1: \documentclass[12pt]{article}
2: \usepackage{amssymb,amsmath}
3: \begin{document}
4: \newtheorem{theorem}{Theorem}
5: \newtheorem{lemma}{Lemma}
6: \newtheorem{proposition}{Proposition}
7: \newtheorem{definition}{Definition}
8: \begin{center}{\Large Introductory Clifford Analysis}
9: \end{center}
10: \begin{center}{by John Ryan}\end{center}
11: \begin{center}{Department of Mathematics, University of Arkansas}
12: \end{center}
13: \begin{center}{Fayetteville, AR 72701, U. S. A.}\end{center}
14: \begin{center}{\Large Introduction}\end{center}
15:
16: \ We want in this chapter to regard Clifford algebras as natural generalizations of the complex number system. First let us note that if $z$ is a complex number then
17: $\overline{z}z=\|z\|^{2}$. Now for a quaternion $q$ we also have
18: $\overline{q}q=\|q\|^{2}$. In this way quaternions may be regarded
19: as a generalization of the complex number system. It seems natural
20: to ask if one can extend basic results of one complex variable
21: analysis on holomorphic function theory to four dimensions using
22: quaternions. The answer is yes. This was developed by the Swiss
23: mathematician Rudolph Fueter in the 1930's and 1940's and also my Moisil and Teodorecu \cite{mt}. See for
24: instance \cite{f}. An excellent review of this work is given in the
25: survey article "Quaternionic analysis" by A. Sudbery, see \cite{su}. There is
26: also earlier work of Dixon \cite{di}.
27: However in previous lectures we have seen that for a vector $x\in
28: R^{n}$ when we consider $R^{n}$ as embedded in the Clifford
29: algebra $Cl_{n}$ then $x^{2}=-\|x\|^{2}$. So it seems reasonable
30: to ask if all that is known in the quaternionic setting extends to
31: the Clifford algebra setting. Again the answer is yes. The earlier aspects
32: of this study was developed by amongst others Richard
33: Delanghe \cite{d}, Viorel Iftimie \cite{i} and David Hestenes \cite{h}. The subject that
34: has grown from these works is now called Clifford analysis.
35:
36: \ In more recent times Clifford analysis has found a wealth of unexpected
37: applications in a number of branches of mathematical analysis particularly
38: classical harmonic analysis, see for instance the work of Alan McIntosh
39: and his collaborators, for instance \cite{lms,lmq}, Marius Mitrea \cite{m1,m2}
40: and papers in \cite{r4}. Links to representation theory and several complex variables may be
41: found in \cite{gm,r1,r2,r3} and elsewhere.
42:
43: \ The purpose of this paper is to present a review of many of the basic
44: aspects of Clifford analysis.
45:
46: \ Alternative accounts of much of this work together with other related
47: results can be found in \cite{bds,dss,gs,gm,ks,o,r4,r5}.
48:
49:
50: \begin{center}{\Large Foundations of Clifford Analysis}\end{center}
51:
52: \ We start by replacing the vector $x=x_{1}e_{1}+\ldots+x_{n}e_{n}$ by the
53: differential operator $D=\Sigma_{j=1}^{n}e_{j}\frac{\partial}{\partial x_{j}}$.
54: One basic but interesting property of $D$ is that $D^{2}=-\triangle_{n}$,
55: the Laplacian $\Sigma_{j=1}^{n}\frac{\partial^{2}}{\partial x_{j}^{2}}$ in
56: ${\bf{R}}^{n}$. The differential operator $D$ will be called a Dirac operator.
57: This is because the classical Dirac operator constructed over four dimensional
58: Minkowski space squares to give the wave operator.
59:
60: \begin{definition}
61: Suppose that $U$ is a domain in ${\bf{R}}^{n}$ and $f$ and $g$ are $C^{1}$
62: functions defined on $U$ and taking values in $Cl_{n}$. Then $f$ is called a
63: left monogenic function if $Df=0$ on $U$ while $g$ is called a right
64: monogenic function on $U$ if $gD=0$ where $gD=\Sigma_{j=1}^{n}
65: \frac{\partial g}{\partial x_{j}}e_{j}$.
66: \end{definition}
67:
68: \ Alternatively left monogenic functions are also called left regular
69: functions and perhaps most appropriately left Clifford holomorphic
70: functions. The term Clifford holomorphic functions or Clifford analytic functions appears to be due to Stephen Semmes, see \cite{se} and elsewhere. We shall most often use the term Clifford holomorphic functions here.
71:
72: \ Examples of such functions include the gradients of real valued harmonic
73: functions on $U$. So if $h$ is harmonic on $U$ and it is also real valued
74: then $Dh$ is a vector valued left monogenic function. It is also a right
75: monogenic function. Such a function is more commonly referred to as a
76: conjugate harmonic function or a harmonic $1$-form. See for instance
77: \cite{sw}. An example of such a function is $G(x)=\frac{x}{\|x\|^{n}}$.
78:
79: \ It should be noted that if $f$ and $g$ are left monogenic functions then
80: due to the lack of commutativity of the Clifford algebra, it is not in
81: general true that their product $f(x)g(x)$ is left monogenic.
82:
83: \ To introduce other possible examples of left monogenic functions suppose
84: that $\mu$ is a $Cl_{n}$ valued measure with compact support $[\mu]$ in
85: ${\bf{R}}^{n}$. Then the convolution $\int_{[\mu]}G(x-y)d\mu(y)$ defines a
86: left monogenic function on the maximal domain lying in
87: ${\bf{R}}^{n}\backslash[\mu]$. The previously defined integral is the
88: Cauchy transform of the measure $[\mu]$.
89:
90: \ Another way to construct examples of left monogenic functions was introduced
91: by Littlewood and Gay in \cite{lg} for the case $n=3$ and independently
92: re-introduced for all $n$ by Sommen \cite{s2}. Suppose $U'$ is a domain in
93: ${\bf{R}}^{n-1}$, the span of $e_{2},\ldots,e_{n}$. Suppose also that
94: $f'(x')$ is a $Cl_{n}$ valued function such that at each point $x'\in U'$
95: there is a multiple series expansion in $x_{2},\ldots,x_{n}$ that converges
96: uniformly on some neighbourhood of $x'$ in $U'$ to $f'$. Such a function is
97: called a real analytic function. The series
98: \[\Sigma_{k=0}^{\infty}\frac{1}{k!}x_{1}^{k}(-e_{1}D'f'(x')=\exp(-x_{1}e_{1}D')
99: f'(x')\]
100: where $D'=\Sigma_{j=2}^{n}e_{j}\frac{\partial}{\partial x_{j}}$, defines a
101: left monogenic function $f$ in some neighborhood $U(f')$ in ${\bf{R}}^{n}$ of
102: $U'$. The left monogenic function $f$ is the Cauchy-Kowalewska extension of
103: $f'$.
104:
105: \ It should be noted that if $f$ is a left monogenic function then
106: $\overline{f}$ and $\tilde{f}$ are both right monogenic functions.
107:
108: \ We now turn to analogues of Cauchy's Theorem and Cauchy's integral formula.
109: \begin{theorem}{\bf{(The Clifford-Cauchy Theorem):}} Suppose that $f$ is
110: a
111: left Clifford holomorphic function on $U$ and $g$ is a right Clifford holomorphic function on $U$.
112: Suppose also that $V$ is a bounded subdomain of $U$ with piecewise
113: differentiable boundary $S$ lying in $U$. Then
114: \begin{equation}
115: \int_{S}g(x)n(x)f(x)d\sigma(x)=0
116: \end{equation}
117: where $n(x)$ is the outward pointing normal vector to $S$ at $x$ and $\sigma$
118: is the Lebesgue measure on $S$.
119: \end{theorem}
120:
121: \ The proof follows directly from Stokes' Theorem. One important point to
122: keep
123: in mind though is that as $Cl_{n}$ is not a commutative algebra then it is
124: important to place the vector $n(x)$ between $f$ and $g$. One then has that
125: \[\int_{S}g(x)n(x)f(x)d\sigma(x)=\int_{V}((g(x)D)f(x)+g(x)(Df(x)))dx^{n}=0.\]
126:
127: \ Suppose that $g$ is the gradient of a real valued harmonic function and
128: $f=1$. Then the real part of Equation 1 gives the following well known integral
129: formula.
130: \[\int<grad g(x),n(x)>d\sigma(x)=0.\]
131:
132: \ We now turn to the analogue of a Cauchy integral formula.
133: \begin{theorem}{\bf{(Clifford-Cauchy Integral Formula):}}
134: Suppose that $U$, $V$, $S$, $f$ and $g$ are all as in Theorem 1 and that
135: $y\in V$. Then
136: \[f(y)=\frac{1}{\omega_{n}}\int_{S}G(x-y)n(x)f(x)d\sigma(x)\]
137: and
138: \[g(y)=\frac{1}{\omega_{n}}\int_{S}g(x)n(x)G(x-y)d\sigma(x)\]
139: where $\omega_{n}$ is the surface area of the unit sphere in ${\bf{R}}^{n}$.
140: \end{theorem}
141: {\bf{Proof:}} The proof follows very similar lines to the argument in one
142: variable complex analysis. We shall establish the formula for $f(y)$ the proof
143: being similar for $g(y)$. First let us take a sphere $S^{n-1}(y,r)$ centered
144: at $y$ and of radius $r$. The radius $r$ is chosen sufficently small so that
145: the closed disc with boundary $S^{n-1}(y,r)$ lies in $V$. Then by the
146: Clifford-Cauchy theorem
147: \[\int_{S}G(x-y)n(x)f(x)d\sigma(x)=
148: \int_{S^{n-1}(y,r)}G(x-y)n(x)f(x)d\sigma(x).\]
149: However on $S^{n-1}(y,r)$ the vector $n(x)=\frac{y-x}{\|x-y\|}$. So
150: $G(x-y)n(x)=\frac{1}{r^{n-1}}$. So
151: \[\int_{S^{n-1}(y,r)}G(x-y)n(x)f(x)d\sigma(x)=\int_{S^{n-1}(y,r)}
152: \frac{1}{r^{n-1}}(f(x)-f(y))d\sigma(x)\]
153: \[+\int_{S^{n-1}(y,r)}\frac{f(y)}{r^{n-1}}
154: d\sigma(x).\]
155: The right side of this previous expression reduces to
156: \[\int_{S^{n-1}(y,r)}\frac{(f(x)-f(y))}{r^{n-1}}d\sigma(x)+f(y)\int_{S^{n-1}}
157: d\sigma(x).\]
158: Now $\int_{S^{n-1}}d\sigma(x)=\omega_{n}$ and by continuity
159: $\lim_{r\rightarrow 0}\int_{S^{n-1}(y,r)}
160: \frac{(f(x)-f(y)}{r^{n-1}}d\sigma(x)=0$. The result follows. $\Box$
161:
162: \ One important feature is to note that Kelvin inversion, ie $x^{-1}=\frac{-x}{\|x\|^{2}}$ whenever $x$ is non-zero, plays a
163: fundamental
164: role in this proof. Moreover the proof is almost exactly the same as the
165: proof of Cauchy's Integral Formula for piecewise $C^{1}$ curves in one
166: variable complex analysis.
167:
168: \ Having obtained a Cauchy Integral Formula in ${\bf{R}}^{n}$ a number of
169: basic results that one might see in a first course in one variable complex
170: analysis carry over more or less automatically to the context described here.
171: This includes a Liouville Theorem and Weierstrass Convergence Theorem. We
172: leave it as an exercise to the interested reader to set up and establish the
173: Clifford analysis analogues of these results. Their statements and proofs can
174: be found in \cite{bds}.
175:
176: \ Theorems 1 and 2 show us that the individual components of the equations
177: $Df=0$ and $gD=0$ comprise generalized Cauchy-Riemann equations. In the
178: particular case where $f$ is just vector valued so
179: $f=\Sigma_{j=1}^{n}f_{j}e_{j}$ then the generalized Cauchy-Riemann equations
180: become
181: $\frac{\partial f_{j}}{\partial x_{i}}=\frac{\partial f_{i}}{\partial x_{j}}$
182: whenever $i\neq j$ and $\Sigma_{j=1}^{n}\frac{\partial f_{j}}{\partial x_{j}}
183: =0$. This system of equations is often referred to as the Riesz system.
184:
185: \ Having obtained an analogue of Cauchy's integral formula in euclidean
186: space we shall now exploit this result to show how many
187: consequences of the classical Cauchy integral carry over to the context
188: described here. We begin with the Mean Value Theorem.
189:
190: \begin{theorem}
191: Suppose that $D(y, R)$ is a closed disc centered at $y$, of radius $R$ and
192: lying in $U$. Then for each left Clifford holomorphic function $f$ on $U$
193: \[f(y)=\frac{1}{R\omega_{n}}\int_{D(y, R)}\frac{f(x)}{\|x-y\|^{n-1}}dx^{n}.\]
194: \end{theorem}
195: {\bf{Proof:}}
196: We have already seen that for each $r\in(0, R)$
197: \[f(y)=\frac{1}{\omega_{n}}\int_{S^{n-1}(y, r)}\frac{f(x)}{\|x-y\|^{n-1}}
198: d\sigma(x),\]
199: where $S^{n-1}(y, r)$ is the $(n-1)$-dimensional sphere centered at $y$ and
200: of radius $r$. We obtain the result by integrating both sides of this
201: expression with respect to the variable $r$ and dividing throughout by $R$.
