The Hodge Star Operator

Similar documents
NOTES ON DIFFERENTIAL FORMS. PART 1: FORMS ON R n

The Symmetric Space for SL n (R)

NOTES ON DIFFERENTIAL FORMS. PART 3: TENSORS

Section 2. Basic formulas and identities in Riemannian geometry

MATH 4030 Differential Geometry Lecture Notes Part 4 last revised on December 4, Elementary tensor calculus

LECTURE 9: MOVING FRAMES IN THE NONHOMOGENOUS CASE: FRAME BUNDLES. 1. Introduction

4.7 The Levi-Civita connection and parallel transport

Physics 110. Electricity and Magnetism. Professor Dine. Spring, Handout: Vectors and Tensors: Everything You Need to Know

Math 52H: Multilinear algebra, differential forms and Stokes theorem. Yakov Eliashberg

CALCULUS ON MANIFOLDS. 1. Riemannian manifolds Recall that for any smooth manifold M, dim M = n, the union T M =

There are two things that are particularly nice about the first basis

Hadamard s Theorem. Rich Schwartz. September 10, The purpose of these notes is to prove the following theorem.

Survey on exterior algebra and differential forms

Extension and Representation of Divergence-free Vector Fields on Bounded Domains. Tosio Kato, Marius Mitrea, Gustavo Ponce, and Michael Taylor

William P. Thurston. The Geometry and Topology of Three-Manifolds

GEOMETRICAL TOOLS FOR PDES ON A SURFACE WITH ROTATING SHALLOW WATER EQUATIONS ON A SPHERE

Linear Algebra II. 7 Inner product spaces. Notes 7 16th December Inner products and orthonormal bases

WARPED PRODUCTS PETER PETERSEN

Math 114: Course Summary

INTRODUCTION TO ALGEBRAIC GEOMETRY

HOMEWORK 2 - RIEMANNIAN GEOMETRY. 1. Problems In what follows (M, g) will always denote a Riemannian manifold with a Levi-Civita connection.

1 First and second variational formulas for area

LECTURE 10: THE PARALLEL TRANSPORT

Lecture 13. Differential forms

Tensor Calculus, Part 2

1.13 The Levi-Civita Tensor and Hodge Dualisation

Differential Forms in R n

CS 468, Lecture 11: Covariant Differentiation

Lecture 3: Vectors. Any set of numbers that transform under a rotation the same way that a point in space does is called a vector.

Math 350 Fall 2011 Notes about inner product spaces. In this notes we state and prove some important properties of inner product spaces.

225A DIFFERENTIAL TOPOLOGY FINAL

Math 225B: Differential Geometry, Final

Introduction and Vectors Lecture 1

08a. Operators on Hilbert spaces. 1. Boundedness, continuity, operator norms

Gravitation: Tensor Calculus

REFLECTIONS IN A EUCLIDEAN SPACE

Caltech Ph106 Fall 2001

satisfying the following condition: If T : V V is any linear map, then µ(x 1,,X n )= det T µ(x 1,,X n ).

Theorema Egregium, Intrinsic Curvature

DIFFERENTIAL FORMS AND COHOMOLOGY

A Primer on Three Vectors

Cartesian Tensors. e 2. e 1. General vector (formal definition to follow) denoted by components

Clifford Algebras and Spin Groups

4 Riemannian geometry

Warped Products. by Peter Petersen. We shall de ne as few concepts as possible. A tangent vector always has the local coordinate expansion

Tensor Analysis in Euclidean Space

PROPERTIES OF SIMPLE ROOTS

Complex manifolds, Kahler metrics, differential and harmonic forms

Differential Geometry MTG 6257 Spring 2018 Problem Set 4 Due-date: Wednesday, 4/25/18

is the desired collar neighbourhood. Corollary Suppose M1 n, M 2 and f : N1 n 1 N2 n 1 is a diffeomorphism between some connected components

= ϕ r cos θ. 0 cos ξ sin ξ and sin ξ cos ξ. sin ξ 0 cos ξ

Math 426H (Differential Geometry) Final Exam April 24, 2006.

(x, y) = d(x, y) = x y.

