Motivation towards Clifford Analysis: The classical Cauchy s Integral Theorem from a Clifford perspective

Similar documents
Rough metrics, the Kato square root problem, and the continuity of a flow tangent to the Ricci flow

A SYMMETRIC FUNCTIONAL CALCULUS FOR NONCOMMUTING SYSTEMS OF SECTORIAL OPERATORS

Square roots of operators associated with perturbed sub-laplacians on Lie groups

Multivector Calculus

Tools from Lebesgue integration

Inequalities of Babuška-Aziz and Friedrichs-Velte for differential forms

1. If 1, ω, ω 2, -----, ω 9 are the 10 th roots of unity, then (1 + ω) (1 + ω 2 ) (1 + ω 9 ) is A) 1 B) 1 C) 10 D) 0

Geometry and the Kato square root problem

Quadratic estimates and perturbations of Dirac type operators on doubling measure metric spaces

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

f k j M k j Among them we recognize the gradient operator on polynomial-valued functions which is part of the complete gradient (see [9]) u j xj.

Square roots of perturbed sub-elliptic operators on Lie groups

The Cylindrical Fourier Transform

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

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)

Archiv der Mathematik Holomorphic approximation of L2-functions on the unit sphere in R^3

Notions such as convergent sequence and Cauchy sequence make sense for any metric space. Convergent Sequences are Cauchy

On higher order Cauchy-Pompeiu formula in Clifford analysis and its applications

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

Geometry and the Kato square root problem

TEST CODE: PMB SYLLABUS

SMSTC (2017/18) Geometry and Topology 2.

Some notes on Coxeter groups

Linear Algebra Massoud Malek

Qualifying Exams I, 2014 Spring

COMPLEX ANALYSIS AND RIEMANN SURFACES

CHAPTER 8. Smoothing operators

Some Rarita-Schwinger Type Operators

Geometry and the Kato square root problem

Fredholmness of Some Toeplitz Operators

Implicit Functions, Curves and Surfaces

10. Smooth Varieties. 82 Andreas Gathmann

Math 699 Reading Course, Spring 2007 Rouben Rostamian Homogenization of Differential Equations May 11, 2007 by Alen Agheksanterian

THEOREMS, ETC., FOR MATH 515

Sec. 1.1: Basics of Vectors

QUATERNIONS AND ROTATIONS

Hilbert Spaces. Hilbert space is a vector space with some extra structure. We start with formal (axiomatic) definition of a vector space.

Some Rarita-Schwinger Type Operators

Reflections and Rotations in R 3

TANGENT VECTORS. THREE OR FOUR DEFINITIONS.

NOTES ON PRODUCT SYSTEMS

The Kato square root problem on vector bundles with generalised bounded geometry

Math 67. Rumbos Fall Solutions to Review Problems for Final Exam. (a) Use the triangle inequality to derive the inequality

The Fueter Theorem and Dirac symmetries

The Kato square root problem on vector bundles with generalised bounded geometry

Ph.D. Qualifying Exam: Algebra I

The following definition is fundamental.

Vector Spaces. Vector space, ν, over the field of complex numbers, C, is a set of elements a, b,..., satisfying the following axioms.

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

SYMPLECTIC MANIFOLDS, GEOMETRIC QUANTIZATION, AND UNITARY REPRESENTATIONS OF LIE GROUPS. 1. Introduction

Chapter 8 Integral Operators

Mathematical Methods wk 1: Vectors

Mathematical Methods wk 1: Vectors

Definitions and Properties of R N

Fix(g). Orb(x) i=1. O i G. i=1. O i. i=1 x O i. = n G

MULTIPLIERS ON SPACES OF FUNCTIONS ON A LOCALLY COMPACT ABELIAN GROUP WITH VALUES IN A HILBERT SPACE. Violeta Petkova

Lecture Notes on Metric Spaces

Cambridge University Press The Mathematics of Signal Processing Steven B. Damelin and Willard Miller Excerpt More information