202: $\Box$
203:
204: \ Let us now turn to explore the real analyticity properties of Clifford holomorphic
205: functions. First it may be noted that when $n$ is even
206: $G(x-y)=(-1)^{\frac{n-2}{2}}(x-y)^{-n+1}$. Also $(x-y)^{-1}=
207: x^{-1}(1-yx^{-1})^{-1}=(1-x^{-1}y)^{-1}x^{-1}$, and $\|x^{-1}y\|=\|yx^{-1}\|
208: =\frac{\|y\|}{\|x\|}$. So for $\|y\|<\|x\|$
209: \[(x-y)^{-1}=x^{-1}(1+yx^{-1}+\ldots+yx^{-1}\ldots yx^{-1}+\ldots)\]
210: \[=(1+x^{-1}y+\ldots+x^{-1}y\ldots x^{-1}y+\ldots)x^{-1}.\]
211: Hence these two sequences converge uniformly to $(x-y)^{-1}$ provided
212: $\|y\|\leq r<\|x\|$ and they converge pointwise to $(x-y)^{-1}$ provided
213: $\|y\|<\|x\|$. One can now take $(-1)^{\frac{n-2}{2}}$ times the $(n-1)$-fold
214: product of the series expansions of $(x-y)^{-1}$ with itself to obtain a
215: series expansion for $G(x-y)$. In this process of multiplying series
216: together in order to maintain the same radius of convergence one needs to
217: group together all linear combinations of monomials in $y_{1},\ldots, y_{n}$
218: that are of the same order. Thus we have deduced that when $n$ is even the
219: multiple Taylor series expansion
220: \[\Sigma_{j=0}^{\infty}(\Sigma_{\stackrel{j_{1},\ldots,j_{n}}{j_{1}+
221: \ldots+j_{n}=j}}
222: \frac{y_{1}^{j_{1}}\ldots y_{n}^{j_{n}}}{j_{1}!\ldots j_{n}!}
223: \frac{\partial^{j}G(x)}{\partial x_{1}^{j_{1}}\ldots \partial
224: x_{n}^{j_{n}}})\]
225: converges uniformly to $G(x-y)$ provided $\|y\|<r<\|x\|$ and converges
226: pointwise to $G(x-y)$ provided $\|y\|<\|x\|$.
227:
228: \ A similar argument may be developed when $n$ is odd.
229:
230: \ Returning to Cauchy's integral formula let us suppose that $f$ is a left
231: Clifford holomorphic function defined in a neighbourhood of the closure of some ball
232: $B(0, R)$. Then
233: \[f(y)=\frac{1}{\omega_{n}}\int_{\partial B(0,R)}G(x-y)n(x)f(x)d\sigma(x)=\]
234: \[\frac{1}{\omega_{n}}\int_{\partial B(0,R)}\Sigma_{j=0}^{\infty}
235: (\Sigma_{\stackrel{j_{1}\ldots j_{n}}{j_{1}+\ldots+j_{n}=j}}
236: \frac{y_{1}^{j_{1}}\ldots y_{n}^{j_{n}}}{j_{1}!\ldots j_{n}!}
237: \frac{\partial^{j}G(x)}{\partial x_{1}^{j_{1}}\ldots\partial x_{n}^{j_{n}}})
238: n(x)f(x)d\sigma(x)\]
239: provided $\|y\|<\|x\|$. As this series converges uniformly on each ball
240: $B(0, r)$ for each $r<R$ then this last integral can be re-written as
241: \[\frac{1}{\omega_{n}}\Sigma_{j=0}^{\infty}\int_{\partial B(0, R)}
242: (\Sigma_{\stackrel{j_{1}\ldots j_{n}}{j_{1}+\ldots+j_{n}=j}}
243: \frac{y_{1}^{j_{1}}\ldots y_{n}^{j_{n}}}{j_{1}!\ldots j_{n}!}
244: \frac{\partial^{j}G(x)}{\partial x_{1}^{j_{1}}\ldots \partial x_{n}^{j_{n}}}
245: n(x)f(x))d\sigma(x).\]
246: As the summation within the parentheses is a finite summation this last
247: expression easily reduces to
248: \[\frac{1}{\omega_{n}}\Sigma_{j=0}^{\infty}
249: (\Sigma_{\stackrel{j_{1}\ldots j_{n}}{j_{1}+\ldots+j_{n}=j}}
250: \frac{y_{1}^{j_{1}}\ldots y_{n}^{j_{n}}}{j_{1}!\ldots j_{n}!}
251: \int_{\partial B(0, R)}\frac{\partial^{j}G(x)}{\partial x_{1}^{j_{1}}\ldots
252: \partial x_{n}^{j_{n}}}n(x)f(x))d\sigma(x).\]
253: On placing
254: \[\frac{1}{\omega_{n}}\int_{\partial B(0, R)}\frac{\partial^{j}G(x)}
255: {\partial x_{1}^{j_{1}}\ldots\partial x_{n}^{j_{n}}}n(x)f(x)d\sigma(x)
256: =a_{j_{1}\ldots j_{n}}\]
257: it may be seen that on $B(0, R)$ the series
258: \[\Sigma_{j=0}^{\infty}
259: (\Sigma_{\stackrel{j_{1}\ldots j_{n}}{j_{1}+\ldots+j_{n}=j}}
260: \frac{x_{1}^{j_{1}}\ldots x_{n}^{j_{n}}}{j_{1}!\ldots j_{n}!}
261: a_{j_{1}\ldots j_{n}})\]
262: converges pointwise to $f(y)$. Convergence is uniform on each ball
263: $B(0, r)$ provided $r<R$.
264:
265: \ Similarly if $g$ is a right Clifford holomorphic function defined in a neighbourhood
266: of the closure of $B(0, R)$ then the series
267: \[\Sigma_{j=0}^{\infty}
268: (\Sigma_{\stackrel{j_{1}\ldots j_{n}}{j_{1}+\ldots+j_{n}=j}}
269: b_{j_{1}\ldots j_{n}}\frac{y_{1}^{j_{1}}\ldots y_{n}^{j_{n}}}
270: {j_{1}!\ldots j_{n}!})\]
271: converges pointwise on $B(0, R)$ to $g(y)$ and converges uniformly on
272: $B(0, r)$ for $r<R$, where
273: \[b_{j_{1}\ldots j_{n}}=\frac{1}{\omega_{n}}\int_{\partial B(0, R)}
274: g(x)n(x)\frac{\partial^{j}G(x)}
275: {\partial x_{1}^{j_{1}}\ldots \partial x_{n}^{j_{n}}}d\sigma(x).\]
276:
277: \ By translating the ball $B(0, R)$ to the ball $B(w, R)$ where
278: $w=w_{1}e_{1}+\ldots+w_{n}e_{n}$ one may readily observe that for any left
279: Clifford holomorphic function $f$ defined in a neighbourhood of the closure of $B(w, R)$
280: the series
281: \[\Sigma_{j=0}^{\infty}
282: (\Sigma_{\stackrel{j_{1}\ldots j_{n}}{j_{1}+\ldots j_{n}=j}}
283: \frac{(y_{1}-w_{1})^{j_{1}}\ldots (y_{n}-w_{n})^{j_{n}}}
284: {j_{1}!\ldots j_{n}!}a'_{j_{1}\ldots j_{n}})\]
285: converges pointwise on $B(w, R)$ to $f(y)$, where
286: \[a'_{j_{1}\ldots j_{n}}=\frac{1}{\omega_{n}}\int_{\partial B(w, R)}
287: \frac{\partial^{j}G(x-w)}{\partial x_{1}^{j_{1}}\ldots x_{n}^{j_{n}}}
288: n(x)f(x)d\sigma(x).\]
289: Again the series converges uniformly on $B(w, r)$ for each $r<R$.
290: A similar series may be readily obtained for any right Clifford holomorphic function
291: defined in a neighbourhood of the closure of $B(w, R)$.
292:
293: \ The types of power series that we have developed for left Clifford holomorphic
294: functions are not entirely satisfactory. In particular, unlike their complex
295: analogues the homogeneous polynomials
296: \[\Sigma_{\stackrel{j_{1}\ldots j_{n}}{j_{1}+\ldots+j_{n}=j}}
297: \frac{x_{1}^{j_{1}}\ldots x_{n}^{j_{n}}}{j_{1}!\ldots j_{n}!}
298: a_{j_{1}\ldots j_{n}}\]
299: are not expressed as a linear combination of left Clifford holomorphic polynomials. To
300: rectify this situation let us first take a closer look at the Taylor
301: expansion for the Cauchy kernel $G(x-y)$ where all the Taylor coefficients are
302: real. Let us first look at the first order terms in the Taylor expansion.
303: This is the expression
304: \[y_{1}\frac{\partial G(x)}{\partial x_{1}}+\ldots+y_{n}
305: \frac{\partial G(x)}{\partial x_{n}}.\]
306: As $G$ is a Clifford holomorphic function then $\frac{\partial G(x)}{\partial x_{1}}=
307: -\Sigma_{j=2}^{n}e_{1}^{-1}e_{j}\frac{\partial G(x)}{\partial x_{j}}$.
308: Therefore the first order terms of the Taylor expansion for $G(x-y)$ can be
309: re-expressed as
310: \[\Sigma_{j=2}^{n}(y_{j}-e_{1}^{-1}e_{j}y_{1})\frac{\partial G(x)}
311: {\partial x_{j}}.\]
312: Moreover, for $2\leq j\leq n$ the first order polynomial
313: $y_{j}-e_{1}^{-1}e_{j}y_{1}$ is a left Clifford holomorphic polynomial.
314: Let us now go to second order terms. Again we will replace the operator
315: $\frac{\partial}{\partial x_{1}}$ by the operator $-\Sigma_{j=2}^{n}
316: e_{1}^{-1}e_{j}\frac{\partial}{\partial x_{j}}$ whenever it arises. Let us
317: consider the term $\frac{\partial^{2}G(x)}{\partial x_{i}\partial x_{j}}$
318: where $i\neq j\neq 1$. We end up with the polynomial
319: $y_{i}y_{j}-y_{i}y_{1}e_{1}^{-1}e_{j}-y_{j}y_{1}e_{1}^{-1}e_{i}
320: =\frac{1}{2}((y_{i}-y_{1}e_{1}^{-1}e_{i})(y_{j}-y_{1}e_{1}^{-1}e_{j})
321: +(y_{j}-y_{1}e_{1}^{-1}e_{j})(y_{i}-y_{1}e_{1}^{-1}e_{i})$.
322: Similarly the polynomial attached to the term $\frac{\partial^{2}G(x)}
323: {\partial x_{i}^{2}}$ is $(y_{i}-y_{1}e_{1}^{-1}e_{i})^{2}$. Using the
324: Clifford algebra anti-commutation relationship $e_{i}e_{j}+e_{j}e_{i}=-2
325: \delta_{ij}$ and on replacing the differential operator $\frac{\partial}
326: {\partial x_{1}}$ by the operator $-\Sigma_{j=2}^{n}e_{1}^{-1}e_{j}
327: \frac{\partial}{\partial x_{j}}$ it may be determined that the power
328: series we previously obtained for $G(x-y)$ can be replaced by the
329: series $\Sigma_{j=0}^{\infty}(\Sigma_{\stackrel{j_{2}\ldots j_{n}}
330: {j_{2}+\ldots+j_{n}=j}}P_{j_{2}\ldots j_{n}}(y)\frac{\partial^{j}G(x)}
331: {\partial x_{2}^{j_{2}}\ldots \partial x_{n}^{j_{n}}})$, where $\|y\|<
332: \|x\|$ and
333: \[P_{j_{2}\ldots j_{n}}(y)=\frac{1}{j!}\Sigma(y_{\sigma(1)}
334: -y_{1}e_{1}^{-1}e_{\sigma(1)})\ldots
335: (y_{\sigma(j)}-y_{1}e_{1}^{-1}e_{\sigma(j)}).\]
336: Here $\sigma(i)\in\{2,\ldots, n\}$ and the previous summation is taken over
337: all permutations of the
338: monomials $(y_{\sigma(i)}-y_{1}e_{1}^{-1}e_{\sigma(i)})$ without repetition.
339: The quaternionic monogenic analogues for these polynomials were introduced
340: by Fueter \cite{f} while the Clifford analogues, $P_{j_{2}\ldots j_{n}}$,
341: described here were introduced by Delanghe in \cite{d}. It should be noted that
342: each polynomial $P_{j_{2}\ldots j_{n}}(y)$ takes its values in the space
343: spanned by $1$, $e_{1}e_{2},\dots, e_{1}e_{n}$. Also each such polynomial is
344: homogeneous of degree $j$. Similar arguments to those
345: just outlined give that $G(x-y)=\Sigma_{j=0}^{\infty}
346: (\Sigma_{\stackrel{j_{2}\ldots j_{n}}{j_{2}+\ldots+j_{n}=j}}
347: \frac{\partial^{j}G(x)}{\partial x_{2}^{j_{2}}\ldots x_{n}^{j_{n}}}
348: \widetilde{P_{j_{2}\ldots j_{n}}}(y)$ provided $\|y\|<\|x\|$.
349: \begin{proposition}
350: Each of the polynomials $P_{j_{2}\ldots j_{n}}(y)$ is a left Clifford holomorphic
351: polynomial.
