1 Wedge Products 2 the Exterior Derivative of Differential Forms

Total Page:16

File Type:pdf, Size:1020Kb

1 Wedge Products 2 the Exterior Derivative of Differential Forms Math 1540 Hastings section Spring 2011 Notes 13 Di¤erential 2-forms on R2: 1 Wedge products Above we de…ned the 2-form dx dy; etc. These are called the “wedge products”of the one-forms dx and dy: If we have^ general one-forms on R2; say p dx + q dy and r dx + s dy; then their wedge product is (ps qr) dx dy: (1) ^ Once we realize that ps qr is a function from R2 to R we see that this was de…ned earlier, as a 2-form , in equations (3) and (4) of notes 12. The formula (1) can be obtained from ordinary algebra: (p dx + q dy) (r dx + s dy) = pr dx dx + ps dx dy + qr dy dx + qs dy dy; ^ ^ ^ ^ ^ where we used the previously established formulas dx dx = dy dy = 0 and dy dx = dx dy: ^ ^ ^ ^ 2 The exterior derivative of di¤erential forms We …rst add to our list of di¤erential forms: De…nition 1 A 0-cell on Rn is a function : 0 Rn: f g ! In practice, we will think of a 0-cell as a point in Rn: De…nition 2 A 0-form on Rn is a function F : Rn R: (There is a reason I used F instead of f; as you will see. ! Thus, if is a 0-cell and F is a 0-form, then F ( (0)) is de…ned as a real number. Further on we will use our usual integral notation for the value of a form on a cell, giving F = F ( (0)) : Z(0) We will probably not use ! to denote a 0-form, but instead the name of the function, such as F: 1 De…nition 3 The exterior derivative of a 0-form F is the functional dF : S1;n R given by ! dF () = F ( (1)) F ( (0)) : (2) We know that not all functionals on S1;n are one-forms. For example, the length of the curve is not a 1-form because it cannot be expressed as an integral in terms 2 of 0 = (0 ; 0 ). However, considering n = 2; if F : R R is smooth, then 1 2 ! 1 @F @F ( (t) ; (t)) 0 (t) + ( (t) ; (t)) 0 (t) dt @x 1 2 1 @y 1 2 2 Z0 1 d = F (1 (t) ; 2 (t)) dt = F ( (1)) F ( (0)) = dF () (3) dt Z0 Hence, dF is a 1-form . Continuing our use of integral notation for 1-forms, we now can write dF () = dF = F ( (1)) F ( (0)) : (4) Z In fact, comparing the formula above with the formula for a general 1-form with name ! = pdx + qdy; namely, 1 ! () = (p (x (t) ; y (t)) x0 (t) + q (x (t) ; y (t)) y0 (t)) dt; (5) Z0 we see that @F @F dF = dx + dy: (6) @x @y De…nition 4 Suppose that ! = p dx + q dy is a 1-form on R2: Then the exterior derivative of ! is d! = dp dx + dq dy; (7) ^ ^ where dp and dq are the exterior derivatives of the zero forms p and q: We have no di¢ culty knowing this is a form, since the wedge product was de…ned so as always to be a form; that is, an integral involving a determinant. We can easily get a more useful formula. Note that p and q are 0-forms on R2: @p @p We can use equation (6) to give dp = dx dx + @y dy; and a similar expression for dq: The, substituting these in (7) and using the rules for wedge products, we get @q @p d! = dx dy: (8) @x @y ^ 2 3 Chains and boundaries. Recall that a k-cell in Rn is a smooth function : Ik Rn: ! n De…nition 5 A k-chain (on R ) is a set of pairs (ai; i) of numbers ai and k-cells n i (on R ): = (a1; 1) ; :::; (aN ; N ) : f g Given a k-chain ; and a k-form !; we de…ne the integral of ! over as follows: N ! = i=1ai !: (9) Z Zi P This is a real number, and we can consider ! as a functional on the set of all k-chains. R As an example, consider the following four 1-cells on R2: 1 (t) = (1; t) ; 2 (t) = (0; t) ; 3 (t) = (t; 1) ; 4 (t) = (t; 0) ; (10) for 0 t 1.