math0304175/hh.tex
1: \input liemacs.tex
2: 
3: 
4: 
5: \def\b{\mathop{\bf b}\nolimits}
6: 
7: 
8: 
9: \def\c{\mathop{\bf c}\nolimits}
10: 
11: 
12: 
13: \def\cH{\mathop{{\bf c}_{\hbox{\fiverm G/H}}}\nolimits}
14: 
15: 
16: 
17: \def\e{\mathop{\bf e}\nolimits}
18: 
19: 
20: 
21: \def\hfH{\hbox{\fiverm H}}
22: 
23: 
24: 
25: \def\hfK{\hbox{\fiverm K}}
26: 
27: 
28: 
29: \def\ssmapright#1{\smash{\mathop{\ssarr}\limits^{#1}}}
30: 
31: \input epsf.tex 
32:  
33:  
34:  
35:  
36:  
37:  
38:  
39: \def\bs{\backslash} 
40:  
41: \def\addots{\mathinner{\mkern1mu\raise1pt\vbox{\kern7pt\hbox{.}}\mkern2mu 
42: \raise4pt\hbox{.}\mkern2mu\raise7pt\hbox{.}\mkern1mu}} 
43:  
44: \pageno=1 
45: \def\up#1{\leavevmode \raise.16ex\hbox{#1}} 
46: \font\smallheadfont=cmr8 at 8truept 
47: \font\smallbfheadfont=cmbx8 at 8truept 
48: %\font\largeheadfont=cmdunh10 at 14.4truept 
49: \font\headfont=cmdunh10 at 12truept 
50: \chardef\ss="19 
51: \def\3{\ss} 
52:  
53: \def\firstpage{\nin 
54: {\obeylines \parindent 0pt } 
55: \vskip2cm 
56: \centerline {\bfone \title} 
57: \ssk 
58: \centerline {\bfone \titletwo} 
59: \gsk 
60: \centerline{\bf\author} 
61: \vskip1.5cm \rm} 
62:  
63:  
64: \def\title{Holomorphic $H$-spherical distribution vectors} 
65: \def\titletwo{in principial series representations} 
66:  
67: \def\author{Simon Gindikin, Bernhard Kr\"otz and 
68: Gestur \'Olafsson} 
69:  
70: \footnote{}{SG was supported in part  by  NSF-grant DMS-0070816} 
71: \footnote{}{BK was supported in part  by NSF-grant DMS-0097314} 
72: \footnote{}{G\'O was supported in part by NSF-grant  DMS-0070607 and DMS-0139783} 
73:  
74:  
75: \def\address 
76: {Simon Gindikin 
77:  
78: Department of Mathematics 
79:  
80: Rutgers University 
81:  
82: New Brunswick, NJ 08903 
83:  
84: USA 
85:  
86: {\tt gindikin@math.rutgers.edu} 
87:  
88: \bsk 
89: \bsk 
90:  
91: Gestur \'Olafsson 
92:  
93: Louisiana State University 
94:  
95: Department of Mathematics 
96:  
97: Baton Rouge, LA 70803 
98:  
99: USA 
100:  
101: {\tt olafsson@math.lsu.edu} 
102:  
103: } 
104:  
105: \def\addresstwo 
106: {Bernhard Kr\"otz 
107:  
108: Department of Mathematics 
109: University of Oregon  
110: Eugene Or 97403-1221  
111: USA  
112:  
113: {\tt kroetz@math.uoregon.edu} 
114: } 
115:  
116:  
117: \firstpage 
118:  
119:  
120: \subheadline{Abstract} 
121: \noindent 
122: Let $G/H$ be a semisimple symmetric space. 
123: The main tool to embed a principal series 
124: representation of  $G$ into $L^2(G/H)$ are 
125: the $H$-invariant distribution vectors. 
126: If $G/H$ is a non-compactly 
127: causal symmetric space, then $G/H$ can be realized as a 
128: boundary component of the complex crown $\Xi$. 
129: In this article we construct a minimal $G$-invariant 
130: subdomain $\Xi_H$ of $\Xi$ with 
131: $G/H$ as Shilov boundary. Let $\pi$ be 
132: a spherical principal series representation 
133: of $G$. We show 
134: that the space of $H$-invariant 
135: distribution vectors of $\pi$, which admit a 
136: holomorphic extension to $\Xi_H$,  
137: is one dimensional. Furthermore we give a spectral definition of 
138: a Hardy space corresponding to those distribution 
139: vectors. In particular we achieve a geometric 
140: realization of a multiplicity free subspace of $L^2(G/H)_{\rm mc}$ 
141: in a space of holomorphic functions. 
142:  
143: \sectionheadline{Introduction} 
144:  
145: \noindent 
146: Holomorphic extensions and boundary value maps have been 
147: valuable tools to solve problems in representation theory 
148: and harmonic analysis on {\it real} symmetric 
149: spaces. Two of the best known constructions  
150: are Hardy spaces with their boundary value maps and  
151: Cauchy-Szeg\"o-kernels, 
152: and Fock space constructions with their corresponding 
153: Segal-Barmann transform. It is in this flavour that we establish a 
154: correspondence between eigenfunctions on a Riemannian symmetric spaces $X=G/K$ 
155: and  a  non-compactly causal (NCC) symmetric spaces $Y=G/H$ in  this paper. 
156: In particular 
157: we, via analytic continuation, relate a {\it spherical function} $\phi_\lambda$ on 
158: $G/K$ to a {\it holomorphic} $H$-invariant distribution on $G/H$. 
159:  
160:  
161: \par Let us explain our results in more detail. On the geometric level we construct a certain minimal 
162: $G$-invariant Stein domain $\Xi_H\subeq X_\C=G_\C/K_\C$ with 
163: the following properties: The Riemannian symmetric space $X$ is embedded into $\Xi_H$ as 
164: a totally real submanifold and the affine non-compactly causal space $Y$ is isomorphic to the 
165: distinguished (Shilov) boundary of $\Xi_H$. The details 
166: of this construction are carried out in Section 1. 
167:  
168: \par The minimal tube $\Xi_H$ is a subdomain of the complex 
169: crown $\Xi\subeq X_\C$ of $X$ -- an object first introduced 
170: in [AG90] which became subject of intense study over the last 
171: few years. A consequence is that all $\D(X)$-eigenfunctions 
172: on $X$ extend holomorphically to $\Xi_H$ [KS01b]. Another key fact 
173: is that $\D(X)\simeq \D (Y)$. Thus by taking limits on the boundary $Y$ we  obtain a realization 
174: of the $\D(X)$-eigenfunctions on $X$ as $\D(Y)$-eigenfunctions on $Y$. 
175: Conversely, eigenfunctions on $Y$ which holomorphically extend to $\Xi_H$ 
176: yield by restriction eigenfunctions on $X$. 
177:  
178:  
179: \par It seems to us that the above mentioned transition between eigenfunctions 
180: on $X$ and $Y$ is most efficiently described using the 
181: techniques from representation theory.  To fix the notation let $(\pi,{\cal H})$ 
182: denote an admissible Hilbert representation of $G$ with 
183: finite length. We write ${\cal H}^K$ for the space of $K$-fixed 
184: vectors and $({\cal H}^{-\infty})^H$ for the space of $H$-fixed 
185: distribution vectors of $\pi$. Using the method of analytic 
186: continuation of representations as developed in [KS01a] we 
187: establish a bijection 
188: $${\cal H}^K\ssmapright{\simeq} ({\cal H}^{-\infty})_{\rm hol}^H , \ \ 
189: v_{\hbox{\fiverm K}}\mapsto v_{\hbox{\fiverm H}}$$ 
190: where $({\cal H}^{-\infty})_{\rm hol}^H\subeq ({\cal H}^{-\infty})^H$ 
191: denotes the subspace characterized through the property that 
192: associated matrix coefficients on $Y$ extend holomorphically 
193: to $\Xi_H$ (cf.\ Theorem 2.1.3, Theorem 2.2.4). This bijection and various 
194: ramifications are the subject proper of Section 2. 
195:  
196: \par In Section 3 we give an application 
197: of our theory towards the geometric realization of the most-continuous spectrum 
198: $L^2(Y)_{\rm mc}$ of $L^2(Y)$. First progress in this direction was achieved in [GK\'O01]. 
199: There, for the cases where $\Xi=\Xi_H$,  we defined a Hardy space 
200: ${\cal H}^2(\Xi)$ on $\Xi$ and showed that there is an isometric boundary value mapping 
201: realizing ${\cal H}^2(\Xi)$ as a 
202: multiplicity one subspace of $L^2(Y)_{\rm mc}$ of full spectrum. 
203: It was an open problem how to define 
204: Hardy spaces for general NCC symmetric spaces $Y$ and to determine  the Plancherel 
205: measure explicitely. We solve this problem by giving a spectral definition of the 
206: Hardy space, i.e., we take the conjectured Plancherel measure and define a Hilbert 
207: space of holomorphic functions  ${\cal H}^2(\Xi_H)$ on $\Xi_H$. The identification 
208: of ${\cal H}^2(\Xi_H)$ as  a Hardy space then follows by establishing 
209: an isometric boundary value mapping $b\: {\cal H}^2(\Xi_H)\into L^2(G/H)_{\rm mc}$. 
210: In particular we achieve a geometric realization of a multiplicity free subspace of $L^2(Y)_{\rm mc}$ 
211: in holomorphic functions. 
212:  
213: \msk It is our pleasure to thank the referee for his very careful  work. He pointed out many  
214: inaccuracies and made useful remarks on the presentation of the paper.   
215:  
216:  
217:  
218: \sectionheadline{1. Complex crowns and the domains $\Xi_H$} 
219:  
220: \noindent 
221: The purpose of this section is to give the geometric preliminaries  
222: of the analytical constructions to come. Our two main  
223: players are a Riemannian symmetric space $G/K$ on the one hand side  
224: and on the other hand a non-compactly causal symmetric space $G/H$.  
225: The two symmetric spaces $G/K$ and $G/H$ are ``connected'' through  
226: a complex $G$-invariant domain $\Xi_H\subeq G_\C/K_\C$ in the  
227: following way: $G/K$ is a totally real submanifold and  
228: $G/H$ constitutes the distinguished (Shilov) boundary of $\Xi_H$.  
229: The domain $\Xi_H$ constructed in this section is an appropriate  
230: subdomain of the complex crown $\Xi$ of the Riemannian symmetric space 
231: $G/K$. 
232:  
233: \par This section is organized as follows. We start by briefly recalling  
234: the defintion and some standard features of non-compactly causal  
235: symmetric spaces . Then we switch to complex crowns $\Xi$  
236: and summarize the main results of [GK02a] on how 
237: to realize $G/H$ in the distinguished boundary of $\Xi$. Finally  
238: we give the construction of the domain $\Xi_H$.  
239:  
240:  
241:  
242: \subheadline{1.1. Non-compactly causal symmetric spaces (NCC)} 
243:  
244: \noindent 
245: In this subsection we recall some facts on non-compactly causal symmetric spaces. 
246: The material is standard and can be found in the monograph 
247: [H\'O96]. 
248:  
249: \ssk  
250: Let $G$ be a connected semisimple  Lie group and $\g$ be its Lie algebra. 
251: Denote by $\g_\C=\g\otimes_\R\C$ the complexification of $\g$. 
252: If $\h$ is a subalgebra of $\g$,  then we denote by $\h_\C$ the  complex subalgebra 
253: of $\g_\C$ generated by $\h$. 
254: We assume that  $G$ is contained 
255: in a complex group $G_\C$ with Lie algebra $\g_\C$. 
256:  
257: \par If $\sigma\:G \to G$ is 
258: an involution, then, by abuse of notation, we use the same letter 
259: for the derived involution on the Lie algebra $\g$ and its complex linear  
260: extension to  $\g_\C$. 
261: \par Let $\theta\:G\to G$ be a Cartan involution and denote by $K<G$ the corresponding 
262: maximal compact subgroup. Let $\k=\{X\in\g\: \theta(X)=X\}$ 
263: and $\p =\{X\in \g\:  \theta (X)=-X\}$. Then 
264: $\k$ is the Lie algebra of $K$. 
265:  
266: \par In the sequel we let $\tau$ denote an involution on $G$ which we may assume to commute 
267: with $\theta$. Let $G^\tau :=\{g\in G\: \tau (g)=g\}$ and 
268: let $H$ be an open subgroup of $G^\tau$. Then $G/H$ is called  a {\it symmetric 
269: spaces}.  On the Lie algebra level 
270: $\tau$ induces a splitting $\g=\h+\q$ with $\h$ the $+1$ and $\q$ the $-1$-eigenspace of 
271: $\tau$. Notice that $\h$ is the Lie 
272: algebra of $H$. The pair $(\g,\h)$ is called a {\it symmetric pair}. We have, as $\theta$ and $\tau$ commute: 
273: $$\eqalign{ \g &=\k+ \p\cr 
274: &=\h+ \q\cr 
275: &=\k\cap \h +\k\cap \q + \p\cap \h + \p\cap \q 
276: } 
277: $$ 
278: \par 
279: The symmetric pair 
280: $(\g ,\h)$ is called {\it irreducible} if the only 
281: $\tau$-invariant ideals in $\g$ are the trivial ones, 
282: $\{0\}$ and $\g$. In this case either $\g$ is 
283: simple or $\g \simeq \h\oplus \h$, with $\h$ simple, and $\tau (X,Y)= 
284: (Y,X)$ the flip. We say that the symmetric space $G/H$ is {\it irreducible} 
285: if the corresponding symmetric pair $(\g ,\h)$ is irreducible. 
286:  
287: \par Let $\emptyset \not= C\subseteq \g$ be an open subset of $\g$. Then 
288: $C$ is said to be {\it hyperbolic} if for all $X\in C$ the map ${\rm ad}(X):\g \to \g$ is 
289: semisimple with real eigenvalues. 
290:  
291: \Definition 1.1.1. {\bf (NCC)} Assume that $G/H$ is an irreducible symmetric space. Then 
292: the following two conditions are equivalent: 
293:  
294: \item{(a)} There exists a non-empty $H$-invariant open hyperbolic convex cone $C 
295: \subseteq \q$ which contains  no affine lines; 
296:  
297: \item{(b)} There exists an element $T_0\in \q\cap \p$, $T_0\not= 0$, 
298: which is fixed by $H\cap K$. 
