Harmonic Differential Forms and the Hodge Decomposition Theorem

Total Page:16

File Type:pdf, Size:1020Kb

Harmonic Differential Forms and the Hodge Decomposition Theorem Harmonic Differential Forms and the Hodge Decomposition Theorem Matthew Romney Abstract This is an expository paper for Rui Fernandes's Spring 2014 Differentiable Manifolds 2 (Math 519) class. Introduction This paper will explain Hodge's theorem giving an orthogonal decomposition of a dif- ferential form on a compact Riemannian manifold. This is one of the central ideas of Hodge theory, developed in the 1930's by the Scottish mathematician William V.D. Hodge and laid out in his 1941 book [1]. It had origins in algebraic geometry, par- ticularly the work of Solomon Lefschetz, but it is immediately relevant to differential geometry. This is due to the corollary that every de Rham cohomology class for a com- pact Riemannian manifold has a unique harmonic representative, where \harmonic" refers to the appropriate generalization of the usual Laplace operator. In the intro- duction to the 1989 edition of Hodge's book, Michael Atiyah describes Hodge's work as one of the great mathematical landmarks of the century. He adds that Hodge was \the victim of his own success," in the sense that Hodge's original work would be over- shadowed as other mathematicians refined and expanded on his ideas and found new applications. A presentation of the Hodge decomposition theorem has two main obstacles. First is the large amount of notation and definitions needed to simply state the result. Second is the difficulty of the proof, which draws heavily on harmonic and functional analysis. This paper will communicate as much of the ideas as is reasonable possible, but this will fall well short of a complete rigorous proof. The interested reader will find a self- contained proof in Warner's 1971 textbook on manifold theory [5]. A recent expository paper of Nikolai Nowaczyk explains Warner's approach in a more accesible form [4]. The current paper is bases on these two sources, along with the standard textbook on smooth manifolds by John Lee [3]. Another recent work for further study is the 2011 book by J. Jost [2], although this book has a more analytic point of view. In the first section we will define the Hodge star operator on an oriented inner product space. Following that, we review the properties of Riemmanian manifolds and extend the definition of the Hodge star operator to this setting. This allows us to define the Laplace-Beltrami operator, which is a suitable generalization of the usual Laplace operator. After a few more definitions, we will finally be able to state the Hodge decomposition theorem and sketch the proof. The last section is devoted to applications to de Rham cohomology. 1 Before proceeding further, we will review the main definitions and notation for ten- sor fields, following that used by Lee. Let V be a vector space and k be a positive integer. A covariant k-tensor on V is a multilinear function α : V k = V × · · · × V ! R. Such a tensor α is said to be alternating if α(v1; : : : ; vi; : : : ; vj; : : : ; vk) = −α(v1; : : : ; vj; : : : ; vi; : : : ; vk) for any pair of indices i; j. Similarly, α is symmetric if α(v1; : : : ; vi; : : : ; vj; : : : ; vk) = α(v1; : : : ; vj; : : : ; vi; : : : ; vk). The space of all alternating covariant k-tensors on V is denoted by Λk(V ∗). From here, we can extend these defini- tions to objects defined on manifolds. We let ΛkT ∗M denote the bundle of alternating k ∗ G k ∗ covariant k-tensors on M; more precisely, Λ T M = Λ (Tp M). A (differential) p2M k-form is a smooth section of ΛkT ∗M. The space of all k-forms is denoted by Ωk(M). 1 M We also use the notation Ω∗(M) := Ωk(M). i=1 1 Hodge Star Operator In this section we will start with an oriented inner product space V of finite dimension n and build up to the definition of the Hodge star operator. The existence of an inner product on V provides a large amount of structure to work with. The most basic consequence is the existence of a positive orthonormal basis (ei), which follows from the Gram-Schmidt process. Next is the existence of the volume element dV . This the n ∗ 0 0 unique element of Λ (V ) satisfying dV (e1; : : : ; en) = 1 for any positive orthonormal 0 basis (ei). Another feature of an inner product is that it provides a canonical isomorphism ∗ between V and V by associating v 2 V with the linear functional Tv given by Tv(w) = hv; wi. Note that there is no such canonical isomorphism between an arbitrary finite