352: \end{proposition}
353: {\bf{Proof:}} As $DP_{j_{2}\ldots j_{n}}(y)=e_{1}(\frac{\partial}
354: {\partial y_{1}}+e_{1}^{-1}\Sigma_{j=2}^{n}e_{j}\frac{\partial}
355: {\partial y_{j}}P_{j_{2}\ldots j_{n}}(y))$ then we shall consider the
356: expression $(\frac{\partial}{\partial y_{1}}+\Sigma_{j=2}^{n}e_{1}^{-1}e_{j}
357: \frac{\partial}{\partial y_{j}})P_{j_{2}\ldots j_{n}}(y)$. This term is equal
358: to
359: \[(\frac{\partial}{\partial y_{j}}+\Sigma_{j=2}^{n}e_{1}^{-1}e_{j}
360: \frac{\partial}{\partial y_{j}})
361: \Sigma(y_{\sigma(1)}-e_{1}^{-1}e_{\sigma(1)}y_{1})\ldots\]
362: \[\ldots(y_{\sigma(i-1)}-e_{1}^{-1}e_{\sigma(i-1)}y_{1})
363: (y_{\sigma(i)}-e_{1}^{-1}
364: e_{\sigma(i)}y_{1})(y_{\sigma(i+1)}-e_{1}^{-1}e_{\sigma(i+1)}y_{1})\ldots\]
365: \[\ldots(y_{\sigma(j)}-e_{1}^{-1}e_{\sigma(j)}y_{1}).\]
366: This is equal to
367: \[\Sigma(y_{\sigma(1)}-e_{1}^{-1}e_{\sigma(1)}y_{1})\ldots
368: (y_{\sigma(i-1)}-e_{1}^{-1}e_{\sigma(i-1)}y_{1})(-e_{1}^{-1}e_{\sigma(i)})
369: (y_{\sigma(i+1)}-e_{1}^{-1}e_{\sigma(i+1)}y_{1})\]
370: \[\ldots (y_{\sigma(j)}-e_{1}^{-1}e_{\sigma(j)}y_{1})+
371: \Sigma e_{1}^{-1}e_{\sigma(i)}
372: (y_{\sigma(1)}-e_{1}^{-1}e_{\sigma(1)}y_{1})\ldots
373: (y_{\sigma(i-1)}-e_{1}^{-1}e_{\sigma(i-1)}y_{1})\]
374: \[(y_{\sigma(i+1)}-e_{1}^{-1}e_{\sigma(i+1)}y_{1})\ldots
375: (y_{\sigma(j)}-e_{1}^{-1}e_{\sigma(j)}y_{1}).\]
376:
377: \ If we multiply the previous term by $y_{1}$ and add to it the following
378: term, which is equal to zero,
379: \[\Sigma(y_{\sigma(1)}-e_{1}^{-1}e_{\sigma(1)}y_{1})\ldots
380: (y_{\sigma(i-1)}-e_{1}^{-1}e_{\sigma(i-1)}y_{1})
381: (y_{\sigma(i)}-y_{\sigma(i)})\]
382: \[(y_{\sigma(i+1)}-e_{1}^{-1}e_{\sigma(i+1)}y_{1})\ldots
383: (y_{\sigma(j)}-e_{1}^{-1}e_{\sigma(i)}y_{1})\]
384: we get, after regrouping terms,
385: \[\Sigma(y_{\sigma(1)}-e_{1}^{-1}e_{\sigma(1)}y_{1})\ldots
386: (y_{\sigma(i-1)}-e_{1}^{-1}e_{\sigma(i-1)}y_{1})
387: (y_{\sigma(i)}-e_{1}^{-1}e_{\sigma(i)}y_{1})\]
388: \[(y_{\sigma(i+1)}-e_{1}^{-1}e_{\sigma(i+1)}y_{1})\ldots
389: (y_{\sigma(j)}-e_{1}^{-1}e_{\sigma(j)}y_{1})\]
390: \[-\Sigma(y_{\sigma(i)}-e_{1}^{-1}e_{\sigma(i)}y_{1})
391: (y_{\sigma(1)}-e_{1}^{-1}e_{\sigma(1)}y_{1})\ldots
392: (y_{\sigma(i-1)}-e_{1}^{-1}e_{\sigma(i-1)}y_{1})\]
393: \[(y_{\sigma(i+1)}-e_{1}^{-1}e_{\sigma(i+1)}y_{1})\ldots
394: (y_{\sigma(j)}-e_{1}^{-1}e_{\sigma(j)}y_{j}).\]
395: As summation is taken over all possible permutations without
396: repetition this last term vanishes. $\Box$
397:
398: \ Using Proposition 1 and the results we previously obtained on series
399: expansions we can obtain the following generalization of Taylor expansions
400: from one variable complex analysis.
401: \begin{theorem}{\bf{(Taylor Series)}}
402: Suppose that $f$ is a left Clifford holomorphic function defined in an open
403: neighbourhood of the closure of the ball $B(w, R)$. Then
404: \[f(y)=\Sigma_{j=0}^{\infty}(\Sigma_{\stackrel{j_{2}\ldots j_{n}}
405: {j_{2}+\ldots+j_{n}=j}}P_{j_{2}\ldots j_{n}}(y-w)a_{j_{2}\ldots j_{n}}),\]
406: where $a_{j_{2}\ldots j_{n}}=\frac{1}{\omega_{n}}\int_{\partial B(w, R)}
407: \frac{\partial^{j}G(x-w)}{\partial x_{2}^{j_{2}}\ldots\partial x_{n}^{j_{n}}}
408: n(x)f(x)d\sigma(x)$ and $\|y-w\|<R$. Convergence is uniform
409: provided $\|x-w\|<r<R$.
410: \end{theorem}
411:
412: \ A simple application of Cauchy's theorem now tells us that the Taylor
413: series that we obtained for $f$ in the previous theorem remains
414: valid on the largest open ball on which $f$ is defined and the largest open
415: ball on which $g$ is defined. Also the previous identities immediately yield
416: the mutual linear independence of the collection of the left Clifford holomorphic
417: polynomials $\{P_{j_{2}\ldots j_{n}}:j_{2}+\ldots j_{n}=j$ and $0\leq j
418: <\infty\}$.
419:
420: \begin{center}{\Large Other Types of Clifford Holomorphic Functions}\end{center}
421:
422: \ Unlike the the classical Cauchy-Riemann operator $\frac{\partial}
423: {\partial\overline{z}}=\frac{\partial}{\partial x}+i\frac{\partial}
424: {\partial y}$ the generalized Cauchy-Riemann operator $D$ that we have
425: introduced here does not have an identity component. Instead we could have
426: considered the differential operator
427: $D'=\frac{\partial}{\partial x_{0}}+\Sigma_{j+1}^{n-1}e_{j}\frac{\partial}
428: {\partial x_{j}}$. Also for $U'$ a domain in $R\oplus R^{n-1}$, the span of
429: $1, e_{1},\ldots, e_{n-1}$, one can consider $Cl_{n-1}$ valued differentiable
430: functions $f'$ and $g'$ defined on $U'$ such that $D'f'=0$ and $g'D'=0$,
431: where $g'D'=\frac{\partial g'}{\partial x_{0}}+\Sigma_{j=1}^{n-1}
432: \frac{\partial g'}{\partial x_{j}}e_{j}$. Traditionally such functions are
433: also called left monogenic and right monogenic functions. To avoid confusion
434: we shall call such functions unital left monogenic and unital right
435: monogenic respectively. In the case where $n=2$ the operator $D'$ corresponds
436: to the usual Cauchy-Riemann operator and unital monogenic functions are the
437: usual holomorphic functions studied in one variable complex analysis.
438: The function $G'(\underline{x})=\frac{\overline{\underline{x}}}
439: {\|\underline{x}\|^{n}}=\underline{x}^{-1}\|\underline{x}\|^{-n+2}$ is an
440: example of a function which is both unital
441: left monogenic and unital right monogenic. It is a simple matter to observe
442: that $f'$ is unital left monogenic if and only if $\tilde{f}'$ is unital
443: right monogenic. However $\overline{f}'$ is not unital right monogenic
444: whenever $f'$ is unital left monogenic. Instead $\overline{f}'$ satisfies
445: the equation $\overline{f}'D'=0$.
446:
447: \ The function theory for unital left monogenic functions is much the same
448: as for left monogenic functions. For instance if $f'$ is unital left
449: monogenic on $U'$ and $g'$ is unital right monogenic on the same domain and
450: $S'$ is a piecewise smooth, compact surface lying in $U'$ and bounding a
451: subdomain $V'$ then $\int_{S'}g'(\underline{x})n(\underline{x})
452: f'(\underline{x})d\sigma(\underline{x})=0$ where $n(\underline{x})$ is the
453: outward pointing normal vector to $S'$ at $\underline{x}$. Also for each
454: $\underline{y}\in V'$ there is the following version of Cauchy's integral
455: formula
456: \[f'(\underline{y})=\frac{1}{\omega_{n}}\int_{S'}G'(\underline{x}-
457: \underline{y})n(\underline{x})f'(\underline{x})d\sigma(\underline{x}).\]
458:
459: \ To get from the operator $D$ to the operator $D'$ one first rewrites $D$
460: as $e_{n}(\frac{\partial}{\partial x_{n}}+\Sigma_{j=1}^{n-1}e_{n}^{-1}e_{j}
461: \frac{\partial}{\partial x_{j}})$. On multiplying on the left by $e_{n}$ and
462: changing the variable $x_{n}$ to $x_{0}$ we get the operator
463: $D"=\frac{\partial}{\partial x_{0}}+\Sigma_{j=1}^{n-1}e_{n}^{-1}e_{j}
464: \frac{\partial}{\partial x_{j}}$. This operator takes its values in the
465: even subalgebra $Cl_{n}^{+}$ of $Cl_{n}$. On applying the isomorphism
466: \[\theta: Cl_{n-1}\rightarrow Cl_{n}^{+}:\theta(e_{j_{1}}\ldots e_{j_{r}})=e_{n}^{-1}e_{j_{1}}\ldots e_{n}^{-1}e_{j_{r}}\]
467: it immediately follows that
468: $\theta(D')=D"$. So if $f'$ is unital left monogenic then $D"\theta(f)=0$.
469: If we change the variable $x_{0}$ of the function $\theta(f(\underline{x}))$
470: to $x_{n}$ we now get a left monogenic function, which we denote by
471: $\theta'(f)(x)$, where $x=x_{1}e_{1}+\ldots+x_{n}e_{n}\in U\subset R^{n}$
472: if and only if $\underline{x}=x_{n}+x_{1}e_{1}+\ldots+x_{n-1}e_{n-1}\in U'
473: \subset R\oplus R^{n-1}$.
474:
475: \ It should be noted that $D'\overline{D'}=\overline{D'}D'=\triangle_{n}$.
476:
477: \ When $n=3$ the algebra$Cl_{3}$ is split by the two projection
478: operators $E_{\pm}=\frac{1}{2}(1\pm e_{1}e_{2}e_{3})$ into a direct sum $E_{+}Cl_{3}\oplus E_{-}Cl_{3}$ and that
479: each of these subalgebras are isomorphic to the quaternion algebra ${\bf{H}}$.
480: In this setting the differential operator $E_{\pm}D'$ can best be written as
481: $\frac{\partial}{\partial t}+i\frac{\partial}{\partial x}+j\frac{\partial}
482: {\partial y}+k\frac{\partial}{\partial z}$ and the operator $E_{\pm}D$ can
483: best be written as $i\frac{\partial}{\partial x}+j\frac{\partial}
484: {\partial y}+k\frac{\partial}{\partial z}$. We shall denote the first of
485: these two operators by $D'_{{\bf{H}}}$ and the second by $D_{{\bf{H}}}$. The
486: operator $D'_{{\bf{H}}}$ is sometimes referred to as the
487: Cauchy-Riemann-Fueter operator. The function theory associated to the
488: differential operators $D_{{\bf{H}}}$ and $D'_{{\bf{H}}}$ is much the same as
489: that associated to the operators $D$ and $D'$. In fact historically the
490: starting point for Clifford analysis was to study the function theoretic
491: aspects of the operators $D'_{{\bf{H}}}$ and $D_{{\bf{H}}}$, see for instance
492: \cite{d,f} and the excellent review article of Sudbery \cite{su}.
493:
494: \ Over all these function theories have proved itself to be much the same as
495: that for the operators $D$ and $D'$. It is a simple enough matter to set up
496: analogues of Cauchy's theorem and Cauchy's integral formula for the
497: quaternionic valued differentiable functions that are either annihilated by
498: $D'_{{\bf{H}}}$ or $D_{{\bf{H}}}$, either acting on the left or on the right.
499: When dealing with the operator $D'_{{\bf{H}}}$ such functions are called
500: quaternionic monogenic. The quaternionic monogenic Cauchy kernel is the
501: function $q^{-1}\|q\|^{-2}$. Consequently for each quaternionic left
502: monogenic function $f(q)$ defined on a domain $U"\subset{\bf{H}}$ and each
503: $q_{0}$ lying in a bounded subdomain with piecewise $C^{1}$ boundary $S"$
504: \[f(q_{0})=\frac{1}{\omega_{3}}\int_{S"}(q-q_{0})^{-1}\|q-q_{0}\|^{-2}n(q)
505: f(q)d\sigma(q).\]
506: Similarly if $g$ is right quaternionic monogenic on $U"$ then
507: \begin{equation}
508: g(q_{0})=\frac{1}{\omega_{3}}\int_{S"}g(q)n(q)(q-q_{0})^{-1}\|q-q_{0}\|^{-2}
509: d\sigma(q).
510: \end{equation}
511:
512: \begin{center}{\Large The Equation $D^{k}f=0$}\end{center}
513:
514: \ It is reasonably well known that if $h$ is a real valued harmonic function
515: defined on a domain $U\subset R^{n}$ then for each $y\in U$ and each
516: compact, piecewise $C^{1}$ surface $S$ lying in $U$ such that $S$ bounds a
517: subdomain $V$ of $S$ and $y\in V$, then
518: \[h(y)=\frac{1}{\omega_{n}}\int_{S}(H(x-y)<n(x),grad h(x)>-<G(x-y),n(x)>h(x))
519: d\sigma(x),\]
520: where $H(x-y)=\frac{1}{(n-2)\|x-y\|^{n-2}}$. This formula is Green's formula
521: for a harmonic function, and it heavily relies on the standard inner product
522: on $R^{n}$. Introducing the Clifford algebra $Cl_{n}$ the right side of
523: Green's formula is the real part of
524: \[\frac{1}{\omega_{n}}\int_{S}(G(x-y)n(x)h(x)-H(x-y)n(x)Dh(x))d\sigma(x).\]
525:
526: \ Assuming that the function $h$ is $C^{2}$ then on applying Stokes' theorem
527: the previous integral becomes
528: \[\frac{1}{\omega_{n}}\int_{S^{n-1}(y, r(y))}
529: (G(x-y)n(x)h(x)-H(x-y)n(x)Dh(x))d\sigma(x),\]
530: where $S^{n-1}(y, r(y))$ is a sphere centered at $y$, of radius $r(y)$ and
531: lying in $V$. On letting the radius $r(y)$ tend to zero the first term of the
532: integral tends to $h(y)$ while the second term tends to zero. Consequently
533: the Clifford analysis version of Green's formula is
534: \[h(y)=\frac{1}{\omega_{n}}\int_{S}(G(x-y)n(x)h(x)-H(x-y)n(x)Dh(x))
535: d\sigma(x).\]
536:
537: \ This formula was obtained under the assumption that $h$ is real valued and
538: $C^{2}$. The fact that we have assumed $h$ to be real valued can easily be
539: observed to be irrelevant, and so we can assume that $h$ is $Cl_{n}$ valued.
540: From now on we shall assume that all harmonic functions take their values in
541: $Cl_{n}$. If $h$ is also a left monogenic function then the Clifford analysis
542: version of Green's formula becomes Cauchy's integral formula.
543:
544:
545: \begin{proposition}
546: Suppose that $f$ is a Clifford holomorphic function on some domain $U$. Then $xf(x)$ is
547: harmonic.
548: \end{proposition}
549: {\bf{Proof:}}
550: $Dxf(x)=-nf(x)-\Sigma_{j=1}^{n}x_{j}\frac{\partial f(x)}{\partial x_{j}}
551: -\Sigma_{\stackrel{j, k}{j\neq k}}x_{k}e_{k}e_{j}
552: \frac{\partial f(x)}{\partial x_{j}}$.