(t I1.) Let 2 = (1; 1) ; ( 1; 2) ; ( 1; 3) ; (1; 4) (11) f g Notice that 1 maps [0; 1] onto the right edge of the unit square, with the orientation given by the unit tangent vector (0; 1) : This points upward. Also, 2 maps [0; 1] onto the left edge of the unit square, also with upward orientation. By pairing 1 with 2; when we compute ! for some 1-form ! = pdx + qdy; we get a term R !: Z2 This is the value of ! on the 1-form (0; 1 t) ; which goes along the left side of the unit square from top to bottom. We have not changed the image of 2; just its 1 orientation. Similar considerations apply to ( 1; 3) and to (1; 4) ; and putting these together, we see that we have computed a line integral around the boundary of I2 going in the counterclockwise direction. Next we wish to de…ne the “boundary”of a k-cell, for k = 1 and 2: We start with a 1-cell : I1 Rn: ! De…nition 6 The boundary of a 1-cell is the 0-chain @ = (1; (1)) ; ( 1; (0)) (12) f g 1 Contrast this with the usual de…nition of 2; which would change its image. 3 3.1 The fundamental theorem of calculus. Suppose that F : R1 R (which equals R1 ). Then as we saw in De…nition 2, F is a 0-form on R1. And! in De…nition 3 we saw that the exterior derivative of F is dF () = dF = F ( (1)) F ( (0)) : (13) Z This is a 1-form, as seen in equation (3), making allowances for the fact that there we were discussing a 0-form on R2. Also, if is a 0-cell, we de…ned the boundary of ; @, in equation (12). Using (9) we get F = F F: (14) Z@ Z(1) Z(0) We have to interpret the right side correctly. Recall that a 0-cell is a point, and a 0-form is a functional on the set of zero cells. As we saw earlier, the usual integral notation for forms gives F = F ( (1)) ; F = F ( (0)) : Z(1) Z(0) Using (13) and (14) we get dF = F: (15) Z Z@ This is purely from the notation; no deep mathematics is involved. However, if F is continuously di¤erentiable, with derivative F 0 = f; then we can consider the 1-form f dx; which by de…nition is given by 1 f dx = f ( (t)) 0 (t) dt: Z Z0 (We are still in one dimension.) Using the change of variable theorem and the fundamental theorem of calculus, we get (1) f = f (x) dx = F ( (1)) F ( (0)) : Z Z(0) From this we see that the 1-form dF is the same as the 1-form F 0 dx: This result does require nontrivial mathematics, namely the fundamental theorem of calculus. 4 3.2 Boundary of a 2-form, Green’s theorem for the unit square. We can now de…ne the “boundary”of a 2-cell. De…nition 7 The boundary of a 2-cell is the 1-chain @ = (1; (1; t)) ; ( 1; (0; t)) ; (1; (t; 0)) ; ( 1; (t; 1)) : (16) f g One example is the 1-chain I gave earlier (equations (10) and (11) ), which is the boundary of the 2-cell (x; y) = (x; y) : Suppose that n = 2; and pdx + qdy is a 1-form on R2: We can ask: What is @ pdx + qdy ? This is the integral over a chain, so we have to use equation (9). In this section we will only consider the case where (x; y) = (x; y) : Thus, the image R of is the unit square I2: Then @ (t) = (1; (1; t)) ; ( 1; (0; t)) ; (1; (t; 0)) ; ( 1; (t; 1)) : f g In that case, (9) becomes pdx + qdy = pdx + qdy pdx + qdy Z@ Z1 Z2 pdx + qdy + pdx + qdy; Z3 Z4 where 1; 2; 3; and 4 are 1-forms in (10). I will evaluate one of these integrals and then give the …nal answer. Since 1 (t) = (1; t) ; the standard line integral formula, as in (5) above, gives 1 1 pdx + qdy = (0 + q (1; t) (1)) dt = q (1; t) dt: Z1 Z0 Z0 Evaluating the other integrals similarly, we get 1 pdx + qdy = (q (1; t) q (0; t) + p (t; 0) p (t; 1)) dt (17) Z@ Z0 We will now evaluate d!; Z 5 where ! = pdx + qdy and where we again choose to be the 2-form (x; y) = (x; y) : Recall from equation (8) that @q @p d! = dx dy: @x @y ^ We then apply the de…nition of the 2-form in De…nition 17 from Notes 12 (equation (3) of those notes). Recall that we are taking (x; y) = (x; y) : Hence @1 @1 1 0 det @x @y = det = 1 @2 @2 0 1 @x @y ! and using Fubini’stheorem, @q @p 1 1 @q 1 1 @p dx dy = dx dy dy dx: @x @y ^ @x @y 0 0 0 0 Z Z 1 Z Z Z = ((q (1; s) q (0; s)) (p (s; 1) p (s; 0))) ds Z0 which is the same as the right side of equation (17).
