Elements of Vector Calculus

Vector Calculus extends the ideas of Calculus to vector fields. As such, one would expect Vector Calculus to be full of remarkable connections and elegant identities. This is, in fact, so -- but only when the right language is chosen for the description. Surprisingly, the language of geometric vectors, which has served us so well in our initial explorations, proves to be ill-equipped at dealing with even some of the most elementary combinations involving vector fields and their derivatives..
To illustrate this point, consider the dot product UV\mathbf{U}\cdot\mathbf{V} of two vector fields U\mathbf{U} and V\mathbf{V}. Since UV\mathbf{U} \cdot\mathbf{V} is a scalar field, it is subject to the gradient operator \mathbf{\nabla}. As a matter of fact, the combination (UV)\mathbf{\nabla }\left( \mathbf{U}\cdot\mathbf{V}\right) is commonly found in numerous applications. And since the gradient satisfies the product rule
(FG)=F G+F G(18.1)\mathbf{\nabla}\left( FG\right) =\mathbf{\nabla}F~G+F~\mathbf{\nabla}G\tag{18.1}
for scalar fields, we expect to find a similar rule for the dot product of vector fields, i.e.
(UV)=UV+UV.(-)\mathbf{\nabla}\left( \mathbf{U}\cdot\mathbf{V}\right) =\mathbf{\nabla U}\cdot\mathbf{V}+\mathbf{U}\cdot\mathbf{\nabla V.} \tag{-}
However, this identity is nonsensical as there is no such thing as U\mathbf{\nabla U} since the gradient cannot be applied to a vector field. Thus, we have failed in one of the most elementary endeavors of Vector Calculus.
To understand the root cause of the problem, let us borrow one of our own paragraphs from Chapter 2 in which we addressed the effectiveness of the language of geometric vectors: As we will quickly discover, working with geometric quantities leads to greater organization of thought. Geometric vectors have fewer algebraic capabilities compared to numbers. This is a positive: the small number of accessible operations focuses our attention on the few that are available. However, when it comes to more sophisticated analyses, what has been a positive will now become a negative. Geometric vectors are simply not rich enough algebraically to accommodate the more advanced demands of Vector Calculus.
There are two possible ways out. The first way, and one that is central to our narrative, is to introduce coordinates in order to switch from vectors to their components. In a microcosm, the immense utility of this approach can be illustrated by the aforementioned gradient of a dot product
(UV).(18.2)\mathbf{\nabla}\left( \mathbf{U}\cdot\mathbf{V}\right) .\tag{18.2}
In terms of components, the dot product UV\mathbf{U}\cdot\mathbf{V} is UiViU_{i}V^{i} and the components of the gradient are given by
k(UiVi).(18.3)\nabla_{k}\left( U_{i}V^{i}\right) .\tag{18.3}
Next, an effortless application of the product rule for the covariant derivative yields
k(UiVi)=kUi Vi+Ui kVi.(18.4)\nabla_{k}\left( U_{i}V^{i}\right) =\nabla_{k}U_{i}~V^{i}+U_{i}~\nabla _{k}V^{i}.\tag{18.4}
Naturally, the gradienttextbf{ }(UV)\mathbf{\nabla}\left( \mathbf{U} \cdot\mathbf{V}\right) itself is produced by contracting k(UiVi)\nabla_{k}\left( U_{i}V^{i}\right) with Zk\mathbf{Z}^{k}, i.e.
(UV)=(kUi Vi+Ui kVi)Zk,(18.5)\mathbf{\nabla}\left( \mathbf{U}\cdot\mathbf{V}\right) =\left( \nabla _{k}U_{i}~V^{i}+U_{i}~\nabla_{k}V^{i}\right) \mathbf{Z}^{k},\tag{18.5}
and the calculation is complete.
Note, however, that the restoration of the geometric invariant (UV)\mathbf{\nabla }\left( \mathbf{U}\cdot\mathbf{V}\right) need not take place at this point if further analysis is to be conducted. And, although the combinations kUi Vi\nabla_{k}U_{i}~V^{i} and Ui kViU_{i}~\nabla_{k}V^{i} are devoid of direct geometric interpretation, they can be easily subjected to further analysis. Meanwhile, the geometric meaningfulness of the eventual results is assured by the tensor framework.
In summary, the identity
k(UiVi)=kUi Vi+Ui kVi(18.6)\nabla_{k}\left( U_{i}V^{i}\right) =\nabla_{k}U_{i}~V^{i}+U_{i}~\nabla _{k}V^{i}\tag{18.6}
has all the simplicity and the structure of its meaningless counterpart
(UV)=UV+UV(-)\mathbf{\nabla}\left( \mathbf{U}\cdot\mathbf{V}\right) =\mathbf{\nabla \mathbf{U}}\cdot\mathbf{V}+\mathbf{U}\cdot\mathbf{\nabla V} \tag{-}
and at the same time enables us to take full advantage of the tensor framework.
The alternative approach, known as dyadic, doubles down on the geometric aspect of vectors in order to avoid the use of coordinate systems at all costs. In the dyadic approach, all objects are assigned some invariant meaning characterized by their interactions with other objects. For example, the object U\mathbf{\nabla U} would be described as a linear transformation that can be applied to any other vector V\mathbf{V} with the result denoted by U(V)\mathbf{\nabla U}\left( \mathbf{V}\right) , (U,V)\left( \mathbf{\nabla U},\mathbf{V}\right) , or simply by UV\mathbf{\nabla U}\cdot\mathbf{V}. No doubt, this operation would be defined in such a way that it produces, without mentioning it explicitly, the vector ZkVikUi\mathbf{Z}^{k}V^{i}\nabla_{k}U_{i}, thus achieving the same result, but axiomatically and without a reference to the components of vectors.
This simple example illustrates the unfortunate side effect of the dyadic approach: the need for new kinds of objects and new kinds of operators. The resulting framework never reaches algebraic closure, i.e. a framework with a finite number of kinds of objects and operators where the result of applying an existing kind of operator to an existing kind of object always produces an object of an existing kind. A visit to the Vector calculus identities page on Wikipedia reveals identities such as
(BA^)=A^(B)+(B)A^,          (18.7)(baT)=a(b)+(b)a, and          (18.8)(BA)=A(AB)+B(AB),          (18.9)\begin{aligned}\mathbf{\nabla}\cdot\left( \mathbf{B}\otimes\mathbf{\hat{A}}\right) & =\mathbf{\hat{A}}\left( \mathbf{\nabla}\cdot\mathbf{B}\right) +\left( \mathbf{B}\cdot\mathbf{\nabla}\right) \mathbf{\hat{A}}\text{,}\ \ \ \ \ \ \ \ \ \ \left(18.7\right)\\\mathbf{\nabla}\cdot\left( \mathbf{ba}^{T}\right) & =\mathbf{a}\left( \mathbf{\nabla}\cdot\mathbf{b}\right) +\left( \mathbf{b\cdot\nabla}\right) \mathbf{a}\text{, and}\ \ \ \ \ \ \ \ \ \ \left(18.8\right)\\\mathbf{\nabla}\left( \mathbf{B}\cdot\mathbf{A}\right) & =\mathbf{\nabla }_{A}\left( \mathbf{A}\cdot\mathbf{B}\right) +\mathbf{\nabla}_{B}\left( \mathbf{A}\cdot\mathbf{B}\right) ,\ \ \ \ \ \ \ \ \ \ \left(18.9\right)\end{aligned}
where each identity clearly attempts to capture the elementary product rule, but is only able to do so by relying on novel kinds of objects and operators. This experience is perfectly captured by Hermann Weyl's quote from his classic Space Time Matter: In trying to avoid continual reference to the components we are obliged to adopt an endless profusion of names and symbols, in addition to an intricate set of rules for carrying out calculations, so that the balance of advantage is considerably on the negative side. An emphatic protest must be entered against these orgies of formalism which are threatening the peace of even the technical scientist.
Let us, then, set the dyadic approach aside and concentrate on the coordinate approach supported by the techniques of Tensor Calculus. In this Chapter, we will discuss four classical differential operators: the gradient, the divergence, the Laplacian, and the curl. Of course, each one of these operators is some combination involving the covariant derivative. This highlights one of the strengths of Tensor Calculus: its highly restricted set of underlying operations. In a Euclidean space, all objects and operations are expressed in terms of addition, multiplication, contraction, and the higher-level construct of the covariant derivative.
When a coordinate system ZiZ^{i} is imposed upon a Euclidean space, we are automatically provided with a number of tensors that can be used in various combinations, including the covariant and the contravariant bases Zi\mathbf{Z}_{i} and Zi\mathbf{Z}^{i}, the metric tensors ZijZ_{ij} and ZijZ^{ij}, and the Levi-Civita symbols εijk\varepsilon_{ijk} and εijk\varepsilon ^{ijk}. These basic elements, along with the covariant derivative i\nabla _{i} will act as jigsaw puzzle pieces in the construction of the invariant differential operators. Interestingly, the metric tensor will rarely appear explicitly in our expressions since, thanks to index juggling, it gets absorbed into the tensor with which it is contracted. Nevertheless, its continual presence is revealed in the placement of indices.

