The Tensor Property

We have finally arrived at the concept of a tensor -- a variant that transforms under a change of coordinates by a special rule. We gave a preview of this cornerstone concept in the previous Chapter. In this Chapter, we will give its complete definition and explore its crucial properties and implications.
As we did in the previous Chapter, let us simultaneously refer the Euclidean space to two coordinate systems: the unprimed coordinates ZiZ^{i} and the primed coordinates ZiZ^{i^{\prime}}. Let
Zi=Zi(Z)(13.21)Z^{i^{\prime}}=Z^{i^{\prime}}\left( Z\right) \tag{13.21}
and
Zi=Zi(Z).(13.22)Z^{i}=Z^{i}\left( Z^{\prime}\right) . \tag{13.22}
be the equations of the coordinate transformation. Recall that the Jacobians JiiJ_{i}^{i^{\prime}} and JiiJ_{i^{\prime}}^{i} are defined as the collections of partial derivatives of those equations, i.e.
Jii(Z)=Zi(Z)Zi(13.31)J_{i}^{i^{\prime}}\left( Z\right) =\frac{\partial Z^{i^{\prime}}\left( Z\right) }{\partial Z^{i}} \tag{13.31}
and
Jii(Z)=Zi(Z)Zi.(13.32)J_{i^{\prime}}^{i}\left( Z^{^{\prime}}\right) =\frac{\partial Z^{i}\left( Z^{\prime}\right) }{\partial Z^{i^{\prime}}}. \tag{13.32}
The precise algebraic definition of a tensor depends on the order of the variant in question. In words, a variant is a tensor if its transformation rule is captured by contracting each of its indices with the appropriate Jacobian. In order to express this definition by an algebraic equation, we will consider a variant TjiT_{j}^{i} with a representative collection of indices and use the equation below to prescribe the rule for transforming each kind of index. It is then understood that in order to apply the definition to a variant with an arbitrary indicial signature, the appropriate rule must be applied to each index. In the future, we will introduce as many as six different kinds of indices and therefore our representative collection of indices will become larger. However, for now, a single superscript and a single subscript with suffice.
A variant TjiT_{j}^{i} is a tensor if its primed and unprimed coordinate manifestations TjiT_{j}^{i} and TjiT_{j^{\prime}}^{i^{\prime}} are related by the equation
Tji=TjiJiiJjj.(14.1)T_{j^{\prime}}^{i^{\prime}}=T_{j}^{i}J_{i}^{i^{\prime}}J_{j^{\prime}}^{j}.\tag{14.1}
This definition is to be understood in the sense that for every index in the indicial signature, there is a contraction with the appropriate Jacobian -- specifically, the one that makes for a valid summation convention. The transformation rule that corresponds to a superscript, and therefore a contraction with JiiJ_{i}^{i^{\prime}}, is called contravariant. The rule that corresponds to a subscript, and therefore a contraction with JiiJ_{i^{\prime}}^{i}, is called covariant.
When applied to a first-order variant TiT^{i} with a superscript, the definition reads
Ti=TiJii.(14.2)T^{i^{\prime}}=T^{i}J_{i}^{i^{\prime}}.\tag{14.2}
Such a variant is called a first-order contravariant tensor. Similarly, for a first-order variant TiT_{i} with a subscript, the definition reads
Ti=TiJii.(14.3)T_{i^{\prime}}=T_{i}J_{i^{\prime}}^{i}.\tag{14.3}
Such a variant is called a first-order covariant tensor. For a higher-order example, a fourth-order variant TjkliT_{jkl}^{i} is a tensor if
Tjkli=TjkliJiiJjjJkkJll.(14.4)T_{j^{\prime}k^{\prime}l^{\prime}}^{i^{\prime}}=T_{jkl}^{i}J_{i}^{i^{\prime} }J_{j^{\prime}}^{j}J_{k^{\prime}}^{k}J_{l^{\prime}}^{l}.\tag{14.4}
It is best described as a fourth-order tensor of the type indicated by its indicial signature.
In the past, we determined placements of indices according to relatively informal guidelines. For example, we assigned a subscript to the covariant basis Zi\mathbf{Z}_{i} since in its definition,
Zi=R(Z)Zi,(9.9)\mathbf{Z}_{i}=\frac{\partial\mathbf{R}\left( Z\right) }{\partial Z^{i}}, \tag{9.9}
the index ii on the right appears on the bottom of a "fraction". The covariant metric tensor ZijZ_{ij}, defined by
Zij=ZiZj,(14.5)Z_{ij}=\mathbf{Z}_{i}\cdot\mathbf{Z}_{j},\tag{14.5}
received two subscripts because that was the only choice that conforms to the rules of the tensor notation. The contravariant components UiU^{i}, defined by
U=UiZi,(14.6)\mathbf{U}=U^{i}\mathbf{Z}_{i},\tag{14.6}
received a superscript in order to make the summation convention work. We recently discovered that our intuitive guidelines lead to placements that proved to be accurate predictors of the manner in which variants transform under a change of coordinates. However, having now given a precise definition of a tensor, we must retroactively reverse the logic. From this point forward, it is the manner in which the tensor transforms that determines the placement of the indices.
Recall that the covariant and contravariant bases, and the covariant and contravariant metric tensors transform according to the rules
Zi=ZiJii          (13.33)Zi=ZiJii          (13.84)Zij=ZijJiiJjj          (13.59)Zij=ZijJiiJjj          (13.87)\begin{aligned}\mathbf{Z}_{i^{\prime}} & =\mathbf{Z}_{i}J_{i^{\prime}}^{i}\ \ \ \ \ \ \ \ \ \ \left(13.33\right)\\\mathbf{Z}^{i^{\prime}} & =\mathbf{Z}^{i}J_{i}^{i^{\prime}}\ \ \ \ \ \ \ \ \ \ \left(13.84\right)\\Z_{i^{\prime}j^{\prime}} & =Z_{ij}J_{i^{\prime}}^{i}J_{j^{\prime}} ^{j}\ \ \ \ \ \ \ \ \ \ \left(13.59\right)\\Z^{i^{\prime}j^{\prime}} & =Z^{ij}J_{i}^{i^{\prime}}J_{j}^{j^{\prime}} \ \ \ \ \ \ \ \ \ \ \left(13.87\right)\end{aligned}
Thus, the use of the terms covariant and contravariant in describing these objects is finally justified.
For tensors of order greater than one, it is cumbersome to properly assign the terms covariant and contravariant, especially if the order of the indices is not explicitly specified. It is more efficient to describe such tensors, as we did above, to be of the kind indicated by the indicial signature. On the other hand, the indicial signature speaks for itself and therefore, in most circumstances, no additional description is needed. We will find that the terms covariant and contravariant are often omitted altogether, except for objects such as the metric tensors, where these terms help to distinguish between two closely related objects.
It is clear from the definition
Tji=TjiJiiJjj(14.1)T_{j^{\prime}}^{i^{\prime}}=T_{j}^{i}J_{i}^{i^{\prime}}J_{j^{\prime}}^{j} \tag{14.1}
that a variant TjiT_{j}^{i} whose elements are defined to be zero in all coordinate systems is a tensor. Furthermore, if a variant that is known to be a tensor, vanishes in one coordinate system, then it vanishes in all coordinate systems. This seemingly simple fact finds frequent applications in practice since it makes it easy to show that a tensor vanishes by finding one special coordinate system in which that is easy to see. In a spectacular application of this idea, we will demonstrate 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}
which holds in any coordinate system in a Euclidean space. The challenging part in proving this identity is demonstrating that the expression on the left is a tensor. Once that is established, however, the identity follows immediately by observing that the Christoffel symbol vanishes in affine coordinates.
Crucially, the definition of a tensor applies to a variant of order zero, i.e. a single number or a single vector. Since variants of order zero have no indices, the definition contains no Jacobians and thus implies that the values of the variant in different coordinate systems are the same. In other words, a tensor of order zero is an invariant. This statement deserves an exclamation point: a tensor of order zero is an invariant! As it will emerge from the upcoming discussion, this observation is not an edge-case technicality but is a fact that is at the heart of the matter and the key to producing geometrically meaningful results.
Stating this fact algebraically requires a slight notational modification. Since a variant TT of order zero has no indices that can accept a prime, we are forced to place it on the symbol TT itself, as in TT^{\prime}, to denote the value of TT in the primed coordinates. In this notation, the fact that TT is a tensor reads
T=T.(14.7)T^{\prime}=T.\tag{14.7}
Importantly, not all variants of order zero are invariants. Recall from Chapter 9, that the volume element is defined to be Z\sqrt{Z}, where ZZ is the determinant of the matrix representing the covariant metric tensor ZijZ_{ij}. Since
Z=det[1001]=1 in Cartesian coordinates(9.59)\sqrt{Z}=\sqrt{\det\left[ \begin{array} {cc} 1 & 0\\ 0 & 1 \end{array} \right] }=1\text{ in Cartesian coordinates} \tag{9.59}
and
Z=det[100r2]=r in polar coordinates,(9.60)\sqrt{Z}=\sqrt{\det\left[ \begin{array} {cc} 1 & 0\\ 0 & r^{2} \end{array} \right] }=r\text{ in polar coordinates,} \tag{9.60}
it is evident that Z\sqrt{Z} is not an invariant and therefore not a tensor.