Canonical systems of basic invariants for unitary reflection groups

Conjugate Harmonic Functions and Clifford Algebras

Functional Analysis HW #5

Differential forms. Proposition 3 Let X be a Riemann surface, a X and (U, z = x + iy) a coordinate neighborhood of a.

arxiv:math/ v1 [math.dg] 6 Jul 1998

be the set of complex valued 2π-periodic functions f on R such that

1 Structures 2. 2 Framework of Riemann surfaces Basic configuration Holomorphic functions... 3

3. The Sheaf of Regular Functions

Chapter 3. Riemannian Manifolds - I. The subject of this thesis is to extend the combinatorial curve reconstruction approach to curves

where m is the maximal ideal of O X,p. Note that m/m 2 is a vector space. Suppose that we are given a morphism

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

Vectors in Function Spaces

A Brief Introduction to Functional Analysis

June 2014 Written Certification Exam. Algebra

CONSIDERATION OF COMPACT MINIMAL SURFACES IN 4-DIMENSIONAL FLAT TORI IN TERMS OF DEGENERATE GAUSS MAP

Math 302 Outcome Statements Winter 2013

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

Outline of the course

L19: Fredholm theory. where E u = u T X and J u = Formally, J-holomorphic curves are just 1

STOKES THEOREM ON MANIFOLDS

Introduction and Vectors Lecture 1

SPECTRAL THEORY EVAN JENKINS

Geometric interpretation of signals: background

Sobolev spaces. May 18

Fréchet differentiability of the norm of L p -spaces associated with arbitrary von Neumann algebras

1.3.1 Definition and Basic Properties of Convolution

Symplectic and Poisson Manifolds

AN EXPOSITION OF THE RIEMANN ROCH THEOREM FOR CURVES

Determinant lines and determinant line bundles

Introduction to Index Theory. Elmar Schrohe Institut für Analysis

Integral Jensen inequality

THEOREMS, ETC., FOR MATH 516

Clifford Algebras and Spin Groups

FFTs in Graphics and Vision. Groups and Representations

22.3. Repeated Eigenvalues and Symmetric Matrices. Introduction. Prerequisites. Learning Outcomes

F (z) =f(z). f(z) = a n (z z 0 ) n. F (z) = a n (z z 0 ) n

FRAMES AND TIME-FREQUENCY ANALYSIS

From Bernstein approximation to Zauner s conjecture

1 Smooth manifolds and Lie groups

SPRING 2008: POLYNOMIAL IMAGES OF CIRCLES

INTRODUCTION TO ALGEBRAIC GEOMETRY

CONVOLUTION OPERATORS IN INFINITE DIMENSION

Transcription:

Motivation towards Clifford Analysis: The classical Cauchy s Integral Theorem from a Clifford perspective Lashi Bandara November 26, 29 Abstract Clifford Algebras generalise complex variables algebraically and analytically. In particular, this includes a generalisation of the notion of holomorphy and Cauchy s Integral Theorem. We give a brief overview of the general theory, before proving the classical Cauchy s Integral Theorem via differential forms and Stokes Theorem, to give a Clifford approach to complex variables that lends itself to generalisation. 1 An overview of Clifford Analysis We let F be R or C. We construct the Clifford Algebra over F as follows. Let {e, e 1,..., e n } be the standard basis for R R n. We define multiplication of these basis elements in the following way: e = 1 e 2 j = e e i e j = e j e i 1 j n 1 i < j n Whenever 1 j 1 < j 2 < < j k n, write S = {j 1, j 2,..., j k } and define e S = e j1 e j2... e jk. For S =, define e = e. The Clifford algebra F (n) is defined as F (n) = F- span {e S : S {1,..., n}} which makes it a 2 n -dimensional algebra. If u, v F (n), then u = S u Se S and v = R u Re R with u S, v R F and uv = S,R u Sv R e S e R. The term u is the scalar part of u. We equip the algebra with an involution. The Clifford conjugate for a basis element e S is the element e S satisfying e S e S = 1 = e = e S e S. Observe then that e S = ±e S, with the sign chosen appropriately. Then, the Clifford conjugate of a general u F (n) is defined as u = S u Se S. A calculation reveals that uv = v u. Also, note that uv = S u Sv S + S R u Sv R e S e R. From this observation we define an inner product u, v = (uv) = S u Sv S. We write = 2 for the associated norm. We can embed R n+1 in F (n) by identifying it with the subspace R- span {e,..., e n } via the map (x,..., x n ) n j= x je j. Then, whenever m n, we can consider R m R n+1 by identifying R m with the subspace span {e 1,..., e m } and so via transitivity, we can embed R m in F (n). 1

