PROOF of DE RHAM's THEOREM 1. Introduction Let M Be a Smooth N

Total Page:16

File Type:pdf, Size:1020Kb

PROOF of DE RHAM's THEOREM 1. Introduction Let M Be a Smooth N PROOF OF DE RHAM'S THEOREM PETER S. PARK 1. Introduction Let M be a smooth n-dimensional manifold. Then, de Rham's theorem states that the de Rham cohomology of M is naturally isomorphic to its singular cohomology with coefficients in R; in particular, de Rham cohomology is a purely topological invariant. This fact is a manifestation of a fundamental relationship between the analytical and topological characteristics of a smooth manifold, where the former include information about solutions to differential equations of the formulation dη = ! for closed forms !, and the latter, information about the \holes" in each given dimension, which is made precise by singular homology. The natural isomorphism will be given by a version of Stokes' theorem, which describes a duality between de Rham cohomology and singular homology. Specifically, the pairing of differential forms and singular chains, which can taken to be smooth, yields a map from the kth de Rham cohomology group to the kth singular cohomology group with coefficients in R for each k ≥ 0. And in 1931, Georges de Rham himself proved that this map is in fact an isomorphism. In this exposition, we will give a proof1 of this beautiful and fundamental result. 2. Smooth Singular Homology Our desired relationship between singular cohomology and de Rham cohomology will be given by the integration of differential forms over singular chains. Specifically, given a singular k-simplex ∗ σ : ∆k ! M and a k-form ! on a manifold M, we wish to integrate over ∆k the pullback form σ !. However, pulling back differential forms requires the map in question to be sufficiently smooth, yet as maps, singular k-simplices are only continuous in general. To address this issue, we will show that singular homology can be computed using only smooth simplices instead of all continuous ones. m For a subset C ⊂ R , we say that a continuous map f : C ! M is smooth at a point p 2 C if there exists an open neighborhood of p in which f has a smooth extension. Furthermore, we say that f is smooth if it is smooth at every point p 2 C. Then, specializing to the case that C is m m equal to the standard k-simplex of R , defined by ∆k ·= [δ0; : : : ; δk] ⊂ R for the standard basis vectors δi, we define a smooth k-simplex of M to be a k-simplex σ : ∆k ! M that is smooth as a 1 map. For the usual free abelian group Ck(M) on all k-chains of M, we define Ck (M) to be the subgroup of Ck(M) generated by the smooth k-simplices of M. Define the smooth k-chains of M 1 to be the elements of Ck (M), i.e., formal sums of finitely many smooth k-simplices. It is clear from the above definitions that the boundary of a smooth (k + 1)-simplex is a smooth k-chain, so the kth smooth singular homology group 1 1 1 Ker(@ : Ck (M) ! Ck−1(M)) Hk (M) = 1 1 Im(@ : Ck+1(M) ! Ck (M)) 1 is well-defined. Since the inclusion i : Ck (M),!Ck(M) satisfies ι ◦ @ = @ ◦ ι, we see that it induces 1 a map ι∗ : Hk (M) ! Hk(M) defined by ι∗[c] = [ι(c)]. 1This exposition is based on the proof provided in [2, Chapter 18], which in turn is based on that given in [1, Section V.9]. For another proof of de Rham's theorem based on sheaf theory, see [3, Chapter 5]. 