Let V, W be two finite-dimensional vector spaces over R. We are going to define a new vector space V W with two properties:

Vectors. January 13, 2013

LECTURE 1: LINEAR SYMPLECTIC GEOMETRY

Poisson Equation on Closed Manifolds

INNER PRODUCT SPACE. Definition 1

CALCULUS ON MANIFOLDS

MILNOR SEMINAR: DIFFERENTIAL FORMS AND CHERN CLASSES

LECTURE 5: COMPLEX AND KÄHLER MANIFOLDS

7 Curvature of a connection

The Cross Product The cross product of v = (v 1,v 2,v 3 ) and w = (w 1,w 2,w 3 ) is

Metric spaces and metrizability

Klaus Janich. Vector Analysis. Translated by Leslie Kay. With 108 Illustrations. Springer

BERGMAN KERNEL ON COMPACT KÄHLER MANIFOLDS

Homework 11 Solutions. Math 110, Fall 2013.

Determinant lines and determinant line bundles

Contravariant and Covariant as Transforms

Math 396. Map making

Exercises in Geometry II University of Bonn, Summer semester 2015 Professor: Prof. Christian Blohmann Assistant: Saskia Voss Sheet 1

Cartesian Tensors. e 2. e 1. General vector (formal definition to follow) denoted by components

Week 6: Differential geometry I

STOKES THEOREM ON MANIFOLDS

Math 443 Differential Geometry Spring Handout 3: Bilinear and Quadratic Forms This handout should be read just before Chapter 4 of the textbook.

LECTURE 16: LIE GROUPS AND THEIR LIE ALGEBRAS. 1. Lie groups

Math (P)Review Part II:

Gradient Estimates and Sobolev Inequality

Seminar on Vector Field Analysis on Surfaces

Math 396. Linear algebra operations on vector bundles

1 Normal (geodesic) coordinates

DIFFERENTIAL GEOMETRY HW 12

Review Sheet for the Final

Divergence Theorems in Path Space. Denis Bell University of North Florida

The Algebra of Tensors; Tensors on a Vector Space Definition. Suppose V 1,,V k and W are vector spaces. A map. F : V 1 V k

fy (X(g)) Y (f)x(g) gy (X(f)) Y (g)x(f)) = fx(y (g)) + gx(y (f)) fy (X(g)) gy (X(f))

Basics of Special Relativity

Math 61CM - Quick answer key to section problems Fall 2018

Lecture Notes a posteriori for Math 201

HOMOGENEOUS AND INHOMOGENEOUS MAXWELL S EQUATIONS IN TERMS OF HODGE STAR OPERATOR

Math 61CM - Solutions to homework 6

MULTILINEAR ALGEBRA MCKENZIE LAMB

MATH 23a, FALL 2002 THEORETICAL LINEAR ALGEBRA AND MULTIVARIABLE CALCULUS Solutions to Final Exam (in-class portion) January 22, 2003

Totally quasi-umbilic timelike surfaces in R 1,2

Foliations of hyperbolic space by constant mean curvature surfaces sharing ideal boundary

2 bases of V. Largely for this reason, we will ignore the 0 space in this handout; anyway, including it in various results later on is always either m

Lecture II: Vector and Multivariate Calculus

Justin Solomon MIT, Spring 2017

1 Directional Derivatives and Differentiability

Eilenberg-Steenrod properties. (Hatcher, 2.1, 2.3, 3.1; Conlon, 2.6, 8.1, )

Transcription:

The Hodge Star Operator Rich Schwartz April 22, 2015 1 Basic Definitions We ll start out by defining the Hodge star operator as a map from k (R n ) to n k (R n ). Here k (R n ) denotes the vector space of alternating k-tensors on R n. Later on, we will extend this definition to alternating tensors on a finite dimensional vector space that is equipped with an inner product. Let I = (i 1,...,i k ) be some increasing multi-index of length k. That is i 1 < i 2 < i 3 <... Let J = (j 1,...,j n k ) be the complementary increasing multi-intex. For instance, if n = 7 and I = (1,3,5) then J = (2,4,6,7). Let K 0 denote the full multi-index (1,...,n). We first define on the usual basis elements: (dx I ) = ǫ(ij) dx J. (1) Here ǫ(ij) is the sign of the permutation which takes IJ to K 0. Sometimes, we will just write dx I in place of (dx I ). In general, we define ( ai ) dx I = a I ( dx I ). (2) Lemma 1.1 For any ω k (R n ) we have ω = ( 1) k(n k) ω. That is ω = ω unless k is odd and n is even, in which case ω = ω. Proof: It suffices to check this on a basis element dx I. Let J be the complementary multi-index as above. Note that dx I = ±dx I. We just have to get the sign right. When I = (1,...,k), the sign does work out. So, for any other choice of I, we just have to show that the sign only depends on n and k. That tells us that the sign works out in all cases. 1