Not every element of F (n) is invertible [MP87]. However, if x R n+1, then it does have an inverse. An easy calculation shows that e j = e j whenever 1 j n. Given an element x R n+1 the conjugate x = x n j=1 x je j. The Kelvin inverse of x is then given by x 1 = x x 2 = x n j=1 x je j. n j= x2 j We highlight and important fact. The Clifford algebra R (1) can be identified with the Complex numbers. This can be seen by identifying e 1 and ı e 1. We emphasise that this is an algebra isomorphism over R. It is straightforward to check that the Clifford conjugate agrees with the classical complex conjugate. Furthermore, every element in R (1) is invertible, because in the case n = 1, there is a vector space isomorphism R (1) =R R 2. We also note that the algebra R (2) can be identified with the Quaternions. A Banach module over F (n) is a Banach space X over F with an operation of multiplication by elements of F (n) with a κ 1 such that xu κ u x and ux κ u x for all x X and u F (n). In some sense, a Banach module is a generalisation of the concept of a space over a field, since in the module we are able to multiply by elements of F (n). Suppose X and Y are Banach modules over F (n). We say that A : X Y is a right module homomorphism if (Ax)u = A(xu) for all x X and u F (n). Certainly, we also have a notion of left module homomorphism. Namely, it is a map B : X Y satisfying (ux)b = u(xb). That this situation arises can be seen by considering the operator A = n j= A je j, with A j L (X, Y). Then, in general, we do not expect xa = Ax. The space of continuous right homomorphisms is denoted L (n) (X, Y) and it is considered as a Banach space with the uniform operator topology. Let X (n) = X F (n). Every ξ X (n) arises in the form ξ = S x S e S, where x S X. For convenience, we omit the and simply write ξ = S x Se S. Multiplication by u F (n) is defined by ξu = S,R u Rx S e S e R and uξ = S,R u Rx S e R e S. We equip ξ X (n) with the norm ξ = ( S x S 2 ) 1 2. It follows then that X (n) is a Banach module with κ = 1. We highlight a trivial fact. Since F is a Banach space over F, the module F F (n) = F(n) is in fact a Banach module over F (n). This observation shows our choice of notation for a Clifford algebra is consistent with that of a Banach module. Furthermore, the space (L (X, Y)) (n) can be identified with L (n) (X (n), Y (n) ) [Jef4, 3.2]. Now we can begin to consider some analytic properties of Clifford valued functions. Our approach is taken from [MP87] and [Jef4]. A more measure theoretic approach can be found in [Mit94, 1.2]. Given R n+1 an open set, any function f : F (n) can be written as f = S f Se S. Often, we regard f as a function f : F (n) F (n) via the canonical embedding. We say f C (, F (n) ) if f S C () for every S. For such a function, letting j denote the partial derivative in the direction j, we define n D = e + j e j. Then, Df = S ( f S e S + n j=1 jf S e j e S ) and fd = S ( f S e S + n j=1 jf S e S e j ). Such a function is said to be left monogenic if Df = and right monogenic if fd = on. In the case of R (1) =R C, we find that Df = fd and Df = are exactly the holomorphic functions. The notion of monogenic generalises holomorphy. 2 j=1