dimensional vector space and its dual. We will typically use (ei) to denote the dual ∗ basis of (ei). In particular, we now have a natural inner product on V defined for basis i j vectors in the obvious way by he ; e i := hei; eji. An important fact is that this inner product on V ∗ can be extended to an inner product on Λk(V ∗) for any 0 ≤ k ≤ n. Proposition 1. Let 0 ≤ k ≤ n. There exists a unique inner product on Λk(V ∗), also denoted by h·; ·i, such that for any orthonormal basis B = (ei) of V ∗, the basis ΛkB is orthonormal with respect to this inner product. In fact, this larger inner product is given explicitly by hv1 ^:::^vk; w1 ^:::^wki = det(hvi; wji) for any basis elements v1 ^ ::: ^ vk; w1 ^ ::: ^ wk 2 ΛkB. Next, the definition of the Hodge star operator is based on the following result. Proposition 2. For all 0 ≤ k ≤ n, there exists a unique isomorphism ∗ :Λk(V ∗) ! Λn−k(V ∗) such that ! ^ ∗η = h!; ηidV for all !; η 2 Λk(V ∗). We define the Hodge star operator to be the isomorphism ∗ in Proposition 2. To un- derstand it better, in the proof of this proposition it is shown that the Hodge star oper- i1 ik j1 jk ator is defined locally by ∗(e ^· · ·^e ) = ±e ^· · ·^e , where (i1; : : : ; ik; j1; : : : ; jn−k) is a permutation of (1; : : : ; n). That is, (i1; : : : ; ik; j1; : : : ; jn−k) = (σ(1); : : : ; σ(n)) for some permutation σ. The choice of sign in this definition is given by sgn(σ). A number of interesting elementary properties can be derived immediately from the definition. One that will be useful for us is the composition rule ∗ ∗ ! = (−1)k(n−k)!. 2 The reader can also easily check that h!; ηi = h∗!; ∗ηi and that ∗(! ^ ∗η) = ∗(η ^ ∗!) for all !; η 2 Λk(V ∗). 2 Riemannian Manifolds Our next goal is to transfer these definitions to the manifold level. To do this, we need a brief overview of Riemannian manifolds. A Riemannian metric on a manifold M is a smooth symmetric covariant 2-tensor field that is positive definite at each point. Such a tensor field gives an inner product at each point p 2 M, which we normally denote by gp(u; v) or hu; vip. Positive definiteness means that gp(u; u) > 0 for all u 6= 0. The smoothness requirement is equivalent to the property that, for any smooth vector fields X; Y 2 X(M), the function p 7! hX(p);Y (p)ip is smooth (see [3], Proposition 12.19, p. 317). A Riemannian manifold is a pair (M; g), where M is a manifold and g is a Riemannian metric defined on M. We will state some of the essential facts for a Riemannian manifold (M; g). These are analogous to the properties of inner product spaces discussed in the last section. First is the existence of a smooth orthonormal frame (E1;:::;En) in a neighborhood of each point, as well as the corresponding coframe (E1;:::;En). From this, we can prove the existence of the Riemannian volume form dVg, which is the unique smooth orientation form satisfying dVg(E1;:::;En) = 1 for every local oriented orthonormal frame (Ei). Note that dVg is a notational convenience and does not imply that the Riemannian volume form is an exact differential form. Also, we have a natural isomor- ∗ phism between the tangent and cotangent bundles. This is given by gb : TM ! T M ∗ by v 2 TpM 7! gb(v) 2 Tp M, where gb(v)(w) = gp(v; w). This is much like the case of [ inner product spaces, although one new feature is the common notation X for gb(X) ] −1 and ! for gb (!). For this reason gb and its inverse are sometimes called the \musical isomorphisms." As expected, for all 0 ≤ k ≤ n the Hodge star operator extends to a mapping ∗ :Ωk(M) ! Ωn−k(M). It is not immediately clear that ∗ is smooth, but in fact this follows from the local representation ∗(Ei1 ^ · · · ^ Eik ) = ±Ej1 ^ · · · ^ Ejk , as this implies that ∗ in local coordinates is either ±1 or 0. An interesting immediate application is in providing a convenient notation for the classical vector field operations of multivariable calculus. The divergence operator can be written as div X = ∗d ∗ X[. In the case that M = 3, the curl operator is given by [ ] curl X = (∗dX ) . In addition, interior multiplication into dVg can be written as X y [ dVg = ∗X . 3 Laplace-Beltrami Operator Our goal is to define the Laplace-Beltrami operator ∆ : Ωk(M) ! Ωk(M). This is a 1 n 1 n generalization of the usual Laplace operator ∆ : C (R ) ! C (R ) defined by n X @2f ∆f = −div(grad f) = − : (1) (@xi)2 i=1 First, we must define the codifferential (or Beltrami operator) δ :Ωk(M) ! Ωk−1(M). This is by the formulas δ! = (−1)n(k+1)+1 ∗ d ∗ ! for 1 ≤ k ≤ n and 3 δ! = 0 for k = 0.