553: Now
554: \[\Sigma_{\stackrel{j, k}{j\neq k}}x_{k}e_{k}e_{j}
555: \frac{\partial f(x)}{\partial x_{j}}=
556: \Sigma_{k=1}^{n}\Sigma_{j\neq k}x_{k}e_{k}e_{j}
557: \frac{\partial f(x)}{\partial x_{j}}.\]
558: As $f$ is left monogenic this last expression simplifies to
559: $\Sigma_{k=1}^{n}x_{k}\frac{\partial f(x)}{\partial x_{k}}$. Moreover
560: $D\Sigma_{j=1}^{n}x_{j}\frac{\partial f(x)}{\partial x_{j}}=0$. Consequently
561: $D^{2}xf(x)=0$. $\Box$
562:
563: \ The previous proof is a generalization of the statement- "if $h(x)$ is a
564: real valued harmonic function then so is $<x,grad h(x)>$".
565:
566: \ In fact in the previous proof we determine that
567: $Dxf(x)=-nf(x)-2\Sigma_{j=1}^{n}x_{j}\frac{\partial f(x)}{\partial x_{j}}$.
568: In the special case where $f(x)=P_{k}(x)$, a left Clifford holomorphic polynomial of
569: order $k$, this equation simplifies to $DxP_{k}(x)=-(n+2k)P_{k}(x)$. Suppose
570: now that $h(x)$ is a harmonic function defined in a neighbourhood of the
571: ball $B(0,R)$. Now $Dh$ is a left Clifford holomorphic function so we know that there
572: is a series $\Sigma_{l=0}^{\infty}P_{l}(x)$ of left Clifford holomorphic polynomials
573: with each $P_{l}$ homogeneous of degree $l$ and such that the series
574: converges locally uniformly on $B(0, R)$ to $Dh(x)$. Now consider
575: the series $\Sigma_{l=0}^{\infty}\frac{-1}{n+2l}P_{l}(x)$. As
576: $\frac{1}{n+2l}\|P_{l}(x)\|<\|P_{l}(x)\|$ then this new series converges
577: locally uniformly on $B(0, R)$ to a left Clifford holomorphic function $f_{1}(x)$.
578: Moreover, $Dxf_{1}(x)=Dh(x)$ on $B(0, R)$. Consequently $h(x)-xf_{1}(x)$ is
579: equal to a left Clifford holomorphic function $f_{2}(x)$ on $B(0, R)$. Thus we have
580: established:
581: \begin{proposition}
582: Suppose that $h$ is a harmonic function defined in a neighbourhood of
583: $B(0, R)$ then there are left Clifford holomorphic functions $f_{1}$ and $f_{2}$ defined
584: on $B(0, R)$ such that $h(x)=xf_{1}(x)+f_{2}(x)$ for each $x\in B(0, R)$.
585: \end{proposition}
586:
587:
588: \ This result remains invariant under translation. As a consequence
589: it shows us that all harmonic functions are real analytic functions. So
590: there is no need to specify whether or not a harmonic function is $C^{2}$.
591: The result also provides an Almansi type decomposition of harmonic functions
592: in terms of Clifford holomorphic functions over any ball in $R^{n}$.
593:
594: \ It should be noted that Proposition 3 remains true if $h$ is only real
595: valued.
596:
597: \ Proposition 3 gives rise to an alternative proof of the Mean Value
598: Theorem
599: for harmonic functions.
600:
601: \begin{theorem}
602: For any harmonic function $h$ defined in a neighbourhood of a ball $B(a, R)$
603: \[h(a)=\frac{1}{\omega_{n}}\int_{\partial B(a, r)}h(x)d\sigma(x)\]
604: for any $r<R$.
605: \end{theorem}
606: {\bf{Proof:}}
607: Proposition 3 tells us that there is a pair of left Clifford holomorphic functions
608: $f_{1}$ and $f_{2}$ such that $h(x)=(x-a)f_{1}(x)+f_{2}(x)$ on $B(a, R)$.
609: So $h(a)=f_{2}(a)$, and we have previously shown that $\frac{1}{\omega_{n}}
610: \int_{\partial B(a, r)}f_{2}(x)d\sigma(x)=f_{2}(a)$. Now
611: $\int_{\partial B(a,r)}(x-a)f_{1}(x)d\sigma(x)=
612: r\int_{\partial B(a, r)}n(x)f_{1}(x)d\sigma(x)=0$. $\Box$
613:
614: \ The following is an immediate consequence of Proposition 3.
615: \begin{proposition}
616: If $h_{l}(x)$ is a harmonic polynomial homogeneous of degree $l$ then
617: \[h_{l}(x)=p_{l}(x)+xp_{l-1}(x)\]
618: where $p_{l}$ is a left Clifford holomorphic polynomial homogeneous of degree $l$ while
619: $p_{l-1}$ is a left monogenic polynomial which is homogeneous of degree $l-1$.
620: \end{proposition}
621:
622: \ It is well known that pairs of homogeneous harmonic polynomials of
623: differing degrees of homogeneity are orthogonal with respect to the usual
624: inner product over the unit sphere. Proposition 4 offers a further
625: refinement
626: to this. Suppose that $f$ and $g$ are $Cl_{n}$ valued functions defined on
627: $S^{n-1}$ and each component of $f$ and $g$ is square integrable. If we define
628: the $Cl_{n}$ inner product of $f$ and $g$ to be
629: \[<f, g>=\frac{1}{\omega_{n}}\int_{S^{n-1}}\overline{f(x)}g(x)d\sigma(x)\]
630: then if $f$ and $g$ are both real valued this inner product is equal to
631: \[\frac{1}{\omega_{n}}\int_{S^{n-1}}f(x)g(x)d\sigma(x)\]
632: which is the usual inner product for real valued square integrable functions
633: defined on $S^{n-1}$. Now
634: \[<xp_{l-1}(x), p_{l}(x)>=
635: -\frac{1}{\omega_{n}}\int_{S^{n-1}}\overline{p}_{l-1}(x)xp_{l}(x)d\sigma(x)\]
636: \[=-\frac{1}{\omega_{n}}\int_{S^{n-1}}\overline{p}_{l-1}(x)n(x)p_{l}(x)
637: d\sigma(x)=0.\]
638: The evaluation of the last integral is an application of Cauchy's theorem.
639:
640: \ Let us denote the space of $Cl_{n}$ valued functions defined on $S^{n-1}$
641: and such that each component is square integrable by $L^{2}(S^{n-1},Cl_{n})$.
642: Clearly the space of real valued square integrable functions defined on
643: $S^{n-1}$ is a subset of $L^{2}(S^{n-1},Cl_{n-1})$. The space
644: $L^{2}(S^{n-1},Cl_{n})$ is a $Cl_{n}$ module.
645:
646: \ We have shown that by introducing the module $L^{2}(S^{n-1}, Cl_{n})$
647: Proposition 4 provides a further orthogonal decomposition of harmonic
648: polynomials using left Clifford holomorphic polynomials. We shall return to this theme
649: later. This decomposition was introduced for the case $n=4$ by Sudbery
650: \cite{su} and independently extended for all $n$ by Sommen \cite{s2}.
651:
652:
653: \ Let us now consider higher order iterates of the Dirac operator $D$. In the
654: same way as we have that $DH(x)=G(x)$ there is a function $G_{3}(x)$ defined
655: on $R^{n}\backslash\{0\}$ such that $DG_{3}(x)=H(x)$. Specifically
656: $G_{3}(x)=C(n, 3)\frac{x}{\|x\|^{n-2}}$ for some dimensional constant
657: $C(n,3)$. Continuing inductively we may find a function $G_{k}(x)$ on $R^{n}
658: \backslash\{0\}$ such that $DG_{k}(x)=G_{k-1}(x)$. Specifically
659: \[G_{k}(x)=C(n,k)\frac{x}{\|x\|^{n-k+1}}\]
660: when $n$ is odd and so is $k$.
661: \[G_{k}(x)=C(n,k)\frac{1}{\|x\|^{n-k}}\]
662: when $n$ is odd and $k$ is even
663: \[G_{k}(x)=C(n,k)\frac{x}{\|x\|^{n-k+1}}\]
664: when $n$ is even, $k$ is odd and $k<n$
665: \[G_{k}(x)=C(n,k)\frac{1}{\|x\|^{n-k}}\]
666: when $n$ is even, $k$ is even and $k<n$
667: \[G_{k}(x)=C(n,k)(x^{k-n}\log\|x\|+A(n,k)x^{k-n})\]
668: when $n$ is even and $k\geq n$. In the last expression $A(n,k)$ is a real
669: constant dependent on $n$ and $k$. $C(n,k)$ is a constant dependent on $n$
670: and $k$ throughout.
671:
672: \ It should be noted that $G_{1}(x)=G(x)$ and $G_{2}(x)=H(x)$. It should
673: also be noted that $D^{k}G_{k}(x)=0$.
674:
675: \ Here is a simple technique for constructing solutions to the equation
676: $D^{k}g=0$ from left Clifford holomorphic functions. The special case $k=2$ was
677: illustrated in Proposition 2.
678: \begin{proposition}
679: Suppose that $f$ is a left Clifford holomorphic function on $U$ then
680: $D^{k}x^{k-1}f(x)=0$.
681: \end{proposition}
682: {\bf{Proof}} The proof is by induction. We have already seen the result to be
683: true in the case $k=2$ in Proposition 2. If $k$ is odd then $Dx^{k-1}f(x)
684: =(k-1)x^{k-2}f(x)$. If $k$ is even then
685: \[Dx^{k-1}f(x)=-n(k-1)x^{k-2}f(x)
686: +x^{k-2}\Sigma_{j=1}^{n}e_{j}x\frac{\partial f(x)}{\partial x_{j}}.\]
687: By arguments presented in Proposition 5 this expression is equal to
688: \[-n(k-1)x^{k-2}f(x)+x^{k-2}\Sigma_{j=1}^{n}x_{j}\frac{\partial f(x)}
689: {\partial x_{j}}.\]
690: The induction hypothesis tells us that the only term we need consider is
691: $x^{k-2}\Sigma_{j=1}^{n}x_{j}\frac{\partial f(x)}{\partial x_{j}}$.
692: However $\Sigma_{j=1}^{n}x_{j}\frac{\partial f(x)}{\partial x_{j}}$ is a left
693: Clifford holomorphic function. So proof by induction is now complete. $\Box$
694:
695: \ In future we shall refer to a function $g:U\rightarrow Cl_{n}$ which
696: satisfies the equation $D^{k}g=0$ as a left $k$-monogenic function. Similarly
697: if $h:U\rightarrow Cl_{n}$ satisfies the equation $hD^{k}=0$ then $h$ is a
698: right $k$-monogenic function. In the case where $k=1$ we return to the
699: setting of left, or right, Clifford holomorphic functions and when $k=2$ we return to the
700: setting of harmonic functions. When $k=4$ the equations $D^{4}g=0$ and
701: $gD^{4}=0$ correspond to the equations $\triangle_{n}^{2}g=0$ and
702: $\triangle_{n}^{2}h=0$. So left or right $4$-monogenic functions are in fact
703: biharmonic functions. In greater generality if $k$ is even then a left or
704: right $k$-monogenic function $f$ automatically satisfies the equation
705: $\triangle_{n}^{\frac{k}{2}}f=0$.
706:
707: \begin{proposition}
708: Suppose that $p$ is a left $k$-monogenic polynomial homogeneous of degree $q$
709: then there are left Clifford holomorphic polynomials $f_{0},\ldots, f_{k-1}$ such that
710: \[p(x)=f_{0}(x)+\ldots+x^{k-1}f_{k-1}(x)\]
711: and each polynomial $f_{j}$ is homogeneous of degree $q-j$ whenever $q-j\geq
712: 0$ and is identically zero otherwise.
713: \end{proposition}
714: {\bf{Proof:}} The proof is via induction on $k$. The case $k=2$ is
715: established
716: immediately after the proof of Proposition 2. Let us now consider $Dp(x)$.
717: This is a left $k-1$-monogenic polynomial homogeneous of degree $q-1$. So by
718: the induction hypothesis $Dp(x)=g_{1}(x)+\ldots+x^{k-2}g_{k-1}(x)$ where
719: each $g_{j}$ is a left Clifford holomorphic polynomial homogeneous of degree $q-j$
720: whenever $q-j\geq 0$ and is equal to zero otherwise. Using Euler's lemma and
721: the observations made after the proof of Proposition 5 one may now find
722: left Clifford holomorphic polynomials $f_{1}(x),\ldots,f_{k-1}(x)$ such that
723: $D(xf_{1}(x)+\ldots+x^{k-1}f_{k-1}(x))=Dp(x)$ and $f_{j}(x)=c_{j}g_{j}(x)$
724: for some $c_{j}\in R$ and for $1\leq j\leq k-1$. It follows that $p(x)-
725: \Sigma_{j=1}^{k-1}x^{j}f_{j}(x)$ is a left Clifford holomorphic polynomial $f_{0}$
726: homogeneous of degree $q$. $\Box$
727:
728: \ One may now use Proposition 6 and the arguments used to establish
729: Proposition 3 to deduce:
730: \begin{theorem}
731: Suppose that $f$ is a left $k$-monogenic function defined in a neighbourhood
732: of the ball $B(0, R)$ then there are left monogenic functions $f_{0},\ldots,
733: f_{k-1}$ defined on $B(0,R)$ such that
734: $f(x)=f_{0}(x)+\ldots+x^{k-1}f_{k-1}(x)$ on $B(0,R)$.
735: \end{theorem}
736:
737: \ Theorem 6 establishes an Almansi decomposition for left $k$-monogenic
738: functions in terms of left Clifford holomorphic functions over any open ball. It also
739: follows from this theorem that each left $k$-monogenic function is a real
740: analytic function. It is also reasonably well known that if $h$ is a
741: biharmonic function defined in a neighbourhood of $B(0,R)$ then there are
742: harmonic functions $h_{1}$ and $h_{2}$ defined on $B(0, R)$ and such that
743: $h(x)=h_{1}(x)+\|x\|^{2}h_{2}(x)$. In the special case where $k=4$ Theorem
744: 6 both establishes this result and refines it.
745:
746: \ As each left $k$-monogenic function is a real analytic function then we
747: can immediately use Stokes' theorem to deduce the following Cauchy-Green
748: type formula.
749: \begin{theorem}
750: Suppose that $f$ is a left $k$-monogenic function defined on some domain
751: $U$ and suppose that $S$ is a piecewise $C^{1}$ compact surface lying in $U$
752: and bounding a bounded subdomain $V$ of $U$. Then for each $y\in V$
753: \[f(y)=\frac{1}{\omega_{n}}\int_{S}(\Sigma_{j=1}^{k}(-1)^{j-1}G_{j}(x-y)n(x)
754: D^{j-1}f(x))d\sigma(x).\]
755: \end{theorem}
756:
757: \begin{center}{\Large Conformal Groups and Clifford Analysis}\end{center}
758:
759: \ Here we will examine the role played
760: by the conformal group within parts of Clifford analysis. Our starting point
761: is to ask what type of diffeomorphisms acting on subdomains of $R^{n}$
762: preserve Clifford holomorphic functions. If a diffeomorphism $\phi$ can transform the
763: class of left Clifford holomorphic functions on one domain $U$ to a class of left
764: Clifford holomorphic functions on the domain $\phi(U)$ and do the same for the class
765: of right Clifford holomorphic functions on $U$ then it must preserve Cauchy's theorem.
766: So if $f$ is left Clifford holomorphic on $U$ and $g$ is right Clifford holomorphic on $U$ and
767: these functions are transformed to $f'$ and $g'$ respectively left and right
768: Clifford holomorphic functions on $\phi(U)$ then
769: \[\int_{S}g(x)n(x)f(x)d\sigma(x)=0=\int_{\phi(S)}g'(y)n(y)f'(y)d\sigma(y)\]
770: where $S$ is a piecewise $C^{1}$ compact surface lying in $U$ and $y=\phi(x)$.
771: An important point to note here is that we need to assume that $\phi$
772: preserves vectors orthogonal to the tangent spaces at $x$ and $\phi(x)$. As
773: the choice of $x$ and $S$ is arbitrary it follows that the diffeomorphism
774: $\phi$ is angle preserving. In other words $\phi$ is a conformal
775: transformation. A theorem of Liouville \cite{l} tells us that for dimensions
776: $3$ and greater the only conformal transformations on domains are
777: M\"{o}bius transformations.
778:
779: \ In order to deal with M\"{o}bius transformations using Clifford algebras
780: we have seen in a previous chapter that one can use Vahlen matrices. We
781: now proceed to show that each M\"{o}bius transformation
782: preserves monogenicity. Sudbery \cite{su} and also Bojarski \cite{b} have
783: established this fact. We will need the following two lemmata.
784:
785: \begin{lemma}
786: Suppose that $\phi(x)=(ax+b)(cx+d)^{-1}$ is a M\"{o}bius transformation
787: then
788: \[G(u-v)=J(\phi, x)^{-1}G(x-y)\tilde{J}(\phi, y)^{-1}\]
789: where $u=\phi(x)$, $v=\phi(y)$ and $J(\phi, x)=\frac{(\widetilde{cx+d})}
790: {\|cx+d\|^{n}}$.
791: \end{lemma}
792: {\bf{Proof}} The proof essentially follows from the fact that
793: \[(x^{-1}-y^{-1})=x^{-1}(y-x)y^{-1}.\]
794: Consequently $\|x^{-1}-y^{-1}\|=\|x\|^{-1}\|x-y\|\|y\|^{-1}$.
795: Also $ax\tilde{a}-ay\tilde{a}=a(x-y)\tilde{a}$.
796:
797: \ If one breaks the transformation down into terms arising from the generators
798: of the M\"{o}bius group and use the previous set of equations then one will
799: readily arrive at the result. $\Box$
800: \begin{lemma}
801: Suppose that $y=\phi(x)=(ax+b)(cx+d)^{-1}$ is a M\"{o}bius transformation
802: and for domains $U$ and $V$ we have $\phi(U)=V$ then
803: \[\int_{S}f(u)n(u)g(u)d\sigma(u)=\int_{\psi^{-1}(S)}f(\psi(x))\tilde{J}(\psi,x)
804: n(x)J(\psi,x)g(\psi(x))d\sigma(x)\]
805: where $u=\psi(x)$, $S$ is a orientable hypersurface lying in $U$ and
806: $J(\psi,x)=\frac{\widetilde{cx+d}}{\|cx+d\|^{n}}$.
807: \end{lemma}
808: {\bf{Outline Proof}} On breaking $\psi$ up into the generators of the
809: M\"{o}bius group the result follows from noting that
810: \[\frac{\partial x^{-1}}{\partial x_{j}}=-x^{-1}e_{j}x^{-1}.\]
811: $\Box$
812:
813: \ It follows from Cauchy's Theorem that if $g(u)$ is a left Clifford holomorphic
814: function in the variable $u$ then $J(\psi,x)f(\psi(x))$ is left Clifford holomorphic
815: in the variable $x$.
816:
817:
818: \ When $\phi(x)$ is the Cayley transformation $y=(e_{n}x+1)(x+e_{n})^{-1}$
819: we can use this transformation to establish a Cauchy-Kowalewska extension
820: in a neighbourhood of the sphere. If $f(x)$ is a real analytic function
821: defined on an open subset $U$ of $S^{n-1}\backslash\{e_{n}\}$ then $l(y)=
822: J(\phi^{-1},y)^{-1}f(\phi(y))$ is a real analytic function on the open set
823: $V=\phi^{-1}(U)$. This function has a Cauchy-Kowalewska extension to a left
824: Clifford holomorphic function $L(y)$ defined on an open neighbourhood $V(g)\subset R^{n}$
825: of $V$. Consequently $F(x)=J(\phi^{-1}, x)L(\phi^{-1}(x))$ is a left
826: Clifford hholomorphic defined on an open neighbourhood $U(f)=\phi^{-1}(V(g))$ of $U$.
827: Moreover $F_{|U}=f$. Combing with similar arguments for the other Cayley
828: transformation $y=(-e_{n}x+1)(x-e_{n})^{-1}$ one can deduce:
829: \begin{theorem}
830: {\bf{(Cauchy-Kowalewska Theorem)}}
831: Suppose that $f$ is a $Cl_{n}$ valued real analytic function defined on
832: $S^{n-1}$. Then there is a unique left Clifford holomorphic function $F$ defined on an
833: open neighbourhood $U(f)$ of $S^{n-1}$ such that $F_{|S^{n-1}}=f$.
834: \end{theorem}
835:
836: \ In fact if $f(u)$ is defined on some domain and satisfies the equation
837: $D^{k}f=0$ then the function $J_{k}(\psi,x)f(\psi(x))$ satisfies the same
838: equation, where $J_{k}(\psi,x)=\frac{\widetilde{cx+d}}{\|cx+d\|^{n-k+1}}$.
839: \begin{theorem}{\bf{(Fueter-Sce Theorem)}}
840: Suppose that $f=u+iv$ is a holomorphic function on a domain $\Omega\subset
841: {\bf{C}}$ and that $\Omega=\overline{\Omega}$ and $f(\overline{z})=
842: \overline{f(z)}$. Then the function $F(\underline{x})=u(x_{1},\|x'\|)
843: +e_{1}^{-1}\frac{x'}{\|x'\|}v(x_{1},\|x'\|)$ is a unital left
844: $n-1$-monogenic function
845: on the domain $\{\underline{x}:x_{1}+i\|x'\|\in\Omega\}$ whenever $n$ is
846: even. Here $x'=x_{2}e_{2}+\ldots+x_{n}e_{n}$.
847: \end{theorem}
848: {\bf{Proof:}}
849: First let us note that $x^{-1}e_{1}$ is left $n-1$ monogenic whenever
850: $n$ is even. It follows that $\frac{\partial^{k}}{\partial
851: x_{1}^{k}}x^{-1}e_{1}=c_{k}x^{-k-1}e_{1}$ is $n-1$ left monogenic for each
852: positive integer $k$. Here $c_{k}$ is some non-zero real number. Using
853: Kelvin inversion it follows that $x^{k}e_{1}$ is left $n-1$ monogenic for
854: each positive integer $k$. By taking translations and Taylor series
855: expansions for the function $f$ the result follows. $\Box$
856:
857: \ This result was first established for the case $n=4$ by Fueter, \cite{f}, see
858: also Sudbery \cite{su}. It was extended to all even dimensions by Sce \cite{sc},
859: though the methods used do not make use of the conformal group. This result has been applied in \cite{q1,q2} to study various types of singular integral operators acting on $L^{p}$ spaces of Lipschitz perturbations of the sphere.
860:
861: \begin{center}{\Large Conformally Flat Spin Manifolds}\end{center}
862:
863: \ The invariance of monogenic functions under M\"{o}bius transformations described in the previous section makes use of a conformal weight factor $J(\psi,x)$. This invariance can be seen as an automorphic form invariance. This leads to a natural generalization of the concept of a Riemann surface to the euclidean setting. A manifold $M$ is said to be conformally flat if there is an atlas ${\cal A}$ of $M$ whose transition functions are M\"{0}bius transformations. For instance via the Cayley transformations $(e_{n+1}x+1)(x+e_{n+1})^{-1}$ and $(-e_{n+1}x+1)(x-e_{n+1})^{-1}$ one can see that the sphere $S^{n}\subset R^{n+1}$ is an example of a conformally flat manifold. Another way of constructing conformally flat manifolds is to take a simply connected domain $U$ of $R^{n}$ and consider a Kleinian subgroup $\Gamma$ of the M\"{o}bius group that acts discontinuously on $U$. Then the factorization $U\backslash \Gamma$ is a conformally flat manifold. For instance let $U=R^{n}$ and let $\Gamma$ be the integer lattice $Z^{k}=Ze_{1}+\ldots+Ze_{k}$ for some positive integer $k\leq n$. In this case $R^{n}\backslash Z^{k}$ gives the cylinder $C_{k}$ and when $k=n$ we get the $n$-torus. Also if we let $U=R^{n}\backslash\{0\}$ and $\Gamma=\{2^{k}:k\in Z\}$ the resulting manifod is $S^{1}\times S^{n-1}$.
864:
865: \ We locally construct a spinor bundle over $M$ by making the identification $(u,X)$ with either $(x,\pm J(\psi,x)X)$ where $u=\psi(x)=(ax+b)(cx+d)^{-1}=(-ax-b)(-cx-d)^{-1}$. If we can compatibly choose the signs then we have created a spinor bundle over the conformally flat manifold. Note, it might be possible to create more than one spinor bundle over $M$. For instance consider the cylinder $C_{k}$ if we make the identification $(x,X)$ with $x+\underline{m},(-1)^{m_{1}+\ldots+m_{l}}X)$ where $l$ is a fixed integer with $l\leq k$ and $\underline{m}=m_{1}e_{1}+\ldots+m_{l}e_{l}+\ldots+m_{k}e_{k}$ then we have created $k$ different spinor bundles $E^{1},\ldots E^{k}$ over $C_{k}$.
866:
867: \ As we have used the conformal weight function function $J(\psi,x)$ to construct the spinor bundle $E$ then it is easy to see that a section $f:M\rightarrow E$ could be called a left monogenic section if it is locally a left monogenic function. It is now natural to ask if one can construct Cauchy integral formulas for such sections. To do this we essentially need to construct a kernel over the euclidean domain $U$ that is periodic with respect to $\Gamma$ and then use the projection map $p:U\rightarrow M$ to construct from this kernel a Cauchy kernel for $U$. In \cite{kr2} we show that the Cauchy kernel for
868: $C_{k}$ with spinor bundle $E^{l}$ is constructed from the kernel
869: \[\cot_{k,l}(x,y)=\Sigma_{\underline{m}\in Z^{l},\underline{n}\in Z^{k-l}}(-1)^{m_{1}+\ldots+m_{l}}G(x-y+\underline{m}+\underline{n}),\]
870: where $\underline{n}=n_{l+1}e_{l+1}+\ldots+n_{n}e_{n}$.
871: While for the conformally flat spin manifold $S^{1}\times S^{n-1}$ with trivial bundle $Cl_{n}$ the Cauchy kernel is constructed from the kernel
872: \[\Sigma_{k=0}^{\infty}G(2^{k}x-2^{k}y)+2^{2-2n}G(x)(\Sigma_{k=-1}^{-\infty}G(2^{-k}x^{-1}-2^{-k}y^{-1}))G(y).\]
873:
874: \ See \cite{k,kr1,kr2} for more details and related results.
875:
876: \ It should be noted that one may set up a Dirac operator over arbitrary Riemannian manifolds ,see for instance \cite{bbw}, and one may set up Cauchy integral formulas for functions annihilated by these Dirac operators, see for instance \cite{c,m2} for details.
877:
878: \begin{center}{\Large Boundary Behaviour and Hardy Spaces}\end{center}
879:
880: \ Possibly the main topic that unites all that has been previously
881: discussed here on Clifford analysis is its applications to boundary value
882: problems. This in turn leads to a study of boundary behaviour of classes
883: of Clifford holomorphic functions and Hardy spaces. Let us look first at one of the
884: simplest cases. Previously we noted that if $\theta$ is a square integrable
885: function defined on the sphere $S^{n-1}$ then there is a harmonic function
886: $h$ defined on the unit ball in $R^{n}$ with boundary value $\theta$
887: almost everywhere. Previously we have seen that $h(x)=f_{1}(x)+xf_{2}(x)$
888: where $f_{1}$ and $f_{2}$ are left Clifford holomorphic. However on $S^{n-1}$ the
889: function $G(x)$ equals $x$. So one can see that on $S^{n-1}$ we have
890: $\theta(x)=f_{1}(x)+g(x)$ almost everywhere. Here $f_{1}$ is left
891: monogenic on the unit ball $B(0,1)$ and $g$ is left Clifford holomorphic on
892: $R^{n}\backslash clB(0,1)$, where $clB(0,1)$ is the closure of the open
893: unit ball. Let $H^{2}(B(0,1))$ denote the space of Clifford holomorphic functions
894: defined on $B(0,1)$ with extension to a square integrable function on
895: $S^{n-1}$ and let $H^{2}(R^{n}\backslash clB(0,1)$ denote the class of
896: left Clifford holomorphic functions defined on $R^{n}\backslash clB(0,1)$ with square
897: integrable extension to the $S^{n-1}$. What we have so far outlined is
898: that $L^{2}(S^{n-1})=H^{2}(B(0,1))\oplus H^{2}(R^{n}\backslash clB(0,1))$,
899: where $L^{2}(S^{n-1})$ is the space of $Cl_{n}$ valued Lebesgue square
900: integrable functions defined on $S^{n-1}$. This is the Hardy $2$-space
901: decomposition of $L^{2}(S^{n-1})$. It is also true if we replace $2$ by
902: $p$ where $1<p<\infty$. We will not go into this detail here as it is
903: beyond the scope of the material covered here.
904:
905: \ Let us now take an alternative look at a way of obtaining this
906: decomposition. This method will generalize to all reasonable surfaces. We
907: will clarify what we mean by a reasonable surface later. Instead of
908: considering an arbitrary square integrable function on $S^{n-1}$ let us
909: instead assume that $\theta$ is a continuously differentiable function. Let
910: us now consider the integral
911: \[\frac{1}{\omega_{n}}\int_{S^{n-1}}G(x-y)n(x)\theta(x)d\sigma(x)\]
912: where $y\in B(0,1)$. This defines a left Clifford holomorphic function on $B(0,1)$.
913: Now let us allow the point $y$ to approach a point $z\in S^{n-1}$ along a
914: differentiable path $y(t)$. Let us also assume that $\frac{dy(t)}{dt}$
915: evaluated at $t=1$, so $y(t)=z$, is not tangential to $S^{n-1}$ at $z$. We
916: can essentially ignore this last point at a first read. We want to
917: evaluate
918: \[\lim_{t\rightarrow
919: 1}\frac{1}{\omega_{n}}\int_{S^{n-1}}G(x-y(t))n(x)\theta(x)d\sigma(x).\]
920: We do this by removing a small ball on $B(0,1)$ from $S^{n-1}$. The ball
921: is centered at $z$ and is of radius $\epsilon$. We denote this ball by
922: $b(z,\epsilon)$. The previous integral now splits into an integral over
923: $b(z,\epsilon)$ and an integral over $S^{n-1}\backslash b(z,\epsilon)$. On
924: $b(z, \epsilon)$ we can express $\theta(x)$ as
925: $(\theta(x)-\theta(z))+\theta(z)$. As $\theta$ is continuously
926: differentiable then $\|\theta(x)-\theta(z)\|<C\|x-z\|$ for some $C\in
927: R^{+}$. It follows that
928: \[\lim_{\epsilon\rightarrow 0}\lim_{t\rightarrow
929: 1}\int_{b(z,\epsilon)}\|G(x-y(t))n(x)(\theta(x)-\theta(z))\|d\sigma(x)=0.\]
930: Moreover, the term
931: \[\lim_{\epsilon\rightarrow 0}\lim_{t\rightarrow
932: 1}\frac{1}{\omega_{n}}\int_{b(z,\epsilon}G(x-y(t))n(x)\theta(z)d\sigma(x)\]
933: can, as $\theta(z)$ is a Clifford holomorphic function, be replaced by the term
934: \[\lim_{\epsilon\rightarrow 0}\lim_{t\rightarrow
935: 1}\int_{B(0,1)\cap\partial
936: B(z,\epsilon)}G(x-y(t))n(x)\theta(z)d\sigma(x).\]
937: By the residue theorem the limit of this integral evaluates to
938: $\frac{1}{2}\theta(z)$.
939:
940: \ We leave it to the interested reader to note that the singular integral
941: or principal valued integral
942: \[\lim_{\epsilon\rightarrow 0}\lim_{t\rightarrow
943: 1}\frac{1}{\omega_{n}}\int_{S^{n-1}\backslash
944: b(z,\epsilon)}G(x-y(t))n(x)\theta(x)d\sigma(x)=\]
945: \[P.V.\frac{1}{\omega_{n}}\int_{S^{n-1}}G(x-z)n(x)\theta(x)d\sigma(x)\]
946: is bounded.
947:
948: \ We have established that
949: \[\lim_{t\rightarrow 1}\int_{S^{n-1}}G(x-y(t))n(x)\theta(x)d\sigma(x)=\]
950: \[\frac{1}{2}\theta(z)+
951: P.V.\frac{1}{\omega_{n}}\int_{S^{n-1}}G(x-z)n(x)\theta(x)d\sigma(x).\]
952:
953: \ If we now assumed that $y(t)$ is a path tending to $z$ on the complement
954: of $clB(0,1)$, then similar arguments give
955: \[\lim_{t\rightarrow 1}\int_{S^{n-1}}G(x-y(t))n(x)\theta(x)d\sigma(x)=\]
956: \[-\frac{1}{2}\theta(z)+
957: P.V.\frac{1}{\omega_{n}}\int_{S^{n-1}}G(x-z)n(x)\theta(x)d\sigma(x).\]
958: We will write these expressions as
959: \[(\pm\frac{1}{2}I+C_{S^{n-1}})\theta.\]
960:
961: \ If we now consider the limit
962: \[\lim_{t\rightarrow 1}\frac{1}{\omega_{n}}\int_{S^{n-1}}G(x-y(t))n(x)
963: (\frac{1}{2}I+C_{S^{n-1}})\theta(x)d\sigma(x)\]
964: we may determine that
965: \[(\frac{1}{2}I+C_{S^{n-1}})^{2}=\frac{1}{2}I+C_{S^{n-1}}.\]
966: Furthermore
967: \[(\frac{1}{2}I+C_{S^{n-1}})(-\frac{1}{2}I+C_{S^{n-1}})=0\]
968: and
969: \[(-\frac{1}{2}I+C_{S^{n-1}})^{2}=-\frac{1}{2}I+C_{S^{n-1}}.\]
970:
971: \ It is known that each function $\psi\in L^{2}(S^{n-1})$ can be
972: approximated by a sequence of functions each with the same properties as
973: $\theta$. This tells us that the previous formulas can be repeated but
974: this time simply for $\theta\in L^{2}(S^{n-1})$. It follows that for such
975: a $\theta$ we have
976: \[\theta=(\frac{1}{2}I+C_{S^{n-1}})\theta+(\frac{1}{2}I-C_{S^{n-1}})\theta.\]
977: This formula gives the Hardy space decomposition of $L^{2}(S^{n-1})$. In
978: fact if one looks more carefully at the previous calculations to obtain
979: these formulas we see that it is not so significant that the surface used
980: is a sphere and we can re-do the calculations for any "reasonable" hypersurface
981: $S$. In this case we get
982: \[\theta=(\frac{1}{2}I+C_{S})\theta+(\frac{1}{2}I-C_{S})\theta\]
983: where $\theta$ now belongs to $L^{2}(S)$ and
984: \[C_{S}\theta=P.V.\frac{1}{\omega_{n}}\int_{S}G(x-y)n(x)\theta(x)d\sigma(x).\]
985: This gives rise to the Hardy space decomposition
986: \[L^{2}(S)=H^{2}(S^{+})\oplus H^{2}(S^{-})\]
987: where $S^{\pm}$ are the two domains that complement the surface $S$ (we
988: are assuming that $S$ divides $R^{n}$ into two complementary domains.
989:
990: \ Last one should address the smoothness of $S$. In some parts of the
991: literature one simply assumes that $S$ is compact and $C^{2}$, while in
992: more advanced and recent aspects of the literature one assumes that $S$
993: has rougher conditions, usually one assumes that the surface is simply
994: Lipschitz continuous, see for instance \cite{lmq,lms,m1}. The formulas given above involving the singular
995: integral operator $C_{S}$ are called Plemelj formulas. It is a simple
996: exercise to see that these formulas are conformally invariant. So using
997: Kelvin inversion or even a Cayley transformation one can see that these
998: formulas and the Hardy space decompositions are also valid on unbounded
999: surfaces and domains. A great deal of modern Clifford analysis has been
1000: devoted to the study of such Hardy spaces and singular integral operators. This is in fact due to an idea of R. Coifman that various hard problems in classical harmonic analysis studied in euclidean space might be more readily handled using tools from Clifford analysis, particularly the singular Cauchy transform and associated Hardy spaces. In particular Coifman speculated that a more direct proof of the celebrated Coifman-McIntosh-Meyer Theorem \cite{cmm}could be derived using Clifford analysis. The Coifman-McIntosh-Meyer Theorem establishes the $L^{2}$ boundedness of the double layer potential operator for Lipschitz graphs in $R^{n}$. Coifman's observation was that the double layer potential operator is the real or scalar part of the singular Cauchy transform arising in Clifford analysis and discussed earlier. If one can establish the $L^{2}$ boundedness of the singular Cauchy transform for a Lipschitz graph in $R^{n}$ then one automatically has the $L^{2}$ boundedness for the double layer potential operator for the same graph. The $L^{2}$ boundedness of the singular Cauchy transform was first established for Lipschitz graphs with small constant by M. Murray \cite{m} and extended to the general case by A. McIntosh, see for instance \cite{mc}, see also \cite{m1}. One very importand reason for needing to know that the double layer potential operator is $L^{2}$ bounded for Lipschitz graphs is to be able to solve bounndary value problems for domains with Lipschitz graphs as boundaries. such boundary value problems would include the Dirichlet problem and Neuaman problem for the Laplacian. See \cite{mc,m1} for more details. Moreover in \cite{w} Clifford analysis and more precisely the Hardy space decomposition mentioned here is specifically used to solve the water wave problem in three dimensions.
1001:
1002: \begin{center}{\Large More on Clifford Analysis on the Sphere}\end{center}
1003:
1004: \ In the previous section We saw that $L^{2}(S^{n-1})$ splits into a direct sum of Hardy spaces for the corresponding complemetary domains $B(0,1)$ and $R^{n}\backslash cl(B(0,1))$. In an earlier section we saw that any left Clifford holomorphic function $f(x)$ can be expressed as a locally uniformly convergent series $\Sigma_{j=0}^{\infty}f_{j}(x)$ where each $f_{j}(x)$ is left Clifford holomorphic and homogeneous of degree $j$. Now following \cite{su} consider the operator
1005: \[D=x^{-1}xD=x^{-1}(\Sigma_{i<k}e_{i}e_{k}(x_{i}\frac{\partial}{\partial x_{k}}-x_{k}\frac{\partial}{\partial x_{i}}-\Sigma_{j=1}^{n}x_{j}\frac{\partial}{\partial x_{j}}).\]
1006: By letting the last term in this expression act on homogeneous polynomials one may determine from Euler's lemma that in fact $\Sigma_{j=1}^{n}x_{j}\frac{\partial}{\partial x_{j}}$ is the radial operator $r\frac{\partial}{\partial r}$. So $r\frac{\partial}{\partial r}f_{j}(x)=jf_{j}(x)$. As each polynomial $f_{k}$ is Clifford holomorphic it follows that each $f_{j}$ is an eigenvector of the spherical Dirac operator $x\Lambda_{n-1}=x\Sigma_{i<k}e_{i}e_{k}(x_{i}\frac{\partial}{\partial x_{k}}-x_{k}\frac{\partial}{\partial x_{i}})$ with eigenvalue $k$. Now using Kelvin inversion we know that $f_{k}$ is homogeneous of degree $k$ and left Clifford holomorphic if and only if $G(x)f_{k}(x^{-1})$ is homogeneous of degree $-n+1-k$ and is left Clifford holomorphic. On restricting $G(x)f_{k}(x^{-1})$ to the unit sphere this function becomes $xf_{k}(x^{-1})$ and this function is an eigenvector for the spherical Dirac operator $x\Lambda_{n-1}$. As each $f\in H^{2}(R^{n}\backslash cl(B(0,1)))$ can be written as $\Sigma_{k=0}^{\infty}G(x)f_{k}(x^{-1})$ where each $f_{k}$ is homogeneous of degree $k$ and is left Clifford holomorphic it follows that if $h\in L^{2}(S^{n-1})$ then $\Lambda_{n-1} xh(x)=(1-n)xh(x)-x\Lambda_{n-1} h(x)$. Similarly if we replace $S^{n-1}$ by the $n$-sphere $S^{n}$ embedded in $R^{n+1}$ then we have the identity $\Lambda_{n} xh(x)=-nxh(x)-x\Lambda_{n}h(x)$ for each $h\in L^{2}(S^{n})$. As all $C^{\infty}$ functions defined on $S^{n}$ belong to $L^{2}(S^{n})$ this identity holds for all such functions too.
1007:
1008: \ It should be noted that for each $x\in S^{n}$ if we restrict the operator $x\Lambda_{n}$ to the tangent bundle $TS^{n}_{x}$ then we obtain the Euclidean Dirac operator acting on this tangent space.
1009:
1010: \ By using the Cayley transformation $x=\psi(y)(e_{n+1}y+1)(y+e_{n+1})^{-1}$ from $R^{n}$ to $S^{n}\backslash\{e_{n+1}\}$ one can transform left Clifford holomorphic functions from domains in $R^{n}$ to functions defined on domains lying on the sphere. Namely if $f(y)$ is left Clifford holomorphic on the domain $U$ lying in $R^{n}$ then we obtain a function $f'(x)=J(\psi^{-1},x)f(\psi^{-1}(x)$ defined on the domain $U'=\psi(U)$ lying on $S^{n}$. Here $J(\psi^{-1},x)=\frac{x+1}{\|x+1\|^{n}}$. Similarly if $g(y)$ is right Clifford holomorphic on $U$ then $g'(x)=g(\psi^{-1}(x)J(\psi^{-1},x)$ is a well defined function on $U'$. Moreover for any smooth, compact hypersurface $S$ bounding a subdomain $V$ of $U$ we have from the conformal invariance of Cauchy's Theorem $\int_{S'}g'(x)n(x)f'(x)d\sigma'(x)=0$ where $S'=\psi(S)$, and $n(x)$ is the unit vector lying in the tangent space $TS^{n}_{x}$ of $S^{n}$ at $x$ and outer normal to
1011: $S'$ at $x$. Furthermore $sigma'$ is the Lebesgue measure on $S'$.
1012:
1013: \ From Lemma 1 it now follows that for each point $y'\in V'=\psi(V)$ we have the following version of Cauchy's Integral Formula:
1014: \[f'(y')=\frac{1}{\omega_{n}}\int_{S'}G(x-y')n(x)f'(x)d\sigma(x)\]
1015: where as before $G(x-y')=\frac{x-y'}{\|x-y'\|^{n}}$, but now $x$ and $y'\in S^{n}$.
1016: It would now appear that the functions $f'$ and $g'$ are solutions to some spherical Dirac equations. We need to isolate this Dirac operator. We shall achieve this by applying the operator $x\Lambda_{n}$ to the kernel $G_{s}(x,y')=G(x-y')$. As $x$ and $y'\in S^{n}$ then $\|x-y'\|^{2}=2-2<x,y'>$, where $<x,y'>$ is the inner product of $x$ and $y'$. So $G_{s}(x,y')=2^{\frac{-n}{2}}\frac{x-y'}{(1-<x,y'>)^{\frac{n}{2}}}$. So in calculating $x\Lambda_{n}G_{s}(x,y')$ we will need to know what $\Lambda_{n}<x,y'>$ evaluates to. It is a simple exercise to determine that $\Lambda_{n}<x,y'>=xy'+<x,y'>$ which is simply the wedge product, $x\wedge y'$, of $x$ with $y'$.
1017: Now let us calculate $x\Lambda_{n}G_{s}(x,y')$.
1018: Now
1019: \[x\Lambda_{n}G_{s}(x,y')=2^{\frac{-n}{2}}(x\Lambda_{n}\frac{x}{(1-<x,y'>)^{\frac{n}{2}}}-x\Lambda_{n}\frac{y'}{(1-<x,y'>)^{\frac{n}{2}}})\]
1020: \[=2^{\frac{n}{2}}(-x\frac{nx}{(1-<x,y'>)^{\frac{n}{2}}}+\Lambda_{n}\frac{1}{(1-<x,y'>)^{\frac{-n}{2}}}-x\Lambda_{n}\frac{1}{(1-<x,y'>)^{\frac{n}{2}}}y').\]
1021: First
1022: \[\Lambda_{n}\frac{1}{(1-<x,y'>)^{\frac{n}{2}}}=\frac{n}{2}\frac{x\wedge y'}{(1-<x,y'>)^{\frac{n+2}{2}}}.\]
1023: So
1024: \[x\Lambda_{n}G_{s}(x,y')=2^{\frac{-n}{2}}\frac{n}{2(1-<x,y>)^{\frac{n+2}{2}}}(2(1-<x,y'>)+x\wedge y'-x(x\wedge y')y').\]
1025: The expression
1026: \[2(1-<x,y'>)+x\wedge y'-x(x\wedge y')y'\]
1027: is equal to
1028: \[2-2<x,y'>+xy'+<x,y'>-x(xy+<x,y'>)y'.\]
1029: This expression simplifies to
1030: \[1-<x,y'>+xy'-xy'<x,y'>\]
1031: which in turn simplifies to
1032: \[(1-<x,y'>)(1+xy')=-x(1-<x,y'>)(x-y').\]
1033: So
1034: \[x\Lambda_{n}G_{s}(x,y')=\frac{-n}{2}xG_{s}(x,y').\]
1035: Hence $x(\Lambda_{n}+\frac{n}{2})G_{s}(x,y')=0$. So the Dirac operator, $D_{s}$, over the sphere is $x(\Lambda_{n}+\frac{n}{2})$. It follows from our Cauchy integral formula for the sphere that $D_{s}f'(x)=0$. For more details on this operator, related operators and their properties see \cite{cnm,lr,r5,r6,v}
1036:
1037: \ Besides the operator $D_{s}$ we also need a Laplacian $\triangle_{s}$ acting on functions defined on domains on $S^{n}$. To do this we will work backwards, and look first for a fundamental solution to $\triangle_{n}$. A strong candidate for such a fundamental solution is the kernel $H_{s}(x,y')=\frac{1}{n-2}\frac{1}{\|x-y'\|^{n-2}}$. By similar considerations to those made in the previous calculation we find that $D_{s}H_{s}(x,y')=-xH_{s}(x,y')+G_{s}(x,y')$. So $(D_{s}+x)H_{s}(x,y')=G_{s}(x,y')$. Therefore we may define our Laplacian $\triangle_{s}$ to be $D_{s}(D_{s}+x)$.
1038:
1039: \begin{definition}
1040: Suppose $h$ is a $Cl_{n}$ valued function defined on a domain $U'$ of $S^{n}$. Then $h$ is called a harmonic function on $U'$ if $\triangle_{s}h=0$.
1041: \end{definition}
1042:
1043: \ In much the same way as one would derive Green's Theorem in $R^{n}$ one now has:
1044: \begin{theorem}
1045: Suppose $U'$ is a domain on $S^{n}$ and $h:U'\rightarrow Cl_{n}$ is a harmonic function on $U'$. Suppose also that $S'$ is a smooth hypersurface lying in $U'$ and that $S'$ bounds a subdomain $V'$ of $U'$ and that $y'\in V'$. Then
1046: \[h(y')=\frac{1}{\omega_{n}}\int_{S'}(G_{s}(x,y')n(x)h(x)+H_{s}(x,y')n(x)D_{s}h(x))d\sigma'(x).\]
1047: \end{theorem}
1048:
1049: \ See \cite{lr} for more details.
1050:
1051: \begin{center}{\Large The Fourier Transform and Clifford Analysis}\end{center}
1052:
1053: \ Closely related to Hardy spaces is the Fourier transform. Here we will consider $R^{n}$ as divided into upper and lower half space, $R^{n+}$ and $R^{n-}$, respectively. Where $R^{n+}=\{x=x_{1}e_{1}+\ldots+x_{n}e_{n}:x_{n}>o\}$ and $R^{n-}=\{X=x_{1}e_{1}+\ldots+x_{n}e_{n}:x_{n}<0\}$. These two domains have $R^{n-1}=$ span$<e_{1},\ldots,e_{n-1}>$ as common boundary. As before we have that $L^{2}(R^{n-1})=H^{2}(R^{n+})\otimes H^{2}(R^{n-})$. Let us now consider a function $\psi\in L^{2}(R^{n-1})$. Then \[\psi(y)=(\frac{1}{2}\psi(y)+\frac{1}{\omega_{n}}PV\int_{R^{n-1}}G(x'-y)e_{n}\psi(x')dx^{n-1})\]
1054: \[+(\frac{1}{2}\psi(y)-\frac{1}{\omega_{n}}PV\int_{R^{n-1}}G(x'-y)e_{n}\psi(x')dx^{n-1})\]
1055: almost everywhere. Here $\frac{1}{2}\psi(y)+\frac{1}{\omega_{n}}PV\int_{R^{n-1}}G(x'-y)e_{n}\psi(x')dx^{n-1}$ is the nontangential limit of $\frac{1}{\omega_{n}}\int_{R^{n-1}}G(x'-y(t))e_{n}\psi(x')dx^{n-1}$ as $y(t)$ tends to $y$ nontangentially through a smooth path in upper half space, and $\frac{1}{2}\psi(y)-\frac{1}{\omega_{n}}PV\int_{R^{n-1}}G(x'-y)e_{n}\psi(x')dx^{n-1}$ is the nontangential limit of $\frac{-1}{\omega_{n}}\int_{R^{n-1}}G(x'-y(t))e_{n}\psi(x')dx^{n-1}$ as $y(t)$ tends nontangentially to $y$ through a smooth path in lower half space.
1056:
1057: \ Consider now the Fourier transform, ${\cal F}(\psi)=\hat{\psi}$, of $\psi$. In order to proceed we need to calculate
1058: \[{\cal F}(\frac{1}{2}\psi\pm\frac{1}{\omega_{n}}PV\int_{R^{n-1}}G(x'-y)e_{n}\psi(x')dx^{n-1}).\]
1059: In particular we need to determine ${\cal F}(\frac{1}{\omega_{n}}PV\int_{R^{n-1}}G(x'-y)e_{n}\psi(x')dx^{n-1})$. Following \cite{bla} it may be determined that this is $\frac{1}{2}i\frac{\zeta}{\|\zeta\|} e_{n}\hat{\psi}(\zeta)$ where $\zeta=\zeta_{1}e_{1}+\ldots+\zeta_{n-1}e_{n-1}$. So
1060: \[{\cal F}(\frac{1}{2}\psi\pm\frac{1}{\omega_{n}}\int_{R^{n-1}}G(x'-y)e_{n}\psi(x')d\sigma(x')=\frac{1}{2}(1\pm i\frac{\zeta}{\|\zeta\|} e_{n}).\]
1061: Now as observed in \cite{blabla}
1062: \[(\frac{1}{2}(1\pm i\frac{\zeta}{\|\zeta\|} e_{n})^{2}=\frac{1}{2}(1\pm i\frac{\zeta}{\|\zeta\|} e_{n})\]
1063: \[\frac{1}{2}(1\pm i\frac{\zeta}{\|\zeta\|} e_{n})\frac{1}{2}(1\mp i\frac{\zeta}{\|\zeta\|} e_{n})=0\]
1064: and
1065: \[i\zeta e_{n}\frac{1}{2}(1\pm i\frac{\zeta}{\|\zeta\|} e_{n})=\|\zeta\|\frac{1}{2}(1\pm i\frac{\zeta}{\|\zeta\|} e_{n}).\]
1066: Now take the Cauchy Kowalewska extension of $e^{i<x',\zeta>}$. We get
1067: \[\exp(i<x',\zeta>-ix_{n}e_{n}\zeta)\]
1068: defined on some neighbourhood in $R^{n}$ of $R^{n-1}$. Now consider
1069: \[\exp(i<x',\zeta>-ix_{n}e_{n}\zeta)\frac{1}{2}(1\pm i\frac{\zeta}{\|\zeta\|} e_{n}).\]
1070: This simplifies to
1071: \[e^{i<x',\zeta>+x_{n}\|\zeta\|}\frac{1}{2}(1+i\frac{\zeta}{\|\zeta\|} e_{n})\]
1072: if $x_{n}>0$ and to
1073: \[e^{i<x',\zeta>-x_{n}\|\zeta\|}\frac{1}{2}(1-i\frac{\zeta}{\|\zeta\|} e_{n})\]
1074: if $x_{n}<0$. The first of these series converges locally uniformly on upper half space while the second series converges locally uniformly on lower half space. We denote these two functions by $e_{\pm}(x,\zeta)$ respectively. So the integrals $\frac{1}{\omega_{n}}\int_{R^{n-1}}e_{\pm}(x,\zeta)\hat{\psi}(\zeta)d\sigma(\zeta)$ define left monogenic functions $\Psi_{\pm}(x)$. These left monogenic functions are defined on upper and lower half space respectively. Moreover $\Psi_{\pm}\in H^{2}(R^{n\pm})$ respectively, and $\Psi_{\pm}$ explicitly give the Hardy space decomposition of $\psi\in L^{2}(R^{n-1})$. This approach to the links between the Fourier transform and Clifford analysis were first carried out in \cite{s3} and later independently rediscovered and applied in \cite{lmq}.
1075:
1076: \ The integrals $\frac{1}{\omega_{n}}\int_{R^{n-1}}e_{\pm}(x,\zeta)d\zeta^{n-1}$ can be expressed in polar co-ordinates as an integral $\frac{1}{\omega_{n}}\int_{S^{n-2}}\int_{0}^{\infty}e^{ir<x',\zeta'>\pm x_{n}r}r^{n-2}\frac{1}{2}(1\pm\frac{\zeta}{\|\zeta\|}e_{n})drdS^{n-2}$, where $\zeta'=\frac{\zeta}{\|\zeta\|}$. In \cite{lmq} Chun Li observed that the integrals $\int_{0}^{\infty}e^{ir<x',\zeta'>\pm x_{n}r}\frac{1}{2}(1\pm \zeta' e_{n})r^{n-2}dr$ are Laplace transforms of the function $f(R)=R^{n-2}$. So the integral
1077: \[\int_{0}^{\infty}e^{ir<x',\zeta'>-x_{n}r}\frac{1}{2}(1-i\zeta' e_{n})dr\]
1078: evalutes to $(-i)^{n}(n-2)!(<x',\zeta'>+ix_{n})^{-n+1}$. Hence the integral
1079: \[\frac{1}{\omega_{n}}\int_{R^{n-1}}e_{+}(x,\zeta)d\zeta^{n-1}\]
1080: becomes
1081: \[\frac{1}{\omega_{n}}\int_{S^{n-2}}(-i)^{n}(n-2)!(<x',\zeta'>+ix_{n})^{-n+1}\frac{1}{2}(1+ie_{n}\zeta')dS^{n-2}.\]
1082:
1083: \ For any complex number $a+ib$ the product $(a+ib)\frac{1}{2}(1+ie_{n}\zeta')$ is equal to $(a-be_{n})\frac{1}{2}(1+ie_{n}\zeta')$. Thus the previous integral becomes
1084: \[\frac{1}{\omega_{n}}\int_{S^{n-2}}(-i)^{n}(<x',\zeta'>-x_{n}e_{n}\zeta')^{-n+1}\frac{1}{2}(1+ie_{n}\zeta')dS^{n-2}.\]
1085: The imaginary, or $iCl_{n}$ part of this integral is the integral of an odd function so when $n$ is even the integral becomes
1086: \[\frac{1}{\omega_{n}}\int_{S^{n-2}}\frac{(n-2)!e_{n}\zeta'}{2(<x',\zeta'>-x_{n}e_{n}\zeta')^{n-1}}dS^{n-2}\]
1087: and when $n$ is odd the integral becomes
1088: \[\frac{-1}{\omega_{n}}\int_{S^{n-2}}\frac{(n-2)!}{(<x',\zeta'>-x_{n}e_{n}\zeta')^{n-1}}dS^{n-2}.\]
1089: These integrals are the plane wave decompositions of the Cauchy kernel for upper half space described by Sommen in \cite{s4}.
1090: \ It should be noted that while introducing the Fourier transform and exploring some of its links with Clifford analysis we have also been forced to complexify the Clifford algebra $Cl_{n}$ so that we now work with the complex Clifford algebra $Cl_{n}(C)$. Further the functions $\frac{1}{2}(1\pm i\frac{\zeta}{\|\zeta\|e_{n}})$ can be seen as defined on spheres lying in the null cone $\{x_{n}e_{n}+iw': w'\in R^{n-1}$ and $x_{n}^{2}-\|w'\|^{2}=0\}$. This leads naturally to the question: what domains in $C^{n}$ do the functions $e_{\pm}(x,\zeta)$ extend to? Here we are replacing the real vector variable $x\in R^{n}$ by a complex vector variable $\underline{z}=z_{1}e_{1}+\ldots+z_{n}e_{n}\in C^{n}$, where $z_{1},\ldots,z_{n}\in C$. On placing $\underline{z}=x+iy$ where $x$ and $y$ are real vector variables, it may be noted that the term is $e^{-<\zeta,y'>-x_{n}\|\zeta\|}$ is well defined for $x_{n}\|\zeta\|>|<\zeta,y'>|$. So in this case $iy'+x_{n}e_{n}$ varies over the interior of the forward null cone $\{iy'+x_{n}e_{n}:x_{n}>0$ and $x_{n}>\|y'\|\}$. So $e_{+}(\underline{z},\zeta)$ is well defined for each $\underline{z}=x+iy=x'+iy'+(x_{n}+iy_{n})e_{n}\in C^{n}$ where $x'\in R^{n-1}$, $y_{n}\in R$, $x_{n}>0$ and $\|y'\|<x_{n}$. Similarly $e_{-}(x,\zeta)$ holomorphically extends to $\{\underline{z}=x'+iy'+(x_{n}+iy_{n})e_{n}:x'\in R^{n-1}$, $x_{n}<0$, $y_{n}\in R$ and $\|y'\|<|x_{n}|\}$. We denote these domains by $C^{\pm}$ respectively. It should be noted that $\Psi^{\pm}$ holomorphically continue to $C^{\pm}$ respectively. We denote these holomorphic continuations of $\Psi^{\pm}$ by $\Psi'^{\pm}$. The domains $C^{\pm}$ are examples of tube domains.
1091:
1092: \begin{center}{\Large Complex Clifford Analysis}\end{center}
1093:
1094: \ In the previous section we ended by showing that for any left Clifford holomorphic function $f\in H^{2}(R^{n,+})$ may be holomorphically continued to a function $f^{\dagger}$ defied on a tube domain $C^{+}$ in $C^{n}$. In this section we will briefly show how this type of holomorphic continuation happens for all Clifford holomorphic functions defined on a domain $U\subset R^{n}$. First let $S$ be a compact smooth hypersurface lying in $U$ and suppose that $S$ bounds a subdomain $V$ of $U$. Cauchy's integral formula gives us
1095: \[f(y)=\frac{1}{\omega_{n}}\int_{S}G(x-y)n(x)f(x)d\sigma(x)\]
1096: for each $y\in V$. Let us now consider complexifying the Cauchy kernel. The function $G(x)$ holomorphically continues to $\frac{\underline{z}}{(\underline{z}^{2})^{\frac{n}{2}}}$. In even dimensions this is a well defined function on $C^{n}\backslash N(0)$ where $N(0)=\{\underline{z}\in C^{n}:\underline{z}^{2}=0\}$. In odd dimensions this lifts to a well defined function on complex $n$-dimensional Riemann surface double covering $C^{n}\backslash N(0)$. Though things work out well in odd dimensions we will for ease just work with the cases where $n$ is even. In holomorphically extending $G(x-y)$ in the variable $y$ we obtain a function $G^{\dagger}(x-\underline{z})=\frac{x-\underline{z}}{((x-\underline{z})^{2})^{\frac{n}{2}}}$. This function is well defined on $C^{n}\backslash N(x)$ where $N(x)=\{\underline{z}\in C^{n}:(x-\underline{z})^{2}=0\}$. So the integral $\frac{1}{\omega_{n}}\int_{S}G^{\dagger}(x-\underline{z})n(x)f(x)d\sigma(x)$ is well defined provided $\underline{z}$ is not in $N(x)$ for any $x\in S$. It may be determined that the set $C^{n}\backslash\cup_{x\in S}N(x)$ is an open set in $C^{n}$. We shall take the component of this open set which contains $V$. We shall denote this connected open subset of $C^{n}$ by $V^{\dagger}$. It now follows that the left Clifford holomorphic function $f(y)$ now has a holomorphic extension $f^{\dagger}(\underline{z})$ to $V^{\dagger}$. Furthermore this function is given by the integral formula
1097: \[f^{\dagger}(\underline{z})=\frac{1}{\omega_{n}}\int_{S}G^{\dagger}(x-\underline{z})n(x)f(x)d\sigma(x).\]
1098:
1099: \ Furthermore the holomorphic function $f^{\dagger}$ is now a solution to the complex Dirac equation $D^{\dagger}f^{\dagger}=0$, where $D^{\dagger}=\Sigma_{j=1}^{n}e_{j}\frac{\partial}{\partial z_{j}}$. By letting allowing the hypersurface to deform and move out to include more of $U$ in its interior we see that $f^{\dagger}$ is a well defined holomorphic function on $U^{\dagger}$ where $U^{\dagger}$ is the component of $C^{n}\backslash\cup_{x\in cl(U)\backslash U}N(x)$ which contains $U$. In the special cases where $U$ is either of $R^{n,\pm}$ then $U^{\dagger}=C^{\pm}$. See \cite{r1,r2,r3} for more details.
1100:
1101: \begin{thebibliography}{999}
1102: \bibitem{a} L. V. Ahlfors, {\em M\"{o}bius transformations in $R^{n}$ expressed
1103: through
1104: $2\times 2$ matrices of Clifford numbers}, Complex Variables, 5, 1986,
1105: 215-224.
1106: \bibitem{abs} M. F. Atiyah, R. Bott and A. Shapiro {\em Clifford modules},
1107: Topology, 3, 1965, 3-38.
1108: \bibitem{b} B. Bojarski {\em Conformally covariant differential operators},
1109: Proceedings, XX th Iranian Math. Congress, Tehran, 1989.
1110: \bibitem{bbw} B. Booss-Bavnbek and K. Wojciechowski, {\em Elliptic Boundary
1111: Problems for Dirac Operators}, Birkhauser, Basel, 1993.
1112: \bibitem{bds} F. Brackx, R. Delanghe and F. Sommen {\em Clifford Analysis},
1113: Pitman, London, 1982.
1114: \bibitem{c} D. Calderbank, {\em Clifford analysis for Dirac operators on manifolds
1115: with boundary}, Max-Plank-Institut f\"{u}r Mathematik preprint, Bonn, 1996.
1116: \bibitem{cmm} R. Coifman, A. McIntosh and Y. Meyer, {\em L'int\'{e}grale de Cauchy d\'{e}finit un op\'{e}rator born\'{e} sur $L^{2}$ pour les courbes lipschitziennes}, Annals of Mathematics, 116, 1982, 361-387.
1117: \bibitem{cnm} J. Cnops and H. Malonek {\em An introduction to Clifford analysis},
1118: Univ. Coimbra, Coimbra, 1995.
1119: \bibitem{d} R. Delanghe, {\em On regular-analytic functions with values in a
1120: Clifford algebra}, Math. Ann., 185, 1970, 91-111.
1121: \bibitem{dss} R. Delanghe, F. Sommen and V. Soucek, {\em Clifford Algebra and
1122: Spinor-Valued Functions}, Kluwer, Dordrecht, 1992.
1123: \bibitem{di} A. C. Dixon, {\em On the Newtonian potential}, Quarterly Journal of
1124: Mathematics, 35, 1904, 283-296.
1125: \bibitem{f} R. Fueter, {\em Die Funktionentheorie der Differentialgleichungen
1126: $\triangle u=0$ und $\triangle \triangle u=0$ mit
1127: Vier Reallen Variablen}, Commentarii Mathematici Helvetici 7, 1934-1935,
1128: 307-330.
1129: \bibitem{gs} K. Guerlebeck and W. Sproessig, {\em Quaternionic and Clifford Calculus
1130: for Physicists and Engineers}, Wiley, Chichester, 1998.
1131: \bibitem{gm} J. Gilbert and M. A. M. Murray, {\em Clifford Algebras and Dirac
1132: Operators in Harmonic Analysis}, CUP, Cambridge, 1991.
1133: \bibitem{h} D. Hestenes, {\em Multivector functions}, J. Math. Anal. Appl., 24,
1134: 1968, 467-473.
1135: \bibitem{i} V. Iftimie, {\em Fonctions hypercomplexes}, Bull. Math. de la Soc. Sci.
1136: Math. de la R. S. de Roumanie 9, 1965, 279-332.
1137: \bibitem{k}R. S. Krausshar, {\em Automorphic forms in Clifford analysis}, Complex Variables, 47, 2002, 417-440.
1138: \bibitem{kr1} R. S. Krausshar and J. Ryan, {\em Clifford and harmonic analysis on cylinders and tori}, to appear.
1139: \bibitem{kr2} R. S. Krausshar and J. Ryan, {\em Some conformally flat spin manifolds, Dirac operators and automorphic forms}, to appear.
1140: \bibitem{ks} V. Kravchenko and M. Shapiro, {\em Integral Representations for Spatial
1141: Models of Mathematical Physics}, Pitman Research Notes in Mathematics, London,
1142: no 351, 1996.
1143: \bibitem{lms} C. Li, A. McIntosh and S. Semmes, {\em Convolution singular integrals
1144: on Lipschitz surfaces}, J. of the AMS, 5, 1992, 455-481.
1145: \bibitem{lmq} C. Li, A. McIntosh and T. Qian, {\em Clifford algebras, Fourier
1146: transforms, and singular convolution operators on Lipschitz surfaces},
1147: Revista Mathematica Iberoamericana, 10, 1994, 665-721.
1148: \bibitem{lg} T. E. Littlewood and C. D. Gay, {\em Analytic spinor fields}, Proc.
1149: Roy. Soc., A313, 1969, 491-507.
1150: \bibitem{l} J. Liouville {\em Extension au cas des trois dimensions de la
1151: question du trace geographique. Applications de l'analyse a geometrie}, G.
1152: monge, Paris, 1850, 609-616.
1153: \bibitem{lr} H. Liu and J. Ryan {\em Clifford analysis techniques for spherical pde}, Journal of Fourier Analysis and its Applications, 8, 2002, 535-564.
1154: \bibitem{mc} A. McIntosh, {\em Clifford algebras, Fourier theory, singular integrals, and harmonic functions on Lipschitz domains}, Clifford Algebras in Analysis and Related Topics, ed J. Ryan, 33-88, CRC Press, Boca Raton, 1996.
1155: \bibitem{m1} M. Mitrea, {\em Clifford Wavelets, Singular Integrals and Hardy Spaces},
1156: Lecture Notes in Mathematics, Springer-Verlag, Heidelberg, No 1575, 1994.
1157: \bibitem{m2} M. Mitrea, {\em Generalized Dirac operators on non-smooth manifolds and
1158: Maxwell's equations}, Journal of Fourier Analysis and its Applications, 7, 2001, 207-256.
1159: \bibitem{mt} Gr. C. Moisil and N. Theodorescu, {\em Fonctions holomorphes dans
1160: l'espace}, Mathematica (Cluj), 5, 1931, 142-159.
1161: \bibitem{m} M. Murray, {\em The cauchy integral, Calderon commutation, and conjugation of singular integrals in $R^{n}$}, Trans. of the AMS, 298, 1985, 497-518.
1162: \bibitem{o} E. Obolashvili, {\em Partial Differential Equations in Clifford Analysis},
1163: Pitman Monographs and Surveys in Pure and Applied Mathematics no 96, Harlow,
1164: 1998.
1165: \bibitem{q1} T. Qian, {\em Singular integrals on star shaped Lipschitz surfaces in the quaternionic space}, Math Ann, 310, 1998, 601-630.
1166: \bibitem{q2} T. Qian, {\em Fourier analysis on starlike Lipschitz surfaces}, Journal of Functional Analysis, 183, 2001, 370-412.
1167: \bibitem{r1} J. Ryan, {\em Cells of Harmonicity and generalized Cauchy integral
1168: formulae}, Proc. London Math. Soc, 60, 1990, 295-318.
1169: \bibitem{r2} J. Ryan, {\em Intrinsic Dirac operators in $C^{n}$}, Advances in
1170: Mathematics, 118, 1996, 99-133.
1171: \bibitem{r3} J. Ryan, {\em Complex Clifford analysis and domains of holomorphy},
1172: Journal of the Australian Mathematical Society, Series A, 48, 1990, 413-433.
1173: \bibitem{r4} J. Ryan, Ed {\em Clifford Algebras in Analysis and Related Topics},
1174: Studies in Advanced Mathematics, CRC Press, Boca Raton, 1995.
1175: \bibitem{r5} J. Ryan, {\em Basic Clifford analysis}, Cubo Matem\/{a}tica
1176: Educacional, 2, 2000, 226-256.
1177: \bibitem{r6} J. Ryan, {\em Dirac operators on spheres and hyperbolae}, Bolletin de la Sociedad Matematica a Mexicana, 3, 1996, 255-270.
1178: \bibitem{sc} M. Sce {\em Osservaziono sulle serie di potenzi nei moduli
1179: quadratici}, Atti. Acad. Nax. Lincei Rend Sci Fis Mat. Nat, 23, 1957,
1180: 220-225.
1181: \bibitem{se} S. Semmes, {\em Some remarks concerning integrals of curvature on curves and surfaces}, to appear.
1182: \bibitem{s1} F. Sommen {\em Spherical monogenics and analytic functionals on the
1183: unit sphere}, Tokyo J. Math., 4, 1981, 427-456.
1184: \bibitem{s2} F. Sommen {\em A product and an exponential function in hypercomplex
1185: function theory}, Appl. Anal., 12, 1981, 13-26.
1186: \bibitem{s3} F. Sommen, {\em Microfunctions with values in a Clifford algebra II}, Scientific Papers of the College of Arts and Sciences, The University of Tokyo, 36, 1986, 15-37.
1187: \bibitem{s4} F. Sommen, {\em Plane wave decomposition of monogenic functions}, Ann. Polon. Math., 49, 1988, 101-114.
1188: \bibitem{sw} E. M. Stein and G. Weiss {\em Introduction to Fourier Analysis on
1189: Euclidean Space}, Princeton University Press, Princeton, 1971.
1190: \bibitem{su} A. Sudbery {\em Quaternionic analysis}, Math. Proc. of the Cambridge
1191: Philosophical Soc., 85, 1979, 199-225.
1192: \bibitem{v} P. Van Lancker, {\em Clifford analysis on the sphere}, Clifford Algebras and their Applications in Mathematical Physics, V. Diedrich et al, editors, Kluwer, Dordrecht, 1998, 201-215.
1193: \bibitem{w} S. Wu, {\em Well-posedness in Sobolev spaces of the full water wave problem in $3$-D}, Journal of the AMS, 12, 1999, 445-495.
1194: \end{thebibliography}
1195: \end{document}
1196: