Riemannian Spaces

A Riemannian space is a coordinate space that is not connected to a Euclidean space or one whose connection to the Euclidean space from which it emerged is ignored. Riemannian spaces make it possible to use the ideas and the internally consistent analytical framework that we have developed without being beholden to a Euclidean geometric reality. The Riemannian framework frees us from adhering to the assumptions underpinning Euclidean geometry and, in fact, from our inability to exceed three dimensions. As such, Riemannian spaces are a powerful metaphorical extension of analytical Euclidean Geometry capable of describing realities that deviate from the Euclidean paradigm. This is a powerful generalization that is essential for us if we believe, as we have for over one hundred years, that the physical space of our existence is, in fact, non-Euclidean
At this point in our narrative, the concept of a Riemannian space seems all but inevitable. After all, with the exception of Chapter 18, we have been operating almost exclusively in the coordinate space for quite some time now, only occasionally checking in with the original Euclidean space. The issues that have preoccupied our attention, in particular the tensor property, required analyses that relied on the properties of the metric tensor ZijZ_{ij}, but not its definition. It seems natural, then, to ask what would happen if ZijZ_{ij} was arbitrarily selected rather than calculated by pairwise inner products of the covariant basis vectors Zi\mathbf{Z}_{i}? The short answer is that the overall framework remains intact and, in fact, becomes richer and more interesting.
We should note, however, that the apparent inevitability of Riemannian space is counter-historical. Indeed, it is the Riemannian ideas that led to the development of Tensor Calculus, the latter being first and foremost a tool for organizing those ideas into an analytical framework. Meanwhile, the fact that Riemannian spaces are being discussed in Chapter 20 is just another illustration that mathematical textbooks usually tell ideas in reverse-historical order.
Let us briefly recount our journey to Riemannian spaces. Our initial point of departure was the concept of a Euclidean space -- the geometric space of our everyday physical existence in which Euclid's axioms are found to have a reasonable degree of internal consistency and their consequences reasonably approximate reality as we perceive it. A Euclidean space is automatically endowed with the concepts of length and angles from which we construct the operation of the dot product of two geometric vectors U\mathbf{U} and V\mathbf{V}, i.e.
UV=lenUlenVcosγ.(2.14)\mathbf{U}\cdot\mathbf{V}=\operatorname{len}\mathbf{U}\operatorname{len} \mathbf{V}\cos\gamma. \tag{2.14}
The imposition of a coordinate system upon a Euclidean space gives rise to the coordinate space -- a region in R1 \mathbb{R} ^{1}, R2 \mathbb{R} ^{2}, or R3 \mathbb{R} ^{3} that represents the coordinates of the points in the Euclidean space and gives rise to a slew of analytical objects, such as the covariant and the contravariant bases Zi\mathbf{Z}_{i} and Zi\mathbf{Z}^{i}, the covariant and the contravariant metric tensors ZijZ_{ij} and ZijZ^{ij}, the volume element Z\sqrt{Z}, the Christoffel symbol Γjki\Gamma_{jk}^{i}, and the Levi-Civita symbols εijk\varepsilon_{ijk} and εijk\varepsilon^{ijk}. These objects are said to be induced from the Euclidean space. The fact that these objects arose in the context of a Euclidean space means that their values are not arbitrary, but rather subject to numerous constraints. For example, the metric tensor ZijZ_{ij} is symmetric, i.e.
Zij=Zji,(9.34)Z_{ij}=Z_{ji}, \tag{9.34}
and positive-definite. Consequently, the Christoffel symbol Γjki\Gamma_{jk}^{i} is symmetric in its subscripts, i.e.
Γjki=Γkji.(12.21)\Gamma_{jk}^{i}=\Gamma_{kj}^{i}. \tag{12.21}
And finally, the foremost of all constraints is the Riemann-Christoffel identity
ΓjmkZiΓimkZj+ΓinkΓjmnΓjnkΓimn=0,(15.127)\frac{\partial\Gamma_{jm}^{k}}{\partial Z^{i}}-\frac{\partial\Gamma_{im}^{k} }{\partial Z^{j}}+\Gamma_{in}^{k}\Gamma_{jm}^{n}-\Gamma_{jn}^{k}\Gamma _{im}^{n}=0, \tag{15.127}
derived at the end of Chapter 15.
Over the course of our narrative, we gradually transitioned from working with vectors in the Euclidean space to working with their components in the corresponding coordinate space. In particular, once all vectors are converted into their components, we are able to leave all vector quantities, including the bases Zi\mathbf{Z}_{i} and Zi\mathbf{Z}^{i}, completely out of the analysis and proceed only with objects that are available in the coordinate space. We have found that all analyses that can be performed geometrically in the Euclidean space can also be performed algebraically in the coordinate space. (In fact, the coordinate space offers far more robust analytical and computational tools than the original Euclidean space where we are constrained to use geometric methods, which are quite limited.) Thus, coordinate spaces have proven to be self-sufficient. However, since a coordinate space analysis exactly parallels what happens in the Euclidean space, its results are always consistent with Euclid's axioms and their consequences.
Our next important observation was the almost all-encompassing role played by the metric tensor ZijZ_{ij}. Thanks to the formula
UV=ZijUiVj,(10.25)\mathbf{U}\cdot\mathbf{V}=Z_{ij}U^{i}V^{j}, \tag{10.25}
the metric tensor ZijZ_{ij} is sufficient for the evaluation of dot products in terms of the components of vectors. This enables it to capture much of the geometry of the original Euclidean space. Furthermore, we discovered that the Christoffel symbol Γjki\Gamma_{jk}^{i} can be expressed in terms of the covariant metric tensor ZijZ_{ij} and its derivatives, i.e.
Γjki=12Zim(ZmjZk+ZmkZjZjkZm).(12.50)\Gamma_{jk}^{i}=\frac{1}{2}Z^{im}\left( \frac{\partial Z_{mj}}{\partial Z^{k}}+\frac{\partial Z_{mk}}{\partial Z^{j}}-\frac{\partial Z_{jk}}{\partial Z^{m}}\right) . \tag{12.50}
Thus, despite the fact that our original definition was vector-based, i.e.
ZiZj=ΓijkZk(12.20)\frac{\partial\mathbf{Z}_{i}}{\partial Z^{j}}=\Gamma_{ij}^{k}\mathbf{Z}_{k} \tag{12.20}
or, equivalently,
Γijk=ZkZiZj,(12.25)\Gamma_{ij}^{k}=\mathbf{Z}^{k}\cdot\frac{\partial\mathbf{Z}_{i}}{\partial Z^{j}}, \tag{12.25}
and thus included references to objects not available in the coordinate space, the metric tensor can serve as an alternative starting point for the Christoffel symbol. Therefore, all operations in the coordinate space, including the covariant derivative
kTji=TjiZkΓkjmTmi+ΓkmiTjm,(15.53)\nabla_{k}T_{j}^{i}=\frac{\partial T_{j}^{i}}{\partial Z^{k}}-\Gamma_{kj} ^{m}T_{m}^{i}+\Gamma_{km}^{i}T_{j}^{m}, \tag{15.53}
can be built up strictly from the covariant metric tensor ZijZ_{ij} and its derivatives. On a practical level, this means that the connection between the original Euclidean space and the coordinate space can be completely severed once the metric tensor ZijZ_{ij} has been calculated by the formula
Zij=ZiZj.(14.5)Z_{ij}=\mathbf{Z}_{i}\cdot\mathbf{Z}_{j}. \tag{14.5}
The only time that a reference to the original Euclidean space may be needed is to gain the geometric interpretation of the final results obtained by the coordinate space analysis.
In summary, in order for a coordinate space to be truly self-sufficient, all that is needed is the metric tensor field Zij(Z)Z_{ij}\left( Z\right) . This insight inevitably raises the intriguing possibility of starting out with a domain in Rn \mathbb{R} ^{n} and choosing Zij(Z)Z_{ij}\left( Z\right) arbitrarily (albeit subject to some desirable conditions, such as symmetry, positive definiteness, and spatial continuity) while preserving the overall analytical framework. In other words, the idea is to apply all of the machinery that we have developed so far to a "coordinate" space with a metric tensor field that is not necessarily induced from a Euclidean space.
Will this algebraic construct exhibit the same level of internal consistency as a Euclidean coordinate space or should we expect to run into insurmountable contradictions? On the one hand, it may be reasonable to hope for internal consistency since we are only changing the input while preserving the framework. On the other hand, the internal logical consistency of a Euclidean coordinate space is buttressed by the internal logical consistency of the Euclidean space itself. By assigning a metric tensor field that is not induced from a Euclidean space, we are no longer able to look to a Euclidean space for an absolute geometric interpretation of our results, a reference that has served as a reliable source of internal consistency.
Fortunately, even in the absence of a Euclidean space, the tensor framework remains. After all, the tensor property has to do with the transformation from one coordinate system to another and, therefore, the concept of a tensor is relative and does not require the absolute reference of a Euclidean space. The concept of a tensor survives virtually unchanged while the ultimate concept of an invariant preserves the meaning of having the same value in all coordinate systems but foregoes the greater implication of having a coordinate-free geometric interpretation. Nevertheless, this is sufficient for providing the new framework with internal logical consistency.
Meanwhile, it is clear that the results of a Riemannian analysis may deviate from the experiences of our everyday physical existence that is consistent with the axioms and the conclusions of Euclidean Geometry. For the most striking example, note there is no longer reason to expect that the Riemann-Christoffel identity
ΓjmkZiΓimkZj+ΓinkΓjmnΓjnkΓimn=0,(15.127)\frac{\partial\Gamma_{jm}^{k}}{\partial Z^{i}}-\frac{\partial\Gamma_{im}^{k} }{\partial Z^{j}}+\Gamma_{in}^{k}\Gamma_{jm}^{n}-\Gamma_{jn}^{k}\Gamma _{im}^{n}=0, \tag{15.127}
will continue to hold as it did in any Euclidean coordinate space. Recall that this identity is an expression of the "straightness" of a Euclidean space and the consequent admissibility of an affine coordinate system. Thus, violation of the above identity for an arbitrarily selected metric tensor field ZijZ_{ij} signals that it does not correspond to any Euclidean coordinate space.