299:  
300: If one of those equivalent conditions are satisfied, then $G/H$ is called 
301: {\it non-compactly causal}, or {\it NCC} for short. 
302: \qed 
303:  
304: \Remark 1.1.2. (a) The element $T_0$ in Definition 1.1.1 is 
305: unique up to multiplication by scalar. We can normalize 
306: $T_0$ such that ${\rm ad}(T_0)$ has spectrum $\{0,1,-1\}$. The eigenspace 
307: corresponding to $0$ is exactly $\g^{\theta\tau}=\k\cap\h +\p\cap \q$. 
308: \par\nin (b) If $G/H$ is NCC and $\a\subset \p\cap \q$ is maximal 
309: abelian, then $T_0\in \a$ by (a). Hence, again by (a), it follows 
310: that $\a$ is also maximal abelian in $\p$ and in $\q$. 
311: \par\nin (c) Let $T_0$ be as above. Then the interior of the convex hull of $\R^+\Ad (H)T_0$ 
312: is a minimal, $H$-invariant open hyperbolic convex cone in $\q$. 
313: \par\nin (d) All the NCC pairs $(\g, \h)$ are classified and we refer 
314: to [H\'O96, Th.\ 3.2.8] for the complete list.\qed 
315:  
316:  
317: \subheadline{1.2. The complex crown of a Riemannian symmetric space} 
318:  
319: \noindent 
320: The NCC spaces are exactly the affine symmetric spaces that can be 
321: realized as a symmetric subspace in the distinguished boundary 
322: of the complex crown $\Xi$ of the Riemannian symmetric space 
323: $G/K$. We will therefore recall some basic facts about 
324: $\Xi$. We refer to [GK02a] and [GK02b] as a standard source.  
325:  
326:  
327: Let the notation be as in Subsection 1.1. Let $\a\subeq \p$ be a maximal 
328: abelian subalgebra. For $\alpha \in \a^*$ let 
329: $\g^\alpha=\{ X\in \g\: (\forall H\in \a)\ [H,X]=\alpha(H)X\}$ and 
330: let $\Sigma:=\{\alpha\in \a^*\:\alpha\not=0,\g^\alpha\not=\{0\}\}$ be 
331: the corresponding set of restricted roots. 
332:  
333:  
334: \par Following [AG90] we define a bounded convex subset of $\a$ by 
335: $$\Omega=\{ X\in \a\: (\forall \alpha\in \Sigma)\ |\alpha(X)|<{\pi\over 2}\}\ .$$ 
336: Denote by $K_\C$ the analytic subgroup of $G_\C$ with Lie algebra $\k_\C$. 
337: Then we define a $G-K_\C$ double coset domain in $G_\C$ by 
338:  
339: $$\tilde \Xi=G\exp(i\Omega)K_\C$$ 
340: and recall that $\tilde\Xi$ is open in $G_\C$ [KS01a]. In particular the domain 
341: $$\Xi=\tilde\Xi/K_\C$$ 
342: is an open $G$-invariant subset of  $G_\C/K_\C$ containing $G/K$ as a totally 
343: real submanifold. We refer to $\Xi$ as the {\it complex crown} of the 
344: Riemannian symmetric space $G/K$ (cf.\ [AG90]). Observe that the definition 
345: of $\Xi$ and $\tilde \Xi$ is independent of the choice of $\a\subeq \p$. 
346: For a subset $\omega\subeq \a$ we define a tube domain in $A_\C=\exp(\a_\C)$ 
347: by 
348: $$T( \omega) =A\exp(i\omega)$$ 
349: and notice that $T (2\Omega)$ is biholomorphic to $\a+i 2\Omega$ via the 
350: exponential map. 
351:  
352: \msk Fix a positive system $\Sigma^+$ of $\Sigma$ and define a subalgebra 
353: $\n$ of $\g$ by 
354: $$\n=\bigoplus_{\alpha\in \Sigma^+} \g^\alpha\ .$$ 
355: Write $N_\C$ for the analytic subgroup of $G_\C$ with Lie algebra $\n_\C$. 
356: Then it follows from [GK02b] that 
357: $$\tilde\Xi\subeq N_\C T(\Omega) K_\C\leqno(1.2.1)$$ 
358: or even more precisely 
359: $$\tilde\Xi=\left[\bigcap_{g\in G} g  N_\C T(\Omega) K_\C \right]_0\leqno (1.2.2)$$ 
360: where the subscript ${}_0$ denotes the connected component of $[\cdot]$ containing $G$. 
361:  
362: \subheadline{1.3. The distinguished boundary of $\Xi$} 
363:  
364: \noindent 
365: Write $\oline \Xi$ and $\partial \Xi$ for the closure respectively the boundary  of $\Xi$ in 
366: $G_\C/K_\C$. The {\it distinguished boundary} of $\Xi$ is a certain finite 
367: union of $G$-orbits in $\partial \Xi$ which features many properties of 
368: a Shilov boundary. It was introduced and investigated in [GK02a] 
369: and the objective of this subsection is to recall its definition and basic 
370: properties. 
371:  
372:  
373: \par  Let 
374: ${\cal W}=N_K(\a)/ Z_K(\a)$ be the Weyl group of $\a$ in $G$. 
375: Write $\oline\Omega$ for the closure of $\Omega$ and notice that 
376: $\oline\Omega$ is a ${\cal W}$-invariant compact convex set. Denote by 
377: $\partial_e\Omega$ the set of extreme points of $\oline \Omega$. Then 
378: there exists $X_1,\ldots ,X_n\in\partial_e\Omega$ such 
379: that 
380: $$\partial_e\Omega={\cal W}(X_1)\amalg\ldots\amalg{\cal W}(X_n)\ .$$ 
381: Define the {\it distinguished boundary}  of $\Xi$ in $G_\C/K_\C$ 
382: by 
383: $$\partial_d\Xi\:=G\exp(i\partial_e\Omega)K_\C/K_\C\ .$$ 
384: We refer to [GK02a] for detailed information about $\partial_d\Xi$ 
385: and recall here only the facts that we need. 
386:  
387: \par For $1\leq j\leq n$  let 
388: $z_j\:=\exp(iX_j)K_\C \in \partial_d\Xi$. If $G_\C$ is not simply connected it 
389: can happen that $G(z_j)=G(z_k)$ for some $j\neq k$. But after relabelling 
390: the $z_j$ we can assume that there is an $m\leq n$ such that 
391: $G(z_j)\not=G(z_k)$ for $1\le j,k\le m$, $j\not= k$, and 
392: $$\partial_d\Xi= G(z_1)\amalg\ldots\amalg G(z_m)\ .$$ 
393: Denote by $H_j$  the isotropy subgroup of $G$ in $z_j$. Then as $G$-spaces we have 
394:  
395: $$\partial_d\Xi= G/H_1\amalg\ldots \amalg G/ H_m \ .$$ 
396: As a consequence of the complete classification of $\partial_d\Xi$ in 
397: [GK02a] we obtain the following fact. 
398:  
399:  
400: \Proposition 1.3.1. For the distinguished boundary  
401: $\partial_d\Xi$ of $\Xi$  the following assertions hold: 
402: \item{(i)} If one of the boundary components 
403: $G/H_j$ of $\partial_d\Xi$ is a symmetric space, then it is a 
404: non-compactly causal symmetric space.  
405: \item{(ii)} Every 
406: non-compactly causal symmetric space of the form $G/H$  
407: is locally isomorphic to a $G$-orbit in the distinguished boundary 
408: of $\partial_d\Xi$ of $\Xi$. \qed 
409:  
410:  
411:  
412: \subheadline{1.4. The domains $\Xi_H$} 
413:  
414: \noindent 
415: We keep the notation from Subsections 1.2 - 1.3. {}From now on we  
416: fix an element $X_{\hbox{\fiverm H}}\in \partial_e\Omega$ 
417: and set $x_{\hbox{\fiverm H}}=\exp(iX_{\hbox{\fiverm H}})$, $z_{\hbox{\fiverm H}}= 
418: x_{\hbox{\fiverm H}}K_\C$. As the notation suggests we denote by $H<G$ the stabilizer  
419: of $z_{\hbox{\fiverm H}}\in \partial_d\Xi$ in $G$.  
420:  
421: \par For the rest of this paper we will employ the following assumptions: 
422: $G_\C$ is simply connected and $G/H\simeq G(z_{\hbox{\fiverm H}})$ is an NCC symmetric space (cf.\  
423: Proposition 1.3.1). Notice that this implies in particular $H=G^\tau$.  
424: Recall the element $T_0$ from Definition 1.1.1 and notice that we have  
425: (up to sign) $X_{\hbox{\fiverm H}}={\pi\over 2}T_0$.  
426:  
427:  
428:  
429: \par We define  a domain $\Omega_{\hbox{\fiverm H}}\subeq \Omega$ 
430: by 
431:  
432: $$\Omega_{\hbox{\fiverm H}}=\Int \left(\conv\{{\cal W}(X_{\hbox{\fiverm H}})\}\right)\ .$$ 
433: Here $\conv\{\cdot\}$ denotes the convex hull of $\{ \cdot\}$ and $\Int(\cdot)$ denotes 
434: the interior of $(\cdot)$. {}From the definition we immediately obtain that: 
435:  
436: \msk 
437: \item{(1.4.1)} $\Omega_{\hbox{\fiverm H}}$ is open in $\a$. 
438: \item{(1.4.2)} $0\in \Omega_{\hbox{\fiverm H}}$ (because $ X_{\hbox{\fiverm H}}\neq 0$ 
439: and ${\cal W}(X_{\hbox{\fiverm H}})$ meets every Weyl chamber). 
440: \item{(1.4.3)} The set of extremal points of $\oline{\Omega_{\hbox{\fiverm H}}}$ is ${\cal W}(X_{\hbox{\fiverm H}})$. 
441: \msk 
442:  
443: Let us illustrate the geometry for one example. 
444:  
445: \Example 1.4.1. Let $G=\Sl(3,\R)$. Then $\a$ is two-dimensional  
446: and $\Sigma$ is a root system of type $A_2$. We have $\partial_e\Omega={\cal W}(X_1)\amalg {\cal W}(X_2)$ and the  
447: corresponding isotropy subgroups are given by $H_1=\SO(1,2)$ and $H_2=\SO(2,1)$.  
448: With $H=H_1$ the geometry of $\Omega$ and $\Omega_H$ is depicted as follows: 
449:  
450: \gsk  
451: \epsfbox{hex.eps}
452: \gsk\gsk 
453: \qed 
454:  
455: \nin Let us define a domain $\Xi_H\subeq G_\C/ K_\C$ by 
456: $$\Xi_H=G\exp(i\Omega_{\hbox{\fiverm H}})K_\C/ K_\C\, .$$ 
457: The domain $\Xi_H$ has the following properties: 
458: \msk 
459: \item{(1.4.4)} $\Xi_H$ is $G$-invariant (clear from the definition). 
460: \item{(1.4.5)} $\Xi_H$ is open in $G_\C/ K_\C$ (follows from (1.4.1) and [AG90]). 
461: \item{(1.4.6)} $G/K\subeq \Xi_H$ is a totally real submanifold (follows from (1.4.2) and (1.4.5)). 
462: \item{(1.4.7)} $\Xi_H$ is Stein (follows from the convexity of $\Omega_{\hbox{\fiverm H}}$ and [GK02b]). 
463: \item{(1.4.8)} $\Xi_H\subeq \Xi$ (because $\Omega_{\hbox{\fiverm H}}\subeq \Omega$). 
464: \item{(1.4.9)} $\Xi_H=\Xi$ iff $\Sigma$ is of type $C_n$ (cf.\ [KS01b]). 
465: \msk 
466:  
467: \nin  Write $\oline{\Xi_H}$ for the closure of $\Xi_H$ in $G_\C/ K_\C$ and define the 
468: {\it distinguished boundary} of $\Xi_H$ by 
469: $$\partial_d\Xi_H=G(z_{\hbox{\fiverm H}})\, .$$ 
470: Notice that $\partial_d\Xi_H\simeq G/H$ as $G$-spaces. Let us remark further that 
471: $\partial_d\Xi_H\subeq \partial_d\Xi$.  
472:  
473: \par The distinguished boundary $\partial_d\Xi_H$ of $\Xi_H$ can be 
474: considered as some sort of Shilov boundary of $\Xi_H$. More precisely, 
475: mimicking the argument in [GK02a, Th.\ 2.3] we obtain that 
476:  
477: $$\sup_{z\in \Xi_H} |f(z)|=\sup_{z\in \partial_d\Xi_H|} |f(z)|\leqno (1.4.10)$$ 
478: for all bounded holomorphic functions $f$ on $\Xi_H$ which continuously extend to $\oline{\Xi_H}$. 
479:  
480:  
481:  
482:  
483: \sectionheadline{2. Holomorphic $H$-spherical distributions} 
484:  
485: \noindent 
486: In the section we assume that $G/H$ is NCC symmetric space 
487: realized as $G(z_{\hbox{\fiverm H}})$ in the distinguished boundary of $\Xi$. 
488: \par Recall 
489: that a representation $(\pi ,V)$ of $G$ is called {\it admissible} if  
490: the multiplicity of each $K$-type is finite and of {\it finite length} if  
491: the associated Harish-Chandra module of $K$-finite vectors $V_K$ is of finite length. 
492: Our aim in this section is to 
493: associate to a non-zero $K$-fixed vector $v_{\hbox{\fiverm K}}$ in an admissible  
494: representation $(\pi,V)$ of finite length a  certain canonical $H$-spherical 
495: distribution vector $v_{\hbox{\fiverm H}}\in (V^{-\infty})^H$. For irreducible representations 
496: $\pi$ the vector $v_{\hbox{\fiverm H}}$ 
497: is unique in the sense that it allows analytic continuation 
498: of generalized matrix coefficients on $G/H$ to holomorphic 
499: functions on $\Xi_H$.  
500: \par We let $X_{\hbox{\fiverm H}},x_{\hbox{\fiverm H}}$, and $z_{\hbox{\fiverm H}}$ be as in 
501: the last subsection and recall that $x_{\hbox{\fiverm H}}^{-1}H_\C x_{\hbox{\fiverm H}}=K_\C$. 
502:  
503: \msk\nin {\bf General Remark:} In all results of this section  which involve the domain 
504: $\Xi_H$ one can replace $\Xi_H$ by the bigger domain $\Xi$. This holds in particular  
505: for the results in Subsection 2.2.  
