Stokes' Theorem on Smooth Manifolds
Total Page:16
File Type:pdf, Size:1020Kb
Stokes’ Theorem on Smooth Manifolds Erik Jörgenfelt Spring 2016 Thesis, 15hp Bachelor of mathematics, 180hp Department of mathematics and mathematical statistics Abstract A proof of Stokes’ theorem on smooth manifolds is given, complete with prerequisite re- sults in tensor algebra and differential geometry. The essay assumes familiarity with multi- variable calculus and linear algebra, as well as a basic understanding of point-set topology. Stokes’ theorem is then applied to the conservation of energy-momentum in general relativity under the existence of so called Killing vectors. Sammanfattning Stokes sats för släta mångfalder bevisas, komplett med nödvändiga resultat från tensoral- gebran och differentialgeometrin. Uppsatsen förutsätter förtrogenhet med flervariabelanalys och linjär algebra, samt en grundläggande förståelse för allmän topologi. Stokes sats appli- ceras sedan till bevarande av energi-momentum i allmän relativitetsteori under existensen av so kallade Killingvektorer. Contents 1 Introduction 1 2 Preliminaries 3 2.1 Tensors.........................................3 2.2 Smooth Manifolds with Boundary..........................9 2.3 Vector Bundles..................................... 14 3 Partitions of Unity 17 4 The Tangent Bundle 25 5 Differential Forms and Stokes’ Theorem 35 5.1 Differential Forms and Integration.......................... 35 5.2 Stokes’ Theorem on Smooth Manifolds........................ 47 6 Conservation of Energy-Momentum 50 6.1 Regular Submanifolds................................. 51 6.2 The Tensor Bundles.................................. 51 6.3 Lie Derivative...................................... 53 6.4 Lorentz Manifolds................................... 54 6.5 Connection....................................... 56 6.6 Hodge Duality..................................... 59 6.7 Conservation of Energy-Momentum and Killing Vectors.............. 60 1 Introduction Stokes’ theorem on smooth manifolds subsumes three important theorems from standard calculus, unifying them and extending the application to any number of dimensions. It does this in the language of differential geometry. According to Katz[6] the unified theorem was first published in 1945 by Élie Cartan in [1], although several previous versions are also listed. It lends its name from the last of the subsumed theorems, which was published by George Stokes as a problem in the Smith’s Prize Exam at Cambridge in 1854[13]. According to Katz he had received a letter with the theorem by Lord Kelvin on July 2. Today we know this theorem as Stokes’ theorem, the Kelvin-Stokes theorem, or the curl theorem. The first published proof is however, again according to Katz, by Hermann Hankel in 1861[4]. The other subsumed theorems are today referred to as Gauss’ theorem or the divergence theorem, and Green’s theorem. Although Gauss did publish three special cases of the former[3], Katz lists Michael Ostrogradsky[10] as the first to state and prove the theorem in a lecture, not published until later, and publish it first in 1831[11]. Green’s theorem was stated first without proof by Augustin Cauchy[2] in 1846, but proved five years later by Bernhard Riemann[12]. Katz also notes that all the aforementioned authors, as well as others publishing related results, were interested in them for applications in physics. Our main theorem, stated below, among other things allows for native application in the formalism of space-time — a 4-dimensional Lorentz manifold — where it is frequently used to work with flux integrals and conservation theorems. In [5] Harrison produces a Stokes’ theorem for non-smooth chains, thus building on the work of Whitney[16], who used TheoremA to define integration over certain non-smooth domains. TheoremA (Stokes’ theorem on smooth manifolds). For any smooth (n − 1)-form ω with compact support on the oriented n-dimensional smooth manifold M with boundary ∂M, we have that Z Z dω = ω, M ∂M where ∂M is given the induced orientation. This essay is structured as follows. In Section2 we cover preliminary results in tensor al- gebra, and define smooth manifolds with boundary and vector bundles. In Section3 we show the existence of a partition of unity subordinate to any open cover of a smooth manifold. In Section4 we construct the tangent bundle over a smooth manifold. In Section5 we construct smooth differential forms, define oriented manifolds, and define integration over oriented smooth manifolds with boundary. We then prove TheoremA. In Section6 we apply TheoremA in the context of general relativity to prove TheoremB: TheoremB. Let S be a Lorentz manifold with the Levi-Civita connection, with dim(S) = n, which we will refer to as spacetime, endowed with a stress-energy tensor, T . If ξ is a Killing i i j vector then J := Tj ξ is a conserved flux, i.e. I