This feature of Riemannian space is, of course, an unequivocal positive. Whether the physical space of our everyday existence is accurately described by Euclidean Geometry has been the subject of continuous skepticism for over two thousand years. According to some models, most notably Einstein's Theory of Relativity, our space violates some of the assumptions underpinning Euclidean Geometry and thus some of its conclusions. Riemannian spaces therefore hold the potential for supporting more general theories of space.
A Riemannian space is a strikingly minimalist algebraic construct. By definition, a Riemannian space is a domain in Rn \mathbb{R} ^{n} along with a metric tensor field Zij(Z)Z_{ij}\left( Z\right) . The integer nn is referred to as the dimension of the space and can have any positive value. We will treat ZijZ_{ij} as the covariant metric tensor, although the Riemannian version of the concept of a tensor, and therefore the meaning of the term covariant, are yet to be clarified. We will always require the metric tensor ZijZ_{ij} to be symmetric, i.e.
Zij=Zji.(20.1)Z_{ij}=Z_{ji}.\tag{20.1}
We will also typically require positive definiteness, although some applications -- including General Relativity -- require non-positive definite metric tensors. For our present purposes we will also assume that Zij(Z)Z_{ij}\left( Z\right) is sufficiently differentiable.
A number of crucial definitions from Euclidean coordinate spaces remain completely intact in the new Riemannian context. The contravariant metric tensor ZijZ^{ij} is defined as the matrix inverse of the covariant metric tensor ZijZ_{ij}, i.e.
ZijZjk=δki.(9.68)Z^{ij}Z_{jk}=\delta_{k}^{i}. \tag{9.68}
The practice of index juggling by contraction with the metric tensors ZijZ_{ij} and ZijZ^{ij} is completely unchanged. The Christoffel symbol Γi,jk\Gamma_{i,jk} is defined by the equation
Γi,jk=12(ZijZk+ZikZjZjkZi)(12.49)\Gamma_{i,jk}=\frac{1}{2}\left( \frac{\partial Z_{ij}}{\partial Z^{k}} +\frac{\partial Z_{ik}}{\partial Z^{j}}-\frac{\partial Z_{jk}}{\partial Z^{i} }\right) \tag{12.49}
while Γjki\Gamma_{jk}^{i} is given by
Γjki=12Zim(ZmjZk+ZmkZjZjkZm).(12.50)\Gamma_{jk}^{i}=\frac{1}{2}Z^{im}\left( \frac{\partial Z_{mj}}{\partial Z^{k}}+\frac{\partial Z_{mk}}{\partial Z^{j}}-\frac{\partial Z_{jk}}{\partial Z^{m}}\right) . \tag{12.50}
From this definition, known as the intrinsic definition, it follows that
ZijZk=Γi,jk+Γj,ik.(12.40)\frac{\partial Z_{ij}}{\partial Z^{k}}=\Gamma_{i,jk}+\Gamma_{j,ik}. \tag{12.40}
The definition of the covariant derivative k\nabla_{k} is completely unchanged, i.e.
kTji=TjiZkΓkjmTmi+ΓkmiTjm,(15.53)\nabla_{k}T_{j}^{i}=\frac{\partial T_{j}^{i}}{\partial Z^{k}}-\Gamma_{kj} ^{m}T_{m}^{i}+\Gamma_{km}^{i}T_{j}^{m}, \tag{15.53}
and all of its properties remain intact.
The Riemann-Christoffel tensor RmijkR_{\cdot mij}^{k}, given by
Rmijk=ΓjmkZiΓimkZj+ΓinkΓjmnΓjnkΓimn,(15.123)R_{\cdot mij}^{k}=\frac{\partial\Gamma_{jm}^{k}}{\partial Z^{i}} -\frac{\partial\Gamma_{im}^{k}}{\partial Z^{j}}+\Gamma_{in}^{k}\Gamma_{jm} ^{n}-\Gamma_{jn}^{k}\Gamma_{im}^{n}, \tag{15.123}
now takes on an even greater importance since it no longer vanishes. In fact, it serves as an important characterization of the Riemannian space and will occupy a central place in our analysis below.
The volume element is Z\sqrt{Z}, where ZZ is the determinant of the covariant metric tensor ZijZ_{ij}. Finally, the Levi-Civita symbols εijk\varepsilon_{ijk} and εijk\varepsilon^{ijk} are given by
εijk=Zeijk and          (17.3)εijk=eijkZ.          (17.4)\begin{aligned}\varepsilon_{ijk} & =\sqrt{Z}e_{ijk}\text{ and}\ \ \ \ \ \ \ \ \ \ \left(17.3\right)\\\varepsilon^{ijk} & =\frac{e^{ijk}}{\sqrt{Z}}. \ \ \ \ \ \ \ \ \ \ \left(17.4\right)\end{aligned}
Recall that the order of the Levi-Civita symbols matches the dimension of the space.
Let us now turn our attention to the differences between the ways in which Euclidean coordinate spaces and Riemannian spaces are constructed. We will find that while most of the identities are exactly the same, their interpretations are different: equations that once served as corollaries will now serve as definitions.
Let us begin by taking a second look at the symmetry requirement
Zij=Zji(20.1)Z_{ij}=Z_{ji} \tag{20.1}
for the covariant metric tensor ZijZ_{ij}. In a Euclidean coordinate space, the symmetry of ZijZ_{ij} is a corollary of its definition, i.e. Zij=ZiZjZ_{ij} =\mathbf{Z}_{i}\cdot\mathbf{Z}_{j}. In a Riemannian space, it is part of the definition itself. The same can be said of the positive definiteness of ZijZ_{ij}.
In a Riemannian space, a vector is any first-order system UiU^{i}. (Below, we will refine this definition by requiring UiU^{i} to be a tensor.) This definition clearly illustrates the loss of the absolute reference that we enjoyed in the Euclidean context. Recall that in a Euclidean coordinate space, a tensor UiU^{i} can be converted into the corresponding invariant geometric vector U\mathbf{U} by the contraction UiZiU^{i}\mathbf{Z}_{i}. This gives UiU^{i} an instant absolute interpretation independent of the coordinate system. Such an interpretation is no longer available. In a Riemannian space, vectors behave in the same way as components of geometric vectors behave in a Euclidean coordinate space. However, beyond this algebraic similitude, the two concepts are distinct and it is essential to accept Riemannian vectors on their own terms, as it was essential to accept Euclidean vectors on their own terms in Chapter 2.
For the concepts of length, angle, and dot product, the logic of Euclidean coordinate spaces is completely reversed, `{a} la the Linear Algebra approach described in Section 2.7. For two vectors UiU^{i} and ViV^{i}, the dot product is defined to be the combination
ZijUiVj.(20.2)Z_{ij}U^{i}V^{j}.\tag{20.2}
Thus, this familiar contraction has changed its role from a corollary to a definition. Although it is called the dot product, there is no symbol for this operation in a Riemannian space that includes a dot. It is easy to show that the above definition satisfies the requisite properties of an inner product in the sense of Linear Algebra.
The length of a vector UiU^{i} is the square root of the dot product with itself, i.e.
ZijUiUj.(20.3)\sqrt{Z_{ij}U^{i}U^{j}}.\tag{20.3}
The angle γ\gamma between two vectors UiU^{i} and ViV^{i} is given by the equation
cosγ=ZijUiVjZijUiUjZijViVj.(20.4)\cos\gamma=\frac{Z_{ij}U^{i}V^{j}}{\sqrt{Z_{ij}U^{i}U^{j}}\sqrt{Z_{ij} V^{i}V^{j}}}.\tag{20.4}
(Note that, technically speaking, each dot product above should use its own set of dummy index names.) The fact that the absolute value of the quantity on the right is less than or equal to 11 is an immediate consequence of the Cauchy-Schwarz inequality. Two vectors UiU^{i} and ViV^{i} are said to be orthogonal if their dot product vanishes, i.e.
ZijUiVj=0.(20.5)Z_{ij}U^{i}V^{j}=0.\tag{20.5}
An ordered set of nn linearly independent vectors is said to be positively oriented, if the determinant of the matrix consisting of the nn vectors is positive.
With the help of index juggling, which, as we have already mentioned, remains completely intact, the above definitions can be expressed in a more concise form. The dot product of UiU^{i} and ViV^{i} is given by
UiVi,(20.6)U_{i}V^{i},\tag{20.6}
the length of a vector UiU^{i} is
UiUi,(20.7)\sqrt{U_{i}U^{i}},\tag{20.7}
and the angle γ\gamma between two vectors UiU^{i} and ViV^{i} satisfies the equation
cosγ=UiViUiUiViVi.(20.8)\cos\gamma=\frac{U_{i}V^{i}}{\sqrt{U_{i}U^{i}}\sqrt{V_{i}V^{i}}}.\tag{20.8}
(Once again, each dot product should have used its own set of dummy index names.) Vectors UiU^{i} and ViV^{i} are orthogonal if
UiVi=0.(20.9)U_{i}V^{i}=0.\tag{20.9}
A set of nn functions Zi(γ)Z^{i}\left( \gamma\right) of a single parameter γ\gamma is referred to as a curve. The integral
L=γ1γ2ZijdZidγdZjdγdγ(20.10)L=\int_{\gamma_{1}}^{\gamma_{2}}\sqrt{Z_{ij}\frac{dZ^{i}}{d\gamma}\frac {dZ^{j}}{d\gamma}}d\gamma\tag{20.10}
or, more explicitly,
L=γ1γ2Zij(Z(γ))dZi(γ)dγdZj(γ)dγdγ,(20.11)L=\int_{\gamma_{1}}^{\gamma_{2}}\sqrt{Z_{ij}\left( Z\left( \gamma\right) \right) \frac{dZ^{i}\left( \gamma\right) }{d\gamma}\frac{dZ^{j}\left( \gamma\right) }{d\gamma}}d\gamma,\tag{20.11}
is, by definition, the length of the curve over the interval from γ1\gamma_{1} to γ2\gamma_{2}.