1 2 PETER S. PARK We would like to show that this map ι∗ is an isomorphism. We will first need to construct for an arbitrary simplex a homotopy to a smooth simplex such that when the homotopy for any simplex is restricted to a boundary face, this yields the homotopy of the simplex given by the boundary face. To do so, we will need the following lemma showing that for a map whose domain is the standard k-simplex, smoothness on each boundary face implies smoothness on the entire boundary simultaneously. m Lemma 2.1. For the standard simplex ∆k ⊂ R , let f : @∆k ! M be a continuous map whose restriction to each boundary face is smooth. Then, f is smooth as a map from the entire boundary @∆k to M. Proof. Recall the notation ∆k = [δ0; : : : ; δk], where the standard basis vectors δ0; : : : ; δk are the ^ vertices of the simplex. For 0 ≤ i ≤ k, let @i∆k = [δ0;:::; δi; : : : ; δk] denote the boundary face opposite of δi. Our hypothesis then is precisely that for every i and x 2 @i∆k, there exists an open ~ ~ neighborhood Ux 3 x and a smooth extension f : Ux ! M of f, i.e., f restricted to Ux \ @i∆k ~ gives f. We then need to show that for every x 2 @∆k, a smooth extension f can be chosen for all ~ boundary faces, i.e., such that f restricted to Ux \ @∆k coincides with f. Since x is contained in at least one boundary face of ∆k but cannot be contained in all of them, we can without loss of generality suppose that the boundary faces containing x are @1∆k;:::;@m∆k for some 1 ≤ j ≤ k, and in particular, that x is not contained in @0∆k. By our hypothesis, there ~ exists an open neighborhood Ui 3 x and a smooth map fi : Ui ! M whose restriction to Ui \ @i∆k Tj coincides with f. Let U = i=1 Ui. We now prove by induction on j that there exists a smooth ~ Sj map f : U ! M whose restriction to U \ i=1 @i∆k coincides with f. Since our claim is local, we can, by replacing U with a subneighborhood if necessary, suppose that f maps U to a coordinate n chart of f(x) that can be diffeomorphically identified with R . The base case j = 1 is trivial. Now, suppose that j > 1 and that there exists a smooth map f~0 : U ! M whose restriction to Sj U \ i=1 @i∆k coincides with f. Note that the boundary face @i∆k is precisely the intersection ~ of the hyperplane xi = 0 with @∆k. Consistent with this, define f : U ! M by 1 m 1 m 1 j−1 j+1 m 1 j−1 j+1 m f~(x ; : : : ; x ) = f~0(x ; : : : ; x )−f~0(x ; : : : ; x ; 0; x ; : : : ; x )+f~j(x ; : : : ; x ; 0; x ; : : : ; x ): ~ For every 1 ≤ i ≤ j − 1, the restriction of f to U \ @i∆k equals f~(x1; : : : ; xi−1; 0; xi+1; : : : ; xn) 1 i−1 i+1 n 1 i−1 i+1 j−1 j+1 n = f~0(x ; : : : ; x ; 0; x ; : : : ; x ) − f~0(x ; : : : ; x ; 0; x ; : : : ; x ; 0; x ; : : : ; x ) 1 i−1 i+1 j−1 j+1 n + f~j(x ; : : : ; x ; 0; x ; : : : ; x ; 0; x ; : : : ; x ) = f(x1; : : : ; xi−1; 0; xi+1; : : : ; xn); ~ i ~ since f0 coincides with f for x 2 ∆k satisfying x = 0 and similarly fj coincides with f for x 2 ∆k j ~ satisfying x = 0. An analogous argument shows that the restriction of f to U \ @j∆k is f~(x1; : : : ; xj−1; 0; xj+1; : : : ; xn) 1 j−1 j+1 n 1 j−1 j+1 n = f~0(x ; : : : ; x ; 0; x ; : : : ; x ) − f~0(x ; : : : ; x ; 0; x ; : : : ; x ) 1 j−1 j+1 n + f~j(x ; : : : ; x ; 0; x ; : : : ; x ) = f(x1; : : : ; xj−1; 0; xj+1; : : : ; xn); which proves the inductive step as needed. We will also need the following, often called Whitney's approximation