Recommended publications
  • Laplacians in Geometric Analysis
    LAPLACIANS IN GEOMETRIC ANALYSIS Syafiq Johar syafi[email protected] Contents 1 Trace Laplacian 1 1.1 Connections on Vector Bundles . .1 1.2 Local and Explicit Expressions . .2 1.3 Second Covariant Derivative . .3 1.4 Curvatures on Vector Bundles . .4 1.5 Trace Laplacian . .5 2 Harmonic Functions 6 2.1 Gradient and Divergence Operators . .7 2.2 Laplace-Beltrami Operator . .7 2.3 Harmonic Functions . .8 2.4 Harmonic Maps . .8 3 Hodge Laplacian 9 3.1 Exterior Derivatives . .9 3.2 Hodge Duals . 10 3.3 Hodge Laplacian . 12 4 Hodge Decomposition 13 4.1 De Rham Cohomology . 13 4.2 Hodge Decomposition Theorem . 14 5 Weitzenb¨ock and B¨ochner Formulas 15 5.1 Weitzenb¨ock Formula . 15 5.1.1 0-forms . 15 5.1.2 k-forms . 15 5.2 B¨ochner Formula . 17 1 Trace Laplacian In this section, we are going to present a notion of Laplacian that is regularly used in differential geometry, namely the trace Laplacian (also called the rough Laplacian or connection Laplacian). We recall the definition of connection on vector bundles which allows us to take the directional derivative of vector bundles. 1.1 Connections on Vector Bundles Definition 1.1 (Connection). Let M be a differentiable manifold and E a vector bundle over M. A connection or covariant derivative at a point p 2 M is a map D : Γ(E) ! Γ(T ∗M ⊗ E) 1 with the properties for any V; W 2 TpM; σ; τ 2 Γ(E) and f 2 C (M), we have that DV σ 2 Ep with the following properties: 1.
    [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]
  • Differential Forms Diff Geom II, WS 2015/16
    J.M. Sullivan, TU Berlin B: Differential Forms Diff Geom II, WS 2015/16 B. DIFFERENTIAL FORMS instance, if S has k elements this gives a k-dimensional vector space with S as basis. We have already seen one-forms (covector fields) on a Given vector spaces V and W, let F be the free vector space over the set V × W. (This consists of formal sums manifold. In general, a k-form is a field of alternating k- P linear forms on the tangent spaces of a manifold. Forms ai(vi, wi) but ignores all the structure we have on the set are the natural objects for integration: a k-form can be in- V × W.) Now let R ⊂ F be the linear subspace spanned by tegrated over an oriented k-submanifold. We start with ten- all elements of the form: sor products and the exterior algebra of multivectors. (v + v0, w) − (v, w) − (v0, w), (v, w + w0) − (v, w) − (v, w0), (av, w) − a(v, w), (v, aw) − a(v, w). B1. Tensor products These correspond of course to the bilinearity conditions Recall that, if V, W and X are vector spaces, then a map we started with. The quotient vector space F/R will be the b: V × W → X is called bilinear if tensor product V ⊗ W. We have started with all possible v ⊗ w as generators and thrown in just enough relations to b(v + v0, w) = b(v, w) + b(v0, w), make the map (v, w) 7→ v ⊗ w be bilinear. b(v, w + w0) = b(v, w) + b(v, w0), The tensor product is commutative: there is a natural linear isomorphism V⊗W → W⊗V such that v⊗w 7→ w⊗v.
    [Show full text]
  • Vector Calculus and Differential Forms with Applications To
    Vector Calculus and Differential Forms with Applications to Electromagnetism Sean Roberson May 7, 2015 PREFACE This paper is written as a final project for a course in vector analysis, taught at Texas A&M University - San Antonio in the spring of 2015 as an independent study course. Students in mathematics, physics, engineering, and the sciences usually go through a sequence of three calculus courses before go- ing on to differential equations, real analysis, and linear algebra. In the third course, traditionally reserved for multivariable calculus, stu- dents usually learn how to differentiate functions of several variable and integrate over general domains in space. Very rarely, as was my case, will professors have time to cover the important integral theo- rems using vector functions: Green’s Theorem, Stokes’ Theorem, etc. In some universities, such as UCSD and Cornell, honors students are able to take an accelerated calculus sequence using the text Vector Cal- culus, Linear Algebra, and Differential Forms by John Hamal Hubbard and Barbara Burke Hubbard. Here, students learn multivariable cal- culus using linear algebra and real analysis, and then they generalize familiar integral theorems using the language of differential forms. This paper was written over the course of one semester, where the majority of the book was covered. Some details, such as orientation of manifolds, topology, and the foundation of the integral were skipped to save length. The paper should still be readable by a student with at least three semesters of calculus, one course in linear algebra, and one course in real analysis - all at the undergraduate level.