20.3.1General symmetries

Let us return to the Riemann-Christoffel tensor
Rmijk=ΓjmkZiΓimkZj+ΓinkΓjmnΓjnkΓimn.(15.123)R_{\cdot mij}^{k}=\frac{\partial\Gamma_{jm}^{k}}{\partial Z^{i}} -\frac{\partial\Gamma_{im}^{k}}{\partial Z^{j}}+\Gamma_{in}^{k}\Gamma_{jm} ^{n}-\Gamma_{jn}^{k}\Gamma_{im}^{n}. \tag{15.123}
It was initially introduced in Chapter 15 in the context of Euclidean coordinate spaces. We were therefore able to conclude that it identically vanishes in all coordinate systems. Now that we are no longer in a Euclidean coordinate space, the Riemann-Christoffel tensor does not vanish and therefore takes on an even greater significance. Recall that the original purpose of the Riemann-Christoffel tensor was to express the commutator ijji\nabla_{i}\nabla_{j}-\nabla_{j}\nabla_{i} in terms of the Christoffel symbols and we discovered that
ijTkjiTk=RmijkTm.(15.124)\nabla_{i}\nabla_{j}T^{k}-\nabla_{j}\nabla_{i}T^{k}=R_{\cdot mij}^{k}T^{m}. \tag{15.124}
This remarkable formula was first given by Gregorio Ricci and Tullio Levi-Civita in their 1901 seminal paper titled M\'{ethodes de calcul differ'{e}ntiel absolu et leurs applications}.
In this Section, we will discuss the most fundamental properties of the Riemann-Christoffel tensor. Most of the details of the demonstrations will be left as exercises.
Let us focus on the version of the Riemann-Christoffel tensor with a lowered first index, i.e.
Rlmij=Zlk(ΓjmkZiΓimkZj+ΓinkΓjmnΓjnkΓimn).(20.12)R_{lmij}=Z_{lk}\left( \frac{\partial\Gamma_{jm}^{k}}{\partial Z^{i}} -\frac{\partial\Gamma_{im}^{k}}{\partial Z^{j}}+\Gamma_{in}^{k}\Gamma_{jm} ^{n}-\Gamma_{jn}^{k}\Gamma_{im}^{n}\right) .\tag{20.12}
By absorbing the metric tensor into the Christoffel symbol, we find that
Rlmij=Γl,jmZiΓl,imZj+Γn,jlΓimnΓn,ilΓjmn.(20.13)R_{lmij}=\frac{\partial\Gamma_{l,jm}}{\partial Z^{i}}-\frac{\partial \Gamma_{l,im}}{\partial Z^{j}}+\Gamma_{n,jl}\Gamma_{im}^{n}-\Gamma _{n,il}\Gamma_{jm}^{n}.\tag{20.13}
The Riemann-Christoffel tensor RlmijR_{lmij} has three fundamental symmetries. The first one is the skew-symmetry in the last two indices, i.e.
Rlmij=Rlmji,(20.14)R_{lmij}=-R_{lmji},\tag{20.14}
which is nearly self-evident from the previous equation. It is not self-evident, but nevertheless true, that RlmijR_{lmij} is also skew-symmetric in the first two indices, i.e.
Rlmij=Rmlij,(20.15)R_{lmij}=-R_{mlij},\tag{20.15}
and symmetric with respect to switching the first two with the last two indices
Rlmij=Rijlm.(20.16)R_{lmij}=R_{ijlm}.\tag{20.16}
Furthermore, RlmijR_{lmij} has an additional symmetry captured by the first Bianchi identity
Rlmij+Rlijm+Rljmi=0.(20.17)R_{lmij}+R_{lijm}+R_{ljmi}=0.\tag{20.17}
As a result of these symmetries, the number of the degrees of freedom in the Riemann-Christoffel tensor equals n2(n21)/12n^{2}\left( n^{2}-1\right) /12. In particular, in a two-dimensional Riemannian space, the Riemann-Christoffel tensor has a single degree of freedom, a fact that we will take full advantage of below and that will lead to the concept of Gaussian curvature. The proofs of the above symmetries as well as the statement regarding the degrees of freedom are left as exercises.

20.3.2The Ricci curvature tensor

The Ricci curvature tensor is the second-order tensor RmjR_{mj} obtained from the Riemann-Christoffel tensor by contracting the first and the third indices, i.e.
Rij=Rikjk.(20.18)R_{ij}=R_{\cdot ikj}^{k}.\tag{20.18}
It is easy to show that the Ricci curvature tensor is symmetric, i.e.
Rij=Rji(20.19)R_{ij}=R_{ji}\tag{20.19}
Its trace RR, i.e.
R=Rmm,(20.20)R=R_{\cdot m}^{m},\tag{20.20}
is known as the scalar curvature.

20.3.3The commutator ijji\nabla_{i}\nabla_{j}-\nabla_{j}\nabla_{i} revisited

Recall that the Riemann-Christoffel tensor originally emerged during the analysis of the commutator
ijji(20.21)\nabla_{i}\nabla_{j}-\nabla_{j}\nabla_{i}\tag{20.21}
in Chapter 15, where we established the following identity for a first-order variant TkT^{k}:
ijTkjiTk=RmijkTm.(15.124)\nabla_{i}\nabla_{j}T^{k}-\nabla_{j}\nabla_{i}T^{k}=R_{\cdot mij}^{k}T^{m}. \tag{15.124}
This identity can be extended to variants of arbitrary order. In the general identity, there is a term on the right for each index of the variant. Thus, as we always do in situations like this, we will give the formula for a variant TjiT_{j}^{i} with a representative collection of indices. The identity reads
ijTlkjiTlk=RmijkTlmRlijmTmk.(20.22)\nabla_{i}\nabla_{j}T_{l}^{k}-\nabla_{j}\nabla_{i}T_{l}^{k}=R_{\cdot mij} ^{k}T_{l}^{m}-R_{\cdot lij}^{m}T_{m}^{k}.\tag{20.22}
Thus, from the point of view of indicial logistics, the Riemann-Christoffel symbol figures on the right are in a way similar to the Christoffel symbol in the definition of the covariant derivative, where the sign of the term and the interplay among indices depends on the flavor of the index.

20.3.4The Gaussian Curvature

The concept of the Gaussian curvature applies in two-dimensional Riemannian spaces. The skew-symmetric properties of the Riemann-Christoffel tensor RlmijR_{lmij}, i.e.
Rlmij=Rlmji          (20.14)Rlmij=Rmlij,          (20.15)\begin{aligned}R_{lmij} & =-R_{lmji}\ \ \ \ \ \ \ \ \ \ \left(20.14\right)\\R_{lmij} & =-R_{mlij}, \ \ \ \ \ \ \ \ \ \ \left(20.15\right)\end{aligned}
reduce it to a single degree of freedom. Indeed, there are only four nonzero elements, i.e.
R1212,  R2121,  R2112,  and  R1221,(20.23)R_{1212}\text{, \ }R_{2121}\text{, \ }R_{2112}\text{, \ and \ }R_{1221},\tag{20.23}
where
R1212=R2121=α and R2112=R1221=α.(20.24)R_{1212}=R_{2121}=\alpha\text{ and }R_{2112}=R_{1221}=-\alpha.\tag{20.24}
Therefore, RlmijR_{lmij} can be captured by the identity
Rlmij=αelmeij,(20.25)R_{lmij}=\alpha e_{lm}e_{ij},\tag{20.25}
where elme_{lm} and eije_{ij} are the permutation systems described in Chapter 16. (Note that we encountered similar identities rooted in the skew-symmetric property in Section 16.10.) The permutation systems elme_{lm} and eije_{ij} can be easily converted into the Levi-Civita symbols
εlm=Zelm  and  εij=Zeij.(17.3)\varepsilon_{lm}=\sqrt{Z}e_{lm}\text{ \ and \ }\varepsilon_{ij}=\sqrt{Z} e_{ij}. \tag{17.3}
Thus, we have
Rlmij=Kεlmεij,(20.26)R_{lmij}=K\varepsilon_{lm}\varepsilon_{ij},\tag{20.26}
where K=α/ZK=\alpha/Z. The quantity KK is known as the Gaussian curvature and is, perhaps, the single most important pointwise characteristic of a two-dimensional Riemannian space.
Multiplying both sides of the equation
Rlmij=Kεlmεij(20.26)R_{lmij}=K\varepsilon_{lm}\varepsilon_{ij} \tag{20.26}
by εlmεij\varepsilon^{lm}\varepsilon^{ij} yields an explicit expression for the Gaussian curvature KK, i.e.
K=14εlmεijRlmij.(20.27)K=\frac{1}{4}\varepsilon^{lm}\varepsilon^{ij}R_{lmij}.\tag{20.27}
It also proves that KK is an invariant.
The Ricci curvature tensor, defined by the identity
Rij=Rikjk,(20.18)R_{ij}=R_{\cdot ikj}^{k}, \tag{20.18}
is given by
Rij=KZij.(20.28)R_{ij}=KZ_{ij}.\tag{20.28}
Therefore, the Gaussian curvature equals half the scalar curvature, i.e.
K=12Rii.(20.29)K=\frac{1}{2}R_{i}^{i}.\tag{20.29}
Proving these identities is left as an exercise.
First, consider the Riemannian space that occupies all of R3 \mathbb{R} ^{3} with the metric tensor ZijZ_{ij} that corresponds to the matrix
[111].(20.30)\left[ \begin{array} {ccc} 1 & & \\ & 1 & \\ & & 1 \end{array} \right] .\tag{20.30}
This space obviously corresponds to a three-dimensional Euclidean space referred to Cartesian coordinates. Such Riemannian spaces are said to be induced from a Euclidean space. Since we have encountered this space before, there is no need to recount the values of ZijZ^{ij}, Γjki\Gamma_{jk}^{i}, and other objects.
