Coordinate Systems and Canonical Forms

Similar documents
Second-Order ODE and the Calculus of Variations

1 Introduction: connections and fiber bundles

The purpose of this lecture is to present a few applications of conformal mappings in problems which arise in physics and engineering.

9 Radon-Nikodym theorem and conditioning

TANGENT VECTORS. THREE OR FOUR DEFINITIONS.

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

1 Differentiable manifolds and smooth maps

Linear Algebra. Min Yan

LECTURE 5: SMOOTH MAPS. 1. Smooth Maps

INVERSE FUNCTION THEOREM and SURFACES IN R n

The theory of manifolds Lecture 3

MATH 423/ Note that the algebraic operations on the right hand side are vector subtraction and scalar multiplication.

LECTURE 28: VECTOR BUNDLES AND FIBER BUNDLES

Curvilinear coordinates

Integration and Manifolds

Chap. 1. Some Differential Geometric Tools

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

Chapter 1. Smooth Manifolds

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))

Chapter 4. Inverse Function Theorem. 4.1 The Inverse Function Theorem

Analysis II: The Implicit and Inverse Function Theorems

8 Nonlinear change of variables

Differential Topology Solution Set #2

Math 6455 Nov 1, Differential Geometry I Fall 2006, Georgia Tech

i = f iα : φ i (U i ) ψ α (V α ) which satisfy 1 ) Df iα = Df jβ D(φ j φ 1 i ). (39)

Math 147, Homework 1 Solutions Due: April 10, 2012

(df (ξ ))( v ) = v F : O ξ R. with F : O ξ O

LECTURE 5: THE METHOD OF STATIONARY PHASE

Math Topology II: Smooth Manifolds. Spring Homework 2 Solution Submit solutions to the following problems:

Vectors Coordinate frames 2D implicit curves 2D parametric curves. Graphics 2008/2009, period 1. Lecture 2: vectors, curves, and surfaces

Intro Vectors 2D implicit curves 2D parametric curves. Graphics 2012/2013, 4th quarter. Lecture 2: vectors, curves, and surfaces

Complex Algebraic Geometry: Smooth Curves Aaron Bertram, First Steps Towards Classifying Curves. The Riemann-Roch Theorem is a powerful tool

LECTURE 2. (TEXED): IN CLASS: PROBABLY LECTURE 3. MANIFOLDS 1. FALL TANGENT VECTORS.

14 Higher order forms; divergence theorem

Math 598 Feb 14, Geometry and Topology II Spring 2005, PSU

Vector fields Lecture 2

Cutting and pasting. 2 in R. 3 which are not even topologically

The theory of manifolds Lecture 2

4.7 The Levi-Civita connection and parallel transport

Richard S. Palais Department of Mathematics Brandeis University Waltham, MA The Magic of Iteration

APPLICATIONS OF DIFFERENTIABILITY IN R n.

Lecture 11 Hyperbolicity.

The theory of manifolds Lecture 3. Tf : TR n TR m

10. Smooth Varieties. 82 Andreas Gathmann

SMSTC (2017/18) Geometry and Topology 2.

Intro Vectors 2D implicit curves 2D parametric curves. Graphics 2011/2012, 4th quarter. Lecture 2: vectors, curves, and surfaces

Chapter 2. Bundles. 2.1 Vector bundles: definitions and examples

8 Change of variables

THEODORE VORONOV DIFFERENTIABLE MANIFOLDS. Fall Last updated: November 26, (Under construction.)

Definition 5.1. A vector field v on a manifold M is map M T M such that for all x M, v(x) T x M.

Math 147, Homework 5 Solutions Due: May 15, 2012

PICARD S THEOREM STEFAN FRIEDL

HILBERT S THEOREM ON THE HYPERBOLIC PLANE

B 1 = {B(x, r) x = (x 1, x 2 ) H, 0 < r < x 2 }. (a) Show that B = B 1 B 2 is a basis for a topology on X.