Recommended publications
  • Best Subordinant for Differential Superordinations of Harmonic Complex-Valued Functions
    mathematics Article Best Subordinant for Differential Superordinations of Harmonic Complex-Valued Functions Georgia Irina Oros Department of Mathematics and Computer Sciences, Faculty of Informatics and Sciences, University of Oradea, 410087 Oradea, Romania; [email protected] or [email protected] Received: 17 September 2020; Accepted: 11 November 2020; Published: 16 November 2020 Abstract: The theory of differential subordinations has been extended from the analytic functions to the harmonic complex-valued functions in 2015. In a recent paper published in 2019, the authors have considered the dual problem of the differential subordination for the harmonic complex-valued functions and have defined the differential superordination for harmonic complex-valued functions. Finding the best subordinant of a differential superordination is among the main purposes in this research subject. In this article, conditions for a harmonic complex-valued function p to be the best subordinant of a differential superordination for harmonic complex-valued functions are given. Examples are also provided to show how the theoretical findings can be used and also to prove the connection with the results obtained in 2015. Keywords: differential subordination; differential superordination; harmonic function; analytic function; subordinant; best subordinant MSC: 30C80; 30C45 1. Introduction and Preliminaries Since Miller and Mocanu [1] (see also [2]) introduced the theory of differential subordination, this theory has inspired many researchers to produce a number of analogous notions, which are extended even to non-analytic functions, such as strong differential subordination and superordination, differential subordination for non-analytic functions, fuzzy differential subordination and fuzzy differential superordination. The notion of differential subordination was adapted to fit the harmonic complex-valued functions in the paper published by S.
    [Show full text]
  • General Relativity Fall 2018 Lecture 10: Integration, Einstein-Hilbert Action
    General Relativity Fall 2018 Lecture 10: Integration, Einstein-Hilbert action Yacine Ali-Ha¨ımoud (Dated: October 3, 2018) σ σ HW comment: T µν antisym in µν does NOT imply that Tµν is antisym in µν. Volume element { Consider a LICS with primed coordinates. The 4-volume element is dV = d4x0 = 0 0 0 0 dx0 dx1 dx2 dx3 . If we change coordinates, we have 0 ! @xµ d4x0 = det d4x: (1) µ @x Now, the metric components change as 0 0 0 0 @xµ @xν @xµ @xν g = g 0 0 = η 0 0 ; (2) µν @xµ @xν µ ν @xµ @xν µ ν since the metric components are ηµ0ν0 in the LICS. Seing this as a matrix operation and taking the determinant, we find 0 !2 @xµ det(g ) = − det : (3) µν @xµ Hence, we find that the 4-volume element is q 4 p 4 p 4 dV = − det(gµν )d x ≡ −g d x ≡ jgj d x: (4) The integral of a scalar function f is well defined: given any coordinate system (even if only defined locally): Z Z f dV = f pjgj d4x: (5) We can only define the integral of a scalar function. The integral of a vector or tensor field is mean- ingless in curved spacetime. Think of the integral as a sum. To sum vectors, you need them to belong to the same vector space. There is no common vector space in curved spacetime. Only in flat spacetime can we define such integrals. First parallel-transport the vector field to a single point of spacetime (it doesn't matter which one).