A related example of a Riemannian space is R4 \mathbb{R} ^{4} with ZijZ_{ij} corresponding to
[1111].(20.31)\left[ \begin{array} {cccc} 1 & & & \\ & 1 & & \\ & & 1 & \\ & & & 1 \end{array} \right] .\tag{20.31}
On the one hand, it is abundantly clear that for a wide range of intents and purposes this Riemannian space is analogous to the one before it. On the other hand, by virtue of having a dimension greater than 33, this Riemannian space does not correspond to any Euclidean coordinate space. This example illustrates one aspect of the generality of Riemannian spaces. Furthermore, this Riemannian space offers us an opportunity to construct an algebraic generalization of Euclidean spaces more closely resembling the geometric space. We will exploit this idea below in Section 20.7. We must acknowledge in advance that this is precisely the kind of association between Euclidean spaces and regular Cartesian grids that we discouraged so strenuously in the beginning of the book. However, such an association is perilous when it is the exclusive interpretation or even the default one and not when it is just one among many.
For a second example, consider the space R3 \mathbb{R} ^{3} with ZijZ_{ij} corresponding to
[1(Z1)21].(20.32)\left[ \begin{array} {ccc} 1 & & \\ & \left( Z^{1}\right) ^{2} & \\ & & 1 \end{array} \right] .\tag{20.32}
Similarly to the previous example, this Riemannian space is related to cylindrical coordinates imposed upon a three-dimensional Euclidean space. We are already familiar with the values of the fundamental elements in this space and will therefore not list them here. Note however, that this space has a singularity at all points corresponding to Z1=0Z^{1}=0 since at those points, the above matrix is not invertible and therefore the contravariant metric tensor ZijZ^{ij} does not exist.
For a third example, consider R2 \mathbb{R} ^{2} with ZijZ_{ij} corresponding to
 [R2R2sin2(Z1)],(20.33)\text{ }\left[ \begin{array} {cc} R^{2} & \\ & R^{2}\sin^{2}\left( Z^{1}\right) \end{array} \right] ,\tag{20.33}
where RR is a constant. Although the general form of ZijZ_{ij} is reminiscent of spherical coordinates, we have not encountered this space before and will therefore summarize the values of the fundamental object. The details of the calculations are left as an exercise.
The volume element Z\sqrt{Z} is given by
Z=R2sin(Z1),(20.34)\sqrt{Z}=R^{2}\sin\left( Z^{1}\right) ,\tag{20.34}
while the contravariant metric tensor ZijZ^{ij} corresponds to
 [R2R2sin2(Z1)].(20.35)\text{ }\left[ \begin{array} {cc} R^{-2} & \\ & R^{-2}\sin^{-2}\left( Z^{1}\right) \end{array} \right] .\tag{20.35}
The nonzero elements of the Christoffel symbol Γjki\Gamma_{jk}^{i} are
Γ221=cos(Z1)sin(Z1)     and    Γ122=Γ212=cotZ1.(20.36)\Gamma_{22}^{1}=-\cos\left( Z^{1}\right) \sin\left( Z^{1}\right) \text{\ \ \ \ \ and\ \ \ \ }\Gamma_{12}^{2}=\Gamma_{21}^{2}=\cot Z^{1}.\tag{20.36}
Finally, the Riemann-Christoffel tensor RkmijR_{kmij} is given by
Rkmij=1R2εkmεij.(20.37)R_{kmij}=\frac{1}{R^{2}}\varepsilon_{km}\varepsilon_{ij}.\tag{20.37}
Thus, the Gaussian curvature KK is given by
K=1R2.(20.38)K=\frac{1}{R^{2}}.\tag{20.38}
The fact that the Riemann-Christoffel tensor RkmijR_{kmij} does not vanish indicates that this Riemannian space does not correspond to any Euclidean coordinate space. Interestingly, it does correspond to a two-dimensional curved surface embedded in a three-dimensional Euclidean space. Of course, embedded surfaces are not discussed in this book and we are therefore not in a position to give a detailed description. However, we ought to mention the most basic facts because it is a very common way in which non-Euclidean Riemannian spaces can arise.
Consider a sphere of radius RR in a three-dimensional Euclidean space. Much like the points in the surrounding space, the points on the sphere can be enumerated by a pair of numbers. In other words, the sphere can be referred to a coordinate system. We will denote the coordinates by S1S^{1} and S2S^{2} or, collectively, SαS^{\alpha} where a different alphabet is used for indices because the dimension of the sphere is different from the dimension of the surrounding space. The surface covariant basis Sα\mathbf{S}_{\alpha} is defined by the equation
Sα=R(S)Sα,(20.39)\mathbf{S}_{\alpha}=\frac{\partial\mathbf{R}\left( S\right) }{\partial S^{\alpha}},\tag{20.39}
and the rest of the framework is constructed by following the Euclidean blueprint. For example, the covariant metric tensor SαβS_{\alpha\beta} is defined by
Sαβ=SαSβ,(20.40)S_{\alpha\beta}=\mathbf{S}_{\alpha}\cdot\mathbf{S}_{\beta},\tag{20.40}
and so on.
Now, when the sphere is referred to spherical coordinates S1,S2=θ,φS^{1},S^{2} =\theta,\varphi illustrated in the figure
(20.41)
the resulting metric tensor SαβS_{\alpha\beta} corresponds to
[R2R2sin2θ],(20.42)\left[ \begin{array} {cc} R^{2} & \\ & R^{2}\sin^{2}\theta \end{array} \right] ,\tag{20.42}
which is precisely the metric tensor in the present example. Therefore, this Riemannian space may be said to be induced from a surface embedded in a Euclidean space. This example shows that some surfaces embedded in Euclidean spaces are themselves examples of non-Euclidean spaces. Note that the Gaussian curvature for embedded surfaces has a beautiful geometric interpretation that will be discussed in a future book.
Before we can discuss the concept of a tensor, we must first talk about coordinate changes. Heretofore, our analysis of coordinate changes has been based on the equations of coordinate transformation
Zi=Zi(Z)          (13.21)Zi=Zi(Z)          (13.22)\begin{aligned}Z^{i^{\prime}} & =Z^{i^{\prime}}\left( Z\right) \ \ \ \ \ \ \ \ \ \ \left(13.21\right)\\Z^{i} & =Z^{i}\left( Z^{\prime}\right) \ \ \ \ \ \ \ \ \ \ \left(13.22\right)\end{aligned}
that related two alternative coordinate systems: the "original" unprimed coordinates ZiZ^{i} and the "new" primed coordinates ZiZ^{i^{\prime}}. However, it is important to understand the crucial role that the Euclidean space played in establishing these equations. Given two alternative coordinate systems ZiZ^{i} and ZiZ^{i^{\prime}}, consider a specific set of unprimed coordinates, such as (Z1,Z2,Z3)=(2,1,5)\left( Z^{1},Z^{2},Z^{3}\right) =\left( 2,1,5\right) . How does one determine the values of the primed coordinates ZiZ^{i^{\prime} } that correspond to the above values of the unprimed coordinates? This is done in two steps. The first step is to identify the physical point PP in the Euclidean space corresponding to unprimed coordinates (2,1,5)\left( 2,1,5\right) . The second step is to observe the primed coordinates corresponding to the same point PP, e.g. (Z1,Z2,Z3)=(3,2,8)\left( Z^{1^{\prime}},Z^{2^{\prime}},Z^{3^{\prime} }\right) =\left( 3,-2,8\right) . By this mechanism, the function Zi(Z)Z^{i^{\prime}}\left( Z\right) maps (2,1,5)\left( 2,1,5\right) to (3,2,8)\left( 3,-2,8\right) , while Zi(Z)Z^{i}\left( Z^{\prime}\right) maps (3,2,8)\left( 3,-2,8\right) back to (2,1,5)\left( 2,1,5\right) .
With the Euclidean space no longer in the picture, we will take the equation
Zi=Zi(Z)(20.43)Z^{i^{\prime}}=Z^{i^{\prime}}\left( Z\right)\tag{20.43}
as the definition, rather than a description, of the coordinate change. The range of the function Zi(Z)Z^{i^{\prime}}\left( Z\right) defines the domain of the "new" primed Riemannian space with coordinates ZiZ^{i^{\prime}}. The inverse mapping function is denoted by Zi(Z)Z^{i}\left( Z^{\prime}\right) , i.e.
Zi=Zi(Z).(20.44)Z^{i}=Z^{i}\left( Z^{\prime}\right) .\tag{20.44}
Naturally, the definitions of the Jacobians JiiJ_{i}^{i^{\prime}} and JiiJ_{i^{\prime}}^{i} remain the same as before:
Jii=Zi(Z)Zi          (13.32)Jii=Zi(Z)Zi.          (13.31)\begin{aligned}J_{i^{\prime}}^{i} & =\frac{\partial Z^{i}\left( Z^{\prime}\right) }{\partial Z^{i^{\prime}}}\ \ \ \ \ \ \ \ \ \ \left(13.32\right)\\J_{i}^{i^{\prime}} & =\frac{\partial Z^{i^{\prime}}\left( Z\right) }{\partial Z^{i}}. \ \ \ \ \ \ \ \ \ \ \left(13.31\right)\end{aligned}
The definition of a tensor under such changes of coordinates will remain the same as before, i.e. TjiT_{j}^{i} is a tensor with a representative collection of indices if TjiT_{j^{\prime}}^{i^{\prime}} and TjiT_{j}^{i} are related by the identity
Tji=TjiJiiJjj.(14.1)T_{j^{\prime}}^{i^{\prime}}=T_{j}^{i}J_{i}^{i^{\prime}}J_{j^{\prime}}^{j}. \tag{14.1}
However, the application of this condition to some objects is not the same as before as the following example will illustrate.