18.2.1The gradient

For a scalar field UU, one way to form a differential invariant is to contract iU\nabla_{i}U with Zi\mathbf{Z}^{i}, resulting in the vector
ZiiU.(18.10)\mathbf{Z}^{i}\nabla_{i}U.\tag{18.10}
We instantly recognize this combination as the gradient U\mathbf{\nabla}U of UU, which can also denoted by the symbol gradU\operatorname{grad}U, i.e.
gradU=ZiiU.(18.11)\operatorname{grad}U=\mathbf{Z}^{i}\nabla_{i}U.\tag{18.11}
Note that a similar combination cannot be formed for a vector field U\mathbf{U} since the product ZiiU \mathbf{Z}^{i}\nabla_{i}\mathbf{U~}is meaningless. This reminds of the fact that we mentioned in Chapter 4 that there is no such thing as the gradient of a vector field. That said, the dot product Zi\mathbf{Z}^{i} and iU\nabla_{i}\mathbf{U}, which is discussed below, is quite meaningful.

18.2.2The Laplacian

Another meaningful invariant combination involving a scalar field UU is iiU\nabla_{i}\nabla^{i}U. This operation is known as the Laplacian of UU and can be denoted by ΔU\Delta U, i.e.
ΔU=iiU.(18.12)\Delta U=\nabla_{i}\nabla^{i}U.\tag{18.12}
Unlike the gradient, the Laplacian can be applied to vector fields. Indeed, the combination
iiU(18.13)\nabla_{i}\nabla^{i}\mathbf{U}\tag{18.13}
represents is a perfectly well-defined tensor expression. For instance, the Laplacian can be applied to the position vector R\mathbf{R} and it is left as an exercise to show that the result is zero. Since the Laplacian of an invariant is itself an invariant, it immediately raises the question of its geometric interpretation. While some invariants do not have a simple geometric interpretation, the Laplacian does, and it will be described later in this Chapter.

18.2.3The divergence

Among combinations that can be applied to first-order tensors UiU^{i} , the most natural one is, undoubtedly,
iUi.(18.14)\nabla_{i}U^{i}.\tag{18.14}
known as the divergence of the tensor field UiU^{i}. Note that the Laplacian of a scalar UU (unrelated to UiU^{i} in the previous sentence) is equivalent to the divergence of the tensor iU\nabla^{i}U.
The divergence iUi\nabla_{i}U^{i} can also be thought of as being associated with the vector field U=UiZi\mathbf{U}=U^{i}\mathbf{Z}_{i}. Thus, the combination iUi\nabla_{i}U^{i} can be referred to as the divergence of the vector field U\mathbf{U} and, in that context, be denoted by U\mathbf{\nabla} \cdot\mathbf{U} or divU\operatorname{div}\mathbf{U}, i.e.
divU=iUi.(18.15)\operatorname{div}\mathbf{U}=\nabla_{i}U^{i}.\tag{18.15}
textbf{ }It is left as a simple exercise to that divU\operatorname{div} \mathbf{U} is given by the equation
divU=ZiiU.(18.16)\operatorname{div}\mathbf{U}=\mathbf{Z}^{i}\cdot\nabla_{i}\mathbf{U.}\tag{18.16}

