Concepcion: Scalar and Vector Potentials, Helmholtz

Total Page:16

File Type:pdf, Size:1020Kb

Concepcion: Scalar and Vector Potentials, Helmholtz Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology Scalar and vector potentials, Helmholtz decomposition, and de Rham cohomology Alberto Valli Department of Mathematics, University of Trento, Italy A. Valli Potentials, Helmholtz decomposition, de Rham cohomology Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology Outline 1 Introduction 2 Scalar and vector potentials in the whole space 3 Scalar and vector potentials in a bounded domain 4 Helmholtz decomposition 5 Homology and de Rham cohomology A. Valli Potentials, Helmholtz decomposition, de Rham cohomology Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology The objects Beyond a doubt, among the \stars" of vector calculus we have the operators grad div curl Aim of this talk is to understand better their properties and their connections with some topological concepts. A. Valli Potentials, Helmholtz decomposition, de Rham cohomology Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology First results First well-known results are (just compute...): curl grad = 0 for each scalar function div curl H = 0 for each vector field H. We can thus write Theorem (1) If H = grad , then curl H = 0 (namely, H is curl-free). Theorem (2) If B = curl A, then div B = 0 (namely, B is divergence-free). A. Valli Potentials, Helmholtz decomposition, de Rham cohomology Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology First results (cont'd) The natural question is: are these conditions sufficient? We will see that the answer depends on the geometry of the region Ω where we are working. A. Valli Potentials, Helmholtz decomposition, de Rham cohomology Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology In the whole space... 3 Let us start from Ω = R . We need some tools. First of all we know [just compute...] that the function 1 1 K(x; y) = (1) 4π jx − yj satisfies −∆x K(x; y) = 0 for x 6= y R @B grad x K(x; 0) · n(x) dSx = −1 ; where B is the ball of center 0 and radius 1, and n the unit outward normal on @B. A. Valli Potentials, Helmholtz decomposition, de Rham cohomology Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology Dirac δ0 distribution [Indeed, in a more advanced mathematical language, the function K(x; y) is the fundamental solution of the −∆ operator, namely, it satisfies −∆x K(x; y) = δ0(x − y) in the distributional sense, δ0 being the Dirac delta distribution centered at 0. Roughly speaking, for each (suitable...) function f the Dirac delta distribution satisfiesZ δ0(x − y)f (y) dy = f (x) : R3 We also know that the function Z u(x) = K(x; y)f (y) dy R3 3 satisfies −∆u = f in R . In fact (formally...) R R −∆u(x) = −∆x [ 3 K(x; y)f (y) dy] = 3 [−∆x K(x; y)]f (y) dy R R = R δ (x − y)f (y) dy = f (x) :] R3 0 A. Valli Potentials, Helmholtz decomposition, de Rham cohomology Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology Scalar and vector potentials Let us come to the determination of a scalar potential for a curl-free vector field H (namely, a scalar function such that grad = H) and of a vector potential A for a divergence-free vector field B (namely, a vector field A such that curl A = B). 