Consider a Cartesian Riemannian space, i.e. Rn \mathbb{R} ^{n} with the metric tensor field ZijZ_{ij} that corresponds to the matrix
[111].(20.45)\left[ \begin{array} {ccc} 1 & & \\ & 1 & \\ & & 1 \end{array} \right] .\tag{20.45}
Consider the following equations of coordinate transformation
Z1(Z1,Z2,Z3)=(Z1)2+(Z2)2          (20.46)Z2(Z1,Z2,Z3)=arctan(Z1,Z2)          (20.47)Z3(Z1,Z2,Z3)=Z3.          (20.48)\begin{aligned}Z^{1^{\prime}}\left( Z^{1},Z^{2},Z^{3}\right) & =\sqrt{\left( Z^{1}\right) ^{2}+\left( Z^{2}\right) ^{2}}\ \ \ \ \ \ \ \ \ \ \left(20.46\right)\\Z^{2^{\prime}}\left( Z^{1},Z^{2},Z^{3}\right) & =\arctan\left( Z^{1},Z^{2}\right)\ \ \ \ \ \ \ \ \ \ \left(20.47\right)\\Z^{3^{\prime}}\left( Z^{1},Z^{2},Z^{3}\right) & =Z^{3}.\ \ \ \ \ \ \ \ \ \ \left(20.48\right)\end{aligned}
Of course, we recognize these equations as a transformation from Cartesian coordinates to cylindrical. For example, the point (1,1,3)\left( 1,1,3\right) maps to the point (2,π/4,3)\left( \sqrt{2},\pi/4,3\right) . Thus, the primed Riemannian space is associated with the domain [0,)×[0,2π)×(,)\left[ 0,\infty\right) \times\left[ 0,2\pi\right) \times\left( -\infty,\infty\right) . The inverse equations of coordinate transformation are, of course,
Z1(Z1,Z2,Z3)=Z1cosZ2          (20.49)Z2(Z1,Z2,Z3)=Z1sinZ2          (20.50)Z3(Z1,Z2,Z3)=Z3.          (20.51)\begin{aligned}Z^{1}\left( Z^{1^{\prime}},Z^{2^{\prime}},Z^{3^{\prime}}\right) & =Z^{1^{\prime}}\cos Z^{2^{\prime}}\ \ \ \ \ \ \ \ \ \ \left(20.49\right)\\Z^{2}\left( Z^{1^{\prime}},Z^{2^{\prime}},Z^{3^{\prime}}\right) & =Z^{1^{\prime}}\sin Z^{2^{\prime}}\ \ \ \ \ \ \ \ \ \ \left(20.50\right)\\Z^{3}\left( Z^{1^{\prime}},Z^{2^{\prime}},Z^{3^{\prime}}\right) & =Z^{3^{\prime}}.\ \ \ \ \ \ \ \ \ \ \left(20.51\right)\end{aligned}
The primed Riemannian space is not yet complete since we have not specified the metric tensor field Zij(Z)Z_{i^{\prime}j^{\prime}}\left( Z^{\prime}\right) . Since our goal is to preserve the tensor framework, Zij(Z)Z_{i^{\prime}j^{\prime} }\left( Z^{\prime}\right) cannot be assigned arbitrarily, nor can it be assigned by mapping the unprimed metric tensor ZijZ_{ij} to the primed space (which, in this example, would mean that ZijZ_{i^{\prime}j^{\prime}} also corresponds to the 3×33\times3 identity matrix). The only logical thing to do is to construct the metric tensor ZijZ_{i^{\prime}j^{\prime}} in the primed coordinate space according to the equation
Zij=ZijJiiJjj.(20.52)Z_{i^{\prime}j^{\prime}}=Z_{ij}J_{i^{\prime}}^{i}J_{j^{\prime}}^{j}.\tag{20.52}
Notice that this equation was not given the number (13.59) of the same equation in Chapter 13 to highlight the fact that it now serves as a definition rather than a corollary. Also note that in this approach, the coordinates ZiZ^{i} are truly primary while all other coordinate systems are secondary. This loss of parity among coordinate systems is caused by the absence of an independent Euclidean space acting as an a priori absolute reference.
To calculate the elements of ZijZ_{i^{\prime}j^{\prime}}, note that
Jii corresponds to J=[cosZ2Z1sinZ20sinZ1Z1cosZ20001].(20.53)J_{i^{\prime}}^{i}\text{ corresponds to }J=\left[ \begin{array} {rrr} \cos Z^{2^{\prime}} & -Z^{1^{\prime}}\sin Z^{2^{\prime}} & 0\\ \sin Z^{1^{\prime}} & Z^{1^{\prime}}\cos Z^{2^{\prime}} & 0\\ 0 & 0 & 1 \end{array} \right] .\tag{20.53}
Therefore, ZijZ_{i^{\prime}j^{\prime}} corresponds to matrix product
[cosZ2Z1sinZ20sinZ2Z1cosZ20001]T[111][cosZ2Z1sinZ20sinZ2Z1cosZ20001],(20.54)\small \left[ \begin{array} {rrr} \cos Z^{2^{\prime}} & -Z^{1^{\prime}}\sin Z^{2^{\prime}} & 0\\ \sin Z^{2^{\prime}} & Z^{1^{\prime}}\cos Z^{2^{\prime}} & 0\\ 0 & 0 & 1 \end{array} \right] ^{T}\left[ \begin{array} {ccc} 1 & & \\ & 1 & \\ & & 1 \end{array} \right] \left[ \begin{array} {rrr} \cos Z^{2^{\prime}} & -Z^{1^{\prime}}\sin Z^{2^{\prime}} & 0\\ \sin Z^{2^{\prime}} & Z^{1^{\prime}}\cos Z^{2^{\prime}} & 0\\ 0 & 0 & 1 \end{array} \right] \normalsize ,\tag{20.54}
which evaluates to the familiar matrix
[1(Z1)21].(20.55)\left[ \begin{array} {ccc} 1 & & \\ & \left( Z^{1^{\prime}}\right) ^{2} & \\ & & 1 \end{array} \right] .\tag{20.55}
In summary, we define the metric tensor ZijZ_{i^{\prime}j^{\prime}} so as to make ZijZ_{ij} transform according to the tensor rule. Similarly, we will augment the notion of a vector UiU^{i} by stipulating that it transforms according to the tensor rule
Ui=UiJii.(20.56)U^{i^{\prime}}=U^{i}J_{i}^{i^{\prime}}.\tag{20.56}
In other words, a Riemannian vector is, by definition, a first-order tensor. In general, for every object defined arbitrarily in the unprimed Riemannian space, we must specify the precise rule by which the object transforms under a change of coordinates to the primed space.
For all other objects, the question of transformation under a change of coordinates can be subjected to the same analysis as before and will reach the same conclusions. This applies to such objects as the contravariant metric tensor ZijZ^{ij}, the volume element Z\sqrt{Z}, the Christoffel symbol Γjki\Gamma_{jk}^{i}, the Riemann-Christoffel tensor RmijkR_{\cdot mij}^{k}, and the Levi-Civita symbols εijk\varepsilon_{ijk} and εijk\varepsilon^{ijk}. Once the metric tensor ZijZ_{i^{\prime}j^{\prime}} is constructed in the primed Riemannian space, the objects Z\sqrt{Z^{\prime}}, Γjki\Gamma_{j^{\prime }k^{\prime}}^{i^{\prime}}, RmijkR_{\cdot m^{\prime}i^{\prime}j^{\prime} }^{k^{\prime}}, εijk\varepsilon_{i^{\prime}j^{\prime}k^{\prime}}, and εijk\varepsilon^{i^{\prime}j^{\prime}k^{\prime}} can be constructed from it. Thus, we can study the rules by which they transform between the two spaces and, because they have the same definitions as they did in Euclidean coordinate spaces, we will, of course, discover that they transform by the exact same rules.
Finally, it must be noted that the concept of a tensor is stronger and more absolute in a Euclidean space than in a Riemannian space. In a Euclidean space, an invariant constructed from tensors represents an object with a clear geometric meaning. Thus, the tensor framework connects the coordinate space to an independent physical reality which serves as a guarantor of the framework's meaningfulness. In a Riemannian space, there is no such independent physical reality that can be used as an absolute reference. Thus, the tensor framework serves as a tool for transferring the principles developed in Euclidean spaces to a new algebraic setting. It brings with it a certain degree in internal consistency. For example, the tensor property continues to be reflexive, symmetric, and transitive as described in Section 14.12. On balance, however, the tensor framework in the context of a Riemannian space should be seen as a metaphorical extension of Euclidean ideas.
A Riemannian space is, first and foremost, an algebraic structure. In a way, its very purpose is to provide a framework where analysis can be performed without a need for geometric intuition. It is somewhat ironic, then, that, being a subset of Rn \mathbb{R} ^{n}, a Riemannian space can be visualized as a Cartesian coordinate grid in a Euclidean space. And not only that, this geometric view of a Riemannian space is actually beneficial.
For example, consider the Riemannian space induced from a Euclidean plane referred to polar coordinates r,θr,\theta.
(20.57)
In the spirit of Riemannian spaces, we will refer to the coordinates as Z1Z^{1} and Z2Z^{2} instead of rr and θ\theta. The induced Riemannian space occupies the domain [0,)×[0,2π)\left[ 0,\infty\right) \times\left[ 0,2\pi\right) . The metric tensor Zij(Z)Z_{ij}\left( Z\right) corresponds to the matrix
[100Z12].(20.58)\left[ \begin{array} {cc} 1 & 0\\ 0 & Z_{1}^{2} \end{array} \right] .\tag{20.58}
When asked to visualize the domain [0,)×[0,2π)\left[ 0,\infty\right) \times\left[ 0,2\pi\right) , all of us, without exception, visualize a semi-infinite rectangular strip of height 2π2\pi. Furthermore, the points corresponding to integer (or any regularly spaced) values of Z1Z^{1} and Z2Z^{2} form a regular grid. The result is nothing but a Euclidean space referred to Cartesian coordinates. We will refer to this space as the arithmetic space associated with the Riemannian space. Of course, any Riemannian space, not just induced ones, may be visualized as an arithmetic space.
(20.59)
The metric tensor, an essential part of the Riemannian space, is not reflected in this picture and thus remains behind the scenes. Surprisingly, the geometry that takes place in the arithmetic space will play an important role in many of our future analyses. For example, the divergence theorem will be proven by first demonstrating it in the arithmetic space.
We must point out the obvious that for induced Riemannian spaces, the arithmetic geometry may look quite different from the original geometry from which the Riemannian space had arisen. For example, consider a curve in a Euclidean plane given by the following equations in polar coordinates
Z1(γ)=γ          (20.60)Z2(γ)=γ.          (20.61)\begin{aligned}Z^{1}\left( \gamma\right) & =\gamma\ \ \ \ \ \ \ \ \ \ \left(20.60\right)\\Z^{2}\left( \gamma\right) & =\gamma.\ \ \ \ \ \ \ \ \ \ \left(20.61\right)\end{aligned}
Of course, we recognize this shape as a spiral
(20.62)
On the other hand, if we consider the curve in the arithmetic space given by the same equations, we will observe a straight line:
(20.63)
Thus, the actual curve and its arithmetic manifestation have two completely different shapes. In other words, the arithmetic space corresponding to an induced Riemannian space is a gross distortion of the associated Euclidean space.