506:  
507:  
508: \subheadline{2.1. The definition of the holomorphic $H$-spherical distribution vector} 
509:  
510:  
511: \noindent 
512: Before we discuss the general case let us assume for the moment 
513: that $V$ is irreducible and finite dimensional. Then 
514: $(\pi, V)$ extends to a holomorphic representation of $G_\C$ 
515: which we also denote by $(\pi, V)$. 
516: If $L<G$ is a subgroup of $G$ then we write $V^L$ for the subspace of 
517: $V$ fixed by $L$. 
518: Then the mapping 
519: $$V^K\to V^H, \ \ v_{\hbox{\fiverm K}}\mapsto v_{\hbox{\fiverm H}}\:=\pi(x_{\hbox{\fiverm H}})v_{\hbox{\fiverm K}}$$ 
520: sets up an isomorphism between the $K$-spherical and $H$-spherical 
521: vectors of $V$. 
522: The obvious problem in the general case is, that $\pi(x_{\hbox{\fiverm H}})v$ is 
523: not necessarily defined as an element in $V$. 
524: \par 
525:  
526: We are now going to develop an appropriate generalization of the mapping 
527: $v_{\hbox{\fiverm K}}\mapsto v_{\hbox{\fiverm H}}$ for an admissible finite length  
528: representation $(\pi ,{\cal H})$ of $G$ in a Hilbert space ${\cal H}$. 
529: Denote by ${\cal H}^\infty$ respectively ${\cal H}^\omega$ the space of smooth respectively 
530: analytic vectors in ${\cal H}$. Their strong anti-duals, i.e., the space 
531: of continuous conjugate linear maps into $\C$, are denoted by ${\cal H}^{-\infty}$ 
532: respectively ${\cal H}^{-\omega}$ and referred 
533: to as the $G$-modules of {\it distribution 
534: vectors}, respectively {\it hyperfunction vectors} of $(\pi, {\cal H})$. 
535: Notice the chain of continuous inclusions ${\cal H}^{\omega}\into {\cal H}^{\infty}\into {\cal H}\into 
536: {\cal H}^{-\infty}\into {\cal H}^{-\omega}$. Here the  inclusion 
537: ${\cal H}\into {\cal H}^{-\infty}$ is the natural one, $v\mapsto (u\mapsto \la v,u\ra)$. We 
538: will also use the notation $u\mapsto \la v,u\ra$ for $v\in {\cal H}^{-\infty}$. Define a representation 
539: $\pi^0$ of $G$ on $\cal H$ by  
540: $$\eqalign{ 
541: \la \pi^0(g) v,u\ra &=\la \pi (g^{-1})^*v,u\ra \cr 
542: &= \la v,\pi (g^{-1})u\ra\cr}\qquad (g\in G; \ u, v\in {\cal H})\ .$$ 
543: Hence the natural representation $\pi^{-\infty}$ of $G$ on ${\cal H}^{-\infty}$ is an extension of the representation 
544: $\pi^0$ on ${\cal H}$ to ${\cal H}^{-\infty}$. The representation $\pi^0$ is 
545: called the {\it conjugate dual representation} of $\pi$. 
546: Notice that $(\pi^0, {\cal H})$ is admissible and of finite length if and 
547: only if the same holds for $(\pi ,{\cal H})$. Note also that $\pi =\pi^0$ if and only if 
548: $\pi $ is unitary. Notice that if $u,v\in {\cal H}$ then $\la v ,u\ra =\overline{\la u, v\ra}$. Accordingly, 
549: if $u\in {\cal H}^{\infty}$ and $v\in {\cal H}^{-\infty}$, then we write 
550: $$\la u,v\ra :=\overline{\la v,u\ra}\, .$$ 
551:  
552:  
553: \par 
554: Denote by ${\cal H}_K$ the $(\g,K)$-module of $K$-finite vectors in ${\cal H}$. Note that by 
555: our assumption that ${\cal H}$ is admissible and of finite length it follows that ${\cal H}_K\subseteq {\cal H}^\omega$. 
556: But usually  we cannot find $H$-fixed vectors in ${\cal H}$ but only 
557: in the larger space of distribution vectors. Another 
558: complication arises as the space of $({\cal H}^{-\infty})^H$ of $H$-invariants in 
559: ${\cal H}^{-\infty}$ is finite dimensional but in general not one-dimensional. 
560: For ``generic'' principal series representations of $G$ one has  
561: $$\dim ({\cal H}^{-\infty})^H=\left| {\cal W}/ {\cal W}_0\right|$$ 
562: with ${\cal W}_0=N_{K\cap H}(\a)/ Z_{K\cap H}(\a)$ the {\it little Weyl group}. 
563: Our correspondence will be that we associate to a 
564: $K$-fixed vector $v_{\hbox{\fiverm K}}$ a unique $H$-fixed distribution vector $v_{\hbox{\fiverm H}}$. 
565: If $\pi$ is irreducible, then this 
566: distribution $v_{\hbox{\fiverm H}}$ is the (up to scalar) unique element of  $({\cal H}^{-\infty})^H$ 
567: for which generalized matrix coefficients extend to holomorphic functions 
568: from $G/H$ to $\Xi$. For the proof we will need the following {\it 
569: Automatic Continuity} Theorem of van den Ban, Brylinski and Delorme  (c.f. [vdBD88, Th. 2.1] and  
570: [BD92, Th.\ 1] for the version used here). 
571: \par If $V$ is a complex vector space, then  let us denote by 
572: $V^{\oline *}$ its algebraic anti-dual. Set 
573: $({\cal H}^{-\infty})^{\h}=\{v\in {\cal H}^{-\infty}\: (\forall X\in \h)\, d\pi^{-\infty}(X)v=0\}$. 
574:  
575:  
576:  
577: \Theorem 2.1.1. {\bf (Automatic Continuity)} 
578: Let $(\pi, {\cal H})$ be an admissible representation of $G$ with finite length. Then 
579: $$({\cal H}^{-\infty})^{H_0}\simeq ({\cal H}_K^{\oline *})^\h\ , \leqno (2.1.1) $$ 
580: meaning every $\h$-fixed anti-linear functional on ${\cal H}_K$ admits 
581: a unique extension to a continuous and $H_0$-fixed anti-linear functional 
582: on ${\cal H}^\infty$. In particular we have that 
583: $$({\cal H}^{-\infty})^{H_0}\simeq ({\cal H}^{-\omega})^{H_0}\ .\ \leqno(2.1.2)$$\qed 
584:  
585: \par Recall the 
586: $G-K_\C$ double coset domain $\tilde \Xi=G\exp(i\Omega)K_\C$ 
587: in $G_\C$ and the complex crown  $\Xi=\tilde\Xi/ K_\C$. Our methods use holomorphic extensions of 
588: representations. We recall therefore some results from [KS01a] \S 4 and in particular Theorem 3.1: 
589:  
590: \Theorem 2.1.2.  Let $(\pi, {\cal H})$ be an admissible Hilbert representation of $G$ 
591: with finite length. Then the following assertions hold: 
592:  
593: \item{(i)} For every $v\in {\cal H}_K$ the orbit 
594: mapping $G\to {\cal H}, \ g\mapsto \pi(g)v$ extends to a $G$-equivariant holomorphic 
595: mapping $\tilde \Xi\to {\cal H}$. 
596:  
597: \item{(ii)} Let $v,w\in {\cal H}_K$. Then the restricted matrix coefficient 
598: $A\to \C, \ a\mapsto \la \pi (a)v, w\ra $ extends to a holomorphic mapping to 
599: the abelian tube domain $T (2\Omega)=A\exp(2i\Omega)$.\qed 
600:  
601: We will from now on assume that $\pi\res_K$ is unitary. This is no restriction 
602: in view of Weyl's unitarity trick. 
603:  
604: \par 
605: As before  we identify $G/H$ with the subset $G(z_{\hbox{\fiverm H}})$ of $\overline{\Xi_H}$. 
606: For $0\leq t<1$ we set $a_t=\exp(itX_{\hbox{\fiverm H}})\in \exp(i\Omega)$ and notice that 
607: $\lim_{t\to 1} a_t =x_{\hbox{\fiverm H}}$. 
608:  
609: \Theorem 2.1.3. Let $(\pi, {\cal H})$  be a $K$-spherical admissible  representation 
610: of $G$ with finite length. Let $v_K\in {\cal H}^K$ be non-zero. Then the anti-linear functional 
611:  
612: $$v_{\hbox{\fiverm H}}^\omega\: {\cal H}^\omega \to \C , \ \ v\mapsto \lim_{t\nearrow 1} \la \pi^0(a_t)v_{\hbox{\fiverm K}}, v\ra$$ 
613: is well defined, non-zero and admits a non-trivial extension to an  
614: $H$-fixed distribution vector of $(\pi, {\cal H})$. \qed 
615:  
616: \ssk \nin {\bf Note:} If $(\pi, {\cal H})$ is unitary, then $\pi^0(a_t)=\pi(a_t)$ and so 
617: $v_{\hbox{\fiverm H}}^\omega(v)=\lim_{t\nearrow 1} \la \pi(a_t)v_{\hbox{\fiverm K}}, v\ra$ for all $v\in {\cal H}^\omega$. 
618:  
619:  
620: \Proof.  We first show that  $v_{\hbox{\fiverm H}}^\omega$ is well defined.  
621: Let $v\in {\cal H}^\omega$. As $v$ is  
622: analytic,  we find an $0<\epsilon <1$ such that  
623: $\pi (a_\eps)v$ is defined. Notice that $\pi^0 (a_{1-\eps})v_{\hbox{\fiverm K}}$ is defined 
624: by Theorem 2.1.2 (i).  
625: Then we have for all $\eps<t<1$ that  
626: $$\la \pi^0(a_t)v_{\hbox{\fiverm K}}, v\ra= \la \pi^0(a_{t-\eps})v_{\hbox{\fiverm K}}, \pi(a_\eps)v\ra$$ 
627: and hence 
628: $$\lim_{t\nearrow 1}\la \pi^0(a_t)v_{\hbox{\fiverm K}}, v\ra=\la \pi^0(a_{1-\eps})v_{\hbox{\fiverm K}},  
629: \pi(a_\eps)v\ra\ .$$ 
630: Thus  $v_{\hbox{\fiverm H}}^\omega$ is defined. 
631:  
632: \par  Next we show that $v_{\hbox{\fiverm H}}^\omega$ is fixed by $H$.  
633: As $G/H$ is NCC, it follows that $H=H_0 Z_{H\cap K}(\a )$ [H\'O96, p.79]. Further it is clear from the definition  
634: that $v_{\hbox{\fiverm H}}^\omega$it is fixed by $Z_{H\cap K}(\a )$. Hence it is  
635: enough to prove that $v_{\hbox{\fiverm H}}^\omega$ 
636: is $H_0$-fixed, i.e. annihilated by $\h$. For that let $Y\in \h$ and 
637: $v\in {\cal H}^\omega$. Then 
638:  
639: $$\eqalign{v_{\hbox{\fiverm H}}^\omega(d\pi(Y)v)&=\lim_{t\nearrow 1}\la \pi^0(a_t)v_{\hbox{\fiverm K}}, d\pi(Y)v\ra= 
640: -\lim_{t\nearrow 1}\la d\pi^0(Y)\pi^0(a_t)v_{\hbox{\fiverm K}}, v\ra\cr 
641: &= - \lim_{t\nearrow 1}\la\pi^0(a_t)d\pi^0(\Ad(a_t)^{-1} Y)v_{\hbox{\fiverm K}}, v\ra\cr 
642: &=  -\lim_{t\nearrow 1}\la\pi^0(a_{t-\eps})d\pi^0(\Ad(a_t)^{-1} Y)v_{\hbox{\fiverm K}}, \pi(a_\eps)v\ra\ .\cr}$$ 
643: {}From Theorem 2.1.2 (i) we now obtain that 
644: $$\pi^0(a_{t-\eps})d\pi^0(\Ad(a_t)^{-1} Y)v_{\hbox{\fiverm K}}\to 
645: \pi^0(a_{1-\eps})d\pi^0(\Ad(x_{\hbox{\fiverm H}})^{-1} Y)v_{\hbox{\fiverm K}}\ .$$ 
646: But $\Ad (x_{\hbox{\fiverm H}})^{-1}Y\in \k_\C$ and hence 
647: $d\pi^0(\Ad (x_{\hbox{\fiverm H}})^{-1}Y)v_{\hbox{\fiverm K}}=0$. 
648: We thus get: 
649: $$\eqalign{v_{\hbox{\fiverm H}}^\omega(d\pi(Y)v) 
650: &=-\lim_{t\nearrow 1}\la\pi^0(a_{t-\eps})d\pi^0(\Ad(a_t)^{-1} Y)v_{\hbox{\fiverm K}}, \pi(a_\eps)v\ra\cr 
651: &= -\la\pi^0(a_{1-\eps})d\pi^0(\Ad(x_{\hbox{\fiverm H}})^{-1} Y)v_{\hbox{\fiverm K}}, \pi(a_\eps)v\ra\cr 
652: &= 0\, .\cr}$$ 
653:  
654:  
655: \par Finally, let us show that $v_{\hbox{\fiverm H}}^\omega\neq 0$. Suppose the contrary, i.e.  
656: $v_{\hbox{\fiverm H}}^\omega=0$. Then it follows  that  
657: $$(\forall u\in {\cal H}_K) \qquad \la v_{\hbox{\fiverm H}}^\omega, u\ra =0\ . \leqno (2.1.3)$$ 
658: Now consider the function  
659: $$f\: \R +i]-2,2[\to \C , \ \ z\mapsto \la \pi^0(\exp(zX_{\hbox{\fiverm H}}))v_{\hbox{\fiverm K}},  
660: v_{\hbox{\fiverm K}}\ra\ .$$ 
661: According to Theorem 2.1.2 (ii) the function $f$ is well defined and holomorphic.  
662: It is clear that $f\not\equiv 0$ as $f(0)=\la  v_{\hbox{\fiverm K}},v_{\hbox{\fiverm K}}\ra>0$. 