3 Consider a vector field H and define in R the function Z (x) = − K(x; y)div H(y) dy : (2) R3 3 Consider a vector field B and define in R the vector field Z A(x) = K(x; y)curl B(y) dy : (3) R3 A. Valli Potentials, Helmholtz decomposition, de Rham cohomology Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology Theorems Theorem (3) Assume that H decays sufficiently fast at infinity and satisfies 3 3 curl H = 0 in R . The function satisfies grad = H in R . Proof. It is easily shown that 1 x − y D K(x; y) = − i i = −D K(x; y) ; xi 4π jx − yj3 yi hence (formally, and using that Di Hj = Dj Hi ...) R R D (x) = − 3 Dx K(x; y)div H(y) dy = 3 Dy K(x; y)div H(y) dy i R i R i P R P R = − 3 Dy Dy K(x; y)H (y) dy = 3 Dy K(x; y)D H (y) dy j R j i j j R j i j P R R = 3 Dy K(x; y)D H (y) dy = − 3 ∆y K(x; y)H (y) dy j R j j i R i = R δ (x − y)H (y) dy = H (x) : R3 0 i i A. Valli Potentials, Helmholtz decomposition, de Rham cohomology Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology Theorems (cont'd) Theorem (4) Assume that B decays sufficiently fast at infinity and satisfies 3 div B = 0 in R . The vector field A satisfies curl A = B (and 3 div A = 0) in R . Proof. We have R D1A2(x) = 3 Dx K(x; y)(D3B1 − D1B3)(y) dy RR 1 = − 3 Dy K(x; y)(D3B1 − D1B3)(y) dy R R 1 = 3 [−Dy Dy K(x; y)B3(y) + Dy Dy K(x; y)B1(y)] dy RR 1 1 3 1 = 3 [−Dy Dy K(x; y)B3(y) − Dy K(x; y)D1B1(y)] dy : R 1 1 3 Similarly, Z D2A1(x) = [Dy2 Dy2 K(x; y)B3(y) + Dy3 K(x; y)D2B2(y)] dy : R3 A. Valli Potentials, Helmholtz decomposition, de Rham cohomology Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology Theorems (cont'd) Since D1B1 + D2B2 = −D3B3, we find R − 3 Dy K(x; y)[D1B1(y) + D2B2(y)] dy R R 3 R = 3 Dy K(x; y)D3B3(y) dy = − 3 Dy Dy K(x; y)B3(y) dy ; R 3 R 3 3 hence R D1A2(x) − D2A1(x) = − 3 ∆y K(x; y)B3(y) dy R R = 3 δ (x − y)B3(y) dy = B3(x) : R 0 Repeating the same computations for the other components, the first part of the thesis follows. On the other hand R D1A1(x) = − 3 Dy K(x; y)(D2B3 − D3B2)(y) dy RR 1 = − 3 [Dy K(x; y)D2B3(y) − Dy K(x; y)D1B2(y)] dy ; R 1 3 and, proceding similary for D2A2 and D3A3, the second part of the thesis is easily verified. A. Valli Potentials, Helmholtz decomposition, de Rham cohomology Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology Leading idea What has been the idea? If satisfies grad = H, then −div H = −div grad = −∆ ; hence we can use the (scalar) integral representation formula in terms of the fundamental solution K; if A satisfies curl A = B (and div A = 0), then curl B = curl curl A = curl curl A − grad div A = −∆A ; hence we can use the (vector) integral representation formula in terms of the fundamental solution K. A. Valli Potentials, Helmholtz decomposition, de Rham cohomology Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology Biot{Savart formulas An alternative (and essentially equivalent) point of view is the one leading to the Biot{Savart formulas: Scalar potential: look for = div grad '. If satisfies grad = H, then H = grad div grad ' = ∆ grad '; hence Z grad ' = − K(x; y)H(y) dy R3 and 1 Z x − y (x) = 3 · H(y) dy : (4) 4π R3 jx − yj A. Valli Potentials, Helmholtz decomposition, de Rham cohomology Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology Biot{Savart formulas (cont'd) Vector potential: look for A = curl Q (with div Q = 0). If A satisfies curl A = B, then B = curl curl Q = −∆Q ; hence Z Q(x) = K(x; y)B(y) dy R3 and 1 Z y − x A(x) = 3 × B(y) dy : (5) 4π R3 jx − yj A. Valli Potentials, Helmholtz decomposition, de Rham cohomology Introduction Scalar and vector potentials in the whole space Scalar and vector potentials in a bounded domain Helmholtz decomposition Homology and de Rham cohomology In a bounded domain... The problem is more complicated in a bounded domain Ω. However, some well-known results are usually presented in any calculus course. Theorem (5) Assume that H satisfies curl H = 0 in Ω and that any closed curve in Ω is the boundary of a suitable surface S ⊂ Ω.