theorem. PROOF OF DE RHAM'S THEOREM 3 Theorem 2.2 (Whitney). Let X be a smooth manifold with or without boundary, Y a smooth manifold without boundary, and f : X ! Y a continuous map smooth on a closed subset C ⊂ X. We have that F is homotopic relative to C to a smooth map X ! Y . Proof. [2, Theorem 6.26]. Now, we are ready to prove our desired lemma that constructs, for any simplex, a homotopy from it to a smooth simplex such that the homotopies respect restriction to boundary faces. k Lemma 2.3. For every k-simplex σ : ∆k ! M from the standard k-simplex of R to M, there exists a continuous map Hσ : ∆k × I ! M satisfying the following: (i) Hσ is a homotopy from σ(x) = Hσ(x; 0) to a smooth k-simplex σ~(x) = Hσ(x; 1). ∼ (ii) For each boundary face inclusion Fi;k : ∆k−1 = @i∆k,!∆k, Hσ◦Fi;k (x; t) = Hσ(Fi;k(x); t) for (x; t) 2 ∆k−1 × I. (iii) For σ a smooth k-simplex, Hσ is the constant homotopy, i.e., Hσ(x; t) = σ(x). Proof. For σ smooth, Hσ can be taken to be the constant homotopy, which is easily seen to satisfy properties (1) and (2). We prove the problem statement for not necessarily smooth σ by induction on the dimension of σ. An arbitrary 0-simplex σ : ∆0 ! M is smooth, so the base case is trivial. Suppose now that for all ` < j, we have constructed for each `-simplex σ0 a corresponding homotopy Hσ0 satisfying the lemma's properties. Consider an arbitrary non-smooth j-simplex σ : ∆k ! M. Define the subset S ·= (∆k × f0g) [ (@∆k × I) ⊂ ∆k × I, and define H0 : S ! M by ( σ(x) if x 2 ∆ ; t = 0; · k H0(x; t) ·= −1 Hσ◦Fi;k (Fi;k (x); t) if x 2 @i∆k: To show that H0 is continuous, we need to show that the functions of the above cases coincide on overlaps so that we can apply the gluing lemma.
Recommended publications
  • Arxiv:2002.06802V3 [Math.AT] 1 Apr 2021 Aao3082,Jpne-Mail: Japan 390-8621, Nagano E Od N Phrases
    A COMPARISON BETWEEN TWO DE RHAM COMPLEXES IN DIFFEOLOGY KATSUHIKO KURIBAYASHI Abstract. There are two de Rham complexes in diffeology. The original one is due to Souriau and the other one is the singular de Rham complex defined by a simplicial differential graded algebra. We compare the first de Rham cohomology groups of the two complexes within the Cech–deˇ Rham spectral sequence by making use of the factor map which connects the two de Rham complexes. As a consequence, it follows that the singular de Rham cohomology algebra of the irrational torus Tθ is isomorphic to the tensor product of the original de Rham cohomology and the exterior algebra generated by a non- trivial flow bundle over Tθ. 1. Introduction The de Rham complex introduced by Souriau [13] is very beneficial in the study of diffeology; see [6, Chapters 6,7,8 and 9]. In fact, the de Rham calculus is applicable to not only diffeological path spaces but also more general mapping spaces. It is worth mentioning that the de Rham complex is a variant of the codomain of Chen’s iterated integral map [3]. While the complex is isomorphic to the usual de Rham complex if the input diffeological space is a manifold, the de Rham theorem does not hold in general. In [11], we introduced another cochain algebra called the singular de Rham com- plex via the context of simplicial sets. It is regarded as a variant of the cubic de Rham complex introduced by Iwase and Izumida in [9] and a diffeological counter- part of the singular de Rham complex in [1, 15, 16].