Finally, note that the term invariant is also used to describe scalar and vector fields that exist in the Euclidean space completely independently of coordinates. These are fields that, so to say, "predate" coordinates. For example, the position vector field R\mathbf{R}, a temperature field TT, and its gradient T\mathbb{\nabla}T can be rightfully described as invariants since their values are independent of coordinates. For these fields, however, the term invariant should not be understood in the sense of a coordinate-independent result of a coordinate-dependent analytical procedure. The invariance of such fields is deeper: these fields are constructed without a reference to any coordinate system.
We will soon show -- in short order! -- that all basic algebraic operations on variants, i.e. addition, multiplication, and contraction, preserve the tensor property. In other words, a sum of two tensors, a product of two tensors, and a contraction of a tensor are all tensors in their own right. Among these operations, contraction is the only one that reduces the order of a variant. It is therefore always the final step in constructing tensors of order zero, i.e. invariants.
The mechanism by which contraction preserves the tensor property is extraordinarily insightful. Later in this Chapter, we will give a general demonstration but, for now, we will reveal that mechanism by considering the narrow example of contracting a second-order tensor TjiT_{j}^{i}. The resulting variant TiiT_{i}^{i} of order zero is commonly referred to as the trace of TjiT_{j}^{i} by analogy with matrices.
Suppose that TjiT_{j}^{i} is a tensor. In other words, its values TjiT_{j^{\prime}}^{i^{\prime}} in the primed coordinate system are related to TjiT_{j}^{i} by the equation
Tji=TjiJiiJjj.(14.1)T_{j^{\prime}}^{i^{\prime}}=T_{j}^{i}J_{i}^{i^{\prime}}J_{j^{\prime}}^{j}. \tag{14.1}
Let TT be the result of contracting TjiT_{j}^{i}, i.e.
T=Tii,(14.8)T=T_{i}^{i},\tag{14.8}
and TT^{\prime} be the result of contracting TjiT_{j^{\prime}}^{i^{\prime}}, i.e.
T=Tii.(14.9)T^{\prime}=T_{i^{\prime}}^{i^{\prime}}.\tag{14.9}
Since,
Tji=TjiJiiJjj,(14.1)T_{j^{\prime}}^{i^{\prime}}=T_{j}^{i}J_{i}^{i^{\prime}}J_{j^{\prime}}^{j}, \tag{14.1}
the contraction TiiT_{i^{\prime}}^{i^{\prime}} is obtained by replacing jj^{\prime} with ii^{\prime} on the right:
Tii=TjiJiiJij.(14.10)T_{i^{\prime}}^{i^{\prime}}=T_{j}^{i}J_{i}^{i^{\prime}}J_{i^{\prime}}^{j}.\tag{14.10}
Note that in the previous identity
Tji=TjiJiiJjj,(14.1)T_{j^{\prime}}^{i^{\prime}}=T_{j}^{i}J_{i}^{i^{\prime}}J_{j^{\prime}}^{j}, \tag{14.1}
the two Jacobians do not interact. In
Tii=TjiJiiJij,(14.10)T_{i^{\prime}}^{i^{\prime}}=T_{j}^{i}J_{i}^{i^{\prime}}J_{i^{\prime}}^{j}, \tag{14.10}
on the other hand, the two Jacobians are connected by a contraction on ii^{\prime}. Since, crucially, the two Jacobians are the matrix inverses of each other, i.e.
JiiJij=δij,(13.48)J_{i}^{i^{\prime}}J_{i^{\prime}}^{j}=\delta_{i}^{j}, \tag{13.48}
we find that
Tii=Tjiδij=Tii,(14.11)T_{i^{\prime}}^{i^{\prime}}=T_{j}^{i}\delta_{i}^{j}=T_{i}^{i},\tag{14.11}
i.e.
T=T.(14.12)T^{\prime}=T.\tag{14.12}
In other words, TT is an invariant, as we set out to show.
This simple example is the culmination of much of our efforts so far. It shows how the tensor framework achieves invariance through the interplay between covariance and contravariance. Tensors are not themselves invariants and are therefore not in and of themselves geometrically meaningful. However, they can be combined to produce geometrically meaningful invariants.
This example also shows how the indicial notation, along with Einstein's summation convention, seamlessly assure that one adheres to operations that preserve the tensor property. The principal guideline of Tensor Calculus is to limit one's analysis to tensors and tensor-preserving operations and thus to be assured of geometrically meaningful results. This seemingly simplistic approach will prove to be extremely effective and will dictate the future course of our investigations.
A further word of appreciation for tensors is in order. What a beautiful object a tensor is! It simultaneously exhibits features that reflect its geometric origins along with some artifacts of the coordinate system. However, the artifacts of the coordinate system are not out of control. They are present in systematic fashion captured algebraically by the Jacobians. By the end of analysis, those artifacts are removed by contraction, leaving us with a geometrically meaningful invariant. Thus, tensors can be thought of as nearly-invariant. So much so that the term invariant is sometimes used in place of tensor. For example, the covariant derivative introduced in Chapter 15 is sometimes referred to as the invariant derivative. The same is the case for the ˙\dot{\nabla} -derivative in the Calculus of Moving Surfaces which is also often referred to as the invariant time derivative.
The sum property states that the sum of two tensors is a tensor in its own right. This, of course, is a nearly trivial statement. After all, the tensor transformation rule amounts to a contraction with a Jacobian which is essentially an elementary sum of elementary products. Thus, the fact that the sum of two tensors is itself a tensor is equivalent to the distributive property of multiplication. Nevertheless, we will give a formal proof of the sum property as it will be help establish a template for a number of future proofs.
Since the precise form of the transformation of a tensor depends on its order, proofs of the tensor property, much like its definition, are usually given for tensors with a reasonably representative collection of indices. This approach is effective as long as it makes it clear that the proof remains valid for tensors with arbitrary indicial signatures. In that spirit, suppose that the variants AjkiA_{jk}^{i} and BjkiB_{jk}^{i} are tensors, i.e.
Ajki=AjkiJiiJjjJkk, and          (14.13)Bjki=BjkiJiiJjjJkk.          (14.14)\begin{aligned}A_{j^{\prime}k^{\prime}}^{i^{\prime}} & =A_{jk}^{i}J_{i}^{i^{\prime} }J_{j^{\prime}}^{j}J_{k^{\prime}}^{k}\text{, and}\ \ \ \ \ \ \ \ \ \ \left(14.13\right)\\B_{j^{\prime}k^{\prime}}^{i^{\prime}} & =B_{jk}^{i}J_{i}^{i^{\prime} }J_{j^{\prime}}^{j}J_{k^{\prime}}^{k}.\ \ \ \ \ \ \ \ \ \ \left(14.14\right)\end{aligned}
Let CjkiC_{jk}^{i} be the sum of AjkiA_{jk}^{i} and BjkiB_{jk}^{i}:
Cjki=Ajki+Bjki.(14.15)C_{jk}^{i}=A_{jk}^{i}+B_{jk}^{i}.\tag{14.15}
Our goal is to show CjkiC_{jk}^{i} is a tensor, i.e.
Cjki=CjkiJiiJjjJkk.(14.16)C_{j^{\prime}k^{\prime}}^{i^{\prime}}=C_{jk}^{i}J_{i}^{i^{\prime}} J_{j^{\prime}}^{j}J_{k^{\prime}}^{k}.\tag{14.16}
By definition, CjkiC_{j^{\prime}k^{\prime}}^{i^{\prime}} is the sum of AjkiA_{j^{\prime}k^{\prime}}^{i^{\prime}} and BjkiB_{j^{\prime}k^{\prime} }^{i^{\prime}}:
Cjki=Ajki+Bjki.(14.17)C_{j^{\prime}k^{\prime}}^{i^{\prime}}=A_{j^{\prime}k^{\prime}}^{i^{\prime} }+B_{j^{\prime}k^{\prime}}^{i^{\prime}}.\tag{14.17}
Substituting the transformation rules for AjkiA_{jk}^{i} and BjkiB_{jk}^{i} yields
Cjki=AjkiJiiJjjJkk+BjkiJiiJjjJkk.(14.18)C_{j^{\prime}k^{\prime}}^{i^{\prime}}=A_{jk}^{i}J_{i}^{i^{\prime}} J_{j^{\prime}}^{j}J_{k^{\prime}}^{k}+B_{jk}^{i}J_{i}^{i^{\prime}}J_{j^{\prime }}^{j}J_{k^{\prime}}^{k}.\tag{14.18}
By the distributive law,
Cjki=(Ajki+Bjki)JiiJjjJkk.(14.19)C_{j^{\prime}k^{\prime}}^{i^{\prime}}=\left( A_{jk}^{i}+B_{jk}^{i}\right) J_{i}^{i^{\prime}}J_{j^{\prime}}^{j}J_{k^{\prime}}^{k}.\tag{14.19}
Since Ajki+Bjki=CjkiA_{jk}^{i}+B_{jk}^{i}=C_{jk}^{i} we find that, indeed,
Cjki=CjkiJiiJjjJkk,(14.16)C_{j^{\prime}k^{\prime}}^{i^{\prime}}=C_{jk}^{i}J_{i}^{i^{\prime}} J_{j^{\prime}}^{j}J_{k^{\prime}}^{k}, \tag{14.16}
as we set out to prove.