    [Show full text]
  • 6.10 the Generalized Stokes's Theorem
    6.10 The generalized Stokes’s theorem 645 6.9.2 In the text we proved Proposition 6.9.7 in the special case where the mapping f is linear. Prove the general statement, where f is only assumed to be of class C1. n m 1 6.9.3 Let U R be open, f : U R of class C , and ξ a vector field on m ⊂ → R . a. Show that (f ∗Wξ)(x)=W . Exercise 6.9.3: The matrix [Df(x)]!ξ(x) adj(A) of part c is the adjoint ma- b. Let m = n. Show that if [Df(x)] is invertible, then trix of A. The equation in part b (f ∗Φ )(x) = det[Df(x)]Φ 1 . is unsatisfactory: it does not say ξ [Df(x)]− ξ(x) how to represent (f ∗Φ )(x) as the ξ c. Let m = n, let A be a square matrix, and let A be the matrix obtained flux of a vector field when the n n [j,i] × from A by erasing the jth row and the ith column. Let adj(A) be the matrix matrix [Df(x)] is not invertible. i+j whose (i, j)th entry is (adj(A))i,j =( 1) det A . Show that Part c deals with this situation. − [j,i] A(adj(A)) = (det A)I and f ∗Φξ(x)=Φadj([Df(x)])ξ(x) 6.10 The generalized Stokes’s theorem We worked hard to define the exterior derivative and to define orientation of manifolds and of boundaries. Now we are going to reap some rewards for our labor: a higher-dimensional analogue of the fundamental theorem of calculus, Stokes’s theorem.
    [Show full text]
  • Differential Geometry: Discrete Exterior Calculus
    Differential Geometry: Discrete Exterior Calculus [Discrete Exterior Calculus (Thesis). Hirani, 2003] [Discrete Differential Forms for Computational Modeling. Desbrun et al., 2005] [Build Your Own DEC at Home. Elcott et al., 2006] Chains Recall: A k-chain of a simplicial complex Κ is linear combination of the k-simplices in Κ: c = ∑c(σ )⋅σ σ∈Κ k where c is a real-valued function. The dual of a k-chain is a k-cochain which is a linear map taking a k-chain to a real value. Chains and cochains related through evaluation: –Cochains: What we integrate –Chains: What we integrate over Boundary Operator Recall: The boundary ∂σ of a k-simplex σ is the signed union of all (k-1)-faces. ∂ ∂ ∂ ∂ +1 -1 The boundary operator extends linearly to act on chains: ∂c = ∂ ∑c(σ )⋅σ = ∑c(σ )⋅∂σ σ∈Κ k σ∈Κ k Discrete Exterior Derivative Recall: The exterior derivative d:Ωk→ Ωk+1 is the operator on cochains that is the complement of the boundary operator, defined to satisfy Stokes’ Theorem. -1 -2 -1 -3.5 -2 0.5 0.5 2 0 0.5 2 0 0.5 2 1.5 1 1 Since the boundary of the boundary is empty: ∂∂ = 0 dd = 0 The Laplacian Definition: The Laplacian of a function f is a function defined as the divergence of the gradient of f: ∆f = div(∇f ) = ∇ ⋅∇f The Laplacian Definition: The Laplacian of a function f is a function defined as the divergence of the gradient of f: ∆f = div(∇f ) = ∇ ⋅∇f Q.
    [Show full text]
  • Arxiv:1412.2393V4 [Gr-Qc] 27 Feb 2019 2.6 Geodesics and Normal Coordinates
    Riemannian Geometry: Definitions, Pictures, and Results Adam Marsh February 27, 2019 Abstract A pedagogical but concise overview of Riemannian geometry is provided, in the context of usage in physics. The emphasis is on defining and visualizing concepts and relationships between them, as well as listing common confusions, alternative notations and jargon, and relevant facts and theorems. Special attention is given to detailed figures and geometric viewpoints, some of which would seem to be novel to the literature. Topics are avoided which are well covered in textbooks, such as historical motivations, proofs and derivations, and tools for practical calculations. As much material as possible is developed for manifolds with connection (omitting a metric) to make clear which aspects can be readily generalized to gauge theories. The presentation in most cases does not assume a coordinate frame or zero torsion, and the coordinate-free, tensor, and Cartan formalisms are developed in parallel. Contents 1 Introduction 2 2 Parallel transport 3 2.1 The parallel transporter . 3 2.2 The covariant derivative . 4 2.3 The connection . 5 2.4 The covariant derivative in terms of the connection . 6 2.5 The parallel transporter in terms of the connection . 9 arXiv:1412.2393v4 [gr-qc] 27 Feb 2019 2.6 Geodesics and normal coordinates . 9 2.7 Summary . 10 3 Manifolds with connection 11 3.1 The covariant derivative on the tensor algebra . 12 3.2 The exterior covariant derivative of vector-valued forms . 13 3.3 The exterior covariant derivative of algebra-valued forms . 15 3.4 Torsion . 16 3.5 Curvature .
