\\\mathbf{Z}^{2} & =\mathbf{j}\ \ \ \ \ \ \ \ \ \ \left(9.105\right)\\\mathbf{Z}^{3} & =\mathbf{k.}\ \ \ \ \ \ \ \ \ \ \left(9.106\right)\end{aligned}
Recall that, in polar coordinates,
Zij corresponds to [1001r2].(9.86)Z^{ij}\text{ corresponds to }\left[ \begin{array} {cc} 1 & 0\\ 0 & \frac{1}{r^{2}} \end{array} \right] . \tag{9.86}
Thus,
[Z1Z2]=[1001r2][Z1Z2],(9.107)\left[ \begin{array} {c} \mathbf{Z}^{1}\\ \mathbf{Z}^{2} \end{array} \right] =\left[ \begin{array} {cc} 1 & 0\\ 0 & \frac{1}{r^{2}} \end{array} \right] \left[ \begin{array} {c} \mathbf{Z}_{1}\\ \mathbf{Z}_{2} \end{array} \right] ,\tag{9.107}
i.e.
Z1=Z1          (9.108)Z2=1r2Z2.          (9.109)\begin{aligned}\mathbf{Z}^{1} & =\mathbf{Z}_{1}\ \ \ \ \ \ \ \ \ \ \left(9.108\right)\\\mathbf{Z}^{2} & =\frac{1}{r^{2}}\mathbf{Z}_{2}.\ \ \ \ \ \ \ \ \ \ \left(9.109\right)\end{aligned}
In words, the vector Z1\mathbf{Z}^{1} coincides with Z1\mathbf{Z}_{1}, while Z2\mathbf{Z}^{2} points in the same direction as Z2\mathbf{Z}_{2} and has length 1/r1/r. The following figure illustrates the contravariant basis in polar coordinates.
(9.110)
It is left as an exercise to show that in cylindrical coordinates,
Z1=Z1          (9.111)Z2=1r2Z2          (9.112)Z3=Z3.          (9.113)\begin{aligned}\mathbf{Z}^{1} & =\mathbf{Z}_{1}\ \ \ \ \ \ \ \ \ \ \left(9.111\right)\\\mathbf{Z}^{2} & =\frac{1}{r^{2}}\mathbf{Z}_{2}\ \ \ \ \ \ \ \ \ \ \left(9.112\right)\\\mathbf{Z}^{3} & =\mathbf{Z}_{3}.\ \ \ \ \ \ \ \ \ \ \left(9.113\right)\end{aligned}
Likewise, it is left as an exercise to show that in spherical coordinates,
Z1=Z1          (9.114)Z2=1r2Z2          (9.115)Z3=1r2sin2θZ3.          (9.116)\begin{aligned}\mathbf{Z}^{1} & =\mathbf{Z}_{1}\ \ \ \ \ \ \ \ \ \ \left(9.114\right)\\\mathbf{Z}^{2} & =\frac{1}{r^{2}}\mathbf{Z}_{2}\ \ \ \ \ \ \ \ \ \ \left(9.115\right)\\\mathbf{Z}^{3} & =\frac{1}{r^{2}\sin^{2}\theta}\mathbf{Z}_{3}.\ \ \ \ \ \ \ \ \ \ \left(9.116\right)\end{aligned}
In Section 3.1, we introduced the concept of orientation for a complete set of linearly independent vectors. This concept can be applied to the covariant basis Zi\mathbf{Z}_{i} and in this way extended to coordinate systems. Namely, the orientation of a coordinate system is identified with the orientation of the covariant basis: a coordinate system is said to be positively oriented or right-handed if the covariant basis is positively oriented, and negatively oriented or left-handed otherwise.
Naturally, the orientation of the coordinate system depends on the order of the coordinates. When the indicial notation is used to denote the coordinates ZiZ^{i}, the order is obvious. However, when special letters are used for special coordinates (such as rr, θ\theta, φ\varphi), the order must be explicitly specified. In all special coordinate systems introduced in this book, the order of the variables was chosen so that the coordinate system is positively oriented.
Generally speaking, the orientation is a local property defined at a point and may change from one region to another. When this occurs, the covariant "basis" vectors fail to be linearly independent on the boundary between the two regions. This signals a singularity that requires special care.
Finally, recall that two sets of vectors have the same orientation if they are related by a matrix with a positive determinant. Since the covariant and the contravariant bases are related by the metric tensor ZijZ_{ij}, which is positive definite matrix and therefore has a positive determinant, the two bases have the same orientation. Thus, we may refer to the orientation of the coordinate basis, without indicating its flavor.
Exercise 9.1Show that in affine coordinates, the basis vector Zi\mathbf{Z}_{i} connects the point with coordinates (Z1,Z2,Z3)\left( Z^{1},Z^{2},Z^{3}\right) with the point whose ii-th coordinate is increased by 11.