Let us now demonstrate the product property of tensors, i.e. that the product of two tensors is a tensor in its own right. As before, we will consider two tensors with specific indicial signatures and, as before, it will be clear that the argument remains valid in general. Suppose that AjiA_{j}^{i} and BlmkB_{lm}^{k} are tensors, i.e.
Aji=AjiJiiJjj and          (14.20)Blmk=BlmkJkkJllJmm,          (14.21)\begin{aligned}A_{j^{\prime}}^{i^{\prime}} & =A_{j}^{i}J_{i}^{i^{\prime}}J_{j^{\prime}} ^{j}\text{ and}\ \ \ \ \ \ \ \ \ \ \left(14.20\right)\\B_{l^{\prime}m^{\prime}}^{k^{\prime}} & =B_{lm}^{k}J_{k}^{k^{\prime} }J_{l^{\prime}}^{l}J_{m^{\prime}}^{m},\ \ \ \ \ \ \ \ \ \ \left(14.21\right)\end{aligned}
and let CjlmikC_{jlm}^{ik} be their product:
Cjlmik=AjiBlmk.(14.22)C_{jlm}^{ik}=A_{j}^{i}B_{lm}^{k}.\tag{14.22}
We must show that CjlmikC_{jlm}^{ik} is a tensor, i.e.
Cjlmik=CjlmikJiiJjjJkkJllJmm.(14.23)C_{j^{\prime}l^{\prime}m^{\prime}}^{i^{\prime}k^{\prime}}=C_{jlm}^{ik} J_{i}^{i^{\prime}}J_{j^{\prime}}^{j}J_{k}^{k^{\prime}}J_{l^{\prime}} ^{l}J_{m^{\prime}}^{m}.\tag{14.23}
In the primed coordinate system, CjlmikC_{j^{\prime}l^{\prime}m^{\prime} }^{i^{\prime}k^{\prime}} is given by the product of AjiA_{j^{\prime} }^{i^{\prime}} and BlmkB_{l^{\prime}m^{\prime}}^{k^{\prime}}:
Cjlmik=AjiBlmk.(14.24)C_{j^{\prime}l^{\prime}m^{\prime}}^{i^{\prime}k^{\prime}}=A_{j^{\prime} }^{i^{\prime}}B_{l^{\prime}m^{\prime}}^{k^{\prime}}.\tag{14.24}
Substituting the transformation rules for AjiA_{j}^{i} and BlmkB_{lm}^{k}, we find
Cjlmik=AjiJiiJjjBlmkJkkJllJmm.(14.25)C_{j^{\prime}l^{\prime}m^{\prime}}^{i^{\prime}k^{\prime}}=A_{j}^{i} J_{i}^{i^{\prime}}J_{j^{\prime}}^{j}B_{lm}^{k}J_{k}^{k^{\prime}}J_{l^{\prime} }^{l}J_{m^{\prime}}^{m}.\tag{14.25}
Rearranging the multiplicative terms yields the desired identity
Cjlmik=CjlmikJiiJjjJkkJllJmm.(14.23)C_{j^{\prime}l^{\prime}m^{\prime}}^{i^{\prime}k^{\prime}}=C_{jlm}^{ik} J_{i}^{i^{\prime}}J_{j^{\prime}}^{j}J_{k}^{k^{\prime}}J_{l^{\prime}} ^{l}J_{m^{\prime}}^{m}. \tag{14.23}
Note that the product property applies to the dot product as well as to the ordinary tensor product. Demonstrating this fact is left as an exercise.
According to the contraction property, a contraction of a tensor is a tensor in its own right. Earlier, we demonstrated a special case of this property when we showed that the contraction TiiT_{i}^{i} of a second-order tensor TjiT_{j}^{i} is an invariant, i.e. a tensor of order zero. At the heart of the proof was the reciprocal relationship between the Jacobians JiiJ_{i}^{i^{\prime}} and JiiJ_{i^{\prime}}^{i}. This result needs to be generalized only slightly to show that a contraction of any tensor is a tensor in its own right.
Once again, the proof proceeds by considering a tensor with a representative collection of indices. Suppose that SjkiS_{jk}^{i} is a tensor, i.e.
Sjki=SjkiJiiJjjJkk,(14.26)S_{j^{\prime}k^{\prime}}^{i^{\prime}}=S_{jk}^{i}J_{i}^{i^{\prime}} J_{j^{\prime}}^{j}J_{k^{\prime}}^{k},\tag{14.26}
and let TkT_{k} be the result of contracting SjkiS_{jk}^{i} on ii and jj:
Tk=Siki.(14.27)T_{k}=S_{ik}^{i}.\tag{14.27}
Or goal is to show that TkT_{k} is a tensor, i.e.
Tk=TkJkk.(14.28)T_{k^{\prime}}=T_{k}J_{k^{\prime}}^{k}.\tag{14.28}
To this end, contract both sides of the identity
Sjki=SjkiJiiJjjJkk,(14.26)S_{j^{\prime}k^{\prime}}^{i^{\prime}}=S_{jk}^{i}J_{i}^{i^{\prime}} J_{j^{\prime}}^{j}J_{k^{\prime}}^{k}, \tag{14.26}
on ii^{\prime} and jj^{\prime}, i.e. replace jj^{\prime} with ii^{\prime} on both sides of the equation:
Siki=SjkiJiiJijJkk.(14.29)S_{i^{\prime}k^{\prime}}^{i^{\prime}}=S_{jk}^{i}J_{i}^{i^{\prime}} J_{i^{\prime}}^{j}J_{k^{\prime}}^{k}.\tag{14.29}
Note that the first two Jacobians on the right are contracted on ii^{\prime}. Since JiiJij=δijJ_{i}^{i^{\prime}}J_{i^{\prime}}^{j}=\delta_{i}^{j} and Sjkiδij=SikiS_{jk} ^{i}\delta_{i}^{j}=S_{ik}^{i}, we find
Siki=SikiJkk.(14.30)S_{i^{\prime}k^{\prime}}^{i^{\prime}}=S_{ik}^{i}J_{k^{\prime}}^{k}.\tag{14.30}
Since Siki=TkS_{i^{\prime}k^{\prime}}^{i^{\prime}}=T_{k^{\prime}} and Siki=TkS_{ik} ^{i}=T_{k}, we have
Tk=TkJkk,(14.28)T_{k^{\prime}}=T_{k}J_{k^{\prime}}^{k}, \tag{14.28}
as we set out to prove.
As we have already discussed, the implications of the contraction property are far reaching. When we are able to contract away all the indices of a tensor, the result is a tensor of order zero, i.e. an invariant. For example, if SklijS_{kl}^{ij} is a tensor, then SijijS_{ij}^{ij} is an invariant. The quantity SjiijS_{ji}^{ij} is also an invariant and, of course, it is very much different from SijijS_{ij}^{ij}.
In Chapter 11, we introduced index juggling, a notational device that allows us to use the same letter for variants related by contraction with a metric tensor. Since the metric tensors are, in fact, tensors, index juggling preserves the tensor property by a combination of the product and contraction properties. Consistent with its name, index juggling invariably changes the placement of the index, e.g.
Ti=ZijTj(14.31)T^{i}=Z^{ij}T_{j}\tag{14.31}
and therefore the type of the tensor. More precisely, the result of performing index juggling on a tensor of the type indicated by its indicial signature is a tensor in its own right of the type indicated by the resulting signature.
It is tempting to think of the Kronecker delta δji\delta_{j}^{i} as an invariant since it has the same values (namely, 11 if i=ji=j and 00 otherwise) in all coordinates. However, labeling it an invariant is not a valid application of that concept since the Kronecker delta is not a variant of order zero. It is a variant of order two and therefore a more relevant question is whether it is a tensor. As a matter of fact, this question is essential since the Kronecker delta will continue to appear frequently in our analysis. And, since we are making a commitment to operate exclusively with tensors, we must make sure that the Kronecker delta is one.
By definition, the Kronecker delta is a tensor if the identity
δji=δjiJiiJjj(14.32)\delta_{j^{\prime}}^{i^{\prime}}=\delta_{j}^{i}J_{i}^{i^{\prime}}J_{j^{\prime }}^{j}\tag{14.32}
is valid. Since δjiJii=Jji\delta_{j}^{i}J_{i}^{i^{\prime}}=J_{j}^{i^{\prime}} and JjiJjj=δjiJ_{j}^{i^{\prime}}J_{j^{\prime}}^{j}=\delta_{j^{\prime}}^{i^{\prime}}, the two sides of the equation match and thus the above identity is confirmed. We therefore conclude that the Kronecker delta is indeed a tensor.