    [Show full text]
  • Topology and Physics 2019 - Lecture 2
    Topology and Physics 2019 - lecture 2 Marcel Vonk February 12, 2019 2.1 Maxwell theory in differential form notation Maxwell's theory of electrodynamics is a great example of the usefulness of differential forms. A nice reference on this topic, though somewhat outdated when it comes to notation, is [1]. For notational simplicity, we will work in units where the speed of light, the vacuum permittivity and the vacuum permeability are all equal to 1: c = 0 = µ0 = 1. 2.1.1 The dual field strength In three dimensional space, Maxwell's electrodynamics describes the physics of the electric and magnetic fields E~ and B~ . These are three-dimensional vector fields, but the beauty of the theory becomes much more obvious if we (a) use a four-dimensional relativistic formulation, and (b) write it in terms of differential forms. For example, let us look at Maxwells two source-free, homogeneous equations: r · B = 0;@tB + r × E = 0: (2.1) That these equations have a relativistic flavor becomes clear if we write them out in com- ponents and organize the terms somewhat suggestively: x y z 0 + @xB + @yB + @zB = 0 x z y −@tB + 0 − @yE + @zE = 0 (2.2) y z x −@tB + @xE + 0 − @zE = 0 z y x −@tB − @xE + @yE + 0 = 0 Note that we also multiplied the last three equations by −1 to clarify the structure. All in all, we see that we have four equations (one for each space-time coordinate) which each contain terms in which the four coordinate derivatives act. Therefore, we may be tempted to write our set of equations in more \relativistic" notation as ^µν @µF = 0 (2.3) 1 with F^µν the coordinates of an antisymmetric two-tensor (i.
    [Show full text]
  • Arxiv:2009.07259V1 [Math.AP] 15 Sep 2020
    A GEOMETRIC TRAPPING APPROACH TO GLOBAL REGULARITY FOR 2D NAVIER-STOKES ON MANIFOLDS AYNUR BULUT AND KHANG MANH HUYNH Abstract. In this paper, we use frequency decomposition techniques to give a direct proof of global existence and regularity for the Navier-Stokes equations on two-dimensional Riemannian manifolds without boundary. Our techniques are inspired by an approach of Mattingly and Sinai [15] which was developed in the context of periodic boundary conditions on a flat background, and which is based on a maximum principle for Fourier coefficients. The extension to general manifolds requires several new ideas, connected to the less favorable spectral localization properties in our setting. Our argu- ments make use of frequency projection operators, multilinear estimates that originated in the study of the non-linear Schr¨odingerequation, and ideas from microlocal analysis. 1. Introduction Let (M; g) be a closed, oriented, connected, compact smooth two-dimensional Riemannian manifold, and let X(M) denote the space of smooth vector fields on M. We consider the incompressible Navier-Stokes equations on M, with viscosity coefficient ν > 0, @ U + div (U ⊗ U) − ν∆ U = − grad p in M t M ; (1) div U = 0 in M with initial data U0 2 X(M); where I ⊂ R is an open interval, and where U : I ! X(M) and p : I × M ! R represent the velocity and pressure of the fluid, respectively. Here, the operator ∆M is any choice of Laplacian defined on vector fields on M, discussed below. arXiv:2009.07259v1 [math.AP] 15 Sep 2020 The theory of two-dimensional fluid flows on flat spaces is well-developed, and a variety of global regularity results are well-known.
    [Show full text]
  • Harmonic Functions
    Lecture 1 Harmonic Functions 1.1 The Definition Definition 1.1. Let Ω denote an open set in R3. A real valued function u(x, y, z) on Ω with continuous second partials is said to be harmonic if and only if the Laplacian ∆u = 0 identically on Ω. Note that the Laplacian ∆u is defined by ∂2u ∂2u ∂2u ∆u = + + . ∂x2 ∂y2 ∂z2 We can make a similar definition for an open set Ω in R2.Inthatcase, u is harmonic if and only if ∂2u ∂2u ∆u = + =0 ∂x2 ∂y2 on Ω. Some basic examples of harmonic functions are 2 2 2 3 u = x + y 2z , Ω=R , − 1 3 u = , Ω=R (0, 0, 0), r − where r = x2 + y2 + z2. Moreover, by a theorem on complex variables, the real part of an analytic function on an open set Ω in 2 is always harmonic. p R Thus a function such as u = rn cos nθ is a harmonic function on R2 since u is the real part of zn. 1 2 1.2 The Maximum Principle The basic result about harmonic functions is called the maximum principle. What the maximum principle says is this: if u is a harmonic function on Ω, and B is a closed and bounded region contained in Ω, then the max (and min) of u on B is always assumed on the boundary of B. Recall that since u is necessarily continuous on Ω, an absolute max and min on B are assumed. The max and min can also be assumed inside B, but a harmonic function cannot have any local extrema inside B.