Exercise 9.2Show that in polar coordinates, the basis vector
Z1=R(r,θ)r,(9.117)\mathbf{Z}_{1}=\frac{\partial\mathbf{R}\left( r,\theta\right) }{\partial r},\tag{9.117}
is a unit vector that points in the outward radial direction. Similarly, show that
Z1=R(r,θ,φ)r(9.118)\mathbf{Z}_{1}=\frac{\partial\mathbf{R}\left( r,\theta,\varphi\right) }{\partial r}\tag{9.118}
in spherical coordinates is a unit vector that points in the outward radial direction.
Exercise 9.3Show that (the matrix corresponding to) the covariant metric tensor ZijZ_{ij} is positive definite. In fact, any matrix consisting of pairwise dot products of linearly independent vectors is positive definite.
Exercise 9.4Consider the identity
ZijZjk=δki(9.68)Z^{ij}Z_{jk}=\delta_{k}^{i} \tag{9.68}
and contract both sides with ZklZ^{kl}, i.e.
ZijZjkZkl=δkiZkl,(9.119)Z^{ij}Z_{jk}Z^{kl}=\delta_{k}^{i}Z^{kl},\tag{9.119}
or,
ZijZjkZkl=Zil.(9.120)Z^{ij}Z_{jk}Z^{kl}=Z^{il}.\tag{9.120}
Using this identity, demonstrate that the contravariant metric tensor is symmetric.
Exercise 9.5For the angle γ\gamma between Z1\mathbf{Z}_{1} and Z2\mathbf{Z}_{2}, show that
cosγ=Z12Z11Z22(9.121)\cos\gamma=\frac{Z_{12}}{\sqrt{Z_{11}Z_{22}}}\tag{9.121}
and therefore
sinγ=Z11Z22Z122Z11Z22.(9.122)\sin\gamma=\sqrt{\frac{Z_{11}Z_{22}-Z_{12}^{2}}{Z_{11}Z_{22}}}.\tag{9.122}
Note that the numerator in the fraction above corresponds to the determinant of the submatix
[Z11Z12Z21Z22](9.123)\left[ \begin{array} {cc} Z_{11} & Z_{12}\\ Z_{21} & Z_{22} \end{array} \right]\tag{9.123}
of the covariant metric tensor.
Exercise 9.6Show that in the nn-dimensional space,
ZijZij=n.(9.124)Z_{ij}Z^{ij}=n.\tag{9.124}
Exercise 9.7Show that inverting the metric tensor also works in the direction opposite of that described in Section 9.5.2, i.e.
Qi=ZijPj(9.71)Q_{i}=Z_{ij}P^{j} \tag{9.71}
implies
Pi=ZijQj.(9.70)P^{i}=Z^{ij}Q_{j}. \tag{9.70}
Exercise 9.8Show that inverting the metric tensor works for systems with indicial signatures of arbitrary complexity, e.g.
Pklirst=ZijQjklrst(9.125)P_{kl}^{irst}=Z^{ij}Q_{jkl}^{rst}\tag{9.125}
implies
Qiklrst=ZijPkljrst.(9.126)Q_{ikl}^{rst}=Z_{ij}P_{kl}^{jrst}.\tag{9.126}
Exercise 9.9Justify each step in the following chain of identities:
ZiZj=ZikZkZjlZl=ZikZjlZkl=Zikδkj=Zij.(9.92)\mathbf{Z}^{i}\cdot\mathbf{Z}^{j}=Z^{ik}\mathbf{Z}_{k}\cdot Z^{jl} \mathbf{Z}_{l}=Z^{ik}Z^{jl}Z_{kl}=Z^{ik}\delta_{k}^{j}=Z^{ij}. \tag{9.92}
Exercise 9.10Justify each step in the following chain of identities:
ZiZj=ZikZkZj=ZikZkj=δji.(9.94)\mathbf{Z}^{i}\cdot\mathbf{Z}_{j}=Z^{ik}\mathbf{Z}_{k}\cdot\mathbf{Z} _{j}=Z^{ik}Z_{kj}=\delta_{j}^{i}. \tag{9.94}
Exercise 9.11Show that the angle between Zi\mathbf{Z}^{i} and Zi\mathbf{Z}_{i} is less than π/2\pi/2 and that the product of the lengths of the two vectors is at least 11.
Exercise 9.12What are the components of Zi\mathbf{Z}^{i} with respect to the covariant basis Zj\mathbf{Z}_{j} and what are the components of Zi\mathbf{Z}_{i} with respect to the contravariant basis Zj\mathbf{Z}^{j}?
Problem 9.1Devise an algorithm for restoring the coordinate system from the metric tensor ZijZ_{ij}. As we discussed in Section 9.3.6, this can be done only to within a rigid transformation. Refer to Problem 2.3 as a suggestion for the first step in the algorithm.
Send feedback to Pavel