CHAPTER 3. Gauss map. In this chapter we will study the Gauss map of surfaces in R 3.

Lecture 8. Connections

LECTURE 15: COMPLETENESS AND CONVEXITY

LECTURE 10: THE PARALLEL TRANSPORT

1 The Differential Geometry of Surfaces

BACKGROUND IN SYMPLECTIC GEOMETRY

Introduction to Computational Manifolds and Applications

Symplectic critical surfaces in Kähler surfaces

13 Spherical geometry

CALCULUS ON MANIFOLDS

Practice Qualifying Exam Questions, Differentiable Manifolds, Fall, 2009.

Outline of the course

Elementary maths for GMT

Surfaces JWR. February 13, 2014

9 Change of variables

Choice of Riemannian Metrics for Rigid Body Kinematics

Notes on multivariable calculus

MULTIDIMENSIONAL SINGULAR λ-lemma

The Geometry of Euler s equation. Introduction

1 Euclidean geometry. 1.1 The metric on R n

THE INVERSE FUNCTION THEOREM FOR LIPSCHITZ MAPS

COMPLEX MULTIPLICATION: LECTURE 13

THEODORE VORONOV DIFFERENTIAL GEOMETRY. Spring 2009

Changing coordinates to adapt to a map of constant rank

NOTES ON DIFFERENTIAL FORMS. PART 3: TENSORS

we have the following equation:

612 CLASS LECTURE: HYPERBOLIC GEOMETRY

THE EULER CHARACTERISTIC OF A LIE GROUP

Quasi-conformal maps and Beltrami equation

UNIVERSITY OF DUBLIN

Lecture 13 The Fundamental Forms of a Surface

STOKES THEOREM ON MANIFOLDS

1 Vector fields, flows and Lie derivatives

Vectors. January 13, 2013

Prague, II.2. Integrability (existence of the Riemann integral) sufficient conditions... 37

HOMEWORK 2 - SOLUTIONS

Numerical Analysis: Solving Nonlinear Equations

7 Curvature of a connection

Section 3.5 The Implicit Function Theorem

M4P52 Manifolds, 2016 Problem Sheet 1

Math 396. Quotient spaces

Chapter 2 Linear Transformations

Problem: A class of dynamical systems characterized by a fast divergence of the orbits. A paradigmatic example: the Arnold cat.

1. Tangent Vectors to R n ; Vector Fields and One Forms All the vector spaces in this note are all real vector spaces.

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

Lifting Smooth Homotopies of Orbit Spaces of Proper Lie Group Actions

Transcription:

Appendix D Coordinate Systems and Canonical Forms D.1. Local Coordinates Let O be an open set in R n. We say that an n-tuple of smooth realvalued functions defined in O, (φ 1,...,φ n ), forms a local coordinate system for O if the map φ : p (φ 1 (p),...,φ n (p)) is a diffeomorphism, that is, if it is a one-to-one map of O onto some other open set U of R n and if the inverse map ψ := φ 1 : U R n is also smooth. The relation of ψ to φ is clearly completely symmetrical, and in particular ψ defines a coordinate system (ψ 1,...,ψ n )inu. D.1.1. Remark. By the Inverse Function Theorem, the necessary and sufficient condition for φ to have a smooth inverse in some neighborhood of a point p is that the Dφ p is an invertible linear map, or equivalently that the differentials (dφ i ) p are linearly independent and hence a basis for (R n ).Inotherwords,givennsmooth real-valued functions (φ 1,...,φ n ) defined near p and having linearly independent differentials at p, they always form a coordinate system in some neighborhood O of p. The most obvious coordinates are the standard coordinates, φ i (p) = p i, with O all of R n (so φ is just the identity map). If e 1,...,e n is the standard basis for R n,thenφ 1,...,φ n is just the dual basis for (R n ). We will usually denote these standard coordinates by (x 1,...,x n ). More generally, given any basis f 1,...,f n for R n,wecanlet(φ 1,...,φ n ) be the corresponding dual basis. Such 247