Let (Σ, M, µ) be a measure space. Then, the integral of a µ-measurable function f over a set is given by f dµ = ( f S dµ)e S. S With this in mind, we have the following result as a first step towards a generalisation of Cauchy s Integral Theorem [FB82, 8.8]. Theorem 1.1 (Existence and uniqueness of fundamental solution). There exists a unique left and right fundamental solution E L 1 loc (Rn+1, F (n) ) in the sense of distributions S (R n+1 ) (n) to the equation DE = ED = δ e. When x, the fundamental solution is given by E(x) = 1 x ω n+1 x n+1 1 where ω n+1 = 1 n+1 2π 2 Γ( n+1 2 ), the area of the unit sphere Sn. Furthermore, E is both left and right monogenic on R n+1 \ {}. A detailed treatment of distributions taking values in Banach modules is found in [FB82, 2]. From the fundamental solution, we can construct a generalised Cauchy kernel. Analogous to the case of complex variables, we write the Cauchy kernel as G x (ω) = E(ω x) for all ω x and the following theorem shows that it is indeed a generalisation of the classical Cauchy kernel. Theorem 1.2 (Monogenic Cauchy s integral theorem). Suppose that R n+1 F (n) is a bounded open set with a smooth boundary, and ν : R n+1 F (n) R n+1 F (n) is the unit outer normal to. Furthermore, let µ denote the surface measure on. Suppose f, g : F (n) where and are open sets. If f is left monogenic, and g is right monogenic. Then the following hold: 1. G x(ω)ν(ω)f(ω) dµ(ω) = f(x) when x and otherwise, 2. g(ω)ν(ω)g x(ω) dµ(ω) = g(x) when x and otherwise, 3. g(ω)ν(ω)f(ω) dµ(ω) =. 2 The case of R (1) We now focus our attention on R (1) and prove Theorem 1.2 for when n = 1 using differential forms to be adequately general. For classical reasons, the directions e and e 1 are respectively associated with the variables x and y. Then, D = x e + y e 1. We emphasise that we always regard R 2 R (1). We firstly show that we have a product rule for D. Proposition 2.1. Let R 2 be an open set and let f, g : R (1) be differentiable. Then D(fg) = (Df)g + f(dg). Proof. Firstly, we write f = f e + f 1 e 1 and g = g e + f 1 e 1. Then, Df = ( x e + y e 1 )(f e + f 1 e 1 ) = ( x f y f 1 )e + ( x f 1 + y f )e 1 3

and fg = (f g f 1 g 1 )e + (f 1 g + f g 1 )e 1. We compute D(fg): Also, D(fg) = [ x (f g f 1 g 1 ) y (f 1 g + f g 1 )]e + [ x (f 1 g + f g 1 ) + y (f g f 1 g 1 )]e 1 = [ x f g + f x g x f 1 g 1 f 1 x g 1 y f 1 g f 1 y g y f g 1 f y g 1 ]e + [ x f 1 g + f 1 x g + x f g 1 + f x g 1 + y f g + f y g y f 1 g 1 f 1 y g 1 ]e 1 (Df)g = [( x f y f 1 )g ( x f 1 + y f )g 1 ]e + [( x f 1 + y f )g + ( x f y f 1 )g 1 ]e 1 and by interchanging f and g, (Dg)f = [( x g y g 1 )f ( x g 1 + y g )f 1 ]e + [( x g 1 + y g )f + ( x g y g 1 )f 1 ]e 1 It is then a simple but tedious task to compare the expressions to conclude D(fg) = (Df)g + f(dg). As a consequence of Theorem 1.1, the Cauchy kernel G p (ω) is simply a translation of of E and we have the following Corollary. Corollary 2.2. Let f = G p and suppose g is left monogenic. Then, D(G p g) = (DG p )g on \{p}. The following is a Stokes type theorem for the operator D. Proposition 2.3. Let, R 2 be open sets such that and is bounded with smooth boundary. Suppose also that is equipped with unit outer normal ν and surface measure µ. Let f C 1 (, R (1) ). Then, D(G p f) dl = G p νf dµ for all p. Proof. First, let θ = (G p f) = (G p ) f (G p ) 1 f 1 and θ 1 = (G p f) 1 = (G p ) f 1 + (G p ) 1 f. It then follows that D(G p f) = ( x θ y θ 1 )e + ( x θ 1 + y θ )e 1 and therefore, [ ] [ ] D(G p f) dl = ( x θ y θ 1 ) dl e + ( x θ 1 + y θ ) dl e 1. Define ξ 1, ξ 2 n 1 ( ) by ξ 1 = θ 1 dx + θ dy and ξ 2 = θ dx + θ 1 dy. Therefore, dξ 1 = ( x θ y θ 1 ) dx dy and dξ 2 = ( x θ 1 + y θ ) dx dy. By the definition of integration of an n-form and by the application of Stokes Theorem, ( x θ y θ 1 ) dl = ( x θ y θ 1 ) dx dy = dξ 1 = ξ 1 = (θ 1 dx + θ dy) and by similar calculation ( x θ 1 + y θ ) dl = ( θ dx + θ 1 dy) 4