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]
  • Direct Derivation of Liénard–Wiechert Potentials, Maxwell's
    Direct derivation of Li´enard{Wiechert potentials, Maxwell's equations and Lorentz force from Coulomb's law Hrvoje Dodig∗ University of Split, Rudera Boˇskovi´ca37, 21000 Split, Croatia Abstract In 19th century Maxwell derived famous Maxwell equations from the knowledge of three experimental physical laws: the electrostatic Coulomb's law, Ampere's force law and Faraday's law of induction. However, theoretical basis for Ampere's force law and Fara- day's law remains unknown to this day. Furthermore, the Lorentz force is considered as summary of experimental phenomena, the theoretical foundation that explains genera- tion of this force is still unknown. To answer these fundamental theoretical questions, in this paper we derive relativisti- cally correct Li´enard{ Wiechert potentials, Maxwell's equations and Lorentz force from two simple postulates: (a) when all charges are at rest the Coulomb's force acts between the charges, and (b) that disturbances caused by charge in motion propagate away from the source with finite velocity. The special relativity was not used in our derivations nor the Lorentz transformation. In effect, it was shown in this paper that all the electrody- namic laws, including the Lorentz force, can be derived from Coulomb's law and time retardation. This was accomplished by analysis of hypothetical experiment where test charge is at rest and where previously moving source charge stops at some time in the past. Then the generalized Helmholtz decomposition theorem, also derived in this paper, was applied to reformulate Coulomb's force acting at present time as the function of positions of source charge at previous time when the source charge was moving.
    [Show full text]
  • The Helmholtz Decomposition of Decreasing and Weakly Increasing Vector Fields
    The Helmholtz’ decomposition of decreasing and weakly increasing vector fields D. Petrascheck and R. Folk Institute for Theoretical Physics University Linz Altenbergerstr. 69, Linz, Austria Abstract Helmholtz’ decomposition theorem for vector fields is presented usually with too strong restric- tions on the fields. Based on the work of Blumenthal of 1905 it is shown that the decomposition of vector fields is not only possible for asymptotically weakly decreasing vector fields, but even for vector fields, which asymptotically increase sublinearly. Use is made of a regularization of the Green’s function and the mathematics of the proof is formulated as simply as possible. We also show a few examples for the decomposition of vector fields including the electric dipole radiation. Keywords: Helmholtz theorem, vector field, electromagnetic radiation arXiv:1506.00235v2 [physics.class-ph] 14 Oct 2015 1 I. INTRODUCTION According to the Helmholtz’ theorem one can decompose a given vector field ~v(~x) into a sum of two vector fields ~vl(~x) and ~vt(~x) where ~vl is irrotational (curl-free) and ~vt solenoidal (divergence-free), if the vector field fulfills certain conditions on continuity and asymptotic decrease (r ). Here ~x is the position vector in three-dimensional space and r = ~x its →∞ | | absolute value. The two parts of the vector field can be expressed as gradient of a scalar potential and curl of a vector potential, respectively. Concerning the validity, the uniqueness of the decomposition and the existence of the respective potentials one finds different conditions. The fundamental theorem for vector fields is historically based on Helmholtz’ work on vortices1,2 and therefore also known as Helmholtz’ decomposition theorem.
    [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]
  • Electromagnetic Fields and Energy
    MIT OpenCourseWare http://ocw.mit.edu Haus, Hermann A., and James R. Melcher. Electromagnetic Fields and Energy. Englewood Cliffs, NJ: Prentice-Hall, 1989. ISBN: 9780132490207. Please use the following citation format: Haus, Hermann A., and James R. Melcher, Electromagnetic Fields and Energy. (Massachusetts Institute of Technology: MIT OpenCourseWare). http://ocw.mit.edu (accessed [Date]). License: Creative Commons Attribution-NonCommercial-Share Alike. Also available from Prentice-Hall: Englewood Cliffs, NJ, 1989. ISBN: 9780132490207. Note: Please use the actual date you accessed this material in your citation. For more information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms 8 MAGNETOQUASISTATIC FIELDS: SUPERPOSITION INTEGRAL AND BOUNDARY VALUE POINTS OF VIEW 8.0 INTRODUCTION MQS Fields: Superposition Integral and Boundary Value Views We now follow the study of electroquasistatics with that of magnetoquasistat­ ics. In terms of the flow of ideas summarized in Fig. 1.0.1, we have completed the EQS column to the left. Starting from the top of the MQS column on the right, recall from Chap. 3 that the laws of primary interest are Amp`ere’s law (with the displacement current density neglected) and the magnetic flux continuity law (Table 3.6.1). � × H = J (1) � · µoH = 0 (2) These laws have associated with them continuity conditions at interfaces. If the in­ terface carries a surface current density K, then the continuity condition associated with (1) is (1.4.16) n × (Ha − Hb) = K (3) and the continuity condition associated with (2) is (1.7.6). a b n · (µoH − µoH ) = 0 (4) In the absence of magnetizable materials, these laws determine the magnetic field intensity H given its source, the current density J.