i J dΣi ≡ 0, Σ for any sufficiently nice closed hypersurface Σ. This essay is based on [7], with some inspiration from [14] and [8]. The proof of TheoremB is inspired by [8]. We prove several supporting results that were left without proof in [7], among 1 these Proposition 2.20, Proposition 2.27, and Proposition 5.1; and state and prove Proposition 5.16, which was not included in [7]. Additionally, our definition of smooth manifolds makes the stronger assumption of second countability over paracompactness, which [7] uses. Second countability is necessary in some other applications such as Sard’s theorem. We thus proceed to show, in Proposition 3.3, that our definition implies paracompactness. Furthermore, we make no mention of what Lee refers to as TpMphys, or roughly “the tangent space according to physicists,” defined via transformation laws. Instead we use what Lee refers to as “the kinematic tangent space,” or TpMkin, here referred to as simply TpM, defined via curves, to define the tangent space at the boundary (Definition 4.15), and in Lemma 5.18. This essay also expands on any arguments in [7] that the author found abstruse; either because they presumed a greater familiarity with the subject than can be expected at this level, or because Lee intentionally left parts of the argument to the reader. Finally, the focus on TheoremA allows the material to be restructured to provide a (hopefully) clearer line of reasoning. 2 Preliminaries 2.1 Tensors Here we study alternating covariant tensors. Tensors can be thought of as generalisations of vectors, in as much as it is linear in each argument, and indeed all vectors are also tensors. We will not concern ourselves with this fact here, but rather construct our tensors from vectors and dual vectors, the latter of which are elements of the dual of a vector space. We therefore first define the dual of a vector space, before introducing consolidated tensors. Although we quickly move on to consider only alternating covariant tensors, the dual space will appear again in Section4 in the form of the cotangent space. In Section6 we briefly touch on more general tensors again. Alternating covariant tensors will form the basis which we build our definition of differential forms (Section5) on. As such they are crucial to our development of TheoremA. The exterior product of two alternating covariant tensors is defined and explored to sufficient depth to provide a basis for each space of alternating covariant tensors, and at the end this allows us to give an alternate definition of the determinant of a linear transformation on a vector space and show that it is equivalent to the matrix definition. The last result not only provides a satisfying link between these results and ones known from elementary linear algebra, but will also be used in Section5 when we investigate oriented manifolds. We begin with the definition of the dual of a vector space, and the dual of any basis: Definition 2.1. The dual space of a vector space V is the set of all linear maps ω : V → R. ∗ We denote it by V , and call its elements dual vectors. If (e1, . , en) is a basis of V , and 1 n ∗ i i 1 n ω , . , ω are elements of V such that ω (ej) = δj, then ω , . , ω is called the dual basis of (e1, . , en). That ω1, . , ωn is a basis of V ∗ follows from the linearity of its elements, and we will immediately move on to define (consolidated) tensors over vector spaces: Definition 2.2. A (consolidated) tensor on a real valued vector space V is a multilinear map ∗ ∗ t : V1 × ... × Vr × V1 × ... × Vs → R, ∗ ∗ where all Vi = V and all Vj = V . We say that t is contravariant of dimension r and covariant r of dimension s, and write t ∈ Ts (V ). In fact, once we have defined the tensor product and shown its associativity — a property we should expect from a product operator — we will only be dealing with covariant tensors, i.e. 0 tensors that are contravariant of dimension zero, denoted Ts (V ) or just Ts(V ). As an example of a covariant tensor, consider the familiar inner product (or “dot-product”) on Rn. It takes two vectors as arguments, is linear in each argument, and returns a real number — a two-dimensional, covariant tensor. The tensor product combines two tensors to produce a third. In principle it adjoins the two tensors. As such it can easily be seen to have an identity element, but is obviously not commutative. Definition 2.3. For tensors t ∈ T r1 (V ) and t ∈ T r2 (V ), we define the tensor product t ⊗t ∈ 1 s1 2 s2 1 2 T r1+r2 by s1+s2 1 r1+r2 t1 ⊗ t2(θ , . , θ ,v1, . vs1+s2 ) := 1 r1 r1+1 r1+r2 t1(θ , . , θ , v1, . , vs1 )t2(θ , . , θ , vs1+1, . , vs1+s2 ). 3 Proposition 2.4. Let t ∈ T r1 (V ), t ∈ T r2 (V ), t ∈ T r3 (V ). Then 1 s1 2 s2 3 s3 (t1 ⊗ t2) ⊗ t3 = t1 ⊗ (t2 ⊗ t3). ri ,...,ri Proof. Applying both sides to (α1, . , αr1+r2+r3 , v , . , v ) and taking t 1 k to mean 1 s1+s2+s3 sj1 ,...,sj` r r t α i1 , .