Let ν(p) = ν (p)e + ν 1 (p)e 1. Then the orthogonal projection of ν given by ν (p) = ν 1 (p)e + ν (p)e 1 and ν Γ(T( )). Letting dµ be the volume form on, dµ(ν ) = 1. Note that { ν 1 (p)e, ν (p)e 1 } form a basis for T p ( ) and it follows that dx = ν 1 dµ and dy = ν dµ. It then follows that, [ ] [ ] D(G p f) dl = ( θ 1 ν 1 + θ ν ) dµ e + (θ ν 1 + θ 1 ν )dµ e 1 = ( θ 1 ν 1 + θ ν )e + (θ ν 1 + θ 1 ν )e 1 dµ = ( (G p ) f 1 ν 1 (G p ) 1 f ν 1 + (G p ) f ν (G p ) 1 f 1 ν )e and + ((G p ) f ν 1 (G p ) 1 f 1 ν 1 + (G p ) f 1 ν + (G p ) 1 f ν )e 1 dµ G p νf = [((G p ) ν (G p ) 1 ν 1 )f ((G p ) 1 ν + (G p ) ν 1 )f 1 ]e + [((G p ) 1 ν + (G p ) ν 1 )f + ((G p ) ν (G p ) 1 ν 1 )f 1 ]e 1. The conclusion then follows by comparing these two calculations and since we can interchange between the surface measure and the surface (volume) form. Combining these results, we can prove the following version of Cauchy s integral theorem. Theorem 2.4. Let, R 2 be open sets such that and is bounded with smooth boundary. Suppose also that is equipped with unit outer normal ν and surface measure µ. Let f : R (1) be left monogenic. Then, f(p) = G p νf dµ for all p. Proof. Since and are open and means that there exists an ε > such that + ε. Let 1 = + 1 2 ε and 2 = + ε. Then, there exists a function f S (R 2 ) (1) such that f = f on 1 and f = on R 2 \ 2. Next, we observe that on 1, f is left monogenic on 1 \ {p} we apply Corollary 2.2 combined with the fact that f S (R 2 ) (1) and f = f on 1, D(G p f) dl = D(G p f) dl = (DG p ) f = f dδ p = f(p) = f(p) Then by application of Proposition 2.3, f(p) = D(G p f) dl = which concludes the proof of the theorem. G p νf dµ We remark that the key elements of the proof were Proposition 2.1 and 2.3. Indeed, if we were able to generalise these two key results, the proof of the main theorem would be a proof for case (1) of Theorem 1.2 without alteration. 5