18.2.4A note on the instructive utility of Tensor Calculus

The logic that has led us to the gradient, the Laplacian, and the divergence illustrates that the tensor framework not only provides us with a framework for analyzing the invariant properties of a given object, but also gives us clear and exhaustive instructions for constructing new invariants. For example, observe that an invariant combination must have an even number of indices and, therefore, 22 is the least nonzero number of indices that an invariant can have. Thus, from this "orchestration" point of view, it is a trivial observation -- and yet it is a profound insight -- that the gradient, the Laplacian, and the divergence are the only possible invariant expressions involving the symbols i\nabla_{i}, UU, U\mathbf{U}, UiU^{i}, and Zi\mathbf{Z}^{i} -- with the implied possibility of index juggling. Of course, since the Laplacian is the divergence of gradient, we could reduce the number of elementary differential operators to two, but it is worthwhile to treat the Laplacian as an equal partner.
More complicated invariant expressions with an even number of indices, such as ijiU Zj\nabla_{i}\nabla_{j}\nabla^{i}U~\mathbf{Z}^{j}, jjiiU\nabla_{j}\nabla^{j} \nabla_{i}\nabla^{i}U, and ijiUj\nabla^{i}\nabla_{j}\nabla_{i}U^{j}, can always be viewed as combinations of the gradient, the Laplacian, and the divergence. For example, the combination ijiU Zj\nabla_{i}\nabla_{j}\nabla^{i}U~\mathbf{Z}^{j} is the Laplacian of the gradient of UU, which becomes apparent when the expression is rewritten as
Zjj(iiU),(18.17)\mathbf{Z}^{j}\nabla_{j}\left( \nabla_{i}\nabla^{i}U\right) ,\tag{18.17}
with parentheses added strictly for stylistic purposes. Meanwhile, jjiiU\nabla _{j}\nabla^{j}\nabla_{i}\nabla^{i}U is the Laplacian of the Laplacian of UU. This combination is known as the biharmonic operator and plays an important role in the theory elasticity. Finally, ijiUj\nabla^{i}\nabla_{j} \nabla_{i}U^{j} when rewritten as
ii(jUj)(18.18)\nabla^{i}\nabla_{i}\left( \nabla_{j}U^{j}\right)\tag{18.18}
is easily identified as the Laplacian of the divergence of UjU^{j}.
Of course, with the introduction of additional objects, further invariant operators can be formed. For example, with the help of the Levi-Civita symbol εij\varepsilon^{ij} in two dimensions, we can form the combination εijiUj\varepsilon^{ij}\nabla_{i}U_{j} which is the two-dimensional version of the classical curl. In three dimensions, the combination εijkij\varepsilon ^{ijk}\nabla_{i}\nabla_{j} leaves a free index that can be contracted with Zk\mathbf{Z}_{k} to form the combination εijkijZk\varepsilon^{ijk}\nabla_{i} \nabla_{j}\mathbf{Z}_{k}, which is the actual curl. We will discuss the curl, both in two and three dimensions, later in the Chapter. Meanwhile, we will turn our attention to the geometric interpretations the Laplacian and divergence, having described the geometric interpretation of the gradient in Chapter 4.
Suppose that a scalar field UU describes an equilibrium temperature distribution in a room. Such a distribution is typically not uniform: it may be colder near a window and warmer near a radiator. Empirically, we know of this distribution that there cannot be a point inside the room that is, say, hotter than all of its immediate neighbors. If such a distribution were to occur in the absence of heat sources -- for example, a moment after a candle is extinguished -- the temperature would begin to average out until an equilibrium as achieved where, once again, no point is hotter than its neighbors. Thus, me may decide to model an equilibrium distribution as one where the temperature at each point is exactly the average of its neighbors in some precise mathematical sense. That precise sense can be provided by the Laplacian. In this Section, we will show that the Laplacian of UU captures the deviation between the value of UU at a point and the average of its values on a small sphere centered at that point. Thanks to this property, an equilibrium temperature distribution may be characterized by zero Laplacian, i.e.
iiU=0.(18.19)\nabla_{i}\nabla^{i}U=0.\tag{18.19}
Functions whose Laplacian vanishes are known as harmonic and occupy an important place in applied mathematics. If a temperature distribution is not equilibrium, then its return to equilibrium is naturally modeled by the heat equation
Ut=iiU(18.20)\frac{\partial U}{\partial t}=\nabla_{i}\nabla^{i}U\tag{18.20}
which essentially states the rate of change in temperature at a given point is proportional to its deviation from the average of its neighbors. This canonical example shows how this interpretation of the Laplacian leads to numerous physical applications.
Let us now derive the precise way in which the Laplacian captures the deviation between the value of UU at a point and the average of its values on a small sphere centered at that point. Consider a scalar field UU in a neighborhood of a point PP in a two-dimensional Euclidean space, i.e. a plane, where a sphere is replaced by a circle. Refer the plane to Cartesian coordinates x,yx,y with the point PP at the origin. Recall that in Cartesian coordinates, the Laplacian is given by the equation
iiU=2Ux2+2Uy2.(15.142)\nabla_{i}\nabla^{i}U=\frac{\partial^{2}U}{\partial x^{2}}+\frac{\partial ^{2}U}{\partial y^{2}}. \tag{15.142}
In the vicinity of PP, we can approximate the function U(x,y)U\left( x,y\right) by the first two terms of its Taylor series, i.e.
U(x,y)U(0,0)+Ux x+Uy y                          +12(2Ux2 x2+22Uxy xy+2Uy2 y2),          (18.21)\begin{aligned}U\left( x,y\right) & \approx U\left( 0,0\right) +\frac{\partial U}{\partial x}~x+\frac{\partial U}{\partial y}~y\ \ \ \ \ \ \ \ \ \ \\& \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ +\frac{1}{2}\left( \frac{\partial^{2} U}{\partial x^{2}}~x^{2}+2\frac{\partial^{2}U}{\partial x\partial y} ~xy+\frac{\partial^{2}U}{\partial y^{2}}~y^{2}\right) ,\ \ \ \ \ \ \ \ \ \ \left(18.21\right)\end{aligned}
where all of the derivatives are evaluated at PP.
Next, consider a circle of radius rr centered at PP.
(18.22)
If the circle is parameterized by the equations
x(θ)=rcosθ          (18.23)y(θ)=rsinθ,          (18.24)\begin{aligned}x\left( \theta\right) & =r\cos\theta\ \ \ \ \ \ \ \ \ \ \left(18.23\right)\\y\left( \theta\right) & =r\sin\theta,\ \ \ \ \ \ \ \ \ \ \left(18.24\right)\end{aligned}
then the average U\left\langle U\right\rangle of the values of UU over the circle is given by the integral
U=12πr02πU(rcosθ,rsinθ) rdθ.(18.25)\left\langle U\right\rangle =\frac{1}{2\pi r}\int_{0}^{2\pi}U\left( r\cos\theta,r\sin\theta\right) ~rd\theta.\tag{18.25}
Substituting the Taylor series approximation for UU and evaluating the resulting integral, we find
UU(0,0)+14r2(2Ux2+2Uy2),(18.26)\left\langle U\right\rangle \approx U\left( 0,0\right) +\frac{1}{4} r^{2}\left( \frac{\partial^{2}U}{\partial x^{2}}+\frac{\partial^{2} U}{\partial y^{2}}\right) ,\tag{18.26}
where we see the expression of the Laplacian emerging on the right. Indeed, the leading term in the difference between U\left\langle U\right\rangle and U(0,0)U\left( 0,0\right) is directly proportional to the Laplacian, i.e.
UU(0,0)14r2iiU.(18.27)\left\langle U\right\rangle -U\left( 0,0\right) \approx\frac{1}{4} r^{2}\nabla_{i}\nabla^{i}U.\tag{18.27}
This equation captures the precise sense in which the Laplacian is a measure of the deviation between the value of a function and the average of its neighbors.
In the following figure, the function illustrated by the contour plot on the left is characterized by a small value of the Laplacian. By contrast, the function on the right has a high value of the Laplacian.
  (18.28)
The geometric interpretation of divergence comes from the divergence theorem, a crucial generalization of the Fundamental Theorem of Calculus from one-dimensional intervals to higher-dimensional domains. The divergence theorem and its applications represent a vast topic that we will discuss in a future book. For the sake of the present discussion, however, we will give the statement of the theorem here.
Consider a closed domain Ω\Omega with boundary SS and outward unit normal N\mathbf{N} with components NiN^{i}.
(18.29)
For a vector field U\mathbf{U} with components UiU^{i}, the divergence theorem reads
ΩiUidΩ=SNiUidS.(18.30)\int_{\Omega}\nabla_{i}U^{i}d\Omega=\int_{S}N_{i}U^{i}dS.\tag{18.30}
In dyadic terms, the divergence theorem appears in the form
ΩdivU dΩ=SUN dS.(18.31)\int_{\Omega}\operatorname{div}\mathbf{U~}d\Omega=\int_{S}\mathbf{U} \cdot\mathbf{N~}dS.\tag{18.31}
The geometric interpretation of the divergence of U\mathbf{U} which we are currently pursuing can be gleaned from the integral on the right.
For simplicity, consider a uniform vector field U\mathbf{U} in the vicinity of a straight boundary SS characterized by the normal N\mathbf{N}.
(18.32)
This simple configuration will help us demonstrate that the dot product UN\mathbf{U}\cdot\mathbf{N} corresponds to the flux of U\mathbf{U} across SS, i.e. the rate at which the fluid mass crosses the interface. We will begin by considering the two extreme examples illustrated in the following figure.
  (18.33)
The left plot shows a flow parallel to the boundary SS. In this case, no fluid crosses SS and therefore the flux is zero. Correspondingly, the dot product UN\mathbf{U}\cdot\mathbf{N} vanishes and therefore so does the integral SUN dS\int_{S}\mathbf{U}\cdot\mathbf{N~}dS. The right plot shows a flow orthogonal to the boundary, which results in the greatest flux. Correspondingly, UN\mathbf{U}\cdot\mathbf{N} attains its maximum value, which equals the magnitude of U\mathbf{U}, and the integral SUN dS\int_{S} \mathbf{U}\cdot\mathbf{N~}dS indeed corresponds to the rate at which the fluid mass crosses the interface SS, i.e. the flux.
Finally, consider a flow with an angle of attack that equals γ\gamma with respect to the normal N\mathbf{N}. Identify a segment along the boundary of length ΔS\Delta S and calculate the amount of fluid that crosses that segment in a period of time Δt\Delta t.
  (18.34)
The fluid that crosses the segment in that time period is contained in the (gray) parallelogram with sides ΔS\Delta S and UΔtU\Delta t, and an angle of π/2γ\pi/2-\gamma. The total amount of fluid contained in this parallelogram corresponds to its area AA given by
A=UΔtΔSsin(π/2γ)=UΔtΔScosγ.(18.35)A=U\Delta t\Delta S\sin\left( \pi/2-\gamma\right) =U\Delta t\Delta S\cos\gamma.\tag{18.35}
Thus, the amount of fluid crossing the segment per unit time and per unit length along the interface is
AΔtΔS=Ucosγ.(18.36)\frac{A}{\Delta t\Delta S}=U\cos\gamma\mathbf{.}\tag{18.36}
Since UcosγU\cos\gamma is precisely UN\mathbf{U}\cdot\mathbf{N}, we have
AΔtΔS=UN.(18.37)\frac{A}{\Delta t\Delta S}=\mathbf{U}\cdot\mathbf{N.}\tag{18.37}
Therefore, the integral
SUN dS(18.38)\int_{S}\mathbf{U}\cdot\mathbf{N~}dS\tag{18.38}
represents the instantaneous amount per unit time of fluid crossing the boundary, which we understand to be the flux. More specifically, since N\mathbf{N} is the outward normal, the integral represents the instantaneous rate of fluid escaping the domain Ω\Omega.
Thanks to the divergence theorem
ΩdivU dΩ=SUN dS,(18.31)\int_{\Omega}\operatorname{div}\mathbf{U~}d\Omega=\int_{S}\mathbf{U} \cdot\mathbf{N~}dS, \tag{18.31}
the volume integral
ΩdivU dΩ(18.39)\int_{\Omega}\operatorname{div}\mathbf{U~}d\Omega\tag{18.39}
also represents the instantaneous rate of fluid escaping the domain Ω\Omega. In order to obtain an intuitive understanding of divU\operatorname{div} \mathbf{U} in the local sense, consider a "small" domain Ω\Omega over which the field U\mathbf{U} is approximately linear and the quantity divU\operatorname{div}\mathbf{U} is therefore approximately constant. If divU\operatorname{div}\mathbf{U} is small then, correspondingly, the net flux across the closed boundary is also small. Such a flow may look like the one in the following plot.
(18.40)
If, on the other hand, divU\operatorname{div}\mathbf{U} is large then we expect a large net flow across the boundary. If divU>0\operatorname{div}\mathbf{U}\gt 0, the corresponding flow may look like the one in the following plot.
(18.41)
Of course, this is an exaggerated flow meant to illustrate the point. Nevertheless, these examples make it apparent where the divergence operator got its name.
In this Section, we will derive the expressions for the divergence and the Laplacian in the most common special coordinate systems.