    [Show full text]
  • Lecture 15. De Rham Cohomology
    Lecture 15. de Rham cohomology In this lecture we will show how differential forms can be used to define topo- logical invariants of manifolds. This is closely related to other constructions in algebraic topology such as simplicial homology and cohomology, singular homology and cohomology, and Cechˇ cohomology. 15.1 Cocycles and coboundaries Let us first note some applications of Stokes’ theorem: Let ω be a k-form on a differentiable manifold M.For any oriented k-dimensional compact sub- manifold Σ of M, this gives us a real number by integration: " ω : Σ → ω. Σ (Here we really mean the integral over Σ of the form obtained by pulling back ω under the inclusion map). Now suppose we have two such submanifolds, Σ0 and Σ1, which are (smoothly) homotopic. That is, we have a smooth map F : Σ × [0, 1] → M with F |Σ×{i} an immersion describing Σi for i =0, 1. Then d(F∗ω)isa (k + 1)-form on the (k + 1)-dimensional oriented manifold with boundary Σ × [0, 1], and Stokes’ theorem gives " " " d(F∗ω)= ω − ω. Σ×[0,1] Σ1 Σ1 In particular, if dω =0,then d(F∗ω)=F∗(dω)=0, and we deduce that ω = ω. Σ1 Σ0 This says that k-forms with exterior derivative zero give a well-defined functional on homotopy classes of compact oriented k-dimensional submani- folds of M. We know some examples of k-forms with exterior derivative zero, namely those of the form ω = dη for some (k − 1)-form η. But Stokes’ theorem then gives that Σ ω = Σ dη =0,sointhese cases the functional we defined on homotopy classes of submanifolds is trivial.
    [Show full text]
  • Bibliography
    Bibliography [AK98] V. I. Arnold and B. A. Khesin, Topological methods in hydrodynamics, Springer- Verlag, New York, 1998. [AL65] Holt Ashley and Marten Landahl, Aerodynamics of wings and bodies, Addison- Wesley, Reading, MA, 1965, Section 2-7. [Alt55] M. Altman, A generalization of Newton's method, Bulletin de l'academie Polonaise des sciences III (1955), no. 4, 189{193, Cl.III. [Arm83] M.A. Armstrong, Basic topology, Springer-Verlag, New York, 1983. [Bat10] H. Bateman, The transformation of the electrodynamical equations, Proc. Lond. Math. Soc., II, vol. 8, 1910, pp. 223{264. [BB69] N. Balabanian and T.A. Bickart, Electrical network theory, John Wiley, New York, 1969. [BLG70] N. N. Balasubramanian, J. W. Lynn, and D. P. Sen Gupta, Differential forms on electromagnetic networks, Butterworths, London, 1970. [Bos81] A. Bossavit, On the numerical analysis of eddy-current problems, Computer Methods in Applied Mechanics and Engineering 27 (1981), 303{318. [Bos82] A. Bossavit, On finite elements for the electricity equation, The Mathematics of Fi- nite Elements and Applications IV (MAFELAP 81) (J.R. Whiteman, ed.), Academic Press, 1982, pp. 85{91. [Bos98] , Computational electromagnetism: Variational formulations, complementar- ity, edge elements, Academic Press, San Diego, 1998. [Bra66] F.H. Branin, The algebraic-topological basis for network analogies and the vector cal- culus, Proc. Symp. Generalised Networks, Microwave Research, Institute Symposium Series, vol. 16, Polytechnic Institute of Brooklyn, April 1966, pp. 453{491. [Bra77] , The network concept as a unifying principle in engineering and physical sciences, Problem Analysis in Science and Engineering (K. Husseyin F.H. Branin Jr., ed.), Academic Press, New York, 1977.