An alternative derivation is based on the identity
δji=ZiZj.(9.93)\delta_{j}^{i}=\mathbf{Z}^{i}\cdot\mathbf{Z}_{j}. \tag{9.93}
Since the contravariant and the covariant bases are tensors, the tensor property of the Kronecker delta follows from the product property.
There are number of interesting invariants involving the Kronecker delta. The simplest one is
δii(14.33)\delta_{i}^{i}\tag{14.33}
which, of course, represents the dimension nn of the relevant Euclidean space. Other invariants involve a second-order tensor AjiA_{j}^{i}. For example,
δijAji=Aii,(14.34)\delta_{i}^{j}A_{j}^{i}=A_{i}^{i},\tag{14.34}
which, as we have already mentioned, is referred to as the trace of AjiA_{j}^{i} by analogy with matrices. Another interesting invariant is
(δikδjlδilδjk)AkiAlj(14.35)\left( \delta_{i}^{k}\delta_{j}^{l}-\delta_{i}^{l}\delta_{j}^{k}\right) A_{k}^{i}A_{l}^{j}\tag{14.35}
and yet another is
(δirδjsδkt+δitδjrδks+δisδjtδkrδirδjtδksδisδjrδktδitδjsδkr)AriAsjAtk.(14.36)\left( \delta_{i}^{r}\delta_{j}^{s}\delta_{k}^{t}+\delta_{i}^{t}\delta _{j}^{r}\delta_{k}^{s}+\delta_{i}^{s}\delta_{j}^{t}\delta_{k}^{r}-\delta _{i}^{r}\delta_{j}^{t}\delta_{k}^{s}-\delta_{i}^{s}\delta_{j}^{r}\delta _{k}^{t}-\delta_{i}^{t}\delta_{j}^{s}\delta_{k}^{r}\right) A_{r}^{i}A_{s} ^{j}A_{t}^{k}.\tag{14.36}
Note, also, that the quantities in parentheses are tensors. These expressions exhibit patterns associated with the determinant, an important concept whose tensor treatment is the subject of Chapter 16.
Suppose that FF is an invariant. It may be a vector field, such as the position vector R\mathbf{R}, or a scalar field, such as a temperature distribution TT. Then the collection of partial derivatives
F(Z)Zi(14.37)\frac{\partial F\left( Z\right) }{\partial Z^{i}}\tag{14.37}
is a first-order covariant tensor. The symbol
iF=F(Z)Zi.(10.125)\nabla_{i}F=\frac{\partial F\left( Z\right) }{\partial Z^{i}}. \tag{10.125}
introduced earlier clearly treats the index as a subscript which is now warranted given the covariant tensor nature of the resulting object. Note that the symbol i\nabla_{i}, as defined above, should not be applied to tensors of order greater than zero since, as we shall see, the result is not necessarily a tensor. This conundrum is fixed by introducing the concept of the covariant derivative which the symbol i\nabla_{i} will eventually denote.
The demonstration of the tensor property of iF\nabla_{i}F is essentially a repeat of the calculation that proved the tensor property of the covariant basis Zi\mathbf{Z}_{i}. This is not surprising since Zi\mathbf{Z}_{i} is a special case of iF\nabla_{i}F for F=RF=\mathbf{R}:
Zi=iR.(9.9)\mathbf{Z}_{i}=\nabla_{i}\mathbf{R.} \tag{9.9}
The variant iF\nabla_{i^{\prime}}F in the primed coordinate system ZiZ^{i^{\prime}} is given by
iF=F(Z)Zi.(14.38)\nabla_{i^{\prime}}F=\frac{\partial F\left( Z^{\prime}\right) }{\partial Z^{i^{\prime}}}.\tag{14.38}
Represent F(Z)F\left( Z^{\prime}\right) as a composition of F(Z)F\left( Z\right) with the equations of coordinate transformation
Zi=Zi(Z),(13.22)Z^{i}=Z^{i}\left( Z^{\prime}\right) , \tag{13.22}
i.e.
F(Z)=F(Z(Z)).(14.39)F\left( Z^{\prime}\right) =F\left( Z\left( Z^{\prime}\right) \right) .\tag{14.39}
Differentiating this identity with respect to ZiZ^{i^{\prime}} yields
F(Z)Zi=F(Z)ZiZi(Z)Zi.(14.40)\frac{\partial F\left( Z^{\prime}\right) }{\partial Z^{i^{\prime}}} =\frac{\partial F\left( Z\right) }{\partial Z^{i}}\frac{\partial Z^{i}\left( Z^{\prime}\right) }{\partial Z^{i^{\prime}}}.\tag{14.40}
Since
F(Z)Zi=iF,     F(Z)Zi=iF,     and     Zi(Z)Zi=Jii,(14.41)\frac{\partial F\left( Z\right) }{\partial Z^{i^{\prime}}}=\nabla _{i^{\prime}}F\text{, \ \ \ \ }\frac{\partial F\left( Z\right) }{\partial Z^{i}}=\nabla_{i}F\text{, \ \ \ \ and \ \ \ \ }\frac{\partial Z^{i}\left( Z^{\prime}\right) }{\partial Z^{i^{\prime}}}=J_{i^{\prime}}^{i},\tag{14.41}
we find
iF=iF Jii,(14.42)\nabla_{i^{\prime}}F=\nabla_{i}F~J_{i^{\prime}}^{i},\tag{14.42}
as we set out to prove.
The fact that iF\nabla_{i}F is a tensor gives new insight into the coordinate expression
F=iF Zi(14.43)\mathbf{\nabla}F=\nabla_{i}F\ \mathbf{Z}^{i}\tag{14.43}
for the gradient F\mathbf{\nabla}F of a scalar field FF. Since iF\nabla_{i}F and Zi\mathbf{Z}^{i} are tensors, the contraction iF Zi\nabla_{i}F\ \mathbf{Z} ^{i} is an invariant.
Consider a tensor Ukij\mathbf{U}_{k}^{ij} with vector elements. Then the contravariant components UkijlU_{k}^{ijl} of Ukij\mathbf{U}_{k}^{ij}, i.e.
Ukij=UkijlZl,(14.44)\mathbf{U}_{k}^{ij}=U_{k}^{ijl}\mathbf{Z}_{l},\tag{14.44}
and the covariant components UklijU_{kl}^{ij}, i.e.
Ukij=UklijZl,(14.45)\mathbf{U}_{k}^{ij}=U_{kl}^{ij}\mathbf{Z}^{l},\tag{14.45}
are tensors in their own right.
This can be demonstrated in a number of ways. The shortest way is to give explicit expressions for the components with the help of the dot product:
Ukijl=ZlUkij          (14.46)Uklij=ZlUkij.          (14.47)\begin{aligned}U_{k}^{ijl} & =\mathbf{Z}^{l}\cdot\mathbf{U}_{k}^{ij}\ \ \ \ \ \ \ \ \ \ \left(14.46\right)\\U_{kl}^{ij} & =\mathbf{Z}_{l}\cdot\mathbf{U}_{k}^{ij}.\ \ \ \ \ \ \ \ \ \ \left(14.47\right)\end{aligned}
The tensor property of UkijlU_{k}^{ijl} and UklijU_{kl}^{ij} then follows from the product property of tensors.
Let us also give another proof that avoids the dot product. When we substitute the transformation rules
Ukij=UkijJiiJjjJkk(14.48)\mathbf{U}_{k}^{ij}=\mathbf{U}_{k^{\prime}}^{i^{\prime}j^{\prime}} J_{i^{\prime}}^{i}J_{j^{\prime}}^{j}J_{k}^{k^{\prime}}\tag{14.48}
and
Zl=ZlJll(13.34)\mathbf{Z}_{l}=\mathbf{Z}_{l^{\prime}}J_{l}^{l^{\prime}} \tag{13.34}
into the identity
Ukij=UkijlZl,(14.44)\mathbf{U}_{k}^{ij}=U_{k}^{ijl}\mathbf{Z}_{l}, \tag{14.44}
we find
UkijJiiJjjJkk=UkijlZlJll.(14.49)\mathbf{U}_{k^{\prime}}^{i^{\prime}j^{\prime}}J_{i^{\prime}}^{i}J_{j^{\prime} }^{j}J_{k}^{k^{\prime}}=U_{k}^{ijl}\mathbf{Z}_{l^{\prime}}J_{l}^{l^{\prime}}.\tag{14.49}
Inverting the three Jacobians on the left yields
Ukij=UkijlJiiJjjJkkJllZl.(14.50)\mathbf{U}_{k^{\prime}}^{i^{\prime}j^{\prime}}=U_{k}^{ijl}J_{i}^{i^{\prime} }J_{j}^{j^{\prime}}J_{k^{\prime}}^{k}J_{l}^{l^{\prime}}\mathbf{Z}_{l^{\prime} }.\tag{14.50}
Since this identity represents the decomposition of Ukij\mathbf{U}_{k^{\prime} }^{i^{\prime}j^{\prime}} with respect to Zl\mathbf{Z}_{l^{\prime}}, the quantity multiplying Zl\mathbf{Z}_{l^{\prime}} on the right must be the component UkijlU_{k^{\prime}}^{i^{\prime}j^{\prime}l^{\prime}}, i.e.