We want to show that ǫ(ij K 0 )ǫ(ji K 0 ) only depends on n and k. Let K denote the multi-index which is the reverse of K. (If K is increasing then K is decreasing.) For the second permutation, we can reverse everything and we would get the same answer. That is: But then ǫ(ji K 0 ) = ǫ(ij K 0 ). ǫ(ij K 0 ) = ǫ(i I)ǫ(J J)ǫ(K 0 K 0 )ǫ(ij K 0 ). Putting everything together, we get ǫ(ij K 0 )ǫ(ji K 0 ) = (ǫ(ij K 0 )) 2 ǫ(i I)ǫ(J J)ǫ(K 0 K 0 ) = ǫ(i I)ǫ(J J)ǫ(K 0 K 0 ). Since I and J are increasing, the two quantities ǫ(i I) and ǫ(j J) only depend on the length of I and J respectively and not on their specific terms. The quantity ǫ(k 0 K 0 ) is the same in all cases. 2 Rotational Symmetry Let SO(n) denote the group of orientation preserving orthogonal matrices. Given any SO(n) we have an induced map : l (R n ) l (R n ). The purpose of this section is to prove that (ω) = ( ω), (3) for all ω k (R n ). This equation expresses the fact that the Hodge star operator is rotationally symmetric. In other words, the star operator commutes with the action of SO(n). Say that an element SO(n) is good if Equation 3 holds for. For instance, the identity element is obviously good. 2

Lemma 2.1 Suppose that A and B are good. Then AB is also good. Proof: The only thing to note is that (AB) = B A. So, now we compute (AB) ( ω) = B A ( ω) = B ( A (ω)) = B A (ω) = (AB) (ω). That s it. Before stating the next result, we observe that it concerns an orthogonal matrix whose determinant is 1. So, the next result does not contradict the main thing we are trying to prove. Lemma 2.2 Let k < n be some index. Let be the orthogonal matrix defined by the rules that (e j ) = e j if j k,k + 1 and (e k ) = e k+1 and (e k+1 ) = e k. Then =. Proof: Itsufficestoprovethisonabasis. LetI beanincreasingmulti-index. We will consider the equation on ω = dx I. There are 4 cases to consider. Suppose that I contains both k and k + 1. Then (ω) = ω and (ω) = ω. On the other hand ω = ±dx J, where J does not involve either k or k +1. Hence ( ω) = ω. Suppose that I contains neither k nor k+1. This case is similar to the previous case. Suppose that I contains k but not k +1. Then (ω) = dx I, where I is obtained from I by swapping out k and putting in k + 1. We have ω = ǫ(ij)dx J and ( ω) = ǫ(ij)dx J, where J is obtained from k by swapping out k + 1 for k. On the other hand, we have (ω) = ǫ(i J )dx J. To finish the proof in this case, we just have to show that ǫ(ij) = ǫ(i J ). But the two permutations just differ by composition with the transposition which swaps k and k +1. So, the equality holds. Suppose that I contains k+1 but not k. This case has a similar proof to the previous case. This takes care of all the cases. 3