A more vivid example of the same effect is the familiar distortion that occurs when the spherical Earth is represented on a flat map. The distortion is especially apparent near the poles where the volume element Z=R2sinθ\sqrt{Z} =R^{2}\sin\theta is particularly small.
  (20.64)
The degree to which the shape of a great circle can change is also a striking illustration of the deformation. The great circle in the figure above is the terminator, i.e. the line that separates day and night.
Let us be reminded however that, despite the discrepancy between the Euclidean and the corresponding arithmetic spaces, the availability of the metric tensor ZijZ_{ij} assures that the arithmetic space retains all of the necessary information to perform Euclidean analysis. For example, suppose that the equation
Zi=Zi(t)(10.36)Z^{i}=Z^{i}\left( t\right) \tag{10.36}
described the trajectory of a material particle. Given these equations of motion we cannot easily imagine the actual Euclidean trajectory of the material particle. Nevertheless, we can calculate the components of its velocity and acceleration and calculate their magnitudes. Specifically, the components of the velocity are given by the equation
Vi=dZi(t)dt(10.36)V^{i}=\frac{dZ^{i}\left( t\right) }{dt} \tag{10.36}
while the components of the acceleration are given by
Ai=dVidt+ΓjkiVjVk.(12.80)A^{i}=\frac{dV^{i}}{dt}+\Gamma_{jk}^{i}V^{j}V^{k}. \tag{12.80}
Note that the availability of the Christoffel symbol is crucial for the latter calculation. With the help of the metric tensor, the magnitude of the velocity is given by
ZijViVj(20.65)\sqrt{Z_{ij}V^{i}V^{j}}\tag{20.65}
while that of the acceleration is given by
ZijAiAj.(20.66)\sqrt{Z_{ij}A^{i}A^{j}}.\tag{20.66}
We can also calculate the total distance LL traveled between the times t1t_{1} and t2t_{2} according to the equation
s=t0t1ZijViVjdt.(20.67)s=\int_{t_{0}}^{t_{1}}\sqrt{Z_{ij}V^{i}V^{j}}dt.\tag{20.67}
In summary, the fact that the arithmetic space is severed from the Euclidean space does not in any way preclude us from doing virtually any kind of analysis short of visualizing the actual shapes.
Finally, going slightly beyond our present scope, note that in terms of the equations of motion Zi(t)Z^{i}\left( t\right) , the components AiA^{i} of acceleration are given by
Ai=d2Zidt2+ΓjkidZjdtdZkdt.(20.68)A^{i}=\frac{d^{2}Z^{i}}{dt^{2}}+\Gamma_{jk}^{i}\frac{dZ^{j}}{dt}\frac{dZ^{k} }{dt}.\tag{20.68}
Thus, if the particle is moving uniformly in the Euclidean space and therefore its acceleration is zero, then the equations of motion are characterized by the identity
d2Zidt2+ΓjkidZjdtdZkdt=0(20.69)\frac{d^{2}Z^{i}}{dt^{2}}+\Gamma_{jk}^{i}\frac{dZ^{j}}{dt}\frac{dZ^{k}}{dt}=0\tag{20.69}
known as the geodesic equation. In a Euclidean space, the shortest curve connecting two points is a straight line and, thus, the geodesic equation tells us whether Zi(t)Z^{i}\left( t\right) represents uniform motion along a straight line.
In a general Riemannian space, the length LL of a section of a curve given by Zi(t)Z^{i}\left( t\right) is defined by the integral
L=t1t2ZijdZidtdZjdtdt.(20.70)L=\int_{t_{1}}^{t_{2}}\sqrt{Z_{ij}\frac{dZ^{i}}{dt}\frac{dZ^{j}}{dt}}dt.\tag{20.70}
It can be shown that the geodesic equation represents a form of the Euler-Lagrange equation for the above integral and thus characterizes the shortest path, again in the Riemannian sense, between two points. The discussion of the geodesic equation is beyond the scope of this book because it is best discussed in the context of the Calculus of Moving Surfaces where it can be given a more general treatment.
At the very outset of our narrative, we defined a Euclidean space as the physical space of our everyday existence in which the axioms of Euclidean geometry are valid. In particular, we have relied heavily on two unique characteristics of Euclidean spaces. First, Euclidean spaces can accommodate straight lines in any direction. This enabled us to introduce, and work freely with, geometric vectors. Secondly, Euclidean spaces admit Cartesian coordinate systems characterized by regular orthogonal coordinate grids. As a result, the metric tensor corresponds to the identity matrix at all points.
Between these two advantages, the former outweighs the latter. The methods of Tensor Calculus almost completely mitigate the effects of complicated coordinate systems. In fact, this subject's commitment to avoiding the use of special features of coordinate systems is one of the keys to its success. On the other hand, the use of geometric vectors was instrumental in stimulating our geometric intuition and helping us organize information in a very appealing way. Compared to the collection of coordinates UαU^{\alpha}, the geometric vector U\mathbf{U} is a tangible object that connects our analytical investigations to our geometric intuition. Furthermore, compared to UαU^{\alpha}, U\mathbf{U} is a variant of lower order and is therefore simpler. Finally, geometric vectors help guide our analytical intuition in manipulating coordinate expressions as our mind becomes trained to look for the geometric-vector interpretation.
Unfortunately, Euclidean spaces, according to our geometric definition, are limited to three or fewer dimensions. In higher dimensions, the only construct available to us is the Riemannian space. However, in a Riemannian space, i.e. a subset of Rn \mathbb{R} ^{n} paired with an arbitrary metric tensor field, there is no such thing as a geometric vector. In the absence of geometric vectors, a first-order tensor UαU^{\alpha} is condemned to variance for a greater part of our analysis. We can determine its invariant dot product with another vector VβV^{\beta} as SαβUαVβS_{\alpha\beta}U^{\alpha}V^{\beta} and its invariant length as the square root of the dot product with itself, but it otherwise remains in its intermediate variant state devoid of an immediate geometric interpretation.
However, having learned the great utility of boldface symbols in organizing our ideas in the Euclidean space, we would like to extend it to Riemannian spaces. This can be accomplished by introducing a construct referred to as the arithmetic Euclidean space in the following way. An arithmetic Euclidean space is a Riemannian space that occupies the entirety of Rn \mathbb{R} ^{n} where the metric tensor ZijZ_{ij} corresponds to the n×nn\times n identity matrix. At all points, the elements of the covariant basis, denoted by Zi\mathbf{Z}_{i}, correspond to the ithi^{\text{th}} column of the identity matrix. For example, in R4 \mathbb{R} ^{4},
Z1=[1000],     Z2=[0100],     Z3=[0010],     Z4=[0001].(20.71)\mathbf{Z}_{1}=\left[ \begin{array} {c} 1\\ 0\\ 0\\ 0 \end{array} \right] ,\ \ \ \ \ \mathbf{Z}_{2}=\left[ \begin{array} {c} 0\\ 1\\ 0\\ 0 \end{array} \right] ,\ \ \ \ \ \mathbf{Z}_{3}=\left[ \begin{array} {c} 0\\ 0\\ 1\\ 0 \end{array} \right] ,\ \ \ \ \ \mathbf{Z}_{4}=\left[ \begin{array} {c} 0\\ 0\\ 0\\ 1 \end{array} \right] .\tag{20.71}
Then, for a tensor UiU^{i} of order one, the invariant U\mathbf{U} is defined by
U=UiZi.(20.72)\mathbf{U}=U^{i}\mathbf{Z}_{i}.\tag{20.72}

20.7.1The Frenet equations in higher dimensions

As an illustration, we will extend the Frenet formulas to higher dimensions. Recall that the Frenet equations read
[TsPsQs]=[σσττ][TPQ],(5.101)\left[ \begin{array} {c} \mathbf{T}_{s}\\ \mathbf{P}_{s}\\ \mathbf{Q}_{s} \end{array} \right] =\left[ \begin{array} {rrr} & \sigma & \\ -\sigma & & \tau\\ & -\tau & \end{array} \right] \left[ \begin{array} {c} \mathbf{T}\\ \mathbf{P}\\ \mathbf{Q} \end{array} \right] , \tag{5.101}
where the subscript ss denotes differentiation with respect to arc length ss.
Part of the what makes the Frenet equations so elegant is the skew-symmetric property of the matrix. It turns out that this feature persists to higher dimensions.
Consider a curve embedded in an nn-dimensional arithmetic Euclidean space. Suppose that the curve is referred to its arc length ss and is specified by the equations
Zi=Zi(s).(20.73)Z^{i}=Z^{i}\left( s\right) .\tag{20.73}
Let T1\mathbf{T}_{1} be the unit tangent to the curve. Then
T1=dZi(s)dsZi.(20.74)\mathbf{T}_{1}=\frac{dZ^{i}\left( s\right) }{ds}\mathbf{Z}_{i}.\tag{20.74}
This definition enables us to avoid introducing the position vector R\mathbf{R}, which can certainly be done but is not necessary. Subsequently, however, we will not refer to the components of T1\mathbf{T}_{1} or any of the other vector quantities going forward.
The identity above defines the initial vector T1\mathbf{T}_{1} from the local frame. Subsequently, the unit vector Tm+1\mathbf{T}_{m+1} is obtained from Tm(s) \mathbf{T}_{m}^{\prime}\left( s\right) \ by applying the Gram-Schmidt algorithm in order to make it orthogonal to each of the preceding vectors, and subsequently factoring out a positive scalar to make it unit length. Finally, the normality of Tn\mathbf{T}_{n} is assured in a slightly different way, that you may already anticipate.