663: But (2.1.3) implies that $f^{(n)}(i)=0$ for all $n\in \N_0$; a contradiction  
664: to $f\not\equiv 0$. \qed  
665:  
666: In the sequel we write $v_{\hbox{\fiverm H}}$ for the $H$-fixed distribution vector 
667: obtained from $v_{\hbox{\fiverm H}}^\omega$. We will call 
668: $v_{\hbox{\fiverm H}}$ the {\it holomorphic $H$-spherical distribution vector} of $(\pi, {\cal H})$ 
669: corresponding to $v_{\hbox{\fiverm K}}$. 
670:  
671:  
672: \Remark 2.1.4. By Theorem 2.1.3 we have 
673: $$v_{\hbox{\fiverm H}}^\omega=\hbox{w}-\lim_{t\nearrow 1} \pi^0(a_t)v_{\hbox{\fiverm K}}, $$ 
674: i.e. $v_{\hbox{\fiverm H}}^\omega$ is the weak limit of $\pi^0(a_t)v_{\hbox{\fiverm K}}$ for $t\to 1$ in the 
675: locally convex space ${\cal H}^{-\omega}$.  
676: It is possible to strengthen this convergence: Let $v\in {\cal H}^\omega$ and $C\subeq G$  
677: a compact subset. Then for all $\eps>0$ there exists $0<s<1$ such that for all 
678: $0<s<t<1$: 
679:  
680: $$\sup_{g\in C} |\la v_{\hbox{\fiverm H}}^\omega, \pi(g)v\ra -\la \pi^0(a_t)v_{\hbox{\fiverm K}}, \pi(g)v\ra|<\eps\ 
681:  . \leqno(2.1.4)$$ 
682: In fact this follows from a simple modification of the first part of the  
683: proof of Theorem 2.1.3: we only have to observe that for $v\in {\cal H}^\omega$ 
684: there exists a $0<\delta<1$ such that $\pi(a_\delta)\pi(g)v$ exists for all  
685: $g\in C$.  
686:  
687:  
688: \par Supported by calculations in the rank one case we conjecture 
689: that one actually has 
690: $ \pi(a_t)v_{\hbox{\fiverm K}}\to v_{\hbox{\fiverm H}}$ in ${\cal H}^{-\infty}$ weakly (and hence strongly by the 
691: Banach-Steinhaus Theorem which applies as ${\cal H}^\infty$ is a Fr\'echet space ). \qed 
692:  
693: Let us illustrate the situation by the discussion of one example. 
694:  
695: \Example 2.1.5. Here we will determine an explicit analytic description 
696: of $v_{\hbox{\fiverm H}}$ for unitary principal series of $G=\Sl(2,\R)$. 
697: Let $(\pi_\lambda, {\cal H}_\lambda)$ denote a unitary spherical 
698: principal series of $G$ with parameter $\lambda\in i\a^*$. 
699: Then $\pi_\lambda^0=\pi_\lambda$ as $\pi_\lambda$ is unitary.  In the sequel 
700: we will identify $\a_\C^*$ with $\C$ in such a way that $\rho\in \a^*$ corresponds to $1$. 
701: With our choice of $\a$ to be 
702: $$\a=\left\{ \pmatrix{s & 0\cr 0& -s\cr}\: s\in \R\right\}$$ 
703: this identification is given by 
704: $$\lambda \mapsto \lambda \pmatrix{1 & 0 \cr 0 & -1\cr}\, .$$ 
705:  
706: \par We will use the noncompact realization of ${\cal H}_\lambda =L^2(\R)$ of 
707: $\pi_\lambda$. Then for $g=\pmatrix {a & b\cr c& d\cr}\in G$ the operator 
708: $\pi_\lambda(g)$ is given by 
709:  
710: $$(\pi_\lambda(g)f)(x)=|bx+d|^{-1-\overline{\lambda}} f\left({ax+c\over bx +d}\right) 
711: \qquad (f\in L^2(\R), x\in \R)\ .$$ 
712: A normalized $K$-spherical vector is then given by 
713:  
714: $$v_{\hbox{\fiverm K}}(x)={1\over \sqrt\pi} (1+x^2)^{-{1\over 2}(1+\overline{\lambda})}\ .$$ 
715: \par Notice that $H=\SO(1,1)$ and  (up to sign) we have  
716: $X_{\hbox{\fiverm H}}=\pmatrix {\pi\over 4 & 0\cr 0& -{\pi\over 4}\cr}$.  
717: Thus for $0\leq t<1$ the element $a_t\in 
718: \exp(i\Omega)$ is given by  
719:  
720: $$a_t=\pmatrix{ e^{i{\pi\over 4}t} & 0 \cr 0 & e^{-i{\pi\over 4}t} \cr}\  .$$ 
721: Then we have 
722: $$v_{\hbox{\fiverm H}}^\omega=\hbox{w}-\lim_{t\nearrow 1} \pi_\lambda(a_t)v_{\hbox{\fiverm K}}$$ 
723: or 
724: $$v_{\hbox{\fiverm H}}^\omega(x)=\hbox{w}- \lim_{t\nearrow 1}{e^{i{\pi\over 4}t 
725: (1+\overline{\lambda})}\over \sqrt\pi} (1+e^{i\pi t}x^2)^{-{1\over 2}(1+\overline{\lambda})}\ .$$ 
726: A simple calculation then shows that $v_{\hbox{\fiverm H}}^\omega$ and $v_{\hbox{\fiverm H}}$ 
727: are given by the locally integrable function 
728:  
729: $$v_{\hbox{\fiverm H}} (x)=\cases {{e^{i{\pi\over 4}(1+\overline{\lambda})}\over \sqrt\pi} 
730: (1- x^2)^{-{1\over 2}(1+\overline{\lambda})} & for $|x|<1$, \cr 
731: 0 & for $|x|=1$, \cr 
732: {e^{-i{\pi\over 4}(1+\overline{\lambda})} \over \sqrt \pi} 
733: (x^2-1)^{-{1\over 2}(1+\overline{\lambda})} & for $|x|>1$\ .\cr} $$ 
734: A basis of $({\cal H}_\lambda^{-\infty})^H$ is 
735: given by $v_{\hbox{\fiverm H,1}}, v_{\hbox{\fiverm H, 2}}$ where 
736: $$v_{\hbox{\fiverm H,1}}(x)=\cases { {1\over \sqrt\pi} (1- x^2)^{-{1\over 2}(1+\overline{\lambda})} & for $|x|<1$, \cr 
737: 0 & for $|x|\geq 1$. \cr}$$ 
738: and 
739: $$v_{\hbox{\fiverm H,2}}(x)=\cases {  {1\over \sqrt\pi}(x^2-1)^{-{1\over 2}(1+\overline{\lambda})} & for $|x|>1$, \cr 
740: 0 & for $|x|\leq 1$. \cr}$$ 
741: These two basis vectors are chosen such  that they have support in the 
742: open $H$-orbits on $\P^1(\R)$, namely $]-1, 1[$ and $\P^1(\R)\bs [-1, 1]$. 
743: Notice that $v_{\hbox{\fiverm H}}$ is a non-trivial linear combination of $v_{\hbox{\fiverm H, 1}}$ and  
744: $v_{\hbox{\fiverm H,2}}$ and that 
745: $v_{\hbox{\fiverm H}}$ has full support on $\R$. 
746: Another interesting feature of this example is that we  have here 
747: $$v_{\hbox{\fiverm H}} =\hbox{w}-\lim_{t\nearrow 1}  
748: \pi_\lambda(a_t)v_{\hbox{\fiverm K}} \qquad \hbox{in ${\cal H}_\lambda^{-\infty}$}$$ 
749: and hence also strongly by the Banach-Steinhaus Theorem (compare with the conjecture  
750: stated at the end of Remark 2.1.4) . \qed 
751:  
752:  
753:  
754:  
755:  
756:  
757: \subheadline{2.2. Holomorphic extension of matrix coefficienst on $G/H$ to $\Xi_H$} 
758:  
759: \noindent 
760: In this subsection we clarify the role of the holomorphic distribution vector 
761: $v_{\hbox{\fiverm H}}$ in view of holomorphic extensions of matrix coefficients from 
762: $G/H$ to $\Xi$.  If $U$ is a complex manifold, then we denote by 
763: ${\cal O}(U)$ the space of holomorphic functions $f:U\to \C$. 
764:  
765: \Definition 2.2.1. Let $f$ be a continuous function on $G/H$. Then we say 
766: that $f$ has a {\it holomorphic extension} to $\Xi_H$ if there exists an 
767: $\tilde f\in {\cal O}(\Xi_H)$ such that for all compact subsets 
768: $C\subeq G$  one has 
769: $$\lim_{t\nearrow 1} \sup_{g\in C} |f(gH) -\tilde f(ga_tK_\C)|=0\ .\leqno(2.2.1)$$ 
770: \qed 
771:  
772: Notice that (2.2.1) implies that 
773: $$f(gH)=\lim_{t\nearrow 1} \tilde f(ga_tK_\C)$$ 
774: for all $g\in G$. 
775: Furthermore the holomorphic extensions are unique by  the following lemma: 
776:  
777:  
778: \Lemma 2.2.2. {\rm\bf (Identity Theorem for holomorphic extensions)} Let $f\in C(G/H)$ 
779: and assume that $f$ has 
780: a holomorphic extension $\tilde f\in {\cal O}(\Xi_H)$. Then $f\equiv 0$ implies 
781: $\tilde f\equiv 0$. In particular, the holomorphic extension $f\in C(G/H)$ 
782: is unique if it exists. 
783:  
784: \Proof. This is easily reduced to the one-dimensional case as 
785: follows. Define an abelian tube domain 
786: $T=\exp(\R X_{\hbox{\fiverm H}} + ]-1,1[ iX_{\hbox{\fiverm H}})$ and set $\partial_s 
787: T =\exp(\R X_{\hbox{\fiverm H}} + iX_{\hbox{\fiverm H}})$. We realize 
788: $T\subeq \Xi_H$ and $\partial_sT\subeq G/H$ through the $T$-orbit , respectively $\partial_s T$-orbit, 
789: through $K_\C\in \Xi_H$, respectively $z_{\hbox{\fiverm H}}\in G/H$. 
790: Let $f\in C(G/H)$ and assume that $f$ has a holomorphic extension $\tilde f$. 
791: Let $\phi=f\res_{\partial_s T}$. Then $\phi$ has a holomorphic extension to 
792: $T$ given by $\tilde \phi=\tilde f\res_T$. By the well known one-dimensional 
793: situation we have $\tilde\phi\equiv 0$ if $\phi\equiv 0$. Thus if $f\equiv 0$, we obtain 
794: $\tilde f\res_T\equiv 0$. Replacing $f$ by $f_g$ where $f_g(xH)=f(gxH)$ for $g\in G$, then 
795: the above discussion implies that 
796: $\tilde f\res_{GT}\equiv 0$ and hence $\tilde f\equiv 0$ as $GT$ contains the totally 
797: real submanifold $G/K$ of $\Xi_H$. \qed 
798:  
799: We assume that the complex conjugation $\g_\C\to\g_\C, \ X\mapsto \oline X$ with respect to 
800: the real form $\g$ of $\g_\C$ lifts to a conjugation $g\mapsto \oline g$ of $G_\C$. 
801: Notice that this is always satisfied if $G_\C$ is the universal complexification of $G$. 
802: Notice also that $\overline{x}\in \tilde{\Xi}$ for all $x\in\tilde{\Xi}$. 
803: Let $u,v\in {\cal H}_K$. Then the functions 
804: $$\tilde{\Xi}\ni x\mapsto\la \pi^0(x)u,v\ra , \la u,\pi(\overline{x}^{-1})v \ra\in \C $$ 
805: are well defined and holomorphic by Theorem 2.1.2. Both of them agree if $x\in G$ and 
806: hence they agree on all of $\tilde{\Xi}$. 
807: %The same argument using Theorem 2.1.3 and (2.1.4) we then obtain 
808: %the following extension result: 
809:  
810: \Proposition 2.2.3. Let $(\pi, {\cal H})$ be a admissible $K$-spherical representation of $G$ 
811: with finite length. Let $v\in {\cal H}^\omega$. Then the following assertions hold: 
812:  
813: \item{(i)} The 
814: matrix coefficient 
815: $$f_{v, v_{\hbox{\fiverm H}}}\: G/H\to \C, \ \ gH\mapsto \la \pi (g^{-1})v, v_{\hbox{\fiverm H}}\ra$$ 
816: admits a holomorphic extension $\tilde f_{v, v_{\hbox{\fiverm H}}}$. 
817: Moreover, we have 
818: $$\tilde f_{v, v_{\hbox{\fiverm H}}}(xK_\C)=\la v, \pi^0(\overline x)v_{\hbox{\fiverm K}}\ra $$ 
819: for all $xK_\C\in \Xi_H$. 
820: \item{(ii)} The 
821: matrix coefficient 
822: $$g_{v_{\hbox{\fiverm H}},v }\: G/H\to \C, \ \ gH\mapsto \la \pi^0(g)v_{\hbox{\fiverm H}},v\ra$$ 
823: admits a holomorphic extension $\tilde g_{v_{\hbox{\fiverm H}},v}$. 
824: Moreover, we have 
825: $$\tilde g_{v_{\hbox{\fiverm H}},v}(xK_\C)=\la \pi^0(x )v_{\hbox{\fiverm K}},v\ra \quad\hbox{\rm and} 
826: \quad \tilde{g}_{v_{\hbox{\fiverm H}},v}(xK_\C) 
827: =\overline{\tilde f_{v, v_{\hbox{\fiverm H}}}(\overline{x}K_\C)}$$ 
828: for all $xK_\C\in \Xi_H$.%\qed 
829:  
830: \Proof. We will only show (i) as the proof for (ii) is the same. By Theorem 2.1.2 (i) it follows that 
831: $$\tilde{f}(xK_\C)=\la v,\pi^0(\bar{x})v_{\hbox{\fiverm K}}\ra$$ 
832: exists and is holomorphic on $\Xi_H$. Let $g\in G$ and $0<t<1$. Then 
833: $$\la \pi (g^{-1})v,\pi^0(a_t)v_{\hbox{\fiverm K}}\ra =\la v,\pi^0(ga_t)v_{\hbox{\fiverm K}}\ra 
834: =\tilde{f}(ga_tK_\C)\, .$$ 
835: Taking the limit at $t\to 1$ and using the remarks just before (2.1.4) it follows that 
836: $$\lim_{t\nearrow 1} \tilde{f}(ga_tK_\C)=\la \pi (g^{-1})v,v_{\hbox{\fiverm H}}\ra$$ 
837: and the convergence is uniform on compact subsets in $G$. 