    [Show full text]
  • Universal Coefficient Theorem for Homology
    Universal Coefficient Theorem for Homology We present a direct proof of the universal coefficient theorem for homology that is simpler and shorter than the standard proof. Theorem 1 Given a chain complex C in which each Cn is free abelian, and a coefficient group G, we have for each n the natural short exact sequence 0 −−→ Hn(C) ⊗ G −−→ Hn(C ⊗ G) −−→ Tor(Hn−1(C),G) −−→ 0, (2) which splits (but not naturally). In particular, this applies immediately to singular homology. Theorem 3 Given a pair of spaces (X, A) and a coefficient group G, we have for each n the natural short exact sequence 0 −−→ Hn(X, A) ⊗ G −−→ Hn(X, A; G) −−→ Tor(Hn−1(X, A),G) −−→ 0, which splits (but not naturally). We shall derive diagram (2) as an instance of the following elementary result. Lemma 4 Given homomorphisms f: K → L and g: L → M of abelian groups, with a splitting homomorphism s: L → K such that f ◦ s = idL, we have the split short exact sequence ⊂ f 0 0 −−→ Ker f −−→ Ker(g ◦ f) −−→ Ker g −−→ 0, (5) where f 0 = f| Ker(g ◦ f), with the splitting s0 = s| Ker g: Ker g → Ker(g ◦ f). Proof We note that f 0 and s0 are defined, as f(Ker(g ◦ f)) ⊂ Ker g and s(Ker g) ⊂ Ker(g ◦f). (In detail, if l ∈ Ker g,(g ◦f)sl = gfsl = gl = 0 shows that sl ∈ Ker(g ◦f).) 0 0 0 Then f ◦ s = idL restricts to f ◦ s = id. Since Ker f ⊂ Ker(g ◦ f), we have Ker f = Ker f ∩ Ker(g ◦ f) = Ker f.
    [Show full text]
  • Introduction to Homology
    Introduction to Homology Matthew Lerner-Brecher and Koh Yamakawa March 28, 2019 Contents 1 Homology 1 1.1 Simplices: a Review . .2 1.2 ∆ Simplices: not a Review . .2 1.3 Boundary Operator . .3 1.4 Simplicial Homology: DEF not a Review . .4 1.5 Singular Homology . .5 2 Higher Homotopy Groups and Hurweicz Theorem 5 3 Exact Sequences 5 3.1 Key Definitions . .5 3.2 Recreating Groups From Exact Sequences . .6 4 Long Exact Homology Sequences 7 4.1 Exact Sequences of Chain Complexes . .7 4.2 Relative Homology Groups . .8 4.3 The Excision Theorems . .8 4.4 Mayer-Vietoris Sequence . .9 4.5 Application . .9 1 Homology What is Homology? To put it simply, we use Homology to count the number of n dimensional holes in a topological space! In general, our approach will be to add a structure on a space or object ( and thus a topology ) and figure out what subsets of the space are cycles, then sort through those subsets that are holes. Of course, as many properties we care about in topology, this property is invariant under homotopy equivalence. This is the slightly weaker than homeomorphism which we before said gave us the same fundamental group. 1 Figure 1: Hatcher p.100 Just for reference to you, I will simply define the nth Homology of a topological space X. Hn(X) = ker @n=Im@n−1 which, as we have said before, is the group of n-holes. 1.1 Simplices: a Review k+1 Just for your sake, we review what standard K simplices are, as embedded inside ( or living in ) R ( n ) k X X ∆ = [v0; : : : ; vk] = xivi such that xk = 1 i=0 For example, the 0 simplex is a point, the 1 simplex is a line, the 2 simplex is a triangle, the 3 simplex is a tetrahedron.
    [Show full text]
  • Homology Groups of Homeomorphic Topological Spaces
    An Introduction to Homology Prerna Nadathur August 16, 2007 Abstract This paper explores the basic ideas of simplicial structures that lead to simplicial homology theory, and introduces singular homology in order to demonstrate the equivalence of homology groups of homeomorphic topological spaces. It concludes with a proof of the equivalence of simplicial and singular homology groups. Contents 1 Simplices and Simplicial Complexes 1 2 Homology Groups 2 3 Singular Homology 8 4 Chain Complexes, Exact Sequences, and Relative Homology Groups 9 ∆ 5 The Equivalence of H n and Hn 13 1 Simplices and Simplicial Complexes Definition 1.1. The n-simplex, ∆n, is the simplest geometric figure determined by a collection of n n + 1 points in Euclidean space R . Geometrically, it can be thought of as the complete graph on (n + 1) vertices, which is solid in n dimensions. Figure 1: Some simplices Extrapolating from Figure 1, we see that the 3-simplex is a tetrahedron. Note: The n-simplex is topologically equivalent to Dn, the n-ball. Definition 1.2. An n-face of a simplex is a subset of the set of vertices of the simplex with order n + 1. The faces of an n-simplex with dimension less than n are called its proper faces. 1 Two simplices are said to be properly situated if their intersection is either empty or a face of both simplices (i.e., a simplex itself). By \gluing" (identifying) simplices along entire faces, we get what are known as simplicial complexes. More formally: Definition 1.3. A simplicial complex K is a finite set of simplices satisfying the following condi- tions: 1 For all simplices A 2 K with α a face of A, we have α 2 K.