Ukijl=UkijlJiiJjjJkkJll,(14.51)U_{k^{\prime}}^{i^{\prime}j^{\prime}l^{\prime}}=U_{k}^{ijl}J_{i}^{i^{\prime} }J_{j}^{j^{\prime}}J_{k^{\prime}}^{k}J_{l}^{l^{\prime}},\tag{14.51}
as we set out to prove. It is left as an exercise to show that the covariant component UklijU_{kl}^{ij} is also a tensor.
The fact that the components of tensors with vector elements are tensors offers an important insight into a topic considered earlier in Chapters 10 and 12. Recall our discussions of the motion of a material particle given by the equations
Zi=Zi(t).(10.36)Z^{i}=Z^{i}\left( t\right) . \tag{10.36}
We demonstrated that the components ViV^{i} of the velocity vector V\mathbf{V} are given by
Vi(t)=dZi(t)dt.(10.37)V^{i}\left( t\right) =\frac{dZ^{i}\left( t\right) }{dt}. \tag{10.37}
Due to the tensor property of the components of a vector, we can now conclude that ViV^{i} is a tensor.
Furthermore, we demonstrated that the components AiA^{i} of the acceleration vector A\mathbf{A} are given by the following expression
Ai=dVidt+ΓjkiVjVk.(12.80)A^{i}=\frac{dV^{i}}{dt}+\Gamma_{jk}^{i}V^{j}V^{k}. \tag{12.80}
Thus, we can conclude that the combination on the right is a tensor. This is a very intriguing realization since the Christoffel symbol, which transforms according to the rule
Γijk=ΓijkJiiJjjJkk+JijkJkk,(13.127)\Gamma_{i^{\prime}j^{\prime}}^{k^{\prime}}=\Gamma_{ij}^{k}J_{i^{\prime}} ^{i}J_{j^{\prime}}^{j}J_{k}^{k^{\prime}}+J_{i^{\prime}j^{\prime}}^{k} J_{k}^{k^{\prime}}, \tag{13.127}
is not a tensor. It is also easy to show, and is left as an exercise, that dVi/dtdV^{i}/dt is not a tensor. Yet, the combination
dVidt+ΓjkiVjVk(14.52)\frac{dV^{i}}{dt}+\Gamma_{jk}^{i}V^{j}V^{k}\tag{14.52}
is a tensor. This observation foreshadows two important developments. The first is the recognition that differentiation is not a tensor-preserving operation, i.e. the result of differentiating a tensor is not a tensor. This fact will be discussed below in this Chapter. The second is the counterbalancing insight that a proper corrective term involving the Christoffel symbol can restore the tensor property. This is the main idea behind the operator of the covariant derivative and, in fact, of numerous other novel differential operators in Tensor Calculus and the Calculus of Moving Surfaces.
Any definition of a new concept should be scrutinized against the possibility of internal contradictions. Suppose, for a moment, that the definition of the tensor property of a variant TjiT_{j^{\prime}}^{i^{\prime}} reads
Tji=2TjiJiiJjj.(14.53)T_{j^{\prime}}^{i^{\prime}}=2T_{j}^{i}J_{i}^{i^{\prime}}J_{j^{\prime}}^{j}.\tag{14.53}
Although it is similar to the actual definition Tji=TjiJiiJjjT_{j^{\prime}}^{i^{\prime} }=T_{j}^{i}J_{i}^{i^{\prime}}J_{j^{\prime}}^{j}, the hypothetical definition above would immediately give rise to several internal contradictions.
For instance, what would be the implications of this definition if ZiZ^{i} and ZiZ^{i^{\prime}} happen to be the same coordinate system? Or, to put it another way, if the equations of coordinate transformation were
Z1(Z)=Z1          (14.54)Z2(Z)=Z2          (14.55)Z3(Z)=Z3.          (14.56)\begin{aligned}Z^{1^{\prime}}\left( Z\right) & =Z^{1}\ \ \ \ \ \ \ \ \ \ \left(14.54\right)\\Z^{2^{\prime}}\left( Z\right) & =Z^{2}\ \ \ \ \ \ \ \ \ \ \left(14.55\right)\\Z^{3^{\prime}}\left( Z\right) & =Z^{3}.\ \ \ \ \ \ \ \ \ \ \left(14.56\right)\end{aligned}
Under this scenario, the Jacobians correspond to the identity matrix and this implies that the elements of TjiT_{j^{\prime}}^{i^{\prime}} equal twice their value, i.e.
Tji=2Tji,(14.57)T_{j^{\prime}}^{i^{\prime}}=2T_{j^{\prime}}^{i^{\prime}},\tag{14.57}
which is a glaring contradiction for any nonzero variant TjiT_{j}^{i}.
Furthermore, the hypothetical definition would not hold up when the roles of the coordinate systems ZiZ^{i} and ZiZ^{i^{\prime}} are reversed. According to the hypothetical definition, we must have
Tji=2TjiJiiJjj.(14.58)T_{j}^{i}=2T_{j^{\prime}}^{i^{\prime}}J_{i^{\prime}}^{i}J_{j^{\prime}}^{j}.\tag{14.58}
On the other hand, inverting the Jacobians in equation
Tji=2TjiJiiJjj,(14.53)T_{j^{\prime}}^{i^{\prime}}=2T_{j}^{i}J_{i}^{i^{\prime}}J_{j^{\prime}}^{j}, \tag{14.53}
yields
Tji=12TjiJiiJjj.(14.59)T_{j}^{i}=\frac{1}{2}T_{j^{\prime}}^{i^{\prime}}J_{i^{\prime}}^{i} J_{j^{\prime}}^{j}.\tag{14.59}
This, once again, is an internal contradiction.
Finally, a third contradiction would be reached if we considered three coordinate systems ZiZ^{i}, ZiZ^{i^{\prime}}, and ZiZ^{i^{\prime\prime}} and compared the transformation of a variant TjiT_{j}^{i} from ZiZ^{i} to ZiZ^{i^{\prime}} and then from ZiZ^{i^{\prime}} to ZiZ^{i^{\prime\prime}} to the transformation directly from ZiZ^{i} to ZiZ^{i^{\prime\prime}}.
Of course, the actual definition of a tensor
Tji=TjiJiiJjj,(14.1)T_{j^{\prime}}^{i^{\prime}}=T_{j}^{i}J_{i}^{i^{\prime}}J_{j^{\prime}}^{j}, \tag{14.1}
does not suffer from these contradictions. It is reflexive, meaning that it is valid when ZiZ^{i} and ZiZ^{i^{\prime}} are the same coordinate systems. It is symmetric, meaning that it consistently applies when roles of the coordinate systems ZiZ^{i} and ZiZ^{i^{\prime}} are reversed. The symmetric property of tensors was effectively demonstrated in Section 13.3.5. Finally, it is transitive, meaning that transformations among three or more coordinate systems are consistent. Specifically, the transitive property states that the equations
Tji=TjiJiiJjj   and   Tji=TjiJiiJjj(14.60)T_{j^{\prime}}^{i^{\prime}}=T_{j}^{i}J_{i}^{i^{\prime}}J_{j^{\prime}} ^{j}\text{ \ \ and \ \ }T_{j^{\prime\prime}}^{i^{\prime\prime}}=T_{j^{\prime} }^{i^{\prime}}J_{i^{\prime}}^{i^{\prime\prime}}J_{j^{\prime\prime}} ^{j^{\prime}}\tag{14.60}
imply
Tji=TjiJiiJjj.(14.61)T_{j^{\prime\prime}}^{i^{\prime\prime}}=T_{j}^{i}J_{i}^{i^{\prime\prime} }J_{j^{\prime\prime}}^{j}.\tag{14.61}
The proof of the transitive property is left as an exercise.
All coordinate space variants that we have encountered so far have arisen naturally as systems associated, in one way or another, with geometric objects in the Euclidean space. However, there also exists an important and rather straightforward way of artificially constructing tensors, known as synthetic tensors, that are not associated with any geometric object.