    [Show full text]
  • LP THEORY of DIFFERENTIAL FORMS on MANIFOLDS This
    TRANSACTIONSOF THE AMERICAN MATHEMATICALSOCIETY Volume 347, Number 6, June 1995 LP THEORY OF DIFFERENTIAL FORMS ON MANIFOLDS CHAD SCOTT Abstract. In this paper, we establish a Hodge-type decomposition for the LP space of differential forms on closed (i.e., compact, oriented, smooth) Rieman- nian manifolds. Critical to the proof of this result is establishing an LP es- timate which contains, as a special case, the L2 result referred to by Morrey as Gaffney's inequality. This inequality helps us show the equivalence of the usual definition of Sobolev space with a more geometric formulation which we provide in the case of differential forms on manifolds. We also prove the LP boundedness of Green's operator which we use in developing the LP theory of the Hodge decomposition. For the calculus of variations, we rigorously verify that the spaces of exact and coexact forms are closed in the LP norm. For nonlinear analysis, we demonstrate the existence and uniqueness of a solution to the /1-harmonic equation. 1. Introduction This paper contributes primarily to the development of the LP theory of dif- ferential forms on manifolds. The reader should be aware that for the duration of this paper, manifold will refer only to those which are Riemannian, compact, oriented, C°° smooth and without boundary. For p = 2, the LP theory is well understood and the L2-Hodge decomposition can be found in [M]. However, in the case p ^ 2, the LP theory has yet to be fully developed. Recent appli- cations of the LP theory of differential forms on W to both quasiconformal mappings and nonlinear elasticity continue to motivate interest in this subject.
    [Show full text]
  • Lecture Notes in Mathematics
    Lecture Notes in Mathematics Edited by A. Dold and B. Eckmann Subseries: Mathematisches Institut der Universit~it und Max-Planck-lnstitut for Mathematik, Bonn - voL 5 Adviser: E Hirzebruch 1111 Arbeitstagung Bonn 1984 Proceedings of the meeting held by the Max-Planck-lnstitut fur Mathematik, Bonn June 15-22, 1984 Edited by E Hirzebruch, J. Schwermer and S. Suter I IIII Springer-Verlag Berlin Heidelberg New York Tokyo Herausgeber Friedrich Hirzebruch Joachim Schwermer Silke Suter Max-Planck-lnstitut fLir Mathematik Gottfried-Claren-Str. 26 5300 Bonn 3, Federal Republic of Germany AMS-Subject Classification (1980): 10D15, 10D21, 10F99, 12D30, 14H10, 14H40, 14K22, 17B65, 20G35, 22E47, 22E65, 32G15, 53C20, 57 N13, 58F19 ISBN 3-54045195-8 Springer-Verlag Berlin Heidelberg New York Tokyo ISBN 0-387-15195-8 Springer-Verlag New York Heidelberg Berlin Tokyo CIP-Kurztitelaufnahme der Deutschen Bibliothek. Mathematische Arbeitstagung <25. 1984. Bonn>: Arbeitstagung Bonn: 1984; proceedings of the meeting, held in Bonn, June 15-22, 1984 / [25. Math. Arbeitstagung]. Ed. by E Hirzebruch ... - Berlin; Heidelberg; NewYork; Tokyo: Springer, 1985. (Lecture notes in mathematics; Vol. 1t11: Subseries: Mathematisches I nstitut der U niversit~it und Max-Planck-lnstitut for Mathematik Bonn; VoL 5) ISBN 3-540-t5195-8 (Berlin...) ISBN 0-387q5195-8 (NewYork ...) NE: Hirzebruch, Friedrich [Hrsg.]; Lecture notes in mathematics / Subseries: Mathematischee Institut der UniversitAt und Max-Planck-lnstitut fur Mathematik Bonn; HST This work ts subject to copyright. All rights are reserved, whether the whole or part of the material is concerned, specifically those of translation, reprinting, re~use of illustrations, broadcasting, reproduction by photocopying machine or similar means, and storage in data banks.