248 D. Coordinate Systems and Canonical Forms coordinates are called Cartesian. Inthiscase,φ is the linear isomorphism of R n that maps e i to f i.ifthef i are orthonormal, then these are called orthogonal Cartesian coordinates and φ is an orthogonal transformation. Why not just always stick with the standard coordinates? One reason is that once we understand a concept in R n in terms of arbitrary coordinates, it is easy to make sense of that concept on a general differentiable manifold. But there is another important reason. Namely, it is often possible to simplify the analysis of a problem considerably by choosing a well-adapted coordinate system. In more detail, various kinds of geometric and analytic objects have numerical descriptions in terms of a coordinate system. This observation by Descartes is of course the basis of the powerful analytic geometry approach to studying geometric questions. Now, the precise numerical description of an object is usually highly dependent on the choice of coordinate system, and it can be more or less complicated depending on that choice. Frequently, there will be certain special adapted coordinates with respect to which the numerical description of the object has a particularly simple so-called canonical form, and facts that are difficult to deduce from the description with respect to general coordinates can be obvious from the canonical form. Here is a well-known simple example. An ellipse in the plane, R 2, is given by an implicit equation of the form ax 2 +by 2 +cxy+dx+ey+ f =0, but if we choose the diffeomorphism that translates the origin to the center of the ellipse and rotates the coordinate axes to be the axes of the ellipse, then in the resulting coordinates ξ,η the implicit equation for the ellipse will have the simpler form α 2 ξ 2 + β 2 η 2 =1. Notice that this diffeomorphism is actually a Euclidean motion, so ξ and η are orthogonal Cartesian coordinates and even the metric properties of the ellipse are preserved by this change of coordinates. If that is not important in some context, we could instead use u := αξ and v := βη as our coordinates and work with the even simpler equation u 2 + v 2 =1. This example illustrates the general idea behind choosing coordinates adapted to a particular object Ω in some open set O. Namely,

D.1. Local Coordinates 249 you should think intuitively of finding a diffeomorphism φ that moves, bends, and twists Ω into an object Ω φ in U = φ(o), one that is in a canonical configuration having a particularly simple description with respect to standard coordinates. While one can apply this technique to all kinds of geometric and analytic objects, here we will concentrate on three of the objects of greatest interest to us, namely real-valued functions, f : O R; smoothcurves,σ : I O; and vector fields V defined in O. First let us consider how to define f φ, σ φ,andv φ. For functions and curves the definition is almost obvious; namely f φ : U R is defined by f φ := f φ 1 (so that f φ (φ 1 (x),...φ n (x)) = f(x 1,...,x n ) for all x O) andσ φ : I U is defined by σ φ := φ σ. Defining the vector field V φ in U from the vector field V in O is slightly more tricky. At p in O, V defines a tangent vector, (p, V (p)), that the differential of φ at p maps to a tangent vector at q = φ(p), V φ (q) :=(q, Dφ p (V (p))). Written explicitly in terms of q (and dropping the first component) gives the somewhat ugly formula V φ (q) :=Dφ φ 1 (q)v (φ 1 (q)), but note that if the diffeomorphism φ is linear (i.e., if the coordinates (φ 1,...,φ n ) are Cartesian), then Dφ p = φ, so the formula simplifies to V φ = φv φ 1. Exercise D 1. Recall that the radial (or Euler) vector field R on R n is defined by R(x) =x, or equivalently, written as a differential operator using standard coordinates, R = n i=1 xi x. Show i that if φ is any linear diffeomorphism of R n, then R φ = R, or equivalently, if (y 1,...,y n ) is any Cartesian coordinate system, then R = n i=1 y i y i. That is, the radial vector field has the remakable property that it looks the same in all Cartesian coordinate systems. Show that any linear vector field L on R n with this property must be a constant multiple of the radial field. (Hint: The only linear transformations that commute with all linear isomorphisms of R n are constant multiples of the identity.) Exercise D 2. Let f, σ, andv be as above. a) Show that if σ is a solution curve of V,thatis,ifσ (t) =V (σ(t)), then σ φ is a solution of V φ.