18.5.1The divergence

Let us start with the divergence
iUi.(18.42)\nabla_{i}U^{i}.\tag{18.42}
By definition, the covariant derivative iUj\nabla_{i}U^{j} is given by
iUj=UjZi+ΓikjUk.(18.43)\nabla_{i}U^{j}=\frac{\partial U^{j}}{\partial Z^{i}}+\Gamma_{ik}^{j}U^{k}.\tag{18.43}
Contracting the indices yields the expression
iUi=UiZi+ΓikiUk(18.44)\nabla_{i}U^{i}=\frac{\partial U^{i}}{\partial Z^{i}}+\Gamma_{ik}^{i}U^{k}\tag{18.44}
that is ready for interpretation in any coordinate system. For future convenience, let us switch the indices ii and kk in the second term, i.e.
iUi=UiZi+ΓkikUi.(18.45)\nabla_{i}U^{i}=\frac{\partial U^{i}}{\partial Z^{i}}+\Gamma_{ki}^{k}U^{i}.\tag{18.45}
Recall that we have encountered the combination Γiki\Gamma_{ik}^{i} before in the equation
ZZk=ZΓiki(16.181)\frac{\partial\sqrt{Z}}{\partial Z^{k}}=\sqrt{Z}\Gamma_{ik}^{i} \tag{16.181}
which was later used to demonstrate the tensor property of the Levi-Civita symbols. Later in this Chapter, we will also use the above equation in the derivation of the Voss-Weyl formula which provides an alternative approach to calculating the divergence in any particular coordinate system.
Let us now begin to interpret the equation
iUi=UiZi+ΓkikUi(18.45)\nabla_{i}U^{i}=\frac{\partial U^{i}}{\partial Z^{i}}+\Gamma_{ki}^{k}U^{i} \tag{18.45}
in various coordinate systems. In affine and, in particular, Cartesian, coordinates x,y,zx,y,z, where the Christoffel symbols vanish, we have
iUi=U1x+U2y+U3z.(18.46)\nabla_{i}U^{i}=\frac{\partial U^{1}}{\partial x}+\frac{\partial U^{2} }{\partial y}+\frac{\partial U^{3}}{\partial z}.\tag{18.46}
In cylindrical coordinates, recall that the nonzero elements of the Christoffel symbol Γjki\Gamma_{jk}^{i} are
Γ221=r    and    Γ122=Γ212=1r.(12.52)\Gamma_{22}^{1}=-r\text{\ \ \ \ and\ \ \ \ }\Gamma_{12}^{2}=\Gamma_{21} ^{2}=\frac{1}{r}. \tag{12.52}
Consequently, the three elements of the contraction Γkik\Gamma_{ki}^{k} are
Γk1k=1r,    Γk2k=0,    and    Γk3k=0.(18.47)\Gamma_{k1}^{k}=\frac{1}{r}\text{,\ \ \ \ }\Gamma_{k2}^{k} =0\text{,\ \ \ \ and\ \ \ \ }\Gamma_{k3}^{k}=0.\tag{18.47}
Therefore, the expression for the divergence in terms of the contravariant components reads
iUi=U1r+1rU1+U2θ+U3z.(18.48)\nabla_{i}U^{i}=\frac{\partial U^{1}}{\partial r}+\frac{1}{r}U^{1} +\frac{\partial U^{2}}{\partial\theta}+\frac{\partial U^{3}}{\partial z}.\tag{18.48}
Note that the first two terms can be effectively combined into a single term, i.e.
iUi=1rr(rU1)+U2θ+U3z.(18.49)\nabla_{i}U^{i}=\frac{1}{r}\frac{\partial}{\partial r}\left( rU^{1}\right) +\frac{\partial U^{2}}{\partial\theta}+\frac{\partial U^{3}}{\partial z}.\tag{18.49}
The Voss-Weyl formula discussed later in this Chapter will clarify for us why the combination
1rr(rU1)(18.50)\frac{1}{r}\frac{\partial}{\partial r}\left( rU^{1}\right)\tag{18.50}
is a natural one.
It is left as an exercise to show that in polar coordinates in a two-dimensional Euclidean space, the divergence is given by the equation
iUi=1rr(rU1)+U2θ.(18.51)\nabla_{i}U^{i}=\frac{1}{r}\frac{\partial}{\partial r}\left( rU^{1}\right) +\frac{\partial U^{2}}{\partial\theta}.\tag{18.51}
Finally, let us turn our attention spherical coordinates. Recall that the nonzero elements of the Christoffel symbol Γjki\Gamma_{jk}^{i} are
Γ221=r          (12.54)Γ331=rsin2θ          (12.55)Γ122=Γ212=1r          (12.56)Γ332=sinθcosθ          (12.57)Γ133=Γ313=1r          (12.58)Γ233=Γ323=1tanθ,          (12.59)\begin{aligned}\Gamma_{22}^{1} & =-r\ \ \ \ \ \ \ \ \ \ \left(12.54\right)\\\Gamma_{33}^{1} & =-r\sin^{2}\theta\ \ \ \ \ \ \ \ \ \ \left(12.55\right)\\\Gamma_{12}^{2} & =\Gamma_{21}^{2}=\frac{1}{r}\ \ \ \ \ \ \ \ \ \ \left(12.56\right)\\\Gamma_{33}^{2} & =-\sin\theta\cos\theta\ \ \ \ \ \ \ \ \ \ \left(12.57\right)\\\Gamma_{13}^{3} & =\Gamma_{31}^{3}=\frac{1}{r}\ \ \ \ \ \ \ \ \ \ \left(12.58\right)\\\Gamma_{23}^{3} & =\Gamma_{32}^{3}=\frac{1}{\tan\theta}, \ \ \ \ \ \ \ \ \ \ \left(12.59\right)\end{aligned}
and thus the three elements of Γkik\Gamma_{ki}^{k} are
Γk1k=2r,    Γk2k=1tanθ,    and    Γk3k=0.(18.52)\Gamma_{k1}^{k}=\frac{2}{r}\text{,\ \ \ \ }\Gamma_{k2}^{k}=\frac{1}{\tan \theta}\text{,\ \ \ \ and\ \ \ \ }\Gamma_{k3}^{k}=0.\tag{18.52}
As a result, in spherical coordinates, we have
iUi=U1r+2rU1+U2θ+1tanθU2+U3φ.(18.53)\nabla_{i}U^{i}=\frac{\partial U^{1}}{\partial r}+\frac{2}{r}U^{1} +\frac{\partial U^{2}}{\partial\theta}+\frac{1}{\tan\theta}U^{2} +\frac{\partial U^{3}}{\partial\varphi}.\tag{18.53}
Analogously to the divergence, some of the terms can be combined, yielding the form
iUi=1r2r(r2U1)+1sinθθ(sinθU2)+U3φ,(18.54)\nabla_{i}U^{i}=\frac{1}{r^{2}}\frac{\partial}{\partial r}\left( r^{2} U^{1}\right) +\frac{1}{\sin\theta}\frac{\partial}{\partial\theta}\left( \sin\theta U^{2}\right) +\frac{\partial U^{3}}{\partial\varphi},\tag{18.54}
which will, once again, be elucidated by the Voss-Weyl formula described below.

18.5.2The Laplacian

Since the contravariant derivative iU\nabla^{i}U is given by
iU=ZijjU,(18.55)\nabla^{i}U=Z^{ij}\nabla_{j}U,\tag{18.55}
the expression for the Laplacian in terms of the covariant derivatives alone reads
iiU=ZijijU.(18.56)\nabla_{i}\nabla^{i}U=Z^{ij}\nabla_{i}\nabla_{j}U.\tag{18.56}
Expanding the covariant derivatives in terms of partial derivatives, we find the form
iiU=Zij(2UZiZjΓijmUZm)(18.57)\nabla_{i}\nabla^{i}U=Z^{ij}\left( \frac{\partial^{2}U}{\partial Z^{i}\partial Z^{j}}-\Gamma_{ij}^{m}\frac{\partial U}{\partial Z^{m}}\right)\tag{18.57}
suitable for evaluation in various coordinate systems.
In Cartesian coordinates x,y,zx,y,z, the Christoffel symbol vanishes and the metric tensors correspond to the identity matrix. Therefore, as we found before, we have
iiU=2Ux2+2Uy2+2Uz2.(15.143)\nabla_{i}\nabla^{i}U=\frac{\partial^{2}U}{\partial x^{2}}+\frac{\partial ^{2}U}{\partial y^{2}}+\frac{\partial^{2}U}{\partial z^{2}}. \tag{15.143}
In more general affine coordinates, the most specific form possible is
iiU=Zij2UZiZj,(18.58)\nabla_{i}\nabla^{i}U=Z^{ij}\frac{\partial^{2}U}{\partial Z^{i}\partial Z^{j} },\tag{18.58}
where the elements of the contravariant metric tensor ZijZ^{ij} are constants.
In cylindrical coordinates r,θ,zr,\theta,z, recall that
Zij corresponds to [11r21](9.87)Z^{ij}\text{ corresponds to }\left[ \begin{array} {ccc} 1 & & \\ & \frac{1}{r^{2}} & \\ & & 1 \end{array} \right] \tag{9.87}
and that the nonzero elements of the Christoffel symbol are
Γ221=r    and    Γ122=Γ212=1r.(12.52)\Gamma_{22}^{1}=-r\text{\ \ \ \ and\ \ \ \ }\Gamma_{12}^{2}=\Gamma_{21} ^{2}=\frac{1}{r}. \tag{12.52}
Since ZijZ^{ij} corresponds to a diagonal matrix, the only surviving terms in
iiU=Zij(2UZiZjΓijmUZm)(18.57)\nabla_{i}\nabla^{i}U=Z^{ij}\left( \frac{\partial^{2}U}{\partial Z^{i}\partial Z^{j}}-\Gamma_{ij}^{m}\frac{\partial U}{\partial Z_{m}}\right) \tag{18.57}
must have i=ji=j. Combining these observations yields
iiU=2Ur2+1rUr+1r22Uθ2+2Uz2.(18.59)\nabla_{i}\nabla^{i}U=\frac{\partial^{2}U}{\partial r^{2}}+\frac{1}{r} \frac{\partial U}{\partial r}+\frac{1}{r^{2}}\frac{\partial^{2}U} {\partial\theta^{2}}+\frac{\partial^{2}U}{\partial z^{2}}.\tag{18.59}
As before, the first two terms can be combined into one, leading to the final form
iiU=1rr(rUr)+1r22Uθ2+2Uz2.(18.60)\nabla_{i}\nabla^{i}U=\frac{1}{r}\frac{\partial}{\partial r}\left( r\frac{\partial U}{\partial r}\right) +\frac{1}{r^{2}}\frac{\partial^{2} U}{\partial\theta^{2}}+\frac{\partial^{2}U}{\partial z^{2}}.\tag{18.60}
In spherical coordinates r,θ,φr,\theta,\varphi, recall that
Zij corresponds to [11r21r2sin2θ],(9.88)Z^{ij}\text{ corresponds to }\left[ \begin{array} {ccc} 1 & & \\ & \frac{1}{r^{2}} & \\ & & \frac{1}{r^{2}\sin^{2}\theta} \end{array} \right] , \tag{9.88}
which is again diagonal, and that the nonzero elements of the Christoffel symbols Γijm\Gamma_{ij}^{m} for which i=ji=j are
Γ221=r          (12.54)Γ331=rsin2θ          (12.56)Γ332=sinθcosθ.          (12.57)\begin{aligned}\Gamma_{22}^{1} & =-r\ \ \ \ \ \ \ \ \ \ \left(12.54\right)\\\Gamma_{33}^{1} & =-r\sin^{2}\theta\ \ \ \ \ \ \ \ \ \ \left(12.56\right)\\\Gamma_{33}^{2} & =-\sin\theta\cos\theta. \ \ \ \ \ \ \ \ \ \ \left(12.57\right)\end{aligned}
Upon substituting these values into the equation
iiU=Zij(2UZiZjΓijmUZm)(18.57)\nabla_{i}\nabla^{i}U=Z^{ij}\left( \frac{\partial^{2}U}{\partial Z^{i}\partial Z^{j}}-\Gamma_{ij}^{m}\frac{\partial U}{\partial Z_{m}}\right) \tag{18.57}
and simplifying, we get
iiU=2Ur2+2rUr+1r22Uθ2+1r2tan2θUθ+1r2sin2θ2Uφ2,(18.61)\nabla_{i}\nabla^{i}U=\frac{\partial^{2}U}{\partial r^{2}}+\frac{2}{r} \frac{\partial U}{\partial r}+\frac{1}{r^{2}}\frac{\partial^{2}U} {\partial\theta^{2}}+\frac{1}{r^{2}\tan^{2}\theta}\frac{\partial U} {\partial\theta}+\frac{1}{r^{2}\sin^{2}\theta}\frac{\partial^{2}U} {\partial\varphi^{2}},\tag{18.61}
where, combining terms as before, we arrive at the final form
iiU=1r2r(r2Ur)+1r2sinθθ(sinθUθ)+1r2sin2θ2Uφ2.(18.62)\nabla_{i}\nabla^{i}U=\frac{1}{r^{2}}\frac{\partial}{\partial r}\left( r^{2}\frac{\partial U}{\partial r}\right) +\frac{1}{r^{2}\sin\theta} \frac{\partial}{\partial\theta}\left( \sin\theta\frac{\partial U} {\partial\theta}\right) +\frac{1}{r^{2}\sin^{2}\theta}\frac{\partial^{2} U}{\partial\varphi^{2}}.\tag{18.62}