    [Show full text]
  • NOTES on DIFFERENTIAL FORMS. PART 3: TENSORS 1. What Is A
    NOTES ON DIFFERENTIAL FORMS. PART 3: TENSORS 1. What is a tensor? 1 n Let V be a finite-dimensional vector space. It could be R , it could be the tangent space to a manifold at a point, or it could just be an abstract vector space. A k-tensor is a map T : V × · · · × V ! R 2 (where there are k factors of V ) that is linear in each factor. That is, for fixed ~v2; : : : ;~vk, T (~v1;~v2; : : : ;~vk−1;~vk) is a linear function of ~v1, and for fixed ~v1;~v3; : : : ;~vk, T (~v1; : : : ;~vk) is a k ∗ linear function of ~v2, and so on. The space of k-tensors on V is denoted T (V ). Examples: n • If V = R , then the inner product P (~v; ~w) = ~v · ~w is a 2-tensor. For fixed ~v it's linear in ~w, and for fixed ~w it's linear in ~v. n • If V = R , D(~v1; : : : ;~vn) = det ~v1 ··· ~vn is an n-tensor. n • If V = R , T hree(~v) = \the 3rd entry of ~v" is a 1-tensor. • A 0-tensor is just a number. It requires no inputs at all to generate an output. Note that the definition of tensor says nothing about how things behave when you rotate vectors or permute their order. The inner product P stays the same when you swap the two vectors, but the determinant D changes sign when you swap two vectors. Both are tensors. For a 1-tensor like T hree, permuting the order of entries doesn't even make sense! ~ ~ Let fb1;:::; bng be a basis for V .
    [Show full text]
  • Killing Spinor-Valued Forms and the Cone Construction
    ARCHIVUM MATHEMATICUM (BRNO) Tomus 52 (2016), 341–355 KILLING SPINOR-VALUED FORMS AND THE CONE CONSTRUCTION Petr Somberg and Petr Zima Abstract. On a pseudo-Riemannian manifold M we introduce a system of partial differential Killing type equations for spinor-valued differential forms, and study their basic properties. We discuss the relationship between solutions of Killing equations on M and parallel fields on the metric cone over M for spinor-valued forms. 1. Introduction The subject of the present article are the systems of over-determined partial differential equations for spinor-valued differential forms, classified as atypeof Killing equations. The solution spaces of these systems of PDE’s are termed Killing spinor-valued differential forms. A central question in geometry asks for pseudo-Riemannian manifolds admitting non-trivial solutions of Killing type equa- tions, namely how the properties of Killing spinor-valued forms relate to the underlying geometric structure for which they can occur. Killing spinor-valued forms are closely related to Killing spinors and Killing forms with Killing vectors as a special example. Killing spinors are both twistor spinors and eigenspinors for the Dirac operator, and real Killing spinors realize the limit case in the eigenvalue estimates for the Dirac operator on compact Riemannian spin manifolds of positive scalar curvature. There is a classification of complete simply connected Riemannian manifolds equipped with real Killing spinors, leading to the construction of manifolds with the exceptional holonomy groups G2 and Spin(7), see [8], [1]. Killing vector fields on a pseudo-Riemannian manifold are the infinitesimal generators of isometries, hence they influence its geometrical properties.