The method works the same way for a tensor of any order. Thus, for the sake of specificity, we will illustrate it with the help of a fourth-order tensor AjkliA_{jkl}^{i}. In some unprimed coordinates ZiZ^{i} let AjkliA_{jkl}^{i} have arbitrary values. Then, in all other coordinate systems ZiZ^{i^{\prime}}, define the values of AjkliA_{j^{\prime}k^{\prime }l^{\prime}}^{i^{\prime}} to be
Ajkli=AjkliJiiJjjJkkJll.(14.62)A_{j^{\prime}k^{\prime}l^{\prime}}^{i^{\prime}}=A_{jkl}^{i}J_{i}^{i^{\prime} }J_{j^{\prime}}^{j}J_{k^{\prime}}^{k}J_{l^{\prime}}^{l}.\tag{14.62}
In order to confirm that the resulting object is a tensor, we must show that for two coordinate systems ZiZ^{i^{\prime}} and ZiZ^{i^{\prime\prime}}, the systems AjkliA_{j^{\prime}k^{\prime}l^{\prime}}^{i^{\prime}} and AjkliA_{j^{\prime \prime}k^{\prime\prime}l^{\prime\prime}}^{i^{\prime\prime}} are related by
Ajkli=AjkliJiiJjjJkkJll.(14.63)A_{j^{\prime\prime}k^{\prime\prime}l^{\prime\prime}}^{i^{\prime\prime} }=A_{j^{\prime}k^{\prime}l^{\prime}}^{i^{\prime}}J_{i^{\prime}}^{i^{\prime \prime}}J_{j^{\prime\prime}}^{j^{\prime}}J_{k^{\prime\prime}}^{k^{\prime} }J_{l^{\prime\prime}}^{l^{\prime}}.\tag{14.63}
This, of course, is precisely the transitive property of the tensor discussed in the previous Section. Even though the present circumstances are slightly different in that the transformation rule is stipulated as a matter of definition, the demonstration of this property would be exactly the same.
This construction of synthetic tensors is somewhat un-tensorlike in that it singles out one coordinate system, ZiZ^{i}, relative to others. Nevertheless, a tensor is a tensor whether it arose from a geometric object or not. The significance of synthetic tensors is that any indexed collection of values in any coordinate system is a manifestation of some tensor. In fact, there is a perfect one-to-one correspondence between tensors and the set of all possible indexed collection of values in a given coordinate system. Synthetic tensors have a number of applications, including one found in the next Section in the proof of the quotient theorem.
We have already established a number of ways in which tensors are formed: by differentiation of invariants, by decomposing vectors with the respect to either the covariant or the contravariant basis, by a synthetic procedure described in the previous Section, and, subsequently, by addition, multiplication, and contraction of other tensors. The quotient theorem provides an additional useful way of recognizing that a particular variant is a tensor. As the name suggests, the quotient theorem is, in a way, the converse of the product property.
Suppose that UjkiU_{jk}^{i} and VklV^{kl} are tensors. Then, by the combination of the product and contraction properties, the variant
Wjil=UjkiVkl(14.64)W_{j}^{il}=U_{jk}^{i}V^{kl}\tag{14.64}
is a tensor in its own right. However, what if the circumstances are such that it is known that WjilW_{j}^{il} and VklV^{kl} are tensors. Can we conclude that UjkiU_{jk}^{i} is a tensor?
If no further stipulations are added, the answer is certainly no. For example, if both VklV^{kl} and WjilW_{j}^{il} are identically zero in all coordinate systems, then the above identity holds for any variant UjkiU_{jk}^{i} and thus we cannot conclude that UjkiU_{jk}^{i} is a tensor. In order to be able to conclude that UjkiU_{jk}^{i} is a tensor, the combination UjkiVklU_{jk}^{i}V^{kl} must be a tensor for all tensors VklV^{kl}. That is the content of the quotient theorem. In general, the quotient theorem states that if three variants UU_{\cdot\cdot}^{\cdot\cdot}, VV_{\cdot\cdot} ^{\cdot\cdot}, and WW_{\cdot\cdot}^{\cdot\cdot} are related by the equation
W=UV ,(14.65)W_{\cdot\cdot}^{\cdot\cdot}=U_{\cdot\cdot}^{\cdot\cdot}V_{\cdot\cdot} ^{\cdot\cdot}\ ,\tag{14.65}
that satisfies all rules of the tensor notation, where the dots indicate otherwise arbitrary indicial signatures and any number of valid contractions on the right, and if for any tensor VV_{\cdot\cdot}^{\cdot\cdot} the result WW_{\cdot\cdot}^{\cdot\cdot} is also a tensor, then UU_{\cdot\cdot} ^{\cdot\cdot} is a tensor in its own right.
Before we turn to the proof, let us mention an analogous problem in Linear Algebra that will serve as a structural blueprint for an argument at the end of the proof. If AA is an n×mn\times m matrix and xx is a vector in the sense of Matrix Algebra (i.e. an m×1m\times1 matrix), under what conditions does the equality Ax=0Ax=0 imply that A=0A=0? It is certainly not enough for Ax=0Ax=0 to hold for some xx: if x=0x=0 then Ax=0Ax=0 holds for all matrices AA. Requiring that xx is nonzero is also not sufficient since for any nonzero vector xx, the equality Ax=0Ax=0 holds for any matrix AA that contains the vector xx in its null space. In fact, Ax=0Ax=0 can hold for all xx in an (m1)\left( m-1\right) -dimensional subspace of Rm \mathbb{R} ^{m}, and AA could still be a nonzero, albeit rank-11, matrix. For instance,
[11111222223333344444]x=[0000](14.66)\left[ \begin{array} {ccccc} 1 & 1 & 1 & 1 & 1\\ 2 & 2 & 2 & 2 & 2\\ 3 & 3 & 3 & 3 & 3\\ 4 & 4 & 4 & 4 & 4 \end{array} \right] x=\left[ \begin{array} {c} 0\\ 0\\ 0\\ 0 \end{array} \right]\tag{14.66}
for all xx of the form
x=α[11000]+β[10100]+γ[10010]+δ[10001].(14.67)x=\alpha\left[ \begin{array} {r} 1\\ -1\\ 0\\ 0\\ 0 \end{array} \right] +\beta\left[ \begin{array} {r} 1\\ 0\\ -1\\ 0\\ 0 \end{array} \right] +\gamma\left[ \begin{array} {r} 1\\ 0\\ 0\\ -1\\ 0 \end{array} \right] +\delta\left[ \begin{array} {r} 1\\ 0\\ 0\\ 0\\ -1 \end{array} \right] .\tag{14.67}
Clearly, such vectors xx represents a 44-dimensional subspace of R5 \mathbb{R} ^{5}.
And so we come to the realization that, in order for Ax=0Ax=0 to imply A=0A=0, it must hold for all xx. To prove that that is sufficient, take advantage of the freedom to choose any xx and consider
x1=[100].(14.68)x_{1}=\left[ \begin{array} {c} 1\\ 0\\ \cdots\\ 0 \end{array} \right] .\tag{14.68}
Then Ax1Ax_{1} is the first column of AA and, since Ax1=0Ax_{1}=0, we conclude that the first column of AA is zero. Similarly, the vector
x2=[010](14.69)x_{2}=\left[ \begin{array} {c} 0\\ 1\\ \cdots\\ 0 \end{array} \right]\tag{14.69}
proves that the second column of AA is zero, and so on. This completes the proof.
We can now return to the proof of the quotient theorem based on our representative example. In the primed coordinate system, the identity
Wjil=UjkiVkl(14.64)W_{j}^{il}=U_{jk}^{i}V^{kl} \tag{14.64}
reads
Wjil=UjkiVkl.(14.70)W_{j^{\prime}}^{i^{\prime}l^{\prime}}=U_{j^{\prime}k^{\prime}}^{i^{\prime} }V^{k^{\prime}l^{\prime}}.\tag{14.70}
Substituting the relationships
Wjil=WjilJiiJllJjj    and    Vkl=VklJkkJll(14.71)W_{j^{\prime}}^{i^{\prime}l^{\prime}}=W_{j}^{il}J_{i}^{i^{\prime}} J_{l}^{l^{\prime}}J_{j^{\prime}}^{j}\text{ \ \ \ and \ \ \ }V^{k^{\prime }l^{\prime}}=V^{kl}J_{k}^{k^{\prime}}J_{l}^{l^{\prime}}\tag{14.71}
into the primed identity, we find
WjilJiiJllJjj=UjkiVklJkkJll.(14.72)W_{j}^{il}J_{i}^{i^{\prime}}J_{l}^{l^{\prime}}J_{j^{\prime}}^{j}=U_{j^{\prime }k^{\prime}}^{i^{\prime}}V^{kl}J_{k}^{k^{\prime}}J_{l}^{l^{\prime}}.\tag{14.72}
Invert each of the Jacobians on the left in order to isolate WjilW_{j}^{il}:
Wjil=UjkiVklJiiJkkJjj,(14.73)W_{j}^{il}=U_{j^{\prime}k^{\prime}}^{i^{\prime}}V^{kl}J_{i^{\prime}}^{i} J_{k}^{k^{\prime}}J_{j}^{j^{\prime}},\tag{14.73}
Subtracting this from the initial identity
Wjil=UjkiVkl,(14.64)W_{j}^{il}=U_{jk}^{i}V^{kl}, \tag{14.64}
we find
0=(UjkiJiiJjjJkkUjki)Vkl.(14.74)0=\left( U_{j^{\prime}k^{\prime}}^{i^{\prime}}J_{i^{\prime}}^{i} J_{j}^{j^{\prime}}J_{k}^{k^{\prime}}-U_{jk}^{i}\right) V^{kl}.\tag{14.74}
Since all the preceding identities are valid for any VklV^{kl}, so is this one. By the argument that we developed for the Linear Algebra problem above, in combination with the fact that any combination of values corresponds to some -- possibly synthetic -- tensor, we can conclude that the quantity in parentheses vanishes. In other words,
Ujki=UjkiJiiJjjJkk,(14.75)U_{j^{\prime}k^{\prime}}^{i^{\prime}}=U_{jk}^{i}J_{i^{\prime}}^{i} J_{j}^{j^{\prime}}J_{k}^{k^{\prime}},\tag{14.75}
as we set out to prove.