838: \qed 
839:  
840: Denote by $({\cal H}^{-\infty})^H_{\rm hol}\subset ({\cal H}^{-\infty})^H$ 
841: the space of $H$-invariant distribution vectors $\eta$ such 
842: that the function 
843: $$G/H\ni x \mapsto g_{\eta ,v}(x):=\la \pi^0(x)\eta ,v\ra$$ 
844: has a holomorphic extension to $\Xi_H$  
845: for all $v\in {\cal H}^\omega$. Notice that for $g\in G$ and $x\in \Xi_H$ we have 
846: $$\tilde g_{\eta ,\pi (g)v}(x)=\tilde g_{\eta ,v}(g^{-1}x)\, .\leqno(2.2.2)$$ 
847: Our next task is to prove a converse of Proposition 2.2.3, namely that 
848: the map 
849: $${\cal H}^K\ni v_{\hbox{\fiverm K}}\mapsto v_{\hbox{\fiverm H}}\in ({\cal H}^{-\infty})^H_{\rm hol}$$ 
850: is an isomorphism. In particular only the $H$-invariant distribution 
851: vectors constructed in Theorem 2.1.3 have holomorphic extension. Thus if 
852: $\dim  {\cal H}^K=1$, as in the case of the principal series 
853: representations of $G$, the space $({\cal H}^{-\infty})^H_{\rm hol}$ is 
854: also one-dimensional, i.e., there is (up to scalar) a unique $H$-spherical distribution vector which allows 
855: holomorphic extension of the  smooth matrix coefficients. 
856: \par 
857: As before we consider the pairing 
858: $$ ({\cal H}^{-\infty})^H\times {\cal H}^\infty \to C^\infty(G/H), \ \ (\eta,v)\mapsto 
859: g_{\eta,v}; \ g_{\eta,v}(gH)=\la \pi^0(g)\eta,v\ra\ .$$ 
860:  
861:  
862: \Theorem 2.2.4. Let $(\pi , {\cal H})$ be an admissible representation of $G$ with finite length. Then the map 
863: $${\cal H}^K\ni v_{\hbox{\fiverm K}}\mapsto v_{\hbox{\fiverm H}}\in ({\cal H}^{-\infty})^H_{\rm hol}$$ 
864: is a linear isomorphism. 
865:  
866: \Proof. It follows from Theorem 2.1.3 that ${\cal H}^K\ni v_{\hbox{\fiverm K}}\mapsto 
867: v_{\hbox{\fiverm H}}\in({\cal H}^{-\infty})^H_{\rm hol}$ is well defined and injective. 
868: It is also clear that the map is linear. It remains to show that the map is onto.  
869: For that let $\eta\in ({\cal H}^{-\infty})^H_{\rm hol}$. Define a conjugate linear map 
870: $\tilde\eta : {\cal H}_K\to \C$ by 
871: $$\tilde\eta(u)= \tilde g_{\eta ,u}(K_\C)\, .$$ 
872: Then it follows from (2.2.2) that $\tilde\eta$ is $K$-invariant. Thus we find a unique  
873: $v_{\hbox{\fiverm K}}\in {\cal H}_K$ such that  
874: $\tilde\eta(u)=\la v_{\hbox{\fiverm K}}, u\ra$  for all $u\in {\cal H}_K$.  
875: In particular,  it follows that  
876: $$(\forall u\in {\cal H}_K)\qquad  
877: \tilde g_{\eta ,u}(K_\C)= \tilde g_{v_{\hbox{\fiverm H}} ,u}(K_\C)\ .\leqno(2.2.3)$$ 
878: \par Fix $w\in {\cal H}_K$. We claim that $\tilde g_{\eta ,w}=\tilde g_{v_{\hbox{\fiverm H}} ,w}$.  
879: In fact, it follows from (2.2.2) and (2.2.3) that all derivatives  
880: of  $\tilde g_{\eta ,w}$ and $\tilde g_{v_{\hbox{\fiverm H}} ,w}$ coincide at $K_\C \in\Xi_H$.  
881: Thus $\tilde g_{\eta ,w}=\tilde g_{v_{\hbox{\fiverm H}} ,w}$ by Taylor's Theorem.  
882:  
883: \par {} It follows from our claim that  $g_{\eta ,w}=g_{v_{\hbox{\fiverm H}} ,w}$ for all 
884: $w\in {\cal H}_K$. In particular, we obtain that   
885: $$(\forall w\in {\cal H}_K)\qquad  
886: \la \eta, w\ra =g_{\eta,w}(H)=g_{v_{\hbox{\fiverm H}}, w}(H)=\la v_{\hbox{\fiverm H}}, w\ra\ , $$ 
887: and so $\eta=v_{\hbox{\fiverm H}}$, concluding the proof of the theorem.\qed  
888:  
889:  
890: \Corollary 2.2.5. {\rm\bf (Multiplicity one)}  Let $(\pi, {\cal H})$ be an irreducible $K$-spherical  
891: Hilbert representation of $G$. Then $\dim ({\cal H}^{-\infty})_{\rm hol}^H =1$.  
892:  
893: \Proof. According to [H84, Ch.\ IV, Th. 4.5(iii)] we have that $\dim {\cal H}^K=1$.  
894: Thus the assertion follows from Theorem 2.2.4.\qed  
895:   
896:  
897:  
898:  
899: \subheadline{2.3. Distributional characters and boundary values} 
900:  
901: \noindent 
902: Denote by $dg$ and $dh$  Haar measures on $G$ and $H$. Notice 
903: that both $G$ and $H$ are unimodular and hence a left Haar measure is 
904: also a right Haar measure. Denote by $dgH$ a invariant measure on $G/H$, which 
905: we will normalize in a moment. 
906: Recall that the mapping  
907: $$C_c^\infty (G)\to C_c^\infty (G/H), \ \ f\mapsto f^H; \ f^H(xH)=\int_H f(xh)\ dh$$ 
908: is continuous and onto. We will normalize the measure $dgH$ in such a way that 
909: $$\int_G f(g) \ dg =\int_{G/H} f^H(gH)\ dgH$$ 
910: holds for all $f\in C_c(G)$.  
911:  
912:  
913: \par In this section $(\pi, {\cal H})$ will denote a {\it unitary}  
914: admissible representation representation of $G$ with finite length.  
915: Further we will assume that ${\cal H}^K\not=\{0\}$. Notice that  
916: $\pi$ unitary implies $\pi=\pi^0$.  
917: \par For $f\in C_c^\infty(G)$ let us recall  
918: the mollifying property: $\pi(f){\cal H}^{-\infty}\subeq 
919: {\cal H}^\infty$. Thus the mapping  
920: $$\Theta_{\pi,v_{\hbox{\fiverm H}}} \: C_c^\infty(G)\to \C, \ \ f\mapsto 
921: \la \pi(f)v_{\hbox{\fiverm H}}, v_{\hbox{\fiverm H}}\ra\, .$$ 
922: is well defined. It is known that $\Theta_\pi=\Theta_{\pi,v_{\hbox{\fiverm H}}}$ is a $H$-bi-invariant  
923: positive definite distribution on $G$. The $H$-bi-invariance implies  
924: that $\Theta_\pi(f)$ does only depend on $f^H$. We can therefore define a $H$-invariant 
925: distribution on $G/H$, also denoted by $\Theta_\pi$,  by 
926: $$\Theta_\pi (f^H)=\Theta_\pi(f)\, .$$ 
927:  
928: \par On the other hand we notice that $\pi (x)v_{\hbox{\fiverm K}}\in {\cal H}^\omega$ 
929: for all $x\in G\exp (i\Omega_H)K_\C$ and hence 
930: $x\mapsto \la v_{\hbox{\fiverm H}},\pi (x)v_{\hbox{\fiverm K}}\ra$ 
931: descends to a well defined and anti-holomorphic function on $\Xi_H$.  
932: Here $v_{\hbox{\fiverm K}}$ and $v_{\hbox{\fiverm H}}$ correspond to each other 
933: according to Theorem 2.2.4. We can therefore define the holomorphic function 
934: $\theta_\pi=\theta_{\pi,v_H}:\Xi_H\to \C$ by 
935: $$\theta_\pi(xK_\C)= \overline{\la v_{\hbox{\fiverm H}}, \pi(x)v_{\hbox{\fiverm K}}\ra }=\la \pi (x)v_{\hbox{\fiverm K}} , 
936: v_{\hbox{\fiverm H}}\ra \, .$$ 
937: Then $\theta_\pi$ is left $H$-invariant. 
938:  
939:  
940: Our next aim is to show that $\Theta_\pi$ is given by the limit 
941: operation and convolution: 
942:  
943: $$\Theta_\pi(f)=\lim_{t\nearrow 1}\int_{G/H} 
944: f(gH)\ \overline{\theta_\pi(g^{-1}a_t)} \ dgH$$ 
945: for suitable regular functions $f$. 
946:  
947: \par 
948: As we have only established the convergence $\pi(a_t)v_{\hbox{\fiverm K}}\to v_{\hbox{\fiverm H}}$ in 
949: ${\cal H}^{-\omega}$ and not in ${\cal H}^{-\infty}$, we cannot work 
950: with test-functions but must use an appropriate  space of 
951: analytic vectors.  
952: For that let us write  
953: $L^1(G)^{\omega, \omega}$ for the space  
954: of analytic vectors for the left-right regular representation of $G\times G$ on $L^1(G)$. 
955: Notice that $L^1(G)^{\omega, \omega}$ is an algebra under convolution  
956: which is  invariant under the natural involution $f\mapsto f^*$ with $f^*(x)=\oline {f(x^{-1})}$. 
957: Its importance lies in the fact that  $\pi(f){\cal H}^{-\omega}\subeq {\cal H}^\omega$ holds for all  
958: $f\in  L^1(G)^{\omega,\omega}$ (cf.\ Proposition A.4.1 in the appendix).  
959:  
960: \par Write $L^1(G/H)^\omega$ for the space of analytic vectors 
961: for the left regular representation of $G$ on $L^1(G/H)$. 
962: According to Proposition A.3.2 below,  the averaging map $f\mapsto f^H$ maps 
963: $L^1(G)^{\omega,\omega}$ into $L^1(G/H)^\omega$. 
964: Finally let us define the space: 
965: $${\cal A}^1(G/H)=\{f^H\in L^1(G/H)^\omega\mid f\in L^1(G)^{\omega,\omega}\}\ .  $$ 
966:  
967: \par {}From our discussion above we conclude that the mapping 
968: $$\Theta_\pi^\omega\: L^1(G)^{\omega,\omega} \to \C, \ \ f\mapsto \la \pi(f)v_{\hbox{\fiverm H}}, v_{\hbox{\fiverm H}}\ra$$ 
969: is well defined and $H$ bi-invariant. In particular  $\Theta_\pi^\omega(f)$ depends only on $f^H$ and therefore 
970: factors to ${\cal A}^1(G/H)$. We denote the corresponding map again by $\Theta_\pi^\omega$. 
971:  
972:  
973:  
974: \Theorem 2.3.1. Let $(\pi,{\cal H})$ be an unitary  admissible  representation 
975: of $G$ of finite length with ${\cal H}^K\not=\{0\}$. Let $v_{\hbox{\fiverm K}}\in {\cal H}^K$. Then  
976:  
977: $$(\forall f\in {\cal A}^1(G/H))\qquad \Theta_\pi^\omega(f)=\lim_{t\nearrow 1}\int_{G/H} 
978: f(gH)\ \overline{\theta_\pi(g^{-1}a_t)}\ dgH\ .$$ 
979:  
980: \Proof. Let $F\in L^1(G)^{\omega,\omega}$ be such that $f=F^H$. 
981: As $\pi(F)v_{\hbox{\fiverm H}}\in {\cal H}^\omega$ (cf.\ Proposition A.4.1), it follows from 
982: Theorem 2.1.3 that 
983: $$\Theta_\pi^\omega(F)=\lim_{t\nearrow 1}\la \pi(F)v_{\hbox{\fiverm H}}, \pi(a_t)v_{\hbox{\fiverm K}}\ra\ .$$ 
984: Thus: 
985:  
986: $$\eqalign{\Theta_\pi^\omega(F) 
987: &=\lim_{t\nearrow 1}\la v_{\hbox{\fiverm H}}, \int_G \oline{F(g^{-1})}\pi(g)\pi(a_t)v_{\hbox{\fiverm K}}\ dg\ra\cr 
988: &=\lim_{t\nearrow 1}\int_G F(g^{-1})\  \la v_{\hbox{\fiverm H}},\pi(g)\pi(a_t)v_{\hbox{\fiverm K}}\ra \ dg\cr 
989: &=\lim_{t\nearrow 1}\int_G F(g)\  \la v_{\hbox{\fiverm H}},\pi(g^{-1})\pi(a_t)v_{\hbox{\fiverm K}}\ra \ dg\cr 
990: &=\lim_{t\nearrow 1}\int_G  F(g) \ \overline{\theta_\pi(g^{-1}a_t)}\ dg\ .\cr}\leqno (2.3.1)$$ 
991: Fix $0\leq t<1$. We claim that  $g\mapsto \theta_\pi(g^{-1}a_t)$ is a bounded  
992: function on $G$. Indeed, for $g\in G$  we have  
993: $$\eqalign{|\theta_\pi(g^{-1}a_t)|&= 
994: |\la \pi(g^{-1}a_t)v_{\hbox{\fiverm K}},v_{\hbox{\fiverm K}}\ra|= 
995: |\la \pi(a_t)v_{\hbox{\fiverm K}},\pi(g)v_{\hbox{\fiverm K}}\ra|\cr  
996: &\leq \|\pi(a_t)v_{\hbox{\fiverm K}}\|\cdot 
997: \|\pi(g)v_{\hbox{\fiverm K}}\|\leq \|\pi(a_t)v_{\hbox{\fiverm K}}\|\cdot \| v_{\hbox{\fiverm K}}\|\cr}$$ 
998: and our claim follows from $\|\pi(a_t)v_{\hbox{\fiverm K}}\|<\infty$.  
999: Combining (2.3.1) with our claim then yields  
1000: $$\eqalign{\Theta_\pi^\omega(F)&= 
1001: \lim_{t\nearrow 1}\int_G  F(g) \ \overline{\theta_\pi(g^{-1}a_t)}\ dg\cr 
1002: & =\lim_{t\nearrow 1}\int_{G/H} f(gH)\  \overline{\theta_\pi(g^{-1}a_t)}\ dgH\ ,\cr}$$ 
1003: completing the proof of the theorem. \qed 
1004:  
1005: We finish this subsection by the following simple remark. 