250 D. Coordinate Systems and Canonical Forms b) Show that if f is a constant of the motion for V (i.e., Vf 0), then f φ is a constant of the motion for V φ. A common reason for using a particular coordinate system is that these coordinates reflect the symmetry properties of a geometrical problem under consideration. While Cartesian coordinates are good for problems with translational symmetry, they are not well adapted to problems with rotational symmetry, and the analysis of such problems can often be simplified by using some sort of polar coordinates. In R 2 we have the standard polar coordinates φ(x, y) = (r(x, y),θ(x, y)) defined in O =the complement of the negative x-axis by r := x 2 + y 2 and θ :=the branch of arctan( y x ) taking values in π to π. In this case U is the infinite rectangle (0, ) ( π, π) and the inverse diffeomorphism is given by ψ(r,θ)=(rcos θ, r sin θ). Similarly, in R 3 we can use polar cylindrical coordinates r,θ,z to deal with problems that are symmetric under rotations about the z-axis or polar spherical coordinates r,θ,φ for problems with symmetry under all rotations about the origin. Exercise D 3. If f(x, y) is a real-valued function, then f φ (r,θ)= f φ 1 (r,θ)=f(rcos θ, r sin θ), so if f is C 1, then, by the chain rule, f φ f f fφ f r (r,θ)= x cos θ + y sin θ, and similarly, θ (r,θ)= xr sin θ + f y r cos θ = y f x terms of the f x i f + x y for a general coordinate system φ.. Generalize this to find formulas for fφ φ i Exercise D 4. Show that a C 1 real-valued function in the plane, f : R 2 R, is invariant under rotation if and only if y f f x = x y. (Hint: The condition for f to be invariant under rotation is that f φ (r,θ) should be a function of r only, i.e., fφ θ 0.) in D.2. Some Canonical Forms Now let us look at the standard canonical form theorems for functions, curves, and vector fields. Recall that if f is a C 1 real-valued function defined in some open set G of R n, then a point p G is called a critical point of f if df p =0 and otherwise it is called a regular point of f. Ifp is a regular

D.2. Some Canonical Forms 251 point, then we can choose a basis l 1,...,l n of (R n ) with l 1 = df p, and by the remark at the beginning of this appendix it follows that (f,l 2,...,l n ) is a coordinate system in some neighborhood O of p. This proves the following canonical form theorem for smooth realvalued functions: D.2.1. Proposition. Let f be a smooth real-valued function definedinanopensetg of R n and let p G be a regular point of f. Then there exists a coordinate system (φ 1,...,φ n ) defined in some neighborhood of p such that f φ = φ 1. Informally speaking, we can say that near any regular point a smooth function looks linear in a suitable coordinate system. Next we will see that a straight line is the canonical form for a smooth curve σ : I R n at a regular point, i.e., a point t 0 I such that σ (t 0 ) =0. D.2.2. Proposition. If t 0 is a regular point of the smooth curve σ : I R n, then there is a diffeomorphism φ of a neighborhood of σ(t 0 ) into R n such that σ φ (t) =γ(t), whereγ : R R n is the straight line t (t, 0,...,0). Proof. Without loss of generality we can assume that t 0 =0. Also, since we can anyway translate σ(t 0 ) to the origin and apply a linear isomorphism mapping σ (t 0 )toe 1 =(1, 0,...,0), we will assume that σ(0) =0 and σ (0) = e 1. Then if we define a map ψ near the origin of R n by ψ(x 1,...,x n )=σ(x 1 )+(0,x 2,...,x n ), it is clear that Dψ 0 is the identity, so by the inverse mapping theorem, ψ maps some neighborhood U of the origin diffeomorphically onto another neighborhood O. By definition, ψ γ(t) =σ(t), so if φ = ψ 1,then σ φ (t) =φ σ(t) =γ(t). Notice a pattern in the canonical form theorems for functions and curves. If we keep away from singularities, then locally a function or curve looks like the simplest example. This pattern is repeated for vector fields. Recall that a singularity of a vector field V is a point p where V (p) =0, and the simplest vector fields are the constant vector fields, such as x 1. The canonical form theorem for vector