Corollary 2.3 Suppose that is good and P is any permutation matrix. Then PP 1 is also good. Proof: Call an orthogonal matrix anti-good if it has the transformation law given in Lemma 2.2. The same argument as in Lemma 2.1 shows that the product of two anti-good matrices is good, and that the product of a good and an anti-good matrix is anti-good. Any permutation matrix is the product of finitely many matrices covered by Lemma 2.2. For this reason, it suffices to prove the result when P is such a matrix. In this case P = P 1, and P is anti-good. But then P 1 is anti-good, and P(P 1 ) is good. Lemma 2.4 Let be the element of SO(n) which has the following action: (e j ) = e j for j = 3,4,5,... (e 1 ) = e 1 cos(θ)+e 2 sin(θ), (e 2 ) = e 1 sin(θ)+e 2 cos(θ). In other words, rotates by θ in the e 1,e 2 plane and fixes the perpendicular directions. Then is good. Proof: It suffices to check this on a basis. We have dx I = ǫ(ij)dx J. For notational convenience, we ll consider the case when ǫ(ij) = 1. The other case has the same kind of proof. So, in short, we are considering the case when dx I = dx J. Note that the restriction of to the e 1,e 2 plane is an orientationpreserving rotation. For this reason (dx I ) = dx I when I contains both 1 and 2, and also when I contains neither 1 nor 2. Likewise (dx J ) = dx J. So, in either of these two cases, (dx I ) = (dx I ) = dx J. Consider the case when I contains 1 but not 2. Then dx I = dx 1 dx I. Here I is obtained from I by omitting 1. Similiarly, we have the equations (dx 1 dx I ) = dx 2 dx J, (dx 2 dx I ) = dx 1 dx J. Here J is obtained from J by omitting 2. The sign change in the second calculation comes from the fact that ǫ(ij) = ǫ(îĵ), where Î and Ĵ are obtained from I and J by swapping 1 and 2. 4

We set C = cos(θ) and S = sin(θ). An easy computation shows that (dx 1 ) = Cdx 1 Sdx 2, (dx 2 ) = Sdx 1 +Cdx 2. These calculations tell us that Similarly (dx I ) = ( (Cdx 1 Sdx 2 ) dx I ) = ( Cdx 1 dx I ) ( Sdx2 dx I ) = Cdx2 dx J +Sdx 1 dx J. ( dx I ) = (dx 2 dx J ) = (Sdx 1 +Cdx 2 ) dx J = Sdx 1 dx J +Cdx 2 dx J. The two expressions agree. There is one more case, when I contains 2 but not 1. This case is similar to the last case, and actually follows from the last case and Lemma 1.1. Lemma 2.5 Let i < j be two indices. Let be the element of SO(n) which has the following action: (e k ) = e k if k i,j. (e i ) = e i cos(θ)+e j sin(θ), (e j ) = e i sin(θ)+e j cos(θ). In other words, rotates by θ in the e i,e j plane and fixes the perpendicular directions. Then is good. Proof: Let P be any permutation matrix which maps 1 to i and 2 to j. Let 12 be the matrix from Lemma 2.4. Then = P 12 P 1. By the preceding lemmas, is good. Say that a matrix from Lemma 2.5 is a basic rotation. Lemma 2.6 Any element of SO(n) is the finite product of basic rotations. 5

Proof: The proof goes by induction on n 2. The case n = 2 is obvious. Let w 1,...,w n be any positively oriented orthonormal basis. Applying basic rotations, we can map w n to e n. By induction, we can map w 1,...,w n 1 to e 1,...,e n using basic rotations which fix e n. Lemma 2.5 says that the basic rotations are all good. Lemma 2.1 says that finite products of basic rotations are good. So, by the last result, all elements of SO(n) are good. This completes the proof of the rotational symmetry of the Hodge star operator. 3 A Consequence of the Symmetry Suppose that w 1,...,w n is some other positively oriented orthonormal basis of R n. We define dw I = dw i1... dw Ik. Here dw 1,...,dw n is the basis of 1-tensors dual to our new orthonormal basis. Here is a consequence of the rotational symmetry of the star operator: Lemma 3.1 dw I = ǫ(ij)dw J, where J is the increasing multi-index complementary to I. Proof: For notational convenience, we ll prove this when ǫ(ij) = 1. The other case is similar. There is an element SO(n) such that (w i ) = e i for all i. Note that (dx i ) = dw i. Hence (dx I ) = dw I and (dx J ) = dw J. Now we compute That s it. (dw I ) = (dx I ) = (dx I ) = (dx J ) = dw J. Lemma 3.1 says that we could have made the basic definition for the star operator with respect to any positively oriented orthonormal basis and we could have gotten the same answer. 6