    [Show full text]
  • VISUALIZING EXTERIOR CALCULUS Contents Introduction 1 1. Exterior
    VISUALIZING EXTERIOR CALCULUS GREGORY BIXLER Abstract. Exterior calculus, broadly, is the structure of differential forms. These are usually presented algebraically, or as computational tools to simplify Stokes' Theorem. But geometric objects like forms should be visualized! Here, I present visualizations of forms that attempt to capture their geometry. Contents Introduction 1 1. Exterior algebra 2 1.1. Vectors and dual vectors 2 1.2. The wedge of two vectors 3 1.3. Defining the exterior algebra 4 1.4. The wedge of two dual vectors 5 1.5. R3 has no mixed wedges 6 1.6. Interior multiplication 7 2. Integration 8 2.1. Differential forms 9 2.2. Visualizing differential forms 9 2.3. Integrating differential forms over manifolds 9 3. Differentiation 11 3.1. Exterior Derivative 11 3.2. Visualizing the exterior derivative 13 3.3. Closed and exact forms 14 3.4. Future work: Lie derivative 15 Acknowledgments 16 References 16 Introduction What's the natural way to integrate over a smooth manifold? Well, a k-dimensional manifold is (locally) parametrized by k coordinate functions. At each point we get k tangent vectors, which together form an infinitesimal k-dimensional volume ele- ment. And how should we measure this element? It seems reasonable to ask for a function f taking k vectors, so that • f is linear in each vector • f is zero for degenerate elements (i.e. when vectors are linearly dependent) • f is smooth These are the properties of a differential form. 1 2 GREGORY BIXLER Introducing structure to a manifold usually follows a certain pattern: • Create it in Rn.
    [Show full text]
  • Nonlocal Exterior Calculus on Riemannian Manifolds
    The Pennsylvania State University The Graduate School Department of Mathematics NONLOCAL EXTERIOR CALCULUS ON RIEMANNIAN MANIFOLDS A Dissertation in Mathematics by Thinh Duc Le c 2013 Thinh Duc Le Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy August 2013 ii The dissertation of Thinh Duc Le was reviewed and approved* by the following: Qiang Du Verne M. Willaman Professor of Mathematics Dissertation Adviser Chair of Committee Long-Qing Chen Distinguished Professor of Material Sciences and Engineering Ping Xu Distinguished Professor of Mathematics Mathieu Stienon Associate Professor of Mathematics Svetlana Katok Director of Graduate Studies, Department of Mathematics *Signatures are on file in the Graduate School. iii Abstract Exterior calculus and differential forms are basic mathematical concepts that have been around for centuries. Variations of these concepts have also been made over the years such as the discrete exterior calculus and the finite element exterior calculus. In this work, motivated by the recent studies of nonlocal vector calculus we develop a nonlocal exterior calculus framework on Riemannian manifolds which mimics many properties of the standard (local/smooth) exterior calculus. However the key difference is that nonlocal \interactions" (functions, operators, fields,...) are not required to be smooth. Also any point/particle can interact directly with any other point/particle in the studied domain (at least in principle). Just as in the standard context, we introduce all necessary elements of exterior calculus such as forms, vector fields, exterior derivatives, etc. We point out the relationships between these elements with the known ones in (local) exterior calcu- lus, discrete exterior calculus, etc.