252 D. Coordinate Systems and Canonical Forms fields, often called the Straightening Theorem, just says that near a nonsingular point a smooth vector field looks like a constant vector field. Let s try to make this more precise. D.2.3. Definition. Let V be a vector field defined in an open set O of R n, and let φ =(φ 1,...,φ n ) be local coordinates in O. Wecall φ the flow-box coordinates for V in O if V φ = φ 1. D.2.4. Remark. Let φ(x, y) =(r(x, y),θ(x, y)) denote polar coordinates in R 2. We saw above that if V = x y y x,thenvφ = θ, so that polar coordinates are flow-box coordinates for V. D.2.5. The Straightening Theorem. If V is a smooth vector field defined in an open set O of R n and p 0 O is not a singularity of V, then there exist flow-box coordinates for V in some neighborhood of p 0. Proof. This is a very strong result; it easily implies both local existence and uniqueness of solutions and smooth dependence on initial conditions. And as we shall now see, these conversely quickly give the Straightening Theorem. Without loss of generality, we can assume that p 0 is the origin and V (0) = e 1 =(1, 0,...,0). Choose ε > 0sothatfor p < ε there is a unique solution curve of V, t σ(t, p), defined for t <εand satisfying σ(0,p)=p. The existence of ε follows from the local existence and uniqueness theorem for solutions of ODE (Appendix B). Let U denote the disk of radius ε in R n and define ψ : U R n by ψ(x) =σ(x 1, (0,x 2,...,x n )). It follows from smooth dependence on initial conditions (Appendix G) that ψ is a smooth map. Exercise D 5. Complete the proof of the Straightening Theorem by first showing that Dψ 0 is the identity map (so ψ does define a local coordinate system near the origin) and secondly showing that V φ = φ 1,whereφ:= ψ 1.(Hint:Notethatσ(0, (0,x 2,...,x n )) = (0,x 2,...,x n ), while x 1 σ(x 1, (0,...,0)) = e 1. The fact that Dψ 0 is the identity is an easy consequence. Since t σ(t, p) isasolution curve of V, it follows that V (ψ(x)) = x 1 ψ(x), and it follows that Dψ p maps ( x 1 ) p to V (ψ(p)). Use this to deduce V φ = φ 1.

D.2. Some Canonical Forms 253 D.2.6. Definition. Let V : R n R n be a smooth vector field. If O is an open set in R n, then a smooth real-valued function f : O R is called a local constant of the motion for V if Vf 0inO. Notice that x 2,x 3,...,x n are constants of the motion for the vector field x 1. Hence, D.2.7. Corollary of the Straightening Theorem. If V is a smooth vector field on R n and p is any nonsingular point of V, then there exist n 1 functionally independent local constants of the motion for V defined in some neighborhood O of p. (Functionally independent means that they are part of a coordinate system.) D.2.8. CAUTION. There are numerous places in the mathematics and physics literature where one can find a statement to the effect that every vector field on R n has n 1 constants of the motion. What is presumably meant is something like the above corollary, but it is important to realize that such statements should not be taken literally there are examples of vector fields with no global constants of the motion except constants. A local constant of the motion is very different from a global one. If V is a vector field in R n and f is a local constant of the motion for V,definedinsomeopensetO, then if σ is a solution of V, f(σ(t)) will be constant on any interval I such that σ(i) O; however it will in general have different constant values on different such intervals.