4 General Definition Suppose now that V is an n-dimensional real vector space and Q is an inner product on V. Suppose also that V is oriented. There is an isometry g : R n V which carries the standard basis on R n to a positively oriented orthonormal basis on V, with respect to Q. Any two such isometries differ by composition with an element of SO(n). Given any ω k (V), we define That is, we do the following: ω = (g 1 ) ( g (ω)). Pull ω back to R n using g. Call this new form η. Take η. Pull η back to V using (g 1 ). This last form is ω. The fact that commutes with elements of SO(n) guarantees that this definition is independent of the choice of g. In fact, we can equally well define in the following way: Choose a positively oriented orthonormal basis v 1,...,v n for V and define (dv I ) = ǫ(ij)dv J. The rotational symmetry guarantees that we would get the same answer with respect to any positively oriented orthonormal basis. The general definition allows us to extend the definition of the star operator from inner product spaces to oriented Riemannian manifolds. Suppose that is a smooth oriented manifold with a Riemannian metric. Then we have a smooth choice of inner product Q p on each tangent space T p (). This allows us to define the Hodge star operator in each tangent space. So if ω Ω k () is some smooth k form, then ω is an n k form on. Lemma 4.1 ω is smooth. That is, ω Ω n k (). Proof: Intheneighborhoodofanypoint, hasasmoothlyvaryingandpositively oriented orthonormal basis. We just take a smoothly varying positively oriented basis, coming from a coordinate chart, and apply the Gram-Schmidt process in each tangent plane. When ω is defined locally with respect to this basis, we can see that ω is smooth. 7

5 A Variant of Stokes Theorem Suppose that is an n dimensional oriented manifold-with-boundary contained in R n. In this situation, has a canonical Riemannian metric, coming from the dot product on R n. The same goes for. A good example to consider is when is a closed ball and is the sphere bounding the ball. Let V be a vector field on, say V = (V 1,...,V n ). We have the usual associated 1-form ω = V i dx i. Note that ω is an (n 1) form on and d ω is an n-form on. Stokes theorem, applied to ω, tells us that d( ω) = We re going to re-interpret each half of this equation. The Left Side: A direct calculation shows that ω. (4) d( ω) = V i / x i dx 1... dx n = div(v) dx 1... dx n. So, the left hand side of Equation 4 equals div(v) dx 1...dx n, the usual integral of the divergence of a vector field. The Right Side: Now let s consider the right hand side of Equation 4. Consider the form ω at a point p of. We can find an oriented orthonormal basis for R n at p, say w 1,...,w n, so that w 1,...,w n 1 is an oriented orthonormal basis for T p ( ). ν = ( 1) n 1 w n is the normal vector that is compatible with Stokes theorem. We can write ω = b i w i. Note that the restriction of dw i to at p is 0 unless i = n. Therefore the restriction of ω to at p equals b n dw n. That is ω = b n ( dw n ). 8

But Finally, b n = ω(w n ) = ( 1) n 1 ω(ν) = ( 1) n 1 V ν. dw n = ( 1) n 1 dw 1... dw n 1. Putting these three equations together, we get ω = V ν dw 1... dw n 1. Our theory of integrating functions on manifolds tells us that the right hand side of Equation 4 is V ν. The Interpretation: Putting everything together, we have div(v) = V ν. (5) On the left hand side, we are integrating with the usual volume measure on Euclidean space, and on the right hand side we are integrating a function on an oriented manifold according to the theory explained in the class and in a previous handout. This is a classical n-dimensional generalization of the usual low dimensional version of Stokes theorem which involves the divergence. 6 Another Variant Again consider the case when is an n-dimensional manifold-with-boundary contained in R n. Let f be a smooth function on. Note that df is the 1- form corresponding to the gradient f. We know from the previous section that div( f) = f ν. A direct calculation shows that div( f) = f := 2 f x 2 i. Here f is the Laplacian of f. This gives us another variant of Stokes Theorem: f = f ν. (6) In physical terms, this result says that the total flux of f through equals the integral of the Laplacian of f on the interior of. 9