    [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]
  • Helmholtz-Hodge Decompositions in the Nonlocal Framework. Well-Posedness Analysis and Applications
    Helmholtz-Hodge decompositions in the nonlocal framework { Well-posedness analysis and applications M. D'Elia Sandia National Laboratories, Albuquerque, NM, [email protected] C. Flores, California State University Channel Islands, Camarillo, CA, [email protected] X. Li, University of North Carolina-Charlotte, NC, [email protected] P. Radu, University of Nebraska-Lincoln, NE, [email protected] Y. Yu, Lehigh University, Bethlehem, PA, [email protected] Abstract. Nonlocal operators that have appeared in a variety of physical models satisfy identities and enjoy a range of properties similar to their classical counterparts. In this paper we obtain Helmholtz-Hodge type decompositions for two-point vector fields in three components that have zero nonlocal curls, zero nonlocal divergence, and a third component which is (nonlocally) curl-free and divergence-free. The results obtained incorporate different nonlocal boundary conditions, thus being applicable in a variety of settings. Keywords. Nonlocal operators, nonlocal calculus, Helmholtz-Hodge decompositions. AMS subject classifications. 35R09, 45A05, 45P05, 35J05, 74B99. 1 Introduction and motivation Important applications in diffusion, elasticity, fracture propagation, image processing, subsurface transport, molecular dynamics have benefited from the introduction of nonlocal models. Phenomena, materials, and behaviors that are discontinuous in nature have been ideal candidates for the introduc- tion of this framework which allows solutions with no smoothness, or even continuity properties. This advantage is counterbalanced by the facts that the theory of nonlocal calculus is still being developed and that the numerical solution of these problems can be prohibitively expensive. In [8] the authors introduce a nonlocal framework with divergence, gradient, and curl versions of nonlocal operators for which they identify duality relationships via L2 inner product topology.
    [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]
  • De Rham Cohomology of Smooth Manifolds
    VU University, Amsterdam Bachelorthesis De Rham Cohomology of smooth manifolds Supervisor: Author: Prof. Dr. R.C.A.M. Patrick Hafkenscheid Vandervorst Contents 1 Introduction 3 2 Smooth manifolds 4 2.1 Formal definition of a smooth manifold . 4 2.2 Smooth maps between manifolds . 6 3 Tangent spaces 7 3.1 Paths and tangent spaces . 7 3.2 Working towards a categorical approach . 8 3.3 Tangent bundles . 9 4 Cotangent bundle and differential forms 12 4.1 Cotangent spaces . 12 4.2 Cotangent bundle . 13 4.3 Smooth vector fields and smooth sections . 13 5 Tensor products and differential k-forms 15 5.1 Tensors . 15 5.2 Symmetric and alternating tensors . 16 5.3 Some algebra on Λr(V ) ....................... 16 5.4 Tensor bundles . 17 6 Differential forms 19 6.1 Contractions and exterior derivatives . 19 6.2 Integrating over topforms . 21 7 Cochains and cohomologies 23 7.1 Chains and cochains . 23 7.2 Cochains . 24 7.3 A few useful lemmas . 25 8 The de Rham cohomology 29 8.1 The definition . 29 8.2 Homotopy Invariance . 30 8.3 The Mayer-Vietoris sequence . 32 9 Some computations of de Rham cohomology 35 10 The de Rham Theorem 40 10.1 Singular Homology . 40 10.2 Singular cohomology . 41 10.3 Smooth simplices . 41 10.4 De Rham homomorphism . 42 10.5 de Rham theorem . 44 1 11 Compactly supported cohomology and Poincar´eduality 47 11.1 Compactly supported de Rham cohomology . 47 11.2 Mayer-Vietoris sequence for compactly supported de Rham co- homology . 49 11.3 Poincar´eduality .