In defining the tensor property, we insisted that the variant transforms according to a specific rule under all coordinate transformations. In many situations, however, it is appropriate to consider a limited class of transformations. In those situations, we may classify a variant as a tensor if it transforms according to the classical tensor rule under that particular class of transformations. As a result, a variant that is not considered a tensor in the broad sense may be, in fact, considered a tensor in the narrow sense associated with the limited class of transformations. Such a variant is referred to as a qualified tensor.
For example, we may limit our attention to linear changes of variables, i.e.
Zi(Z)=AiiZi,(14.76)Z^{i^{\prime}}\left( Z\right) =A_{i}^{i^{\prime}}Z^{i},\tag{14.76}
where AiiA_{i}^{i^{\prime}} is a constant second-order system. Then,
Jii=Aii(14.77)J_{i}^{i^{\prime}}=A_{i}^{i^{\prime}}\tag{14.77}
and therefore
Jiji=0.(14.78)J_{ij}^{i^{\prime}}=0.\tag{14.78}
Since the second-order Jacobians are related by
Jjki+JjkiJiiJjjJkk=0.(13.115)J_{j^{\prime}k^{\prime}}^{i}+J_{jk}^{i^{\prime}}J_{i^{\prime}}^{i} J_{j^{\prime}}^{j}J_{k^{\prime}}^{k}=0. \tag{13.115}
we also have
Jjki=0.(14.79)J_{j^{\prime}k^{\prime}}^{i}=0.\tag{14.79}
As a result, the Christoffel symbol Γijk\Gamma_{ij}^{k}, whose general transformation rule reads
Γijk=ΓijkJiiJjjJkk+JijkJkk,(13.127)\Gamma_{i^{\prime}j^{\prime}}^{k^{\prime}}=\Gamma_{ij}^{k}J_{i^{\prime}} ^{i}J_{j^{\prime}}^{j}J_{k}^{k^{\prime}}+J_{i^{\prime}j^{\prime}}^{k} J_{k}^{k^{\prime}}, \tag{13.127}
transforms according to
Γijk=ΓijkJiiJjjJkk(14.80)\Gamma_{i^{\prime}j^{\prime}}^{k^{\prime}}=\Gamma_{ij}^{k}J_{i^{\prime}} ^{i}J_{j^{\prime}}^{j}J_{k}^{k^{\prime}}\tag{14.80}
under linear changes of coordinates. Therefore, it is a tensor with respect to linear changes of coordinates.
Another class of transformations that is frequently considered is orientation-preserving transformations. Such transformations will play an important role in the discussion of the Levi-Civita symbol introduced in Chapter 17.
In the previous Chapter, we showed that the Christoffel symbol Γijk\Gamma _{ij}^{k} transforms according to the rule
Γijk=ΓijkJiiJjjJkk+JijkJkk(13.127)\Gamma_{i^{\prime}j^{\prime}}^{k^{\prime}}=\Gamma_{ij}^{k}J_{i^{\prime}} ^{i}J_{j^{\prime}}^{j}J_{k}^{k^{\prime}}+J_{i^{\prime}j^{\prime}}^{k} J_{k}^{k^{\prime}} \tag{13.127}
and is therefore not a tensor, owing to the additional term on the right. Recall that in order to derive the transformation rule for Γijk\Gamma_{ij}^{k}, we first sought the transformation rule for
Γij=ZiZj(12.2)\mathbf{\Gamma}_{ij}=\frac{\partial\mathbf{Z}_{i}}{\partial Z^{j}} \tag{12.2}
and discovered that
Γij=ΓijJiiJjj+ZiJiji.(13.122)\mathbf{\Gamma}_{i^{\prime}j^{\prime}}=\mathbf{\Gamma}_{ij}J_{i^{\prime}} ^{i}J_{j^{\prime}}^{j}+\mathbf{Z}_{i}J_{i^{\prime}j^{\prime}}^{i}. \tag{13.122}
Thus, Γij\mathbf{\Gamma}_{ij} proved to be the first non-tensor variant in our narrative. And, as a matter of fact, it reveals a common reason for the loss of the tensor property: differentiation.
Generally speaking, the derivative of a tensor of order greater than zero is not a tensor. Let us illustrate this loss of the tensor property for the derivative of a first-order contravariant tensor TiT^{i}. In Chapter 13, we demonstrated the same for the derivative of the covariant basis Zi\mathbf{Z}_{i}. Not surprisingly, the following demonstration will be based on a similar calculation.
Consider the variant
Ti(Z)Zj,(14.81)\frac{\partial T^{i}\left( Z\right) }{\partial Z^{j}},\tag{14.81}
i.e. the collection of the partial derivatives of TiT^{i} with respect to ZjZ^{j}. In the primed coordinates, it corresponds to
Ti(Z)Zj,(14.82)\frac{\partial T^{i^{\prime}}\left( Z^{\prime}\right) }{\partial Z^{j^{\prime}}},\tag{14.82}
i.e. the collection of partial derivatives of TiT^{i^{\prime}} with respect to ZjZ^{j^{\prime}}. The relationship between these two variants can be obtained by differentiating the identity
Ti=TiJii.(14.83)T^{i^{\prime}}=T^{i}J_{i}^{i^{\prime}}.\tag{14.83}
As we have already done on a number of occasions, refer all the elements of the above identity to the primed coordinates ZiZ^{i^{\prime}},
Ti(Z)=Ti(Z(Z))Jii(Z(Z)),(14.84)T^{i^{\prime}}\left( Z^{\prime}\right) =T^{i}\left( Z\left( Z^{\prime }\right) \right) J_{i}^{i^{\prime}}\left( Z\left( Z^{\prime}\right) \right) ,\tag{14.84}
and differentiate both sides with respect to ZjZ^{j^{\prime}}. An application of the product rule yields
Ti(Z)Zj=Ti(Z(Z))ZjJii+TiJii(Z(Z))Zj.(14.85)\frac{\partial T^{i^{\prime}}\left( Z^{\prime}\right) }{\partial Z^{j^{\prime}}}=\frac{\partial T^{i}\left( Z\left( Z^{\prime}\right) \right) }{\partial Z^{j^{\prime}}}J_{i}^{i^{\prime}}+T^{i}\frac{\partial J_{i}^{i^{\prime}}\left( Z\left( Z^{\prime}\right) \right) }{\partial Z^{j^{\prime}}}.\tag{14.85}
A subsequent application of the chain rule yields
Ti(Z)Zj=Ti(Z)ZjZj(Z)ZjJii+TiJii(Z)ZjZj(Z)Zj.(14.86)\frac{\partial T^{i^{\prime}}\left( Z^{\prime}\right) }{\partial Z^{j^{\prime}}}=\frac{\partial T^{i}\left( Z\right) }{\partial Z^{j}} \frac{\partial Z^{j}\left( Z^{\prime}\right) }{\partial Z^{j^{\prime}}} J_{i}^{i^{\prime}}+T^{i}\frac{\partial J_{i}^{i^{\prime}}\left( Z\right) }{\partial Z^{j}}\frac{\partial Z^{j}\left( Z^{\prime}\right) }{\partial Z^{j^{\prime}}}.\tag{14.86}
Finally, since
Zj(Z)Zj=Jjj     and    Jii(Z)Zj=Jiji(14.87)\frac{\partial Z^{j}\left( Z^{\prime}\right) }{\partial Z^{j^{\prime}} }=J_{j^{\prime}}^{j}\text{ \ \ \ \ and \ \ \ }\frac{\partial J_{i}^{i^{\prime }}\left( Z\right) }{\partial Z^{j}}=J_{ij}^{i^{\prime}}\tag{14.87}
we find
TiZj=TiZjJiiJjj+TiJijiJjj.(14.88)\frac{\partial T^{i^{\prime}}}{\partial Z^{j^{\prime}}}=\frac{\partial T^{i} }{\partial Z^{j}}J_{i}^{i^{\prime}}J_{j^{\prime}}^{j}+T^{i}J_{ij}^{i^{\prime} }J_{j^{\prime}}^{j}.\tag{14.88}
As was the case with Γij\mathbf{\Gamma}_{i^{\prime}j^{\prime}} and Γijk\Gamma _{ij}^{k}, the first term on the right,
TiZjJiiJjj,(14.89)\frac{\partial T^{i}}{\partial Z^{j}}J_{i}^{i^{\prime}}J_{j^{\prime}}^{j},\tag{14.89}
is precisely what we would expect of a tensor. However, the presence of the second term TiJijiJjjT^{i}J_{ij}^{i^{\prime}}J_{j^{\prime}}^{j} shows that Ti/Zj\partial T^{i}/\partial Z^{j} is not a tensor.