Let us investigate the construction of T2\mathbf{T}_{2} from T1(s)\mathbf{T} _{1}^{\prime}\left( s\right) . Since T1(s)T1(s)=1\mathbf{T}_{1}\left( s\right) \cdot\mathbf{T}_{1}\left( s\right) =1, differentiating both sides with respect to ss yields 2T1(s)T1(s)=02\mathbf{T}_{1}\left( s\right) \cdot\mathbf{T} _{1}^{\prime}\left( s\right) =0 from which it follows that T1(s)\mathbf{T} _{1}^{\prime}\left( s\right) is orthogonal to T1\mathbf{T}_{1}. Let κ1\kappa_{1} be the length of T1(s)\mathbf{T}_{1}^{\prime}\left( s\right) and let T2\mathbf{T}_{2} be the unit vector that points in the same direction as T1\mathbf{T}_{1}^{\prime}, i.e.
κ1T2=T1(20.75)\kappa_{1}\mathbf{T}_{2}=\mathbf{T}_{1}^{\prime}\tag{20.75}
Let us now move on to the next step and analyze T2\ \mathbf{T}_{2}^{\prime} which, by a previously used argument, is orthogonal to T2\mathbf{T}_{2}, i.e.
T2T1=0.(20.76)\mathbf{T}_{2}\cdot\mathbf{T}_{1}=0.\tag{20.76}
Differentiating this identity with respect to ss yields
T2T1+T2T1=0.(20.77)\mathbf{T}_{2}^{\prime}\cdot\mathbf{T}_{1}+\mathbf{T}_{2}\cdot\mathbf{T} _{1}^{\prime}=0.\tag{20.77}
since T1=κ1T2\mathbf{T}_{1}^{\prime}=\kappa_{1}\mathbf{T}_{2} and therefore T2T1=κ1T2T2=κ1\mathbf{T}_{2}\cdot\mathbf{T}_{1}^{\prime}=\kappa_{1}\mathbf{T}_{2} \cdot\mathbf{T}_{2}=\kappa_{1}, we find
T2T1=κ1(20.78)\mathbf{T}_{2}^{\prime}\cdot\mathbf{T}_{1}=-\kappa_{1}\tag{20.78}
Thus, the vector
T2+κ1T1(20.79)\mathbf{T}_{2}^{\prime}+\kappa_{1}\mathbf{T}_{1}\tag{20.79}
is orthogonal to both T1\mathbf{T}_{1} and T2\mathbf{T}_{2} and can therefore be used to define the unit vector T3\mathbf{T}_{3} and the corresponding absolute curvature κ2\kappa_{2} by the equation
κ2T3=T2+κ1T1.(20.80)\kappa_{2}\mathbf{T}_{3}=\mathbf{T}_{2}^{\prime}+\kappa_{1}\mathbf{T}_{1}.\tag{20.80}
This identity can also be written in the form
T2=κ2T3κ1T1(20.81)\mathbf{T}_{2}^{\prime}=\kappa_{2}\mathbf{T}_{3}-\kappa_{1}\mathbf{T}_{1}\tag{20.81}
that will serve as the second Frenet formula.
We now embark on the inductive step of the procedure. The vector Tm\mathbf{T}_{m} is determined by the condition that it is unit length, i.e.
TmTm=1,(20.82)\mathbf{T}_{m}\cdot\mathbf{T}_{m}=1,\tag{20.82}
and is orthogonal to all preceding vectors Tk\mathbf{T}_{k}, i.e.
TmTk=0 for all k<m(20.83)\mathbf{T}_{m}\cdot\mathbf{T}_{k}=0\text{ for all }k \lt m\tag{20.83}
Furthermore, we presume that for all kk between 11 and m1m-1, we have the Frenet formulas
Tk=κkTk+1κk1Tk1.(20.84)\mathbf{T}_{k}^{\prime}=\kappa_{k}\mathbf{T}_{k+1}-\kappa_{k-1}\mathbf{T} _{k-1}.\tag{20.84}
Differentiating the orthogonality condition, yields
TmTk+TmTk=0.(20.85)\mathbf{T}_{m}^{\prime}\cdot\mathbf{T}_{k}+\mathbf{T}_{m}\cdot\mathbf{T} _{k}^{\prime}=0.\tag{20.85}
Substituting the condition Tk=κkTk+1κk1Tk1\mathbf{T}_{k}^{\prime}=\kappa_{k}\mathbf{T} _{k+1}-\kappa_{k-1}\mathbf{T}_{k-1} yields
TmTk+κkTmTk+1κk1TmTk1=0.(20.86)\mathbf{T}_{m}^{\prime}\cdot\mathbf{T}_{k}+\kappa_{k}\mathbf{T}_{m} \cdot\mathbf{T}_{k+1}-\kappa_{k-1}\mathbf{T}_{m}\cdot\mathbf{T}_{k-1}=0.\tag{20.86}
The dot product TmTk1\mathbf{T}_{m}\cdot\mathbf{T}_{k-1} vanishes for all k<mk \lt m by orthogonality. Meanwhile, the dot product TmTk+1\mathbf{T}_{m}\cdot \mathbf{T}_{k+1} vanishes for all k<m1k \lt m-1 by orthogonality and equals 11 for k=m1k=m-1. Therefore, the above equation tells us that
TmTk=0 for k<m1(20.87)\mathbf{T}_{m}^{\prime}\cdot\mathbf{T}_{k}=0\text{ for }k \lt m-1\tag{20.87}
and
TmTm1=κm1.(20.88)\mathbf{T}_{m}^{\prime}\cdot\mathbf{T}_{m-1}=-\kappa_{m-1}.\tag{20.88}
In other words, Tm\mathbf{T}_{m}^{\prime} is orthogonal to T1,,Tm2\mathbf{T} _{1},\ldots,\mathbf{T}_{m-2}, and, of course, Tm\mathbf{T}_{m}, but not Tm1\mathbf{T}_{m-1}. The vector
Tm+κm1Tm1(20.89)\mathbf{T}_{m}^{\prime}+\kappa_{m-1}\mathbf{T}_{m-1}\tag{20.89}
is, therefore, orthogonal to all mm vectors T1,,Tm\mathbf{T}_{1},\ldots ,\mathbf{T}_{m} and can be used to introduce the unit vector Tm+1\mathbf{T} _{m+1} and the absolute curvature κm\kappa_{m}:
κmTm+1=Tm+κm1Tm1.(20.90)\kappa_{m}\mathbf{T}_{m+1}=\mathbf{T}_{m}^{\prime}+\kappa_{m-1}\mathbf{T} _{m-1}.\tag{20.90}
Rewriting this equation in the form
Tm=κmTm+1κm1Tm1.(20.91)\mathbf{T}_{m}^{\prime}=\kappa_{m}\mathbf{T}_{m+1}-\kappa_{m-1}\mathbf{T} _{m-1}.\tag{20.91}
gives us the mthm^{\text{th}} Frenet equation.
We can continue in this fashion until m=nm=n. Following the pattern with the Frenet equations in three dimensions, the unit vector Tn\mathbf{T}_{n} is chosen so that the set of nn vectors T1,,Tn\mathbf{T}_{1},\ldots,\mathbf{T}_{n} is positively oriented. Then κn1\kappa_{n-1} is chosen so that the identity
κn1Tn=Tn1+κn2Tn2,(20.92)\kappa_{n-1}\mathbf{T}_{n}=\mathbf{T}_{n-1}^{\prime}+\kappa_{n-2} \mathbf{T}_{n-2},\tag{20.92}
is satisfied and, thus, κn1\kappa_{n-1} can be either positive or negative and can rightfully be called the torsion. If κn1>0\kappa_{n-1}\gt 0 the curve is called right-handed and κn1<0\kappa_{n-1} \lt 0 then it is called left-handed.
Finally, we note that by the same argument as above, Tn\mathbf{T}_{n}^{\prime} is orthogonal to all Tm\mathbf{T}_{m} except Tn1\mathbf{T}_{n-1} and that
TnTn1=κn1.(20.93)\mathbf{T}_{n}^{\prime}\cdot\mathbf{T}_{n-1}=-\kappa_{n-1}.\tag{20.93}
Thus,
Tn+κn1Tn1(20.94)\mathbf{T}_{n}^{\prime}+\kappa_{n-1}\mathbf{T}_{n-1}\tag{20.94}
is orthogonal to all nn vectors T1,,Tn\mathbf{T}_{1},\cdots,\mathbf{T}_{n}. Since our analysis is taking place in nn dimensions, we conclude that this vector equals 0\mathbf{0}, i.e.
Tn+κn1Tn1=0.(20.95)\mathbf{T}_{n}^{\prime}+\kappa_{n-1}\mathbf{T}_{n-1}=\mathbf{0.}\tag{20.95}
Thus, the nthn^{\text{th}} Frenet equation reads
Tn=κn1Tn1.(20.96)\mathbf{T}_{n}^{\prime}=-\kappa_{n-1}\mathbf{T}_{n-1}\mathbf{.}\tag{20.96}
In summary, the Frenet equations read
[T1T2Tn1Tn]=[κ1κ1κ2κn1κn1κn1][T1T2Tn1Tn].(20.97)\left[ \begin{array} {l} \mathbf{T}_{1}^{\prime}\\ \mathbf{T}_{2}^{\prime}\\ \cdots\\ \mathbf{T}_{n-1}^{\prime}\\ \mathbf{T}_{n}^{\prime} \end{array} \right] =\left[ \begin{array} {ccccc} & \kappa_{1} & & & \\ -\kappa_{1} & & \kappa_{2} & & \\ & \ddots & & \ddots & \\ & & -\kappa_{n-1} & & \kappa_{n-1}\\ & & & -\kappa_{n-1} & \end{array} \right] \left[ \begin{array} {l} \mathbf{T}_{1}\\ \mathbf{T}_{2}\\ \cdots\\ \mathbf{T}_{n-1}\\ \mathbf{T}_{n} \end{array} \right] .\tag{20.97}
Exercise 20.1Show that from the intrinsic definition of the Christoffel symbol, i.e.