    [Show full text]
  • The Decomposition Theorem, Perverse Sheaves and the Topology Of
    The decomposition theorem, perverse sheaves and the topology of algebraic maps Mark Andrea A. de Cataldo and Luca Migliorini∗ Abstract We give a motivated introduction to the theory of perverse sheaves, culminating in the decomposition theorem of Beilinson, Bernstein, Deligne and Gabber. A goal of this survey is to show how the theory develops naturally from classical constructions used in the study of topological properties of algebraic varieties. While most proofs are omitted, we discuss several approaches to the decomposition theorem, indicate some important applications and examples. Contents 1 Overview 3 1.1 The topology of complex projective manifolds: Lefschetz and Hodge theorems 4 1.2 Families of smooth projective varieties . ........ 5 1.3 Singular algebraic varieties . ..... 7 1.4 Decomposition and hard Lefschetz in intersection cohomology . 8 1.5 Crash course on sheaves and derived categories . ........ 9 1.6 Decomposition, semisimplicity and relative hard Lefschetz theorems . 13 1.7 InvariantCycletheorems . 15 1.8 Afewexamples.................................. 16 1.9 The decomposition theorem and mixed Hodge structures . ......... 17 1.10 Historicalandotherremarks . 18 arXiv:0712.0349v2 [math.AG] 16 Apr 2009 2 Perverse sheaves 20 2.1 Intersection cohomology . 21 2.2 Examples of intersection cohomology . ...... 22 2.3 Definition and first properties of perverse sheaves . .......... 24 2.4 Theperversefiltration . .. .. .. .. .. .. .. 28 2.5 Perversecohomology .............................. 28 2.6 t-exactness and the Lefschetz hyperplane theorem . ...... 30 2.7 Intermediateextensions . 31 ∗Partially supported by GNSAGA and PRIN 2007 project “Spazi di moduli e teoria di Lie” 1 3 Three approaches to the decomposition theorem 33 3.1 The proof of Beilinson, Bernstein, Deligne and Gabber .
    [Show full text]
  • Quelques Souvenirs Des Années 1925-1950 Cahiers Du Séminaire D’Histoire Des Mathématiques 1Re Série, Tome 1 (1980), P
    CAHIERS DU SÉMINAIRE D’HISTOIRE DES MATHÉMATIQUES GEORGES DE RHAM Quelques souvenirs des années 1925-1950 Cahiers du séminaire d’histoire des mathématiques 1re série, tome 1 (1980), p. 19-36 <http://www.numdam.org/item?id=CSHM_1980__1__19_0> © Cahiers du séminaire d’histoire des mathématiques, 1980, tous droits réservés. L’accès aux archives de la revue « Cahiers du séminaire d’histoire des mathématiques » im- plique l’accord avec les conditions générales d’utilisation (http://www.numdam.org/conditions). Toute utilisation commerciale ou impression systématique est constitutive d’une infraction pé- nale. Toute copie ou impression de ce fichier doit contenir la présente mention de copyright. Article numérisé dans le cadre du programme Numérisation de documents anciens mathématiques http://www.numdam.org/ - 19 - QUELQUES SOUVENIRS DES ANNEES 1925-1950 par Georges de RHAM Arrivé à la fin de ma carrière, je pense à son début. C’est dans ma année 1. en 1924, que j’ai décidé de me lancer dans les mathématiques. En 1921, ayant le bachot classique, avec latin et grec, attiré par la philosophie, j’hésitais d’entrer à la Facul- té des Lettres. Mais finalement je me décidai pour la Faculté des Sciences, avec à mon programme l’étude de la Chimie, de la Physique et surtout, pour finir, la Biologie. Je ne songeais pas aux Mathématiques, qui me semblaient un domaine fermé où je ne pourrais rien faire. Pourtant, pour comprendre des questions de Physique, je suis amené à ouvrir des livres de Mathématiques supérieures. J’entrevois qu’il y a là un domaine immense, qui ex- cité ma curiosité et m’intéresse à tel point qu’après cinq semestres à l’Université, j’ abandonne la Biologie pour aborder résolument les Mathématiques.