Similarly, for a covariant tensor TiT_{i}, it is left as an exercise to show that
TiZj=TiZjJiiJjj+TiJiji,(14.90)\frac{\partial T_{i^{\prime}}}{\partial Z^{j^{\prime}}}=\frac{\partial T_{i} }{\partial Z^{j}}J_{i^{\prime}}^{i}J_{j^{\prime}}^{j}+T_{i}J_{i^{\prime }j^{\prime}}^{i},\tag{14.90}
thus demonstrating that Ti/Zj\partial T_{i}/\partial Z^{j} is also not a tensor.
The realization that the derivative of a tensor is not a tensor in its own right is a spoke in the wheel of our emerging tensor framework. Thus, we must immediately direct all of our energies towards regaining control over differentiation. The solution comes in the form of a new differential operator, known as the covariant derivative. It is the subject of the next Chapter.
If I may share a personal preference with regard to the term tensor, I believe that the term proper variant is superior. It is less esoteric, more descriptive, more mellifluous, and uses words that already exist in the English language. Variants that are not tensors can then be described as improper.
As a matter of fact, I believe that the name Tensor Calculus itself helps prevent the subject from attaining the ubiquity that it deserves. It makes the subject sound overly abstract and narrow. The original name given to the subject by its inventors, Gregorio Ricci and Tullio Levi-Civita, was the Absolute Differential Calculus. It is a far better name. The term absolute describes the fact that the framework produces results that are independent of coordinates -- thus, absolute. Here's wishing that the term tensor is retired in favor of proper variant and that the name Absolute Differential Calculus makes a return.
Exercise 14.1Show that for tensors AijA_{ij} and BijB^{ij}, the sum
Aij+Bij(-)A_{ij}+B^{ij} \tag{-}
is not a tensor -- nor is the sum of any two tensors with incompatible indicial signatures. This conclusion serves as a vindication of the rules of the tensor notation described in Chapter 7.
Exercise 14.2Show that if all elements of two tensors are equal in one coordinate system, e.g. Sjki=TjkiS_{jk}^{i}=T_{jk}^{i}, then this is also the case in all coordinate systems.
Exercise 14.3Show that if a tensor AijA_{ij} is symmetric in one coordinate system, i.e.
Aij=Aji,(14.91)A_{ij}=A_{ji},\tag{14.91}
then it is symmetric in all coordinate systems, i.e.
Aij=Aji.(14.92)A_{i^{\prime}j^{\prime}}=A_{j^{\prime}i^{\prime}}.\tag{14.92}
Similarly, if a tensor AijA_{ij} is skew-symmetric in one coordinate system, i.e.
Aij=Aji,(14.93)A_{ij}=-A_{ji},\tag{14.93}
then it is skew-symmetric in all coordinate systems, i.e.
Aij=Aji.(14.94)A_{i^{\prime}j^{\prime}}=-A_{j^{\prime}i^{\prime}}.\tag{14.94}
Exercise 14.4Show that a linear combination of tensors
αAjki+βBjki,(14.95)\alpha A_{jk}^{i}+\beta B_{jk}^{i},\tag{14.95}
where α\alpha and β\beta are scalar fields, is a tensor.
Exercise 14.5Show that the product property extends to the dot product, i.e. the dot product of two tensors with vector elements is a tensor in its own right.
Exercise 14.6In Section 14.11, we showed, without a reference to the dot product, the tensor property of the contravariant components UkijlU_{k}^{ijl} of a tensor Ukij\mathbf{U}_{k}^{ij}. Do the same for the covariant components UklijU_{kl}^{ij} of Ukij\mathbf{U}_{k}^{ij}.
Exercise 14.7Demonstrate the transitive property of tensors in the sense described in Section 14.12.
Exercise 14.8Consider the transformation rule
Aijk=AijkJiiJjjJkk+JijkJkk(14.96)A_{i^{\prime}j^{\prime}}^{k^{\prime}}=A_{ij}^{k}J_{i^{\prime}}^{i} J_{j^{\prime}}^{j}J_{k}^{k^{\prime}}+J_{i^{\prime}j^{\prime}}^{k} J_{k}^{k^{\prime}}\tag{14.96}
which we shall call Christoffel-like since it describes the transformation of the Christoffel symbol Γijk\Gamma_{ij}^{k}, i.e.
Γijk=ΓijkJiiJjjJkk+JijkJkk.(13.127)\Gamma_{i^{\prime}j^{\prime}}^{k^{\prime}}=\Gamma_{ij}^{k}J_{i^{\prime}} ^{i}J_{j^{\prime}}^{j}J_{k}^{k^{\prime}}+J_{i^{\prime}j^{\prime}}^{k} J_{k}^{k^{\prime}}. \tag{13.127}
Show that the Christoffel-like transformation rule is reflexive, symmetric, and transitive in the sense described in Section 14.12.
Exercise 14.9Suppose that AA is a 2×22\times2 matrix corresponding to the system AjiA_{j} ^{i}. Show that
(δkiδljδkjδli)AikAjl=2detA,(14.97)\left( \delta_{k}^{i}\delta_{l}^{j}-\delta_{k}^{j}\delta_{l}^{i}\right) A_{i}^{k}A_{j}^{l}=2\det A,\tag{14.97}
where AA is the matrix corresponding to AjiA_{j}^{i}. Conclude that the determinant of the matrix corresponding to a tensor AjiA_{j}^{i} is an invariant. Clearly, this is not the case of tensors of the form AijA_{ij} and AijA^{ij} as evidenced by the covariant and contravariant metric tensors.
Exercise 14.10Suppose that TiT_{i} is a covariant tensor. Show that
TiZj=TiZjJiiJjj+TiJiji.(14.90)\frac{\partial T_{i^{\prime}}}{\partial Z^{j^{\prime}}}=\frac{\partial T_{i} }{\partial Z^{j}}J_{i^{\prime}}^{i}J_{j^{\prime}}^{j}+T_{i}J_{i^{\prime }j^{\prime}}^{i}. \tag{14.90}
Thus, the derivative of a covariant tensor is not a tensor in its own right.
Exercise 14.11For a tensor TijT_{ij}, show that
TijZk=TijZkJiiJjjJkk+TijJikiJjj+TijJiiJjkj,(14.98)\frac{\partial T_{i^{\prime}j^{\prime}}}{\partial Z^{k^{\prime}}} =\frac{\partial T_{ij}}{\partial Z^{k}}J_{i^{\prime}}^{i}J_{j^{\prime}} ^{j}J_{k^{\prime}}^{k}+T_{ij}J_{i^{\prime}k^{\prime}}^{i}J_{j^{\prime}} ^{j}+T_{ij}J_{i^{\prime}}^{i}J_{j^{\prime}k^{\prime}}^{j},\tag{14.98}
and therefore conclude that the result is not a tensor.
Exercise 14.12Show that if TiT_{i} is a tensor, then the combination
Sij=TiZjTjZi(14.99)S_{ij}=\frac{\partial T_{i}}{\partial Z^{j}}-\frac{\partial T_{j}}{\partial Z^{i}}\tag{14.99}
is also a tensor.
Exercise 14.13For a material particle moving according to the equations of motion
Zi=Zi(t)(10.36)Z^{i}=Z^{i}\left( t\right) \tag{10.36}
derive the transformation rule for the variant
Vi=dZi(t)dt,(10.37)V^{i}=\frac{dZ^{i}\left( t\right) }{dt}, \tag{10.37}
by a direct calculation and show that it is a contravariant tensor.
Exercise 14.14By a direct calculation, derive the transformation rule for the variant
dVidt(14.100)\frac{dV^{i}}{dt}\tag{14.100}
and show that it is not a tensor.
Exercise 14.15Show that the combination
Ai=dVidt+ΓjkiVjVk(12.80)A^{i}=\frac{dV^{i}}{dt}+\Gamma_{jk}^{i}V^{j}V^{k} \tag{12.80}
is a tensor.
Exercise 14.16In fact, for any tensor TiT^{i} defined along the trajectory
Zi=Zi(t)(10.36)Z^{i}=Z^{i}\left( t\right) \tag{10.36}
show that
δTiδt=dTidt+VjΓjkiTk(14.101)\frac{\delta T^{i}}{\delta t}=\frac{dT^{i}}{dt}+V^{j}\Gamma_{jk}^{i}T^{k}\tag{14.101}
is a tensor.
Send feedback to Pavel