1006:  
1007: \Lemma 2.3.2. Suppose that $\pi$ is irreducible. Then $\Theta_\pi$ is an eigendistribution 
1008: of the algebra $\D(G/H)$ of invariant differential operators on $G/H$. 
1009:  
1010: \Proof. Recall the surjective homomorphisms ${\cal U}(\g_\C)^{\frak h} 
1011: \to \D(G/H)$ 
1012: and ${\cal U}(\g_\C)^{\frak k}\to \D(G/K)$. We have $x_{\hbox{\fiverm H}}^{-1}H_\C x_{\hbox{\fiverm H}} 
1013: =K_\C$. Hence, ${\rm Ad}(x_{\hbox{\fiverm H}}^{-1})$ defines an isomorphism 
1014: $${\rm Ad}(x_{\hbox{\fiverm H}}^{-1})\:  
1015: {\cal U}(\g_\C)^{\frak h}\to {\cal U}(\g_\C)^{\frak k}\, .$$ 
1016: \par In order to prove the lemma it is sufficient to show  
1017: that $v_{\hbox{\fiverm H}}$ is an eigenvector for each  
1018: $d\pi^0(u)$, $u\in  {\cal U}(\g_\C)^{\frak h}$.  
1019: \par Notice that for each $\tilde u\in  {\cal U}(\g_\C)^{\frak k}$ 
1020: there exists a constant $c(\tilde u)$ such that  
1021: $d\pi^0(\tilde u)v_{\hbox{\fiverm K}}  
1022: = c(\tilde u) v_{\hbox{\fiverm H}}$.  
1023: For $u\in{\cal U}(\g_\C)^{\frak h}$ we now obtain that 
1024:  
1025: $$\eqalign{d\pi^0(u)v_{\hbox{\fiverm H}}&= 
1026: \lim_{t\nearrow 1}d\pi^0(u)\pi^0(a_t)v_{\hbox{\fiverm K}}\cr 
1027: &=\lim_{t\nearrow 1}\pi^0(a_t) (d\pi^0(\Ad (a_t^{-1})u) 
1028: v_{\hbox{\fiverm K}}\cr 
1029: &=\lim_{t\nearrow 1}\pi^0(a_t) (d\pi^0(\Ad (x_{\hbox {\fiverm H}} 
1030: )^{-1}u)v_{\hbox{\fiverm K}}\cr 
1031: &= c(\Ad(x_{\hbox {\fiverm H}}^{-1})u)v_{\hbox{\fiverm H}}\ .\cr}$$ 
1032: This completes the proof of  the lemma.\qed 
1033:  
1034: \subheadline{2.4. Principal series representations} 
1035:  
1036: \noindent 
1037: In this section we consider the case where $\pi=\pi_\lambda$ is a {\it spherical principal series representation}. 
1038: In particular we will be discuss  the dependence of $v_{\hbox{\fiverm H}}$ 
1039: on the spectral parameter $\lambda$. 
1040:  
1041: \par Let us first recall some well known  facts about the principal series representations. 
1042: For $\alpha\in \Sigma$ let $m_\alpha=\dim \g^\alpha$ and 
1043: $\rho\:={1\over 2}\sum_{\alpha\in \Sigma^+} m_\alpha \alpha$. Write 
1044: $M=Z_K(\a)$ and denote by $\kappa\: G\to K$ and $a\: G\to A$ the projections onto $K$, resp. $A$, 
1045: associated to the Iwasawa decomposition $G=NAK$. Note that $a$ and $\kappa$ have 
1046: unique holomorphic extension 
1047: to $\tilde\Xi$ also denoted by $a$ and $\kappa$ (cf.\ (1.2.1) and [KS01a]). As we are assuming that 
1048: $G\subseteq G_\C$ with $G_\C$ simply connected and $H=G^\tau$, we have $M=Z_H(A)$. In particular 
1049: $M\subseteq H\cap K$. 
1050:  
1051: \par Define a minimal parabolic subgroup of $G$ by $P_{\rm min}=MAN$. 
1052: For $\lambda\in \a_\C^*$ let 
1053: $${\cal D}_\lambda=\{f\in C^\infty(G)\: (\forall g\in G)(\forall man\in P) 
1054: \ f(mang)=a^{\rho-\lambda} f(g)\}\ .$$ 
1055: The group $G$ acts on ${\cal D}_\lambda$ by right translation. Denote the 
1056: corresponding representation by $\pi_\lambda^\infty$, i.e., $\pi_\lambda^\infty(g)f(x)=f(xg)$. 
1057: Denote by $\tilde{\cal H}_\lambda$ the completion of $\cal D_\lambda$ in the norm 
1058: corresponding to the inner product 
1059: $$\la f,g\ra =\int_K f(k)\overline{g(k)}\, dk\, .$$ 
1060: Then $\pi_\lambda^\infty$ extends to a representation $\pi_\lambda$  of $G$ in ${\cal H}_\lambda$.  
1061: We refer to $(\tilde {\cal H}_\lambda ,\pi_\lambda)$ as the {\it spherical principal series representation of $G$ 
1062: with parameter $\lambda$}. The principal series representations $(\pi_\lambda, \tilde {\cal H}_\lambda)$ 
1063: are admissible and of finite length; they are irreducible for generic 
1064: parameters $\lambda$ and unitary if $\lambda\in i\a^*$. It is also well known that 
1065: $\tilde{\cal H}_\lambda^\infty =\cal{D}_\lambda$ excusing our above notation $\pi_\lambda^\infty$. 
1066:  
1067: \par The restriction map $\tilde{\cal{H}}_\lambda\ni f\mapsto f\res_{K}\in L^2(K)$ is 
1068: injective by the left $NA$-covariance of $f$. Furthermore $f\res_{K}$ is left $M$-invariant. Hence 
1069: the restriction map defines an isometry $\tilde{\cal H}_\lambda\hookrightarrow L^2(M\bs K)$. On 
1070: the other hand if $F\in L^2(M\bs K)$ then we can define $f\in  \tilde{\cal H}_\lambda$ by 
1071: $f(nak)=a^{\rho -\lambda}F(Mk)$. Hence 
1072: $\tilde{\cal H}_\lambda \simeq L^2(M\bs K)$. In this realization we 
1073: have 
1074: $$[\pi_\lambda(g)f](Mk)=a(kg)^{\rho -\lambda}f(M\kappa (kg))\, .$$ 
1075: Hence the Hilbert space is the same for all $\lambda$ but the formula for the representation 
1076: depends on $\lambda$. 
1077: Notice that for $k\in K$ this simplifies to 
1078: $(\pi_\lambda(k)f)(Mx)= f(Mxk)$.  We write ${\cal H}_\lambda =L^2(M\bs K)$ to indicate 
1079: the role of $L^2(M\bs K)$ as the representation space for $\pi_\lambda$ and call it 
1080: the {\it compact realization of} $\pi_\lambda$. 
1081: As a consequence of this discussion, we see that $v_{{\hbox{\fiverm K}, \scriptscriptstyle{\lambda}}}=\1_{M\bs K}$ is a 
1082: normalized $K$-fixed vector in 
1083: ${\cal H}_\lambda$ and in fact ${\cal H}_\lambda^K=\C v_{{\hbox{\fiverm K}, \scriptscriptstyle{\lambda}}}$. 
1084: We write $v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}$ 
1085: and $v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}^\omega$ 
1086: instead of $v_{\hbox{\fiverm H}}$ and $v_{\hbox{\fiverm H}}^\omega$ to 
1087: indicate the dependence of $\lambda$. 
1088:  
1089: \par We have $\pi_\lambda^0=\pi_{-\oline \lambda}$. Thus in the compact realization we have 
1090:  
1091: $$[\pi_\lambda^0(a_t)v_{{\hbox{\fiverm K}, \scriptscriptstyle{\lambda}}}](Mk)=a(ka_t)^{\rho+\oline \lambda}\leqno(2.4.1)$$ 
1092: for all $0\leq t<1$ and so 
1093: $$\la v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}^\omega, v\ra =\lim_{t\nearrow 1} \int_{M\bs K} a(ka_t)^{\rho+\oline \lambda} 
1094: \, \oline{ v(Mk)} \ dMk \qquad (v\in {\cal H}_\lambda^\omega)\ .\leqno(2.4.2)$$ 
1095: \par In the sequel it will be importnat  
1096: that ${\cal H}_\lambda^\omega=C^\omega(M\bs K)$ is independent of $\lambda$.  
1097: The following theorem specifies the dependence of $v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}$  on $\lambda$: 
1098:  
1099: \Theorem 2.4.1. The mapping  
1100: $$\a_\C^*\to \coprod_{\lambda\in \a_\C^*} ({\cal H}_\lambda^{-\infty})^H, 
1101: \ \ \lambda\mapsto v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}$$ 
1102: is weakly anti-holomorphic in the sense that for all $v\in C^\omega(M\bs K)$ the mapping  
1103: $$\a_\C^*\to\C, \ \ \lambda\mapsto \la v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}, v\ra $$ 
1104: is anti-holomorphic.  
1105:  
1106:  
1107: \Proof. Let $v\in C^\omega(M\bs K)$. It is convenient to consider  
1108: $v$ as an $M$-invariant function on $K$. As $v$ is analytic, there exists 
1109: an open $K\times K$-invariant neighborhood ${\cal U}$ of $K$ in $K_\C$ such that  
1110: $v$ extends to a holomorphic $M$-invariant function $\tilde v$ on  ${\cal U}$.  
1111:  
1112: \par Let $\eps>0$. Then it follows from the compactness of $Ka_\eps$ and (1.2) that  
1113: we can choose $\eps>0$ small enough such that $\kappa(Ka_\eps)\subeq {\cal U}$.  
1114:  
1115: \par Now consider $v$ as an element of ${\cal H}_\lambda^\omega$. We claim  
1116: that $\pi_\lambda(a_\eps)v$ exists. In fact, using our introductory remarks,  we have   
1117:  
1118: $$[\pi_\lambda(a_\eps)v](Mk)=a(ka_\eps)^{\rho-\lambda} \tilde v(\kappa(ka_\eps))\ .$$ 
1119:  
1120: \par With (2.4.1) we now compute  
1121:  
1122: $$\eqalign{\la v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}, v\ra&=\la \pi_\lambda(a_{1-\eps})^0v_{{\hbox{\fiverm K}, \scriptscriptstyle{\lambda}}}, \pi_\lambda(a_\eps)v\ra \cr  
1123: &=\int_{M\bs K} a(ka_{1-\eps})^{\rho+\oline \lambda}\cdot  
1124:  \oline {a(ka_\eps)^{\rho-\lambda}}\cdot \oline {\tilde v(\kappa(ka_\eps))}\ dMk\ .\cr}\leqno(2.4.3)$$ 
1125: By our remarks at the beginning of the proof, we have  
1126: $$\sup_{k\in K} |\tilde v(\kappa(ka_\eps))|<\infty\ .\leqno(2.4.4)  $$ 
1127: Notice that (1.2) implies that both $a(Ka_{1-\eps})$  and $a(Ka_\eps)$ are compact  
1128: subsets of $T(\Omega)$. Thus, if  $C\subeq \a_\C^*$ is  a compact subset, then 
1129:  
1130: $$\sup_{\lambda\in C}\sup_{k\in K} |a(ka_{1-\eps})^{\rho+\oline \lambda}|<\infty,  
1131: \qquad\hbox{and}\qquad   \sup_{\lambda\in C}\sup_{k\in K} |a(ka_\eps)^{\rho -\lambda}|<\infty 
1132: \ .\leqno(2.4.5)$$ 
1133: Therefore, if we use the estimates (2.4.4) and (2.4.5), it follows from (2.4.3) that  
1134: $\lambda\mapsto \la v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}, v\ra $ 
1135: is anti-holomorphic.\qed  
1136:  
1137:  
1138:  
1139:  
1140:  
1141:  
1142:  
1143: \subheadline{2.5. Integral representation and asymptotic behaviour of $\theta_\pi$} 
1144:  
1145: \noindent 
1146: Previously we have defined a $H$-invariant holomorphic function  
1147: $\theta_\pi$ for unitary representations $\pi$. For non-unitary $\pi$ we define 
1148: $\theta_\pi$ by  
1149:  
1150: $$\theta_\pi(xK_\C)=\la \pi(x)v_{\hbox{\fiverm K}}, v_{\hbox{\fiverm H}}\ra \qquad (xK_\C\in \Xi_H)\ .$$ 
1151: Clearly, $\theta_\pi$ is a holomorphic function on $\Xi_H$. Moreover, if $\pi$ is unitary, then  
1152: $\theta_\pi$ is  $H$-invariant.  
1153:  
1154:  
1155: \par In this subsection we will give an integral representation of  
1156: the functions $\theta_\pi$ for principal series represntations $\pi$.  
1157: This  will also allow us to read off the asymptotic behaviour of $\theta_\pi$.  
1158:  
1159: \ssk Recall the definition of the spherical function $\phi_\lambda$ of parameter  
1160: $\lambda\in\a_\C^*$ by  
1161:  
1162: $$\phi_\lambda(g)=\la \pi_\lambda(g)v_{{\hbox{\fiverm K}, \scriptscriptstyle{\lambda}}},  
1163: v_{{\hbox{\fiverm K}, \scriptscriptstyle{\lambda}}}\ra=\int_K a(kg)^{\rho-\lambda}\ dk\qquad (g\in G)\ .$$ 
1164:  
1165: It follows from Theorem 2.1.2 that $\phi_\lambda$ admits a holomorphic extension 
1166: to $\tilde \Xi_H$ (or $\Xi_H$ if we wish to consider $\phi_\lambda$ as a function on $G/K$). 
1167: Also $\phi_\lambda\res_A$ extends holomorphically to the tube $T(2\Omega)=A\exp(2i\Omega)$. 
1168: Notice that 
1169: $z_{\hbox{\fiverm H}}\in T(2\Omega)$. All 
1170: mentioned holomorphic extensions of $\phi_\lambda$ are also denoted by $\phi_\lambda$. 