    [Show full text]
  • Arxiv:2101.10594V1 [Hep-Th] 26 Jan 2021 Rpsda H Mrec Fti Dai Re Ormf H S the Ramify to Order in Gravity
    [math-ph] ⋆-Cohomology, Connes-Chern Characters, and Anomalies in General Translation-Invariant Noncommutative Yang-Mills Amir Abbass Varshovi∗ Department of Mathematics, University of Isfahan, Isfahan, IRAN. School of Mathematics, Institute for Research in Fundamental Sciences (IPM), Tehran, IRAN. Abstract: Topological structure of translation-invariant noncommutative Yang-Mills theoris are studied by means of a cohomology theory, so called ⋆-cohomology, which plays an intermediate role between de Rham and cyclic (co)homology theory for noncommutative algebras and gives rise to a cohomological formulation comparable to Seiberg-Witten map. Keywords: Translation-Invariant Star Product, Noncommutative Yang-Mills, Spectral Triple, Chern Character, Connes-Chern Character, Family Index Theory, Topological Anomaly, BRST. I. INTRODUCTION Noncommutative geometry is one of the most prominent topics in theoretical physics. Through with last three decades it was extensively believed that the fundamental forces of the nature could be interpreted with more success via the machinery of noncommutative geometry and its different viewpoints [1–4]. Moyal noncommutative fields, inspired by fascinating formulations of strings, were proposed as the emergence of this idea in order to ramify the singular behaviors of quantum field theories, especially the quantum gravity.1 Appearing UV/IR mixing as a pathological feature of the Moyal quantum fields led to a concrete generalization of Moyal product as general translation-invariant noncommutative star products.2 arXiv:2101.10594v1 [hep-th] 26 Jan 2021 However, the topology and the geometry of noncommutative field theories with general translation- invariant star products have not been studied thoroughly yet. Actually, on the one hand it is a problem in noncommutative geometry, and on the other hand it is a physical behavior correlated to the topology and the geometry of the underlying spacetime and corresponding fiber bundles.
    [Show full text]
  • Constrained Optimization on Manifolds
    Constrained Optimization on Manifolds Von der Universit¨atBayreuth zur Erlangung des Grades eines Doktors der Naturwissenschaften (Dr. rer. nat.) genehmigte Abhandlung von Juli´anOrtiz aus Bogot´a 1. Gutachter: Prof. Dr. Anton Schiela 2. Gutachter: Prof. Dr. Oliver Sander 3. Gutachter: Prof. Dr. Lars Gr¨une Tag der Einreichung: 23.07.2020 Tag des Kolloquiums: 26.11.2020 Zusammenfassung Optimierungsprobleme sind Bestandteil vieler mathematischer Anwendungen. Die herausfordernd- sten Optimierungsprobleme ergeben sich dabei, wenn hoch nichtlineare Probleme gel¨ostwerden m¨ussen. Daher ist es von Vorteil, gegebene nichtlineare Strukturen zu nutzen, wof¨urdie Op- timierung auf nichtlinearen Mannigfaltigkeiten einen geeigneten Rahmen bietet. Es ergeben sich zahlreiche Anwendungsf¨alle,zum Beispiel bei nichtlinearen Problemen der Linearen Algebra, bei der nichtlinearen Mechanik etc. Im Fall der nichtlinearen Mechanik ist es das Ziel, Spezifika der Struktur zu erhalten, wie beispielsweise Orientierung, Inkompressibilit¨atoder Nicht-Ausdehnbarkeit, wie es bei elastischen St¨aben der Fall ist. Außerdem k¨onnensich zus¨atzliche Nebenbedingungen ergeben, wie im wichtigen Fall der Optimalsteuerungsprobleme. Daher sind f¨urdie L¨osungsolcher Probleme neue geometrische Tools und Algorithmen n¨otig. In dieser Arbeit werden Optimierungsprobleme auf Mannigfaltigkeiten und die Konstruktion von Algorithmen f¨urihre numerische L¨osungbehandelt. In einer abstrakten Formulierung soll eine reelle Funktion auf einer Mannigfaltigkeit minimiert werden, mit der Nebenbedingung
    [Show full text]
  • De Rham Cohomology in This Section We Will See How to Use Differential
    De Rham Cohomology In this section we will see how to use differential forms to detect topological prop- erties of smooth manifolds. Question. Is there a smooth function F : U → R such that ∂F ∂F (1) = f , and = f ; i.e. dF = f dx + f dx ? ∂x1 1 ∂x2 2 1 1 2 2 ∂2F ∂2F Since ∂x2∂x1 = ∂x1∂x2 , we must have ∂f ∂f (2) 1 = 2 ; i.e. the form f dx + f dx is closed. ∂x2 ∂x1 1 1 2 2 The appropriate question is therefore whether F exists, assuming that f = (f1,f2) satisfies (2). Is condition (2) also sufficient? Example 1. Consider the function f : R2 → R2 given by −x2 x1 f(x1,x2)= 2 2 , 2 2 x1 + x2 x1 + x2 Then (2) is satisfied. However, there is no function F : R2 \{0}→R that satisfies (1). Indeed, assume there were such an function F satisfying (2), then 2π d Z F (cos θ, sin θ)dθ = F (1, 0) − F (1, 0) = 0. 0 dθ which contradicts the fact that d ∂F ∂F F (cos θ, sin θ)= (− sin θ)+ cos θ dθ ∂x ∂y = − f1(cos θ, sin θ) sin θ + f2(cos θ, sin θ) cos θ =1. 2 Theorem 1. Let U ⊂ R be open and star-shaped w.r.t. the point x0.For 2 any smooth function (f1,f2):U → R that satisfies (2), there exists a function F : U → R satisfying (1). 2 Proof. Assume w.l.o.g. that x0 =0∈ R . Consider the function F : U → R, 1 Z F (x1,x2)= (x1f1(tx1,tx2)+x2f2(tx1,tx2))dt.
    [Show full text]
  • Exterior Derivative
    Exterior derivative On a differentiable manifold, the exterior derivative extends the concept of the differential of a function to differential forms of higher degree. e exterior derivative was first described in its current form by Élie Cartan in 1899; it allows for a natural, metric-independent generalization of Stokes' theorem, Gauss's theorem, and Green's theorem from vector calculus. If a k-form is thought of as measuring the flux through an infinitesimal k-parallelotope, then its exterior derivative can be thought of as measuring the net flux through the boundary of a (k + 1)-parallelotope. Contents Definition In terms of axioms In terms of local coordinates In terms of invariant formula Examples Stokes' theorem on manifolds Further properties Closed and exact forms de Rham cohomology Naturality Exterior derivative in vector calculus Gradient Divergence Curl Invariant formulations of grad, curl, div, and Laplacian See also Notes References Definition e exterior derivative of a differential form of degree k is a differential form of degree k + 1. If f is a smooth function (a 0-form), then the exterior derivative of f is the differential of f . at is, df is the unique 1-form such that for every smooth vector field X, df (X) = dX f , where dX f is the directional derivative of f in the direction of X. ere are a variety of equivalent definitions of the exterior derivative of a general k-form. In terms of axioms e exterior derivative is defined to be the unique ℝ-linear mapping from k-forms to (k + 1)-forms satisfying the following properties: 1.
    [Show full text]