18.5.3A note on the absolute nature of tensor analysis

In this Section, we are experiencing one important feature of tensor analysis. Namely, tensor expressions can be used to produce coordinate-dependent expressions directly, i.e. without a reference to another coordinate system. This feature is best characterized by the adjective absolute which was used in the original name of our subject, the Absolute Differential Calculus.
An alternative approach, used in virtually all introductory Calculus textbooks, is to define objects in the context of a Cartesian coordinate system and then to derive their expressions in other coordinate systems by a change of variables from the Cartesian coordinates. For example, the Laplacian would be defined by the expression
2Ux2+2Uy2+2Uz2(18.63)\frac{\partial^{2}U}{\partial x^{2}}+\frac{\partial^{2}U}{\partial y^{2} }+\frac{\partial^{2}U}{\partial z^{2}}\tag{18.63}
in some Cartesian coordinates. Then it would be demonstrated that this expression yields the same value in all Cartesian coordinates. Finally, a (laborious and error-prone) change of variables from Cartesian to spherical coordinates would be used to derive the expression
1r2r(r2Ur)+1r2sinθθ(sinθUθ)+1r2sin2θ2Uφ2.(18.64)\frac{1}{r^{2}}\frac{\partial}{\partial r}\left( r^{2}\frac{\partial U}{\partial r}\right) +\frac{1}{r^{2}\sin\theta}\frac{\partial} {\partial\theta}\left( \sin\theta\frac{\partial U}{\partial\theta}\right) +\frac{1}{r^{2}\sin^{2}\theta}\frac{\partial^{2}U}{\partial\varphi^{2}}.\tag{18.64}
It is safe to say that, in this regard, the absolute approach of Tensor Calculus is superior. Furthermore, Tensor Calculus does not stop there. The Voss-Weyl formula, which we are about to describe, makes the Tensor Calculus approach even more efficient and robust.
The Voss-Weyl formula, named after Aurel Voss and Hermann Weyl, represents an alternative approach to calculating the divergence. It reads
iUi=1ZZi(ZUi),(18.65)\nabla_{i}U^{i}=\frac{1}{\sqrt{Z}}\frac{\partial}{\partial Z^{i}}\left( \sqrt{Z}U^{i}\right) ,\tag{18.65}
where Z\sqrt{Z} is, of course, the volume element. The advantage of the Voss-Weyl formula is that provides a way of calculating the divergence, and therefore the Laplacian, without a reference to the Christoffel symbol.
To prove the Voss-Weyl formula, recall the formula for the divergence
iUi=UiZi+ΓkikUi(18.45)\nabla_{i}U^{i}=\frac{\partial U^{i}}{\partial Z^{i}}+\Gamma_{ki}^{k}U^{i} \tag{18.45}
and the fact that we have encountered the combination Γkik\Gamma_{ki}^{k} previously in the formula
ZZi=ZΓkik(16.181)\frac{\partial\sqrt{Z}}{\partial Z^{i}}=\sqrt{Z}\Gamma_{ki}^{k} \tag{16.181}
for the partial derivative of the volume element Z\sqrt{Z}. Solving for Γkik\Gamma_{ki}^{k}, we have
Γkik=1ZZZi.(18.66)\Gamma_{ki}^{k}=\frac{1}{\sqrt{Z}}\frac{\partial\sqrt{Z}}{\partial Z^{i}}.\tag{18.66}
Substituting this identity into the above formula for divergence, we find
iUi=UiZi+1ZZZiUi.(18.67)\nabla_{i}U^{i}=\frac{\partial U^{i}}{\partial Z^{i}}+\frac{1}{\sqrt{Z}} \frac{\partial\sqrt{Z}}{\partial Z^{i}}U^{i}.\tag{18.67}
Upon factoring out 1/Z1/\sqrt{Z}, we find
iUi=1Z(ZUiZi+ZZiUi),(18.68)\nabla_{i}U^{i}=\frac{1}{\sqrt{Z}}\left( \sqrt{Z}\frac{\partial U^{i} }{\partial Z^{i}}+\frac{\partial\sqrt{Z}}{\partial Z^{i}}U^{i}\right) ,\tag{18.68}
where we observe, by (a reverse application of) the product rule, that the quantity in parentheses equals
(ZUi)Zi.(18.69)\frac{\partial\left( \sqrt{Z}U^{i}\right) }{\partial Z^{i}}.\tag{18.69}
Thus we arrive at the desired result
iUi=1ZZi(ZUi).(18.65)\nabla_{i}U^{i}=\frac{1}{\sqrt{Z}}\frac{\partial}{\partial Z^{i}}\left( \sqrt{Z}U^{i}\right) . \tag{18.65}
A direct application of the Voss-Weyl formula to the Laplacian iiU\nabla _{i}\nabla^{i}U yields
iiU=1ZZi(ZiU).(18.70)\nabla_{i}\nabla^{i}U=\frac{1}{\sqrt{Z}}\frac{\partial}{\partial Z^{i}}\left( \sqrt{Z}\nabla^{i}U\right) .\tag{18.70}
Since the contravariant derivative iU\nabla^{i}U is given by
iU=ZijjU=ZijUZj,(18.71)\nabla^{i}U=Z^{ij}\nabla_{j}U=Z^{ij}\frac{\partial U}{\partial Z^{j}},\tag{18.71}
the final expression for the Laplacian in terms of partial derivatives reads
iiU=1ZZi(ZZijUZj).(18.72)\nabla_{i}\nabla^{i}U=\frac{1}{\sqrt{Z}}\frac{\partial}{\partial Z^{i}}\left( \sqrt{Z}Z^{ij}\frac{\partial U}{\partial Z^{j}}\right) .\tag{18.72}
The Voss-Weyl formula makes it an even simpler task to derive the expressions for the divergence and the Laplacian in special coordinate systems. For example, taking the most difficult example of spherical coordinates r,θ,φr,\theta,\varphi, recall that
Z=r2sinθ.(9.62)\sqrt{Z}=r^{2}\sin\theta. \tag{9.62}
and that
Zij corresponds to [11r21r2sin2θ].(9.88)Z^{ij}\text{ corresponds to }\left[ \begin{array} {ccc} 1 & & \\ & \frac{1}{r^{2}} & \\ & & \frac{1}{r^{2}\sin^{2}\theta} \end{array} \right] . \tag{9.88}
Since the above matrix is diagonal, the Laplacian is given by a sum of three terms, i.e.
iiU=1Zr(ZZ11Ur)+1Zθ(ZZ22Uθ)+1Zφ(ZZ33Uφ),(18.73)\nabla_{i}\nabla^{i}U=\frac{1}{\sqrt{Z}}\frac{\partial}{\partial r}\left( \sqrt{Z}Z^{11}\frac{\partial U}{\partial r}\right) +\frac{1}{\sqrt{Z}} \frac{\partial}{\partial\theta}\left( \sqrt{Z}Z^{22}\frac{\partial U}{\partial\theta}\right) +\frac{1}{\sqrt{Z}}\frac{\partial}{\partial\varphi }\left( \sqrt{Z}Z^{33}\frac{\partial U}{\partial\varphi}\right) ,\tag{18.73}
which leads directly to the final expression
iiU=1r2r(r2Ur)+1r2sinθθ(sinθUθ)+1r2sinθ2Uφ2.(18.62)\nabla_{i}\nabla^{i}U=\frac{1}{r^{2}}\frac{\partial}{\partial r}\left( r^{2}\frac{\partial U}{\partial r}\right) +\frac{1}{r^{2}\sin\theta} \frac{\partial}{\partial\theta}\left( \sin\theta\frac{\partial U} {\partial\theta}\right) +\frac{1}{r^{2}\sin\theta}\frac{\partial^{2} U}{\partial\varphi^{2}}. \tag{18.62}
Note that the structure of the Voss-Weyl shows why combinations such as
1r2r(r2Ur)(18.74)\frac{1}{r^{2}}\frac{\partial}{\partial r}\left( r^{2}\frac{\partial U}{\partial r}\right)\tag{18.74}
represent a natural way of grouping partial derivatives.
In summary, the Voss-Weyl formula offers an elegant way of evaluating the divergence, and therefore the Laplacian, with less effort while maintaining greater structure within the expressions.
We will now add the Levi-Civita symbols to the mix, which will give rise to the curl. Since the order of a Levi-Civita symbol depends on the dimension of the space, one obtains different kinds of objects in different dimensions. The classical curl is associated with three dimensions.
In a three-dimensional Euclidean space, consider a vector field U\mathbf{U} with components UiU^{i}. The curl of U\mathbf{U}, dyadically denoted by curlU\operatorname{curl}\mathbf{U} or ×U\mathbf{\nabla}\times\mathbf{U}, is given by
curlU=εijkiUjZk.(18.75)\operatorname{curl}\mathbf{U}=\varepsilon^{ijk}\nabla_{i}U_{j}\mathbf{Z}_{k}.\tag{18.75}
The components VkV^{k} of V=curlU\mathbf{V}=\operatorname{curl}\mathbf{U} are given by
Vk=εijkiUj.(18.76)V^{k}=\varepsilon^{ijk}\nabla_{i}U_{j}.\tag{18.76}
The invariant nature of the curl immediately follows from the fact that it is defined in terms of tensors and tensor-preserving operations. As always, we must note that the Levi-Civita symbol εijk\varepsilon^{ijk}, as classically defined, is a tensor only with respect to orientation-preserving coordinate transformations. Therefore, the curl of U\mathbf{U}, as defined by the above equations, is also a tensor only with respect to orientation-preserving coordinate transformations. It could be made a full tensor by introducing the orientation scalar Π\Pi, discussed in Section 17.10.
Interestingly, and perhaps somewhat unexpectedly, the covariant derivative i\nabla_{i} in the equation
curlU=εijkiUjZk(18.75)\operatorname{curl}\mathbf{U}=\varepsilon^{ijk}\nabla_{i}U_{j}\mathbf{Z}_{k} \tag{18.75}
can be replaced with the partial derivative /Zi\partial/\partial Z^{i}. Indeed, recall that the Christoffel symbol is symmetric, i.e.
Γijk=Γjik,(12.21)\Gamma_{ij}^{k}=\Gamma_{ji}^{k}, \tag{12.21}
and therefore
εijkΓijm=0(18.77)\varepsilon^{ijk}\Gamma_{ij}^{m}=0\tag{18.77}
as a double contraction of an alternating system with a symmetric system. Since
iUj=UjZiΓijmUm,(15.20)\nabla_{i}U_{j}=\frac{\partial U_{j}}{\partial Z^{i}}-\Gamma_{ij}^{m}U_{m}, \tag{15.20}
we have
Vk=εijkiUj=εijk(UjZiΓijmUm)=εijkUjZi.(18.78)V^{k}=\varepsilon^{ijk}\nabla_{i}U_{j}=\varepsilon^{ijk}\left( \frac{\partial U_{j}}{\partial Z^{i}}-\Gamma_{ij}^{m}U_{m}\right) =\varepsilon^{ijk} \frac{\partial U_{j}}{\partial Z^{i}}.\tag{18.78}
Thus,
curlU=εijkUjZiZk,(18.79)\operatorname{curl}\mathbf{U}=\varepsilon^{ijk}\frac{\partial U_{j}}{\partial Z^{i}}\mathbf{Z}_{k},\tag{18.79}
as we set out to show.
A great mnemonic device for the curl can be formulated with the help of the determinant. Namely, the curl is given by the formula
curlU=1ZZ1Z2Z3123U1U2U3,(18.80)\operatorname{curl}\mathbf{U}=\frac{1}{\sqrt{Z}}\left\vert \begin{array} {ccc} \mathbf{Z}_{1} & \mathbf{Z}_{2} & \mathbf{Z}_{3}\\ \nabla_{1} & \nabla_{2} & \nabla_{3}\\ U_{1} & U_{2} & U_{3} \end{array} \right\vert ,\tag{18.80}
where it is understood that "multiplying" by the covariant derivative means applying it. Since, as we just discovered, the covariant derivatives can be replaced with partial derivatives, the above formula can be rewritten as
curlU=1ZZ1Z2Z3Z1Z2Z3U1U2U3.(18.81)\operatorname{curl}\mathbf{U}=\frac{1}{\sqrt{Z}}\left\vert \begin{array} {ccc} \mathbf{Z}_{1} & \mathbf{Z}_{2} & \mathbf{Z}_{3}\\ \frac{\partial}{\partial Z^{1}} & \frac{\partial}{\partial Z^{2}} & \frac{\partial}{\partial Z^{3}}\\ U_{1} & U_{2} & U_{3} \end{array} \right\vert .\tag{18.81}
We observe, therefore, the curl is given by the same expression in term of partial derivatives in all coordinate systems, save for the factor of 1/Z1/\sqrt{Z}. Thus, there is no need for documenting the particular expressions for the curl in specific coordinate systems.
The geometric interpretation of the curl is that it is a local measure of swirliness in the vector field U\mathbf{U}. Much like with the divergence, this interpretation comes from an integration theorem. Namely, Stokes' theorem which is a form of the divergence theorem that applies to a curved surface patch SS immersed in a vector field U\mathbf{U}. It will be described, along with the divergence theorem, in a future book, but its statement will be given here for the purposes of our present discussion.
(18.82)
If N\mathbf{N} is the unit normal on a patch SS and T\mathbf{T} is the unit tangent along the boundary Γ\Gamma, then the Stokes' theorem reads
ScurlU N dS=ΓUT dΓ.(18.83)\int_{S}\operatorname{curl}\mathbf{U~\cdot N~}dS=\int_{\Gamma}\mathbf{U} \cdot\mathbf{T~}d\Gamma.\tag{18.83}
As with the divergence, the interpretation of curlU\operatorname{curl}\mathbf{U} comes from the contour integral on the right as we consider a small patch.
In fact, let us consider a small circular patch SS and note that the tangent T\mathbf{T} exemplifies swirliness.
(18.84)
At each point on the boundary Γ\Gamma, the dot product UT\mathbf{U} \cdot\mathbf{T} in the integrand of the contour integral ΓUT dΓ\int_{\Gamma }\mathbf{U}\cdot\mathbf{T~}d\Gamma is proportional to the magnitude of U\mathbf{U} and, importantly, to the cosine of the angle between U\mathbf{U} and T\mathbf{T}, i.e. the degree to which U\mathbf{U} is aligned with T\mathbf{T}. Thus, the contour integral rewards the "likeness" between U\mathbf{U} and T\mathbf{T} and thus rewards the swirliness in U\mathbf{U}.
Suppose that the space is referred to Cartesian coordinates so that the circle patch SS is in the xx-yy and with its center at the origin. Consider the three different vector fields given by the equations
U1=yixj+0k,           (18.85)U2=xi+yj+0k, and          (18.86)U3=yixj+0k          (18.87)\begin{aligned}\mathbf{U}_{1} & =y\mathbf{i}-x\mathbf{j}+0\mathbf{k}\text{, }\ \ \ \ \ \ \ \ \ \ \left(18.85\right)\\\mathbf{U}_{2} & =x\mathbf{i}+y\mathbf{j}+0\mathbf{k}\text{, and}\ \ \ \ \ \ \ \ \ \ \left(18.86\right)\\\mathbf{U}_{3} & =y\mathbf{i}-x\mathbf{j}+0\mathbf{k}\ \ \ \ \ \ \ \ \ \ \left(18.87\right)\end{aligned}
and illustrated in the following figure.
    (18.88)
It is left as an exercise to show that
curlU1=2k,          (18.89)curlU2=0, and          (18.90)curlU3=0.          (18.91)\begin{aligned}\operatorname{curl}\mathbf{U}_{1} & =2\mathbf{k}\text{,}\ \ \ \ \ \ \ \ \ \ \left(18.89\right)\\\operatorname{curl}\mathbf{U}_{2} & =\mathbf{0}\text{, and}\ \ \ \ \ \ \ \ \ \ \left(18.90\right)\\\operatorname{curl}\mathbf{U}_{3} & =\mathbf{0}\text{.}\ \ \ \ \ \ \ \ \ \ \left(18.91\right)\end{aligned}
Due to the swirly character of the vector field U1\mathbf{U}_{1} in the first plot, we observe that along the boundary, U\mathbf{U} and T\mathbf{T} are consistently aligned. As a result, the dot product UT\mathbf{U}\cdot\mathbf{T} is positive everywhere and every point makes a positive contribution to the integral ΓUT dΓ\int_{\Gamma}\mathbf{U}\cdot\mathbf{T~}d\Gamma.
Meanwhile, in the other two plots, U\mathbf{U} fails to exhibit any degree swirliness. In the second plot, U\mathbf{U} is orthogonal to T\mathbf{T} at each point resulting in the zero contour integral. In the third plot, U\mathbf{U} varies from being largely aligned with T\mathbf{T} at some points to counter-aligned at other points. Consequently, the contributions to the contour integral from the different sections of the boundary cancel each other, once again resulting in a zero integral.
As a final remark, note that swirliness may be difficult to judge with the naked eye. For example, even the swirly field in the first plot does not appear swirly away from the origin. However, it is equally swirly at all points, as indicated by the constant value of curlU\operatorname{curl}\mathbf{U}. In order for the swirliness to become visually apparent on a small patch, one needs to subtract out the average value of the field.
In a two-dimensional space, the Levi-Civita symbol εij\varepsilon^{ij} has two indices and, thus, there are not enough indices to produce a first-order tensor out of the combination iUj\nabla_{i}U_{j}. Nevertheless, we can still consider the combination εijiUj\varepsilon^{ij}\nabla_{i}U_{j} which produces a scalar invariant VV, i.e.
V=εijiUj.(18.92)V=\varepsilon^{ij}\nabla_{i}U_{j}.\tag{18.92}
As in the case of the three-dimensional curl, the covariant derivatives can be replaced with partial derivatives and the result can be captured by the determinant equation
V=1ZZ1Z2U1U2=1Z(U2Z1U1Z2).(18.93)V=\frac{1}{\sqrt{Z}}\left\vert \begin{array} {cc} \frac{\partial}{\partial Z^{1}} & \frac{\partial}{\partial Z^{2}}\\ U_{1} & U_{2} \end{array} \right\vert =\frac{1}{\sqrt{Z}}\left( \frac{\partial U_{2}}{\partial Z^{1} }-\frac{\partial U_{1}}{\partial Z^{2}}\right) .\tag{18.93}
Much like the curl in three dimensions, VV measures the degree of swirliness in the field U\mathbf{U}. For the vector field
U(x,y)=yi+xj,(18.94)\mathbf{U}\left( x,y\right) =-y\mathbf{i}+x\mathbf{j,}\tag{18.94}
which exemplifies swirliness, VV is given by
V=xyyx=2.(18.95)V=\left\vert \begin{array} {cc} \frac{\partial}{\partial x} & \frac{\partial}{\partial y}\\ -y & x \end{array} \right\vert =2.\tag{18.95}
The identities of Vector Calculus are rarely used in all but the most elementary applications. Most analyses encounter situations not covered by the common Vector Calculus identities involving grad\operatorname{grad}, div\operatorname{div}, curl\operatorname{curl} and the Laplacian. As we already discussed, even the analysis of a combination as fundamental as (UV)\mathbf{\nabla}\left( \mathbf{U}\cdot\mathbf{V}\right) requires the introduction of new operators. Thus, at best, the Vector Calculus identities are incomplete.
In Tensor Calculus, the need for analogous identities simply does not arise. All analyses are built up from the metrics and the covariant derivative, and proceed according to the limited set of rules governing the handful of available operations. It is certainly true that any analysis will reference previously established relationships, such as
εijkεrsk=δirδjsδisδjr.(18.96)\varepsilon_{ijk}\varepsilon^{rsk}=\delta_{i}^{r}\delta_{j}^{s}-\delta_{i} ^{s}\delta_{j}^{r}.\tag{18.96}
However, using such relationships is simply a matter of convenience since all such relationships can be derived from first principles by the same limited set of rules. In other words, none of these relationships can be considered primary in the sense that they must be a priori derived by means outside of the present framework. Thus, every analysis in Tensor Calculus can, and often does, proceed from first principles and is guaranteed not to encounter a combination, such as (UV)\mathbf{\nabla}\left( \mathbf{U} \cdot\mathbf{V}\right) in Vector Calculus, that cannot be reduced to already-available more primary objects. Ultimately, every analysis in a Euclidean space can be performed in terms of the covariant derivative and the metrics, which, in turn, can be expressed in terms of partial derivatives and the position vector. (In a Riemannian space, the covariant metric tensor replaces the position vector as the primary object and therefore every analysis can be reduced to partial derivatives and the metric tensor.)
Nevertheless, for illustrative purposes, we would like to use the techniques of Tensor Calculus to demonstrate some of the most common identities of Vector Calculus, such as
curlgradU=0,          (18.97)divcurlU=0, and          (18.98)curlcurlU=graddivUΔU.          (18.99)\begin{aligned}\operatorname{curl}\operatorname{grad}U & =\mathbf{0}\text{,}\ \ \ \ \ \ \ \ \ \ \left(18.97\right)\\\operatorname{div}\operatorname{curl}\mathbf{U} & =0\text{, and}\ \ \ \ \ \ \ \ \ \ \left(18.98\right)\\\operatorname{curl}\operatorname{curl}\mathbf{U} & =\operatorname{grad} \operatorname{div}\mathbf{U}-\Delta\mathbf{U.}\ \ \ \ \ \ \ \ \ \ \left(18.99\right)\end{aligned}
Other identities that may be stated in dyadic terms but whose proof requires coordinate-based calculations, are left as exercises for the reader.
Let us begin with the first two identities and show that both follow easily from the skew-symmetric property of the Levi-Civita symbol. Since the components of gradU\operatorname{grad}U are jU\nabla_{j}U, the components of curlgradU\operatorname{curl}\operatorname{grad}U are
varepsilon^{ijk}nabla_{i}nabla_{j}U. Recall that, in a Euclidean space, the covariant derivatives commute, i.e. nabla_{i}nabla_{j}U=nabla_{j}nabla_{i}U. Consequently, the preceding expression represents a double contraction between a skew-symmetric system, εijk\varepsilon^{ijk}, and a symmetric system, ijU\nabla_{i}\nabla_{j}U, and therefore vanishes, as we set out to show.
Turning our attention now to divcurlU\operatorname{div}\operatorname{curl}\mathbf{U} , the components of curlU\operatorname{curl}\mathbf{U} are εijkiUj\varepsilon ^{ijk}\nabla_{i}U_{j} and therefore divcurlU\operatorname{div}\operatorname{curl} \mathbf{U} is given by the expression
k(εijkiUj).(18.100)\nabla_{k}\left( \varepsilon^{ijk}\nabla_{i}U_{j}\right) .\tag{18.100}
By the metrinilic property, the Levi-Civita symbol "passes through" the covariant derivative, i.e.
k(εijkiUj)=εijkkiUj,(18.101)\nabla_{k}\left( \varepsilon^{ijk}\nabla_{i}U_{j}\right) =\varepsilon ^{ijk}\nabla_{k}\nabla_{i}U_{j},\tag{18.101}
and we once again have a double contraction (on ii and kk) between a skew-symmetric system, εijk\varepsilon^{ijk}, and a symmetric system, kiUj\nabla_{k}\nabla_{i}U_{j}. Therefore, the result is zero as we set out to show.
Finally, let us consider curlcurlU\operatorname{curl}\operatorname{curl}\mathbf{U}, which will lead to a more intricate analysis. Let
V=curlU     and     W=curlcurlU=curlV.(18.102)\mathbf{V}=\operatorname{curl}\mathbf{U}\text{\ \ \ \ \ and\ \ \ \ \ } \mathbf{W}=\operatorname{curl}\operatorname{curl}\mathbf{U} =\operatorname{curl}\mathbf{V.}\tag{18.102}
The components VkV^{k} of V\mathbf{V} are given by
Vk=εijkiUj.(18.103)V^{k}=\varepsilon^{ijk}\nabla_{i}U_{j}.\tag{18.103}
Thus, the components WsW_{s} of W\mathbf{W} are given by
Ws=εrksr(εijkiUj).(18.104)W_{s}=\varepsilon_{rks}\nabla^{r}\left( \varepsilon^{ijk}\nabla_{i} U_{j}\right) .\tag{18.104}
Once again, the Levi-Civita symbol "passes through" the covariant derivative r\nabla^{r}, i.e.
Ws=εrksr(εijkiUj)=εrksεijkriUj.(18.105)W_{s}=\varepsilon_{rks}\nabla^{r}\left( \varepsilon^{ijk}\nabla_{i} U_{j}\right) =\varepsilon_{rks}\varepsilon^{ijk}\nabla^{r}\nabla_{i}U_{j}.\tag{18.105}
The combination εrksεijk\varepsilon_{rks}\varepsilon^{ijk} is given by
εrksεijk=δrjδsiδriδsj(18.106)\varepsilon_{rks}\varepsilon^{ijk}=\delta_{r}^{j}\delta_{s}^{i}-\delta_{r} ^{i}\delta_{s}^{j}\tag{18.106}
as it is left as an exercise to show. Therefore,
Ws=εrksεijkriUj=(δrjδsiδriδsj)riUj.(18.107)W_{s}=\varepsilon_{rks}\varepsilon^{ijk}\nabla^{r}\nabla_{i}U_{j}=\left( \delta_{r}^{j}\delta_{s}^{i}-\delta_{r}^{i}\delta_{s}^{j}\right) \nabla ^{r}\nabla_{i}U_{j}.\tag{18.107}
Upon applying the distributive law and absorbing the Kronecker symbols, we find
Ws=rsUrrrUs(18.108)W_{s}=\nabla^{r}\nabla_{s}U_{r}-\nabla^{r}\nabla_{r}U_{s}\tag{18.108}
Finally, switching the order of the covariant derivatives in each term (for largely aesthetic reasons), we arrive at the final expression
Ws=srUrrrUs.(18.109)W_{s}=\nabla_{s}\nabla^{r}U_{r}-\nabla_{r}\nabla^{r}U_{s}.\tag{18.109}
Expressed in dyadic terms, the above identity reads
curlcurlU=graddivUΔU,(18.99)\operatorname{curl}\operatorname{curl}\mathbf{U}=\operatorname{grad} \operatorname{div}\mathbf{U}-\Delta\mathbf{U,} \tag{18.99}
as we set out to show.
Exercise 18.1Show that for a vector field U\mathbf{U} and a scalar field ff, the divergence satisfies the product rule
(fU)=fU+fU.(18.110)\mathbf{\nabla}\cdot\left( f\mathbf{U}\right) =\mathbf{\nabla} f\cdot\mathbf{U}+f\mathbf{\nabla}\cdot\mathbf{U.}\tag{18.110}
Exercise 18.2If UU and VV are scalar fields and f(u,v)f\left( u,v\right) is a function of two variables, show that
f(U,V)=fuU+fvV.(18.111)\mathbf{\nabla}f\left( U,V\right) =\frac{\partial f}{\partial u} \mathbf{\nabla}U+\frac{\partial f}{\partial v}\mathbf{\nabla}V.\tag{18.111}
Exercise 18.3By referring to the Christoffel symbol, show that in polar coordinates in a two-dimensional Euclidean space, the divergence is given by
iUi=1rr(rU1)+U2θ.(18.51)\nabla_{i}U^{i}=\frac{1}{r}\frac{\partial}{\partial r}\left( rU^{1}\right) +\frac{\partial U^{2}}{\partial\theta}. \tag{18.51}
Exercise 18.4Show that the Laplacian of the position vector R\mathbf{R} vanishes, i.e.
iiR=0.(18.112)\nabla_{i}\nabla^{i}\mathbf{R}=\mathbf{0.}\tag{18.112}
Exercise 18.5Use the above formula to show that
R=n,(18.113)\mathbf{\nabla}\cdot\mathbf{R}=n,\tag{18.113}
where nn is the dimension of the Euclidean space. For an alternative derivation of this formula, see Exercise 15.5.
Exercise 18.6If rr is the length of the position vector R\mathbf{R}, show that
(Rrn)=0,(18.114)\mathbf{\nabla}\cdot\left( \frac{\mathbf{R}}{r^{n}}\right) =0,\tag{18.114}
where nn is the dimension of the space.
Exercise 18.7Show that for a constant vector field U\mathbf{U},
(UR)=U.(18.115)\mathbf{\nabla}\left( \mathbf{U\cdot R}\right) =\mathbf{U.}\tag{18.115}
Exercise 18.8Show that the covariant derivative applied to a cross product of two fields U\mathbf{U} and V\mathbf{V} satisfies the product rule
i(U×V)=iU×V+U×iV.(18.116)\nabla_{i}\left( \mathbf{U}\times\mathbf{V}\right) =\nabla_{i} \mathbf{U}\times\mathbf{V}+\mathbf{U}\times\nabla_{i}\mathbf{V.}\tag{18.116}

18.11.1Applications of the Voss-Weyl formula

Exercise 18.9Show that in polar coordinates, where
Z=r,(9.60)\sqrt{Z}=r,\tag{9.60}
the expression for the divergence reads
iUi=1r(rU1)r+U2θ.(18.117)\nabla_{i}U^{i}=\frac{1}{r}\frac{\partial\left( rU^{1}\right) }{\partial r}+\frac{\partial U^{2}}{\partial\theta}.\tag{18.117}
Exercise 18.10Show that the Laplacian of the function
U(r,θ)=rncosnθ(18.118)U\left( r,\theta\right) =r^{n}\cos n\theta\tag{18.118}
vanishes.
Exercise 18.11Show that in cylindrical coordinates, where
Z=r,(9.61)\sqrt{Z}=r,\tag{9.61}
the expression for the divergence reads
iUi=1r(rU1)r+U2θ+U3z.(18.119)\nabla_{i}U^{i}=\frac{1}{r}\frac{\partial\left( rU^{1}\right) }{\partial r}+\frac{\partial U^{2}}{\partial\theta}+\frac{\partial U^{3}}{\partial z}.\tag{18.119}
Exercise 18.12Show that in spherical coordinates, where
Z=r2sinθ,(9.62)\sqrt{Z}=r^{2}\sin\theta,\tag{9.62}
the expression for the divergence reads
iUi=1r2(r2U1)r+1sinθ(sinθ U2)θ+U3φ.(18.120)\nabla_{i}U^{i}=\frac{1}{r^{2}}\frac{\partial\left( r^{2}U^{1}\right) }{\partial r}+\frac{1}{\sin\theta}\frac{\partial\left( \sin\theta ~U^{2}\right) }{\partial\theta}+\frac{\partial U^{3}}{\partial\varphi}.\tag{18.120}
Exercise 18.13Show that in cylindrical coordinates, the expression for the Laplacian reads
iiU=1rr(rUr)+1r22Uθ2+2Uz2.(18.121)\nabla_{i}\nabla^{i}U=\frac{1}{r}\frac{\partial}{\partial r}\left( r\frac{\partial U}{\partial r}\right) +\frac{1}{r^{2}}\frac{\partial^{2} U}{\partial\theta^{2}}+\frac{\partial^{2}U}{\partial z^{2}}.\tag{18.121}
Exercise 18.14Show that in spherical coordinates, the expression for the Laplacian reads
iiU=1r2r(r2Ur)+1r2sinθθ(sinθUθ)+1r2sin2θ2Uφ2.(18.122)\nabla_{i}\nabla^{i}U=\frac{1}{r^{2}}\frac{\partial}{\partial r}\left( r^{2}\frac{\partial U}{\partial r}\right) +\frac{1}{r^{2}\sin\theta} \frac{\partial}{\partial\theta}\left( \sin\theta\frac{\partial U} {\partial\theta}\right) +\frac{1}{r^{2}\sin^{2}\theta}\frac{\partial^{2} U}{\partial\varphi^{2}}.\tag{18.122}

18.11.2Other exercises{}

Exercise 18.15In three dimensions, show that
×R=0,(18.123)\mathbf{\nabla}\times\mathbf{R}=\mathbf{0,}\tag{18.123}
where R\mathbf{R} is the position vector. Meanwhile, in two dimensions,
×R=0.(18.124)\nabla\times\mathbf{R}=0\mathbf{.}\tag{18.124}
Exercise 18.16Show that for a constant vector field U\mathbf{U},
×(U×R)=2U.(18.125)\mathbf{\nabla}\times\left( \mathbf{U\times R}\right) =2\mathbf{U.}\tag{18.125}
Exercise 18.17Show the identity
εrksεijk=δrjδsiδriδsj(18.106)\varepsilon_{rks}\varepsilon^{ijk}=\delta_{r}^{j}\delta_{s}^{i}-\delta_{r} ^{i}\delta_{s}^{j} \tag{18.106}
from Section 18.10 by justifying each step in the following chain of identities:
εrksεijk=δrksijk=δrskijk=δrsij=δrjδsiδriδsj.(18.126)\varepsilon_{rks}\varepsilon^{ijk}=\delta_{rks}^{ijk}=-\delta_{rsk} ^{ijk}=-\delta_{rs}^{ij}=\delta_{r}^{j}\delta_{s}^{i}-\delta_{r}^{i}\delta _{s}^{j}.\tag{18.126}
Exercise 18.18Show that in a two-dimensional Euclidean space,
εijijU=0.(18.127)\varepsilon^{ij}\nabla_{i}\nabla_{j}U=0.\tag{18.127}
In words, the curl of the gradient is zero.
Exercise 18.19Show that the one-dimensional "curl" defined as
εii(18.128)\varepsilon^{i}\nabla_{i}\tag{18.128}
corresponds to d/dsd/ds, i.e. the derivative with respect to the arc-length.
Exercise 18.20Explain why the invariant combination
εijij(18.129)\varepsilon^{ij}\nabla_{i}\nabla_{j}\tag{18.129}
does not lead to an interesting new differential operator.
Exercise 18.21Decipher and then derive the following identities.
U×(×V)=V(UV)(U)V          (18.130)(UV)=U×(×V)+V×(×U)+(U)V+(V)U          (18.131)(U×V)=(×U)V(×V)U          (18.132)×(U×V)=(V)UV(U)+U(V)(U)V          (18.133)\begin{aligned}\mathbf{U}\times\left( \mathbf{\nabla}\times\mathbf{V}\right) & =\mathbf{\nabla}_{\mathbf{V}}\left( \mathbf{U}\cdot\mathbf{V}\right) -\left( \mathbf{U}\cdot\mathbf{\nabla}\right) \mathbf{V}\ \ \ \ \ \ \ \ \ \ \left(18.130\right)\\\mathbf{\nabla}\left( \mathbf{U}\cdot\mathbf{V}\right) & =\mathbf{U} \times\left( \mathbf{\nabla}\times\mathbf{V}\right) +\mathbf{V}\times\left( \mathbf{\nabla}\times\mathbf{U}\right) +\left( \mathbf{U}\cdot \mathbf{\nabla}\right) \mathbf{V}+\left( \mathbf{V}\cdot\mathbf{\nabla }\right) \mathbf{U}\ \ \ \ \ \ \ \ \ \ \left(18.131\right)\\\mathbf{\nabla}\cdot\left( \mathbf{U}\times\mathbf{V}\right) & =\left( \mathbf{\nabla}\times\mathbf{U}\right) \cdot\mathbf{V}-\left( \mathbf{\nabla }\times\mathbf{V}\right) \cdot\mathbf{U}\ \ \ \ \ \ \ \ \ \ \left(18.132\right)\\\mathbf{\nabla}\times\left( \mathbf{U}\times\mathbf{V}\right) & =\left( \mathbf{V}\cdot\mathbf{\nabla}\right) \mathbf{U}-\mathbf{V}\left( \mathbf{\nabla}\cdot\mathbf{U}\right) +\mathbf{U}\left( \mathbf{\nabla} \cdot\mathbf{V}\right) -\left( \mathbf{U}\cdot\mathbf{\nabla}\right) \mathbf{V}\ \ \ \ \ \ \ \ \ \ \left(18.133\right)\end{aligned}
Send feedback to Pavel