1171:  
1172:  
1173:  
1174: \par In the sequel we abbreviate and write $\theta_\lambda$ instead 
1175: of $\theta_{\pi_\lambda}$. The next result is  immediate 
1176: from the definitions,  Theorem 2.1.2 and the formula (2.4.1). 
1177:  
1178: \Theorem 2.5.1.  Let $(\pi_\lambda, {\cal H}_\lambda)$ be a principal series representation  
1179: with parameter $\lambda\in \a_\C^*$. Then for all $xK_\C\in \Xi_H$ we 
1180: have 
1181: $$\eqalign{ 
1182: \theta_\lambda(xK_\C)&=\lim_{t\nearrow 1} \int_K a(kx)^{\rho-\lambda} \oline{a(ka_t)^{\rho+\oline \lambda}} \ dk\cr 
1183: &=\lim_{t \nearrow 1}\varphi_\lambda (a_tx)\, .}$$ 
1184: Furthermore, 
1185: $$\theta_\lambda(aK_\C)=\phi_\lambda(x_{\hfH}a)$$ 
1186: for all $a\in T(\Omega)=A\exp (i\Omega)$. Here 
1187: $\phi_\lambda$ denotes the holomorphically extended spherical function to $T(2\Omega)$. 
1188:  
1189: \Proof. Let $x\in\Xi_H$. Then we have 
1190: $$\eqalign{ 
1191: \theta_\lambda (x)&=\la \pi_\lambda (x)v_{\hfK},v_{\hfH}\ra\cr 
1192: &=\lim_{t\nearrow 1}\la \pi_\lambda (x)v_{\hfK},\pi^0_{\lambda} (a_t)v_{\hfK}\ra\cr 
1193: &=\lim_{t\nearrow 1}\int_K a (kx)^{\rho -\lambda}\, \overline{a(ka_t)^{\rho +\overline{\lambda}}}\, dk\, .\cr }$$ 
1194: But we can also write the third line as 
1195: $$\eqalign{ 
1196: \theta_\lambda (x)&=\lim_{t\nearrow 1}\la \pi_\lambda (x)v_{\hfK},\pi^0_{\lambda}(a_t)v_{\hfK}\ra \cr 
1197: &=\lim_{t\nearrow 1}\la \pi_\lambda(a_tx)v_{\hfK},v_{\hfK}\ra\cr 
1198: &=\lim_{t\nearrow 1}\varphi_\lambda (a_t x) 
1199: \, .\cr}$$ 
1200: The last statement follows now from Theorem 2.1.2, part (ii). 
1201: \qed 
1202: To discuss the asymptotic expansions of $\theta_\lambda$ along a 
1203: positive  Weyl chamber we first have to recall 
1204: some facts on the Harish-Chandra expansion of the spherical functions. 
1205: For that let $\a_+=\{ X\in \a\: (\forall \alpha\in \Sigma^+) \ \alpha(X)>0\}$ and set 
1206: $A^+=\exp(\a_+)$. Further we define $\Lambda=\N_0[\Sigma^+]$. 
1207: If $\mu\in \Lambda$, then we define a meromorphic function 
1208: $\Gamma_\mu(\lambda)$ in the parameter $\lambda\in \a_\C^*$ by 
1209: $\Gamma_0(\lambda)=1$ and then recursively by 
1210:  
1211: $$\Gamma_\mu(\lambda)={2\over \la \mu, \mu-\lambda\ra} \sum_{\alpha\in \Sigma^+} 
1212: m_\alpha\sum_{k\in \N} \Gamma_{\mu-2k\alpha} \la \mu+\rho-2k\alpha-\lambda,\alpha\ra\ .$$ 
1213: We call $\lambda\in \a_\C^*$ generic if $\Gamma_\mu(\cdot)$ is holomorphic 
1214: at $\lambda$ for all $\mu\in \Lambda$. 
1215: For generic $\lambda\in \a_\C^*$ we define the Harish-Chandra $\Phi$-function 
1216: on $A^+$ by 
1217: $$\Phi_\lambda(a)=a^{\lambda-\rho} \sum_{\mu\in \Lambda} \Gamma_\mu(\lambda) a^{-\mu} \qquad (a\in A^+)\ .$$ 
1218: This series is locally absolutely convergent. In particular, we see that 
1219: $\Phi_\lambda$ extends to a holomorphic function on 
1220: $A^+\exp(2i\Omega)\subeq A_\C$ which we also denote by $\Phi_\lambda$. 
1221: Finally, with $\c(\lambda)$ the familiar Harish-Chandra $\c$-function on $G/K$, we have 
1222: for all generic parameters $\lambda\in \a_\C^*$ that 
1223: $$\phi_\lambda(a)=\sum_{w\in {\cal W}} \c(w\lambda)\Phi_ {w\lambda}(a)\qquad (a\in A^+)\ .$$ 
1224: Combining these facts with Theorem 2.5.1 we now obtain that: 
1225:  
1226: \Theorem 2.5.2. Let $\lambda\in \a_\C^*$ be a generic parameter. Then the following 
1227: assertions hold. 
1228:  
1229: \item{(i)} For $a\in A^+\exp(i\Omega)$ we have 
1230: $$\theta_\lambda(aK_\C)=\phi_\lambda(z_{\hbox{\fiverm H}}a)=\sum_{w\in W} \c(w\lambda) (z_{\hbox{\fiverm H}}a)^{w\lambda 
1231: -\rho}\Phi_{w\lambda}(z_{\hbox{\fiverm H}}a)\ .$$ 
1232: \item{(ii)}  Suppose that $\la \Re\lambda,\alpha\ra>0$ for all $\alpha\in \Sigma^+$. 
1233: Fix $Y\in\a_+$.  Then  
1234: $$\lim_{t\to \infty} e^{t(\rho -\lambda)(Y)} \theta_\lambda(\exp(tY)K_\C) = 
1235: \c(\lambda)\cdot z_{\hbox{\fiverm H}}^{\lambda -\rho} \ .$$\qed 
1236:  
1237:  
1238: \subheadline{2.6. $H$-orbit coefficients of the holomorphic distribution vector} 
1239:  
1240: \noindent 
1241: As we have remarked already earlier the space $({\cal H}_\lambda^{-\infty })^H$ 
1242: has dimension $\left |{\cal W}/ {\cal W}_0\right|$ for generic $\lambda$. 
1243: One can parametrize  $({\cal H}_\lambda^{-\infty })^H$ through the open 
1244: $H$-orbits in the flag manifold $P_{\rm min}\bs G$. These orbits haven been parametrized 
1245: by Rossmann and Matsuki (cf. [M79]); they are given by 
1246: $$P_{\rm min} wH \qquad (w\in  {\cal W}/ {\cal W}_0)\ .$$ 
1247:  
1248: \par For $\lambda\in \a_\C^*$ and $w\in  {\cal W}/ {\cal W}_0$ define 
1249: a right $H$-invariant function on $G$ by 
1250:  
1251: $$\eta_{\lambda, w}(x)=\cases{ a^{\rho+\oline \lambda} & for $x=manwh\in MAN wH$ \cr 
1252: 0 & otherwise\ .\cr}$$ 
1253:  
1254: For $\lambda\in \a^*$ we will use the notation $\lambda <<0$ if 
1255: $\la \lambda,\alpha\ra<<0$ for all $\alpha\in \Sigma^+$. 
1256: Then it is known that for $\lambda<<0$ the functions $\eta_{\lambda, w}$ are 
1257: continuous and define $H$-fixed distribution vectors of $\pi_\lambda$ 
1258: [\'O87]. 
1259: Moreover,  the distributions $\eta_{\lambda, w}$ admit continuation 
1260: in $\lambda$ to a weakly anti-meromorphic function on $\a_\C^*$. For generic $\lambda\in \a_\C^*$ 
1261: we have 
1262: $$({\cal H}_\lambda^{-\infty})^H=\bigoplus_{w\in {\cal W}/{\cal W}_0} \C \eta_{\lambda,w}$$ 
1263: and the mapping 
1264: $$j_\lambda\: \C^{\left |{\cal W}/{\cal W}_0\right|}\to ({\cal H}_\lambda^{-\infty})^H, 
1265: \ \ (c_w)_w\mapsto \sum_{w\in {\cal W}/{\cal W}_0} c_w \eta_{\lambda,w}$$ 
1266: is a bijection for generic $\lambda$, weakly anti-meromorphic in $\lambda$ [vdB88]. 
1267: For $\lambda<<0$ the inverse of $j_\lambda$ 
1268: is given by the evaluation mapping 
1269: $${\rm ev}\: ({\cal H}_\lambda^{-\infty})^H\to \C^{\left |{\cal W}/{\cal W}_0\right|}, 
1270: \ \ \eta\mapsto (\eta(w))_w\ .$$ 
1271:  
1272: \par On the other hand we know that the weakly holomorphic distribution 
1273: vector $v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}$ 
1274: depends weakly anti-holomorphically on  $\lambda\in \a_\C^*$ (cf.\ Theorem 2.4.1). 
1275: The next theorem gives us the coefficients of $v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}$ 
1276: in terms of the basis $(\eta_{\lambda,w})$ of $({\cal H}_\lambda^{-\infty})^H$. 
1277: We note that for $\lambda<<0$ the distribution $v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}$ is given through 
1278: the bounded measurable function 
1279:  
1280: $$v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}(k)=\lim_{t\nearrow 1} a(ka_t)^{\rho+\oline \lambda}\ .\leqno(2.6.1)$$ 
1281:  
1282:  
1283: \Theorem 2.6.1. For generic parameters $\lambda\in \a_\C^*$ we have 
1284:  
1285: $$v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}=\sum_{w\in {\cal W}/ {\cal W}_0} 
1286:  z_{\hbox{\fiverm H}}^{w^{-1}(\rho +\oline\lambda)}\cdot \eta_{\lambda,w}\ .$$ 
1287:  
1288: \Proof. For $\lambda\in \a_\C^*$ generic write 
1289: $$v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}=\sum_{w\in {\cal W}/ {\cal W}_0} c_{\lambda, w} \cdot  \eta_{\lambda,w}$$ 
1290: for the basis expansion. As  the coefficients $  c_{\lambda, w}$ depend weakly 
1291: anti-meromorphically on $\lambda$, it is sufficient to show that 
1292: $c_{\lambda, w}=z_{\hbox{\fiverm H}}^{w^{-1}(\rho 
1293: +\oline \lambda)}$ 
1294: for $\lambda<<0$. Then (2.6.1) implies that 
1295:  
1296: $$c_{\lambda,w}=v_{\hfH}(w)=\lim_{t\nearrow 1} v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}(wa_t)=a(wz_{\hbox{\fiverm H}})^{\rho+\oline \lambda}= 
1297: z_{\hbox{\fiverm H}}^{w^{-1}(\rho+\oline \lambda)}\ , $$ 
1298: as was to be shown. \qed 
1299:  
1300: \Remark 2.6.2. Let us go back to Example 2.1.5 for $G=\Sl(2,\R)$. Here 
1301: ${\cal W}_0=\{\1\}$ and so ${\cal W}/{\cal W}_0=\{ \1, w\}$ where $w$ is the non-trivial element in 
1302: the Weyl group which acts by multiplication by $-1$. 
1303: The distributions $v_{\hbox{\fiverm H,1}}$ and  $v_{\hbox{\fiverm H,2}}$ from Example 2.1.5 are given in 
1304: the above notation by  
1305:  
1306:   
1307: $$v_{\hbox{\fiverm H,1}}=\eta_{\lambda,\1}\qquad\hbox{and}\qquad v_{\hbox{\fiverm H,2}}=\eta_{\lambda,w}\ .$$ 
1308: In Example 2.1.5 we did show that 
1309: $$v_{\hbox{\fiverm H}}=v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}=e^{i{\pi\over 4} 
1310: (1+\overline{\lambda})}v_{\hbox{\fiverm H,1}}+ e^{-i{\pi\over 4}(1+\overline{\lambda})} v_{\hbox{\fiverm H,2}}\. $$ 
1311: As $z_{\hbox{\fiverm H}}^{\rho+\overline{\lambda}}=e^{i{\pi\over 4}(1+\overline{\lambda})}$, we hence see that 
1312: the above formula is a special case of Theorem 2.6.1.\qed 
1313:  
1314:  
1315: \subheadline{2.7. Transformation under the intertwining matrix} 
1316:  
1317: \noindent 
1318: For $w\in {\cal W}$ and $\lambda\in \a_\C^*$ generic  we have an intertwining operator 
1319: $$A(\lambda, w\lambda)\: (\pi_\lambda, {\cal H}_\lambda^\infty)\to (\pi_{w\lambda}, 
1320: {\cal H}_{w\lambda}^\infty)\ .$$ 
1321: Notice that the Hilbert space ${\cal H}_\lambda$ is independent of $\lambda$ because 
1322: we use the compact realzation. We can therefore speak about meromorphic maps 
1323: from $\a_\C^*$ into the space of bounded operators from 
1324: ${\cal H}_\lambda$ into ${\cal H}_\mu$.  In this sense 
1325: it is well known that the map $\a_\C^*\ni \lambda\mapsto  A(\lambda, w\lambda)$ is meromorphic. 
1326: Dualizing, we obtain an anti-meromorphic family of intertwining operators 
1327: $$A(\lambda, w\lambda)^*\: {\cal H}_{w\lambda}^{-\infty}\to {\cal H}_\lambda^{-\infty}\ . $$ 
1328: Restricting $A(\lambda ,w\lambda)^*$ to the space of $H$-invariant distribution 
1329: vectors we obtain a linear bijection, say 
1330:  
1331: $$A_H(\lambda, w\lambda)^*\: ({\cal H}_{w\lambda}^{-\infty})^H\to ({\cal H}_\lambda^{-\infty})^H\ .$$ 
1332: Often one refers to  $A_H^*(\lambda, w\lambda)$ as the {\it intertwining  matrix}. 
1333: In terms of the basic distribution vectors $(\eta_{w\lambda, w'})_{w'}$ respectively 
1334: $(\eta_{\lambda, w'})_{w'}$ 
1335: the operator  $A_H(\lambda, w\lambda)^*$ has an unknown, seemingly complicated expression. 
1336: In this section we will show that the the intertwining  matrix 
1337: maps the holomorphic distribution vector 
1338: $v_{{\hbox{\fiverm H}, \scriptscriptstyle{w\lambda}}}$ 
1339: to a multiple of $v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}$. 
1340: In order to describe this multiple more precisely we need more notations. 