Γi,jk=12(ZijZk+ZikZjZjkZi),(12.49)\Gamma_{i,jk}=\frac{1}{2}\left( \frac{\partial Z_{ij}}{\partial Z^{k}} +\frac{\partial Z_{ik}}{\partial Z^{j}}-\frac{\partial Z_{jk}}{\partial Z^{i} }\right) , \tag{12.49}
it follows that
ZijZk=Γi,jk+Γj,ik.(12.40)\frac{\partial Z_{ij}}{\partial Z^{k}}=\Gamma_{i,jk}+\Gamma_{j,ik}. \tag{12.40}
Exercise 20.2Show that the Riemann-Christoffel tensor is skew-symmetric in its last two indices, i.e.
Rlmij=Rlmji(20.14)R_{lmij}=-R_{lmji} \tag{20.14}
Exercise 20.3Using the formula
Γi,jk=12(ZijZk+ZikZjZjkZi),(12.49)\Gamma_{i,jk}=\frac{1}{2}\left( \frac{\partial Z_{ij}}{\partial Z^{k}} +\frac{\partial Z_{ik}}{\partial Z^{j}}-\frac{\partial Z_{jk}}{\partial Z^{i} }\right) , \tag{12.49}
present the covariant Riemann-Christoffel tensor
Rlmij=Γl,jmZiΓl,imZj+Γn,jlΓimnΓn,ilΓjmn(20.13)R_{lmij}=\frac{\partial\Gamma_{l,jm}}{\partial Z^{i}}-\frac{\partial \Gamma_{l,im}}{\partial Z^{j}}+\Gamma_{n,jl}\Gamma_{im}^{n}-\Gamma _{n,il}\Gamma_{jm}^{n} \tag{20.13}
in the form
Rlmij=12(2ZljZmZi+2ZmiZlZj2ZmjZlZi2ZliZmZj)+Γn,jlΓimnΓn,ilΓjmn.(20.98)R_{lmij}=\frac{1}{2}\left( \frac{\partial^{2}Z_{lj}}{\partial Z^{m}\partial Z^{i}}+\frac{\partial^{2}Z_{mi}}{\partial Z^{l}\partial Z^{j}}-\frac {\partial^{2}Z_{mj}}{\partial Z^{l}\partial Z^{i}}-\frac{\partial^{2}Z_{li} }{\partial Z^{m}\partial Z^{j}}\right) +\Gamma_{n,jl}\Gamma_{im}^{n} -\Gamma_{n,il}\Gamma_{jm}^{n}.\tag{20.98}
Exercise 20.4From the form obtained in the previous exercise, show that the Riemann-Christoffel tensor is skew-symmetric in its first two indices, i.e.
Rlmij=Rmlij,(20.15)R_{lmij}=-R_{mlij}, \tag{20.15}
and that it is symmetric with respect to switching the first two indices with the last two, i.e.
Rlmij=Rijlm.(20.16)R_{lmij}=R_{ijlm}. \tag{20.16}
Exercise 20.5Show that the skew-symmetric property
Rlmij=Rmlij(20.15)R_{lmij}=-R_{mlij} \tag{20.15}
can also be deduced from
Rlmij=Rlmji(20.14)R_{lmij}=-R_{lmji} \tag{20.14}
and
Rlmij=Rijlm.(20.16)R_{lmij}=R_{ijlm}. \tag{20.16}
Exercise 20.6 Show the first Bianchi identity
Rlmij+Rlijm+Rljmi=0.(20.17)R_{lmij}+R_{lijm}+R_{ljmi}=0. \tag{20.17}
Exercise 20.7 Show the second Bianchi identity
kRlmij+iRlmjk+jRlmki=0.(20.99)\nabla_{k}R_{lmij}+\nabla_{i}R_{lmjk}+\nabla_{j}R_{lmki}=0.\tag{20.99}
Exercise 20.8Show that
Rmijm=0(20.100)R_{\cdot mij}^{m}=0\tag{20.100}
and, similarly,
Rlm  i   i=0.(20.101)R_{lm\ \cdot\ i}^{\ \cdot\ \cdot\ i}=0.\tag{20.101}
Exercise 20.9Show that the skew-symmetries of the Riemann-Christoffel tensor imply that it vanishes in a one-dimensional space. Alternatively demonstrate this fact by an explicit calculation on the basis of the definition
Rmijk=ΓjmkZiΓimkZj+ΓinkΓjmnΓjnkΓimn.(15.123)R_{\cdot mij}^{k}=\frac{\partial\Gamma_{jm}^{k}}{\partial Z^{i}} -\frac{\partial\Gamma_{im}^{k}}{\partial Z^{j}}+\Gamma_{in}^{k}\Gamma_{jm} ^{n}-\Gamma_{jn}^{k}\Gamma_{im}^{n}. \tag{15.123}
Exercise 20.10Show that the Ricci curvature tensor defined by the identity
Rij=Rikjk(20.18)R_{ij}=R_{\cdot ikj}^{k} \tag{20.18}
is symmetric, i.e.
Rij=Rji.(20.19)R_{ij}=R_{ji}. \tag{20.19}
Exercise 20.11Show the general commutator formula
ijTlkjiTlk=RmijkTlmRlijmTmk.(20.22)\nabla_{i}\nabla_{j}T_{l}^{k}-\nabla_{j}\nabla_{i}T_{l}^{k}=R_{\cdot mij} ^{k}T_{l}^{m}-R_{\cdot lij}^{m}T_{m}^{k}. \tag{20.22}
A good approach is to consider a first-order variant
Tk=TmkSm,(20.102)T^{k}=T_{m}^{k}S^{m},\tag{20.102}
and to use the already-established formula for the commutator ijTkjiTk\nabla _{i}\nabla_{j}T^{k}-\nabla_{j}\nabla_{i}T^{k} as the starting point.
Exercise 20.12Show that in a two-dimensional Riemannian space, the Ricci tensor curvature is given by
Rij=KZij.(20.28)R_{ij}=KZ_{ij}. \tag{20.28}
Therefore, the Gaussian curvature KK is half the scalar curvature, i.e.
K=12Rii.(20.103)K=\frac{1}{2}R_{i}^{i}.\tag{20.103}
Exercise 20.13In a two-dimensional Riemannian space, suppose that the metric tensor field is diagonal, i.e.
Z12=Z21=0.(20.104)Z_{12}=Z_{21}=0.\tag{20.104}
This, for instance, is the case for a Riemannian space induced from a Euclidean plane referred to an orthogonal coordinate system. Show that
K=12Z(Z1(1ZZ22Z1)+Z2(1ZZ11Z2)).(20.105)K=-\frac{1}{2\sqrt{Z}}\left( \frac{\partial}{\partial Z^{1}}\left( \frac {1}{\sqrt{Z}}\frac{\partial Z_{22}}{\partial Z^{1}}\right) +\frac{\partial }{\partial Z^{2}}\left( \frac{1}{\sqrt{Z}}\frac{\partial Z_{11}}{\partial Z^{2}}\right) \right) .\tag{20.105}
In particular, if Z111Z_{11}\equiv1, then
K=1Z222Z22Z1Z1.(20.106)K=-\frac{1}{\sqrt{Z_{22}}}\frac{\partial^{2}\sqrt{Z_{22}}}{\partial Z^{1}\partial Z^{1}}.\tag{20.106}
Exercise 20.14Suppose that the metric tensor ZijZ_{ij} that corresponds to the matrix
 [R2R2sin2(Z1)],(20.107)\text{ }\left[ \begin{array} {cc} R^{2} & \\ & R^{2}\sin^{2}\left( Z^{1}\right) \end{array} \right] ,\tag{20.107}
where RR is a constant. Show that the nonzero elements of the Christoffel symbol Γjki\Gamma_{jk}^{i} are
Γ221=cosZ1sinZ1     and    Γ122=Γ212=cotZ1.(20.108)\Gamma_{22}^{1}=-\cos Z^{1}\sin Z^{1}\text{\ \ \ \ \ and\ \ \ \ }\Gamma _{12}^{2}=\Gamma_{21}^{2}=\cot Z^{1}.\tag{20.108}
Furthermore, the Gaussian curvature is given by
K=1R2(20.109)K=\frac{1}{R^{2}}\tag{20.109}
and therefore the Riemann-Christoffel tensor RkmijR_{kmij} is given by
Rkmij=1R2εkmεij.(20.110)R_{kmij}=\frac{1}{R^{2}}\varepsilon_{km}\varepsilon_{ij}.\tag{20.110}
Exercise 20.15Show that for a metric tensor ZijZ_{ij} that corresponds to the matrix
[100(Z1)2n],(20.111)\left[ \begin{array} {cc} 1 & 0\\ 0 & \left( Z^{1}\right) ^{2n} \end{array} \right] ,\tag{20.111}
the Gaussian curvature K(Z)K\left( Z\right) is given by
K(Z)=n(1n)(Z1)2(20.112)K\left( Z\right) =\frac{n\left( 1-n\right) }{\left( Z^{1}\right) ^{2}}\tag{20.112}
and, therefore,
Rkmij=n(1n)(Z1)2εijεkm.(20.113)R_{kmij}=\frac{n\left( 1-n\right) }{\left( Z^{1}\right) ^{2}} \varepsilon_{ij}\varepsilon_{km}.\tag{20.113}
Thus, for n=1n=1, for which ZijZ_{ij} corresponds to polar coordinates, we have, as expected, Rkmij=0R_{kmij}=0. For n>1n\gt 1, on the other hand, the corresponding metric tensor cannot be induced from a Euclidean space.
Exercise 20.16In a three-dimensional space, let
Trs=14εkmrεijsRkmij.(20.114)T^{rs}=\frac{1}{4}\varepsilon^{kmr}\varepsilon^{ijs}R_{kmij}.\tag{20.114}
Show that TrsT^{rs} is symmetric, i.e.
Trs=Tsr.(20.115)T^{rs}=T^{sr}.\tag{20.115}
Furthermore, RkmijR_{kmij} can be recovered from TrsT^{rs} by the formula
Rkmij=Trsεkmrεijs.(20.116)R_{kmij}=T^{rs}\varepsilon_{kmr}\varepsilon_{ijs}.\tag{20.116}
Problem 20.1Show that, owing to the symmetries present in the Riemann-Christoffel tensor, the number of available degrees of freedom in an nn-dimensional Riemannian space is
(n21)n212.(20.117)\frac{\left( n^{2}-1\right) n^{2}}{12}.\tag{20.117}
Send feedback to Pavel