    [Show full text]
  • An Introduction to Space–Time Exterior Calculus
    mathematics Article An Introduction to Space–Time Exterior Calculus Ivano Colombaro 1,* , Josep Font-Segura 1 and Alfonso Martinez 1 1 Department of Information and Communication Technologies, Universitat Pompeu Fabra, 08018 Barcelona, Spain; [email protected] (J.F.-S.); [email protected] (A.M.) * Correspondence: [email protected]; Tel.: +34-93-542-1496 Received: 21 May 2019; Accepted: 18 June 2019; Published: 21 June 2019 Abstract: The basic concepts of exterior calculus for space–time multivectors are presented: Interior and exterior products, interior and exterior derivatives, oriented integrals over hypersurfaces, circulation and flux of multivector fields. Two Stokes theorems relating the exterior and interior derivatives with circulation and flux, respectively, are derived. As an application, it is shown how the exterior-calculus space–time formulation of the electromagnetic Maxwell equations and Lorentz force recovers the standard vector-calculus formulations, in both differential and integral forms. Keywords: exterior calculus, exterior algebra, electromagnetism, Maxwell equations, differential forms, tensor calculus 1. Introduction Vector calculus has, since its introduction by J. W. Gibbs [1] and Heaviside, been the tool of choice to represent many physical phenomena. In mechanics, hydrodynamics and electromagnetism, quantities such as forces, velocities and currents are modeled as vector fields in space, while flux, circulation, divergence or curl describe operations on the vector fields themselves. With relativity theory, it was observed that space and time are not independent but just coordinates in space–time [2] (pp. 111–120). Tensors like the Faraday tensor in electromagnetism were quickly adopted as a natural representation of fields in space–time [3] (pp.
    [Show full text]
  • 3.2. Differential Forms. the Line Integral of a 1-Form Over a Curve Is a Very Nice Kind of Integral in Several Respects
    11 3.2. Differential forms. The line integral of a 1-form over a curve is a very nice kind of integral in several respects. In the spirit of differential geometry, it does not require any additional structure, such as a metric. In practice, it is relatively simple to compute. If two curves c and c! are very close together, then α α. c c! The line integral appears in a nice integral theorem. ≈ ! ! This contrasts with the integral c f ds of a function with respect to length. That is defined with a metric. The explicit formula involves a square root, which often leads to difficult integrals. Even! if two curves are extremely close together, the integrals can differ drastically. We are interested in other nice integrals and other integral theorems. The gener- alization of a smooth curve is a submanifold. A p-dimensional submanifold Σ M ⊆ is a subset which looks locally like Rp Rn. We would like to make sense of an expression⊆ like ω. "Σ What should ω be? It cannot be a scalar function. If we could integrate functions, then we could compute volumes, but without some additional structure, the con- cept of “the volume of Σ” is meaningless. In order to compute an integral like this, we should first use a coordinate system to chop Σ up into many tiny p-dimensional parallelepipeds (think cubes). Each of these is described by p vectors which give the edges touching one corner. So, from p vectors v1, . , vp at a point, ω should give a number ω(v1, .
    [Show full text]
  • Differential Forms: Unifying the Theorems of Vector Calculus
    Differential Forms: Unifying the Theorems of Vector Calculus In class we have discussed the important vector calculus theorems known as Green's Theorem, Divergence Theorem, and Stokes's Theorem. Interestingly enough, all of these results, as well as the fundamental theorem for line integrals (so in particular the fundamental theorem of calculus), are merely special cases of one and the same theorem named after George Gabriel Stokes (1819-1903). This all-including theorem is stated in terms of differential forms. Without giving exact definitions, let us use the language of differential forms to unify the theorems we have learned. A striking pattern will emerge. 0-forms. A scalar field (i.e. a real-valued function) is also called a 0-form. 1-forms. Recall the following notation for line integrals (in 3-space, say): Z Z b Z F(r) · dr = P x0(t)dt +Q y0(t)dt +R z0(t)dt = P dx + Qdy + Rdz; C a | {z } | {z } | {z } C dx dy dz where F = P i + Qj + Rk. The expression P (x; y; z)dx + Q(x; y; z)dy + R(x; y; z)dz is called a 1-form. 2-forms. In evaluating surface integrals we can introduce similar notation: ZZ F · n dS S ZZ @r × @r @r @r @u @v = F · × dudv Γ @r @r @u @v @u × @v ZZ @y @y ZZ @x @x ZZ @x @x @u @v @u @v @u @v = P @z @z dudv − Q @z @z dudv + R @y @y dudv Γ @u @v Γ @u @v Γ @u @v | {z } | {z } | {z } dy^dz dx^dz dx^dy ZZ = P dy ^ dz − Qdx ^ dz + Rdx ^ dy: S We call P (x; y; z)dy ^ dz − Q(x; y; z)dx ^ dz + R(x; y; z)dx ^ dy a 2-form.
    [Show full text]