In the proof of Proposition 2.3, we took the smooth unit outer normal ν and then rotated it counter-clockwise by π/2 to obtain the perpendicular ν Γ(T( )). In this case, we only have a single direction (up to orientation) to choose from. To extract out such a ν in a general setting, we would need to proceed by projecting ν to the directions at each p. Such would be a basis for T p ( ), and we would need to combine these in a way to obtain dµ(ν ) = 1. Also, notice that the Clifford multiplication captures the rotation of ν to obtain the corresponding tangential vector. Thus, we expect this should give us insight into how to construct the correct differential form in the general case. We have yet to show the relationship between monogeniety and holomorphy, and also between the classical Cauchy Integral Theorem in Complex variables. Our discussion from this point onwards will be specific to R (1). Proposition 2.5. f is a left monogenic function if and only if f is right monogenic. Proof. A simple calculation shows that Df = ( x f y f 1 )e + ( x f 1 + y f )e 1 = fd. This justifies us calling a function monogenic rather than left/right monogenic. In light of this observation, the integral in Theorem 2.4 also holds for the right monogenic case. This is actually due to the fact that multiplication in R (1) is commutative. Had we used this fact, the proof of Proposition 2.3 would have been simplified greatly. However, this would have been at the cost of losing scope of the general perspective. We have the following important observation which illustrates the fact that monogeniety is holomorphy in a different (but isomorphic) algebraic setting. Proposition 2.6. A function f is monogenic if and only if f and f 1 satisfy the Cauchy-Riemann equations. Proof. The proof of this is remarkably easy. Notice that = Df = ( x f y f 1 )e +( x f 1 + y f )e 1 if and only if = x f y f 1 and = x f 1 + y f which are exactly the Cauchy-Riemann equations. Now, let I : R (1) =R C denote the usual algebra isomorphism which is given by I(e ) = 1 and I(e 1 ) = ı. In light of this notation, a function f is monogenic if and only if IfI 1 is holomorphic. We require the following key change of coordinates formula for boundaries parametrised by smooth curves. Note that Γ(T( )) denotes sections over the bundle T( ). Proposition 2.7. Let, R 2 be open sets such that and suppose that is smooth and that γ : [, 1] is is a unit speed parametrisation of. As before, let ν be the unit outer normal and µ the surface measure on. Then, 1 θν dµ = e 1 (θ γ) γ dt for all θ C 1 (, R (1) ). Proof. First, we note that γ = γ e + γ 1 e 1 Γ(T( )) and γ = 1. The unit outer normal is then the rotation of γ by π/2 and so it follows that ν γ = γ e γ 1 e 1. With the observation that 6

ν γ = e 1 γ and since γ is parametrised via arc length, θν dµ = 1 which concludes the proof. (θ γ) (ν γ) dt = 1 1 (θ γ) ( e 1 γ) dt = e 1 (θ γ) γ dt The following theorem then illustrates that Theorem 2.4 is really the classical Cauchy Integral Theorem in another language. Proposition 2.8. Let, R 2 be open sets such that and suppose that is smooth and that γ : [, 1] is is a unit speed parametrisation of. As before, let ν be the unit outer normal and µ the surface measure on. Suppose f : R (1) is monogenic and let f : C C be given by f = IfI 1. Then, where γ = Iγ. ( f)(ξ) = I G I 1 (ξ)fν dµ = 1 f(ζ) 2πi γ ζ ξ dζ Proof. Fix ξ, let p = I 1 ξ and, let θ = G p f = G I 1 ξf. Also, by Theorem 1.1, G p (x) = 1 2π x p x p 2 IG I 1 ξi 1 (ζ) = 1 2π 1 ζ ξ and since multiplication in R (1) is commutative, we apply Proposition 2.7 to find: I G I 1 ξfν = ı = 1 ı = 1 ı 1 1 1 = 1 2πı I(G I 1f γ) I γ dt (IG I 1I 1 )(Iγ)(IfI 1 )(Iγ) dt (IG I 1I 1 )( γ) f( γ) dt f γ ζ ξ dζ. References [FB82] F. Sommen F. Brackx, R. Delanghe. Clifford Analysis. Pitman, 1982. [Jef4] Brian Jeffries. Spectral Properties of Noncommuting Operators. Springer-Verlag, 24. [Mit94] Marius Mitrea. Clifford Wavelets, Singular Integrals, and Hardy Spaces. Springer-Verlag, 1994. [MP87] Alan McIntosh and Alan Pryde. A functional calculus for several commuting operators. Indiana Univ. Math. J., 36 no. 2:421 439, 1987. 7