    [Show full text]
  • De Rham Cohomology
    Chapter 5 De Rham Cohomology “Hydrodynamics procreated complex analysis, partial differential equations, Lie groups and algebra theory, cohomology theory and scientific computing.” Vladimir Arnold In Chapter 3, we discussed closed and exact forms. As a reminder, a smooth k-form w on a smooth manifold M is closed if dw = 0 on M, and is exact if w = dh for some smooth (k − 1)-form h defined on the whole M. By the fact that d2 = 0, an exact form must be closed. It is then natural to ask whether every closed form is exact. The answer is no in general. Here is a counter- example. Let M = R2nf(0, 0)g, and define y x w := − dx + dy. x2 + y2 x2 + y2 It can be computed easily that dw = 0 on M, and so w is closed. However, we can show that w is not exact. Consider the unit circle C parametrized by (x, y) = (cos t, sin t) where 0 < t < 2p, and also the induced 1-form i∗w (where i : C ! M). By direct computation, we get: I Z 2p sin t cos t i∗w = − d(cos t) + d(sin t) = 2p. C 0 cos2 t + sin2 t cos2 t + sin2 t If w were exact, then w = d f for some smooth function f : M ! R. Then, we would have: I I I Z 2p d(i∗ f ) i∗w = i∗(d f ) = d(i∗ f ) = dt. C C C 0 dt Since t = 0 and t = 2p represent the same point on C, by Fundamental Theorem of Calculus, we finally get: I i∗w = 0 C which is a contradiction! Therefore, w is not exact on R2nf(0, 0)g.
    [Show full text]
  • Direct Derivation of Liénard–Wiechert Potentials, Maxwell's
    mathematics Article Direct Derivation of Liénard–Wiechert Potentials, Maxwell’s Equations and Lorentz Force from Coulomb’s Law Hrvoje Dodig † Department of Electrical and Information Technology, Faculty of Maritime Studies, University of Split, 21000 Split, Croatia; [email protected] † Current address: Faculty of Maritime Studies, University of Split, Rudera¯ Boškovi´ca37, 21000 Split, Croatia. Abstract: In this paper, the solution to long standing problem of deriving Maxwell’s equations and Lorentz force from first principles, i.e., from Coulomb’s law, is presented. This problem was studied by many authors throughout history but it was never satisfactorily solved, and it was never solved for charges in arbitrary motion. In this paper, relativistically correct Liénard–Wiechert potentials for charges in arbitrary motion and Maxwell equations are both derived directly from Coulomb’s law by careful mathematical analysis of the moment just before the charge in motion stops. In the second part of this paper, the electrodynamic energy conservation principle is derived directly from Coulomb’s law by using similar approach. From this energy conservation principle the Lorentz force is derived. To make these derivations possible, the generalized Helmholtz theorem was derived along with two novel vector identities. The special relativity was not used in our derivations, and the results show that electromagnetism as a whole is not the consequence of special relativity, but it is rather the consequence of time retardation. Keywords: coulomb’s law; liénard–wiechert potentials; maxwell equations; lorentz force Citation: Dodig, H. Direct Derivation 1. Introduction of Liénard–Wiechert Potentials, Maxwell’s Equations and Lorentz In his famous Treatise [1,2], Maxwell derived his equations of electrodynamics based Force from Coulomb’s Law.
    [Show full text]