    [Show full text]
  • 23. Harmonic Functions Recall Laplace's Equation
    23. Harmonic functions Recall Laplace's equation ∆u = uxx = 0 ∆u = uxx + uyy = 0 ∆u = uxx + uyy + uzz = 0: Solutions to Laplace's equation are called harmonic functions. The inhomogeneous version of Laplace's equation ∆u = f is called the Poisson equation. Harmonic functions are Laplace's equation turn up in many different places in mathematics and physics. Harmonic functions in one variable are easy to describe. The general solution of uxx = 0 is u(x) = ax + b, for constants a and b. Maximum principle Let D be a connected and bounded open set in R2. Let u(x; y) be a harmonic function on D that has a continuous extension to the boundary @D of D. Then the maximum (and minimum) of u are attained on the bound- ary and if they are attained anywhere else than u is constant. Euivalently, there are two points (xm; ym) and (xM ; yM ) on the bound- ary such that u(xm; ym) ≤ u(x; y) ≤ u(xM ; yM ) for every point of D and if we have equality then u is constant. The idea of the proof is as follows. At a maximum point of u in D we must have uxx ≤ 0 and uyy ≤ 0. Most of the time one of these inequalities is strict and so 0 = uxx + uyy < 0; which is not possible. The only reason this is not a full proof is that sometimes both uxx = uyy = 0. As before, to fix this, simply perturb away from zero. Let > 0 and let v(x; y) = u(x; y) + (x2 + y2): Then ∆v = ∆u + ∆(x2 + y2) = 4 > 0: 1 Thus v has no maximum in D.
    [Show full text]
  • 1.2 Topological Tensor Calculus
    PH211 Physical Mathematics Fall 2019 1.2 Topological tensor calculus 1.2.1 Tensor fields Finite displacements in Euclidean space can be represented by arrows and have a natural vector space structure, but finite displacements in more general curved spaces, such as on the surface of a sphere, do not. However, an infinitesimal neighborhood of a point in a smooth curved space1 looks like an infinitesimal neighborhood of Euclidean space, and infinitesimal displacements dx~ retain the vector space structure of displacements in Euclidean space. An infinitesimal neighborhood of a point can be infinitely rescaled to generate a finite vector space, called the tangent space, at the point. A vector lives in the tangent space of a point. Note that vectors do not stretch from one point to vector tangent space at p p space Figure 1.2.1: A vector in the tangent space of a point. another, and vectors at different points live in different tangent spaces and so cannot be added. For example, rescaling the infinitesimal displacement dx~ by dividing it by the in- finitesimal scalar dt gives the velocity dx~ ~v = (1.2.1) dt which is a vector. Similarly, we can picture the covector rφ as the infinitesimal contours of φ in a neighborhood of a point, infinitely rescaled to generate a finite covector in the point's cotangent space. More generally, infinitely rescaling the neighborhood of a point generates the tensor space and its algebra at the point. The tensor space contains the tangent and cotangent spaces as a vector subspaces. A tensor field is something that takes tensor values at every point in a space.
    [Show full text]
  • Vector Calculus and Multiple Integrals Rob Fender, HT 2018
    Vector Calculus and Multiple Integrals Rob Fender, HT 2018 COURSE SYNOPSIS, RECOMMENDED BOOKS Course syllabus (on which exams are based): Double integrals and their evaluation by repeated integration in Cartesian, plane polar and other specified coordinate systems. Jacobians. Line, surface and volume integrals, evaluation by change of variables (Cartesian, plane polar, spherical polar coordinates and cylindrical coordinates only unless the transformation to be used is specified). Integrals around closed curves and exact differentials. Scalar and vector fields. The operations of grad, div and curl and understanding and use of identities involving these. The statements of the theorems of Gauss and Stokes with simple applications. Conservative fields. Recommended Books: Mathematical Methods for Physics and Engineering (Riley, Hobson and Bence) This book is lazily referred to as “Riley” throughout these notes (sorry, Drs H and B) You will all have this book, and it covers all of the maths of this course. However it is rather terse at times and you will benefit from looking at one or both of these: Introduction to Electrodynamics (Griffiths) You will buy this next year if you haven’t already, and the chapter on vector calculus is very clear Div grad curl and all that (Schey) A nice discussion of the subject, although topics are ordered differently to most courses NB: the latest version of this book uses the opposite convention to polar coordinates to this course (and indeed most of physics), but older versions can often be found in libraries 1 Week One A review of vectors, rotation of coordinate systems, vector vs scalar fields, integrals in more than one variable, first steps in vector differentiation, the Frenet-Serret coordinate system Lecture 1 Vectors A vector has direction and magnitude and is written in these notes in bold e.g.
    [Show full text]