1341:  
1342: \par For $w\in {\cal W}$ define a subgroup of $\oline N=\theta(N)$ by 
1343: $$\oline N_w=\oline N\cap wNw^{-1}\ .$$ 
1344: For $\Re \lambda<<0$ we define functions 
1345: $$\c_w(\lambda)=\int_{\oline N_w} a(\oline n)^{\rho-\lambda} \ d\oline n\ .$$ 
1346: If $w=w_0$ is the longest element in ${\cal W}$, then we write 
1347: $\c(\lambda)$ instead of $\c_{w_0}(\lambda)$ and remark that 
1348: $\c(\lambda)$ is the familiar Harish-Chandra $c$-function on $G/K$. 
1349: The functions $\c_w(\lambda)$ admit meromorphic continuation to $\a_\C^*$ 
1350: and can be explicitely computed (Gindikin-Karpelevic formula). 
1351:  
1352: \par With this notation the intertwinig operators $A(\lambda,w\lambda)$  for $\lambda<<0$ 
1353: are defined by 
1354: $$[A(\lambda, w\lambda)f](x) =\int_{\oline N_w} f(\oline n w x) \ d\oline n 
1355: \qquad (f\in {\cal D}_\lambda, x\in G)\ .$$ 
1356:  
1357:  
1358: \Theorem 2.7.1. Assume that $\lambda\in \a_\C^*$ is generic. Then 
1359: $$A_H(\lambda, w\lambda)^*v_{{\hbox{\fiverm H}, \scriptscriptstyle{w\lambda}}} 
1360: = \c_w(\oline\lambda)\cdot v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}\ .$$ 
1361:  
1362:  
1363: \Proof. It is well known - and follows immediately from 
1364: the definition -  that  $A(\lambda, w\lambda)v_{{\hbox{\fiverm K}, \scriptscriptstyle{\lambda}}}=\c_w(\lambda) 
1365: v_{{\hfK,w\lambda}}$. But then, as $A(\lambda, w\lambda )^*v_{\hfK,w\lambda}$ is $K$-invariant, we also get 
1366: $$\eqalign{ 
1367: \la A(\lambda ,w\lambda )^*v_{\hfK,w\lambda},v_{\hfK,\lambda}\ra &= \la v_{\hfK,w\lambda},A(\lambda,w\lambda)v_{\hfK,\lambda}\ra\cr 
1368: &=\la v_{\hfK,w\lambda },\c_w(\lambda )v_{\hfK,w\lambda}\ra\cr 
1369: &=\overline{\c_w (\lambda )}\, .\cr} 
1370: $$ 
1371: Noticing that $\overline{\c_w(\lambda)}=\c_w(\overline{\lambda})$ it follows that 
1372: $$A(\lambda ,w\lambda )^*v_{\hfK,w\lambda}=\c_w(\overline{\lambda})v_{\hfK,\lambda}\, .$$ 
1373: Finally, using that $A(\lambda ,w\lambda )^*$ is an intertwining operator, we get for 
1374: a $K$-finite  $u$ : 
1375: $$\eqalign{ 
1376: \la A_H^*(\lambda ,w\lambda)v_{\hfH,w\lambda},u\ra &= 
1377: \la v_{\hfH,w\lambda},A(\lambda ,w\lambda )u\ra\cr 
1378: &=\lim_{t\nearrow 1}\la \pi_{w\lambda }^0(a_t)v_{\hfK,w\lambda},A(\lambda,w\lambda )u\ra\cr 
1379: &=\lim_{t\nearrow 1}\la v_{\hfK,w\lambda},\pi_{w\lambda }(a_t)^{-1}A(\lambda,w\lambda )u\ra\cr 
1380: &=\lim_{t\nearrow 1}\la v_{\hfK,w\lambda},A(\lambda,w\lambda )\pi_{w\lambda }(a_t)^{-1}u\ra\cr 
1381: &=\lim_{t\nearrow 1}\la A(\lambda,w\lambda )^*v_{\hfK,w\lambda},\pi_{w\lambda }(a_t)^{-1}u\ra\cr 
1382: &= \c_w(\overline{\lambda})\lim_{t\nearrow 1}\la \pi_{\lambda }(a_t)v_{\hfK,\lambda },u\ra\cr 
1383: &=\c_w (\overline{\lambda}) \la v_{\hfH,\lambda },u\ra \, .\cr} 
1384: $$ 
1385: Hence $A_H(\lambda,w\lambda)^*v_{{\hbox{\fiverm H}, \scriptscriptstyle{w\lambda}}}= \c_w(\oline \lambda)\cdot 
1386: v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}$, as was to be shown. \qed 
1387:  
1388:  
1389: \subheadline{2.8. Relation to the horospherical picture} 
1390:  
1391: \noindent 
1392: In this subsection we explain the construction of the 
1393: holomorphic distribution vector $v_{\hbox{\fiverm H}}$ from 
1394: the horospherical point of view. 
1395:  
1396: \par Let $\kappa\in \a_\C^*$ be a complex parameter. 
1397: We call the holomorphic function 
1398: $$ \tilde \psi _\kappa(nak)= a^\kappa=\exp(< \kappa , \log a>), 
1399: \qquad n\in N_{\Bbb C}, a\in T(\Omega), k\in K_\Bbb C$$ on $\tilde 
1400: \Xi $ (see\ (1.2.1)) the {\it holomorphic horospherical function 
1401: with parameter $\kappa$}. 
1402:  
1403: \par This function can be pushed down to $\Xi$ as a holomorphic locally $N_\C$-invariant function 
1404: $\psi_\kappa$. The function $\psi_\kappa$ is a holomorphic extension of the usual 
1405: horospherical ($N$-invariant, $A$-homogeneous) function on $G/K$, 
1406: corresponding to spherical principal representations related to $\kappa$.  
1407: For certain  values of the parameter $\kappa$ the function $\psi_\kappa$ has boundary distribution values 
1408: $\psi_{\kappa, \hbox{\fiverm H}}$ on $G/H$. To 
1409: understand the structure of those distributions, let us remark that a Zariski open part 
1410: of G/H is the disjoint union of domains $Y_j= NAy_j$ where $y_1, 
1411: \cdots, y_k$ correspond to the vertices of $\Omega_H$ 
1412: (${\cal W}$-equivalent). On each $Y_j$ we have an $N$-invariant, $A$-homogeneous  
1413: distribution with parameter $\kappa$ (such 
1414: distributions on $Y_j$ are unique up to a multiplicative 
1415: constant). 
1416:  
1417: The function $\tilde \psi_\kappa$  can be also pushed down on a domain 
1418: $D_H$ in $N_\C M_\C \bs G_\C$ as a holomorphic function 
1419: $\psi_\kappa$. This function with parameters holomorphically 
1420: extends $K$-invariant vectors in the principal spherical 
1421: representations on $MN\bs G $. The boundary values of this holomorphic 
1422: function give an $H$-invariant distribution, the domains $Y_j$ 
1423: correspond to the $H$-orbits on $MN\bs G$  and we have the corresponding 
1424: decomposition of $\psi_\kappa$. 
1425:  
1426:  
1427:  
1428:  
1429:  
1430: \sectionheadline{3. An application: Hardy spaces for NCC symmetric spaces} 
1431:  
1432: \noindent 
1433: In this section we apply  our theory 
1434: developed in Section 2 to associate 
1435: to every NCC symmetric space $G/H$ a Hardy space ${\cal H}^2(\Xi_H)$. 
1436: The Hardy space is a $G$-invariant Hilbert space 
1437: of holomorphic functions on $\Xi_H$ featuring a boundary value mapping which gives 
1438: an isometric embedding of  ${\cal H}^2(\Xi_H)$ into the most-continuous 
1439: spectrum $L^2(G/H)_{\rm mc}$ of $L^2(G/H)$. Hence we give a 
1440: realization of a part of $L^2(G/H)_{\rm mc}$ in a space 
1441: of holomorphic function on $\Xi_H$, generalizing and extending our 
1442: previous results from [GK\'O01] to all NCC spaces. 
1443:  
1444: \par This section is organized as follows: After a brief digression on 
1445: Hilbert spaces of holomorphic functions on $\Xi_H$, we  give 
1446: an adhoc definition of the Hardy space through the spectral measure. Then, 
1447: after recalling the theory of the most-continuous spectrum, we will show that 
1448: there is a boundary value mapping embedding ${\cal H}^2(\Xi_H)$ isometrically 
1449: into $L^2(G/H)_{\rm mc}$. 
1450:  
1451:  
1452:  
1453: \subheadline{3.1. $G$-invariant Hilbert spaces of holomorphic functions on $\Xi_H$} 
1454:  
1455: \noindent 
1456: In this section we briefly recall the abstract theory of $G$-invariant 
1457: Hilbert spaces of holomorphic functions, specialized to the complex manifold $\Xi_H$ 
1458: (see  also [FT99] and [K99] for the general theory). 
1459:  
1460: \par In the sequel we will consider ${\cal O}(\Xi_H)$ as a Fr\'echet space with the 
1461: topology of compact convergence. We let $G$ act on ${\cal O}(\Xi_H)$ by the 
1462: left regular representation $L$: 
1463: $$(L(g)f)(z)=f(g^{-1}z)\qquad (g\in G, f\in {\cal O}(\Xi_H), z\in \Xi_H)\ .\leqno(3.1.1)$$ 
1464: By a {\it $G$-invariant Hilbert space 
1465: of holomorphic functions on $\Xi_H$} we understand a Hilbert space ${\cal H}\subeq {\cal O}(\Xi_H)$ 
1466: such that: 
1467:  
1468:  
1469: \ssk 
1470: \item{(IH1)} The inclusion ${\cal H}\into {\cal O}(\Xi_H)$ is continuous. 
1471: \item{(IH2)} The Hilbert space ${\cal H}$ is invariant under $L$ and the the 
1472: corresponding representation of $G$ is unitary. 
1473: \ssk 
1474: It follows from (IH1) that for every $z\in \Xi_H$ the  point 
1475: evaluation ${\cal H}\to \C, \ f\mapsto f(z)$, 
1476: is continuous. Thus, there exists a ${\cal K}_z\in {\cal H}$ such that 
1477: $\la f, {\cal K}_z\ra =f(z)$ holds for every $f\in {\cal H}$. In 
1478: this way we obtain a function 
1479: $${\cal K}\: \Xi_H \times \Xi_H\to \C, \ \ (z,w)\mapsto 
1480: {\cal K}(z,w)=\la {\cal K}_w, {\cal K}_z\ra\ .$$ 
1481: The function ${\cal K}$  is holomorphic in the first variable and anti-holomorphic in the second variable. 
1482: It follows from (IH2) that ${\cal K}$ is $G$-invariant, i.e., ${\cal K}(gz,gw)= 
1483: {\cal K}(z,w)$ holds for all $g\in G$ and all $z,w\in \Xi_H$. 
1484: We call ${\cal K}$ the {\it Cauchy-Szeg\"o kernel} of ${\cal H}$ and note that 
1485: ${\cal H}$ is determined by ${\cal K}$. 
1486:  
1487: \msk  To describe the spectral resolution of ${\cal K}$ denote by $\hat G_s$  the $K$-spherical unitary dual of $G$. 
1488: We view $\hat G_s$ as a subset of $\a_\C^*/ {\cal W}$ using the 
1489: parametrization of the spherical principal series. Notice that the topology on 
1490: $\hat G_s$ coincides with the topology induced from  
1491: $\a_\C^*/{\cal W}$. 
1492: Slightly abusing our notation from Subsection 2.4, we denote by 
1493: $(\pi_\lambda, {\cal H}_\lambda)$ a representative of $\lambda\in \hat G_s$.  
1494: For $\lambda\in \hat G_s$ define the $G$-invariant kernel ${\cal K}_\lambda$ by 
1495: $${\cal K}_\lambda(xK_\C, yK_\C)=\la \pi_\lambda(\oline y)v_{{\hbox{\fiverm K}, \scriptscriptstyle{\lambda}}}, 
1496:  \pi_\lambda(\oline x)v_{{\hbox{\fiverm K}, \scriptscriptstyle{\lambda}}}\ra 
1497: \qquad (xK_\C, yK_\C\in \Xi_H)\ .$$ 
1498: Write $\Xi_H^{\rm opp}$ for $\Xi_H$ but endowed with the  
1499: opposite complex structure. We recall that the map  
1500: $$\hat G_s \to {\cal O}(\Xi_H\times \Xi_H^{\rm opp}),  
1501: \ \ \lambda\mapsto {\cal K}_\lambda$$ 
1502: is continuous [KS01b, Sect. 5]  
1503: (it follows from the fact that the spherical  
1504: functions $\phi_\lambda$ and their holomorphic continuations   
1505: are continuous in $\lambda$).  
1506: Then by [KS01b, Th.\ 5.1] there exists a unique Borel measure $\mu$ on 
1507: $\hat G_s$ such that 
1508:  
1509: $${\cal K}(z,w)=\int_{\hat G_s} {\cal K}_\lambda(z,w)\ d\mu(\lambda) \qquad (z,w\in \Xi_H)\leqno(3.1.2)$$ 
1510: with the  right hand side converging absolutely on compact subsets of $\Xi_H\times \Xi_H$. 
1511: Equivalently phrased, the mapping 
1512:  
1513: $$\Phi\: \int_{\hat G_s}^\oplus {\cal H}_\lambda\ d\mu(\lambda)\to {\cal H} 
1514: , \ \ s=(s_\lambda)_\lambda\mapsto \left(xK_\C\mapsto \int_{\hat G_s} \la \pi_\lambda(x^{-1})s_\lambda, 
1515: v_{{\hbox{\fiverm K}, \scriptscriptstyle{\lambda}}}\ra \ d\mu(\lambda)\right)\leqno (3.1.3)$$ 
1516: is a $G$-equivariant unitary isomorphism. In the sequel we refer to the measure $\mu$ as 
1517: the {\it Plancherel measure} of ${\cal H}$. 
1518:  
1519: \ssk  In [KS01b] 
1520: a criterion was given on a Borel measure $\mu$ on $\hat G_s$ to be a Plancherel measure 
1521: for an invariant Hilbert space ${\cal H}={\cal H}(\mu)$ on $\Xi$. This criterion can be easily adapted to 
1522: invariant Hilbert spaces on $\Xi_H$. Let us provide the necessary modifications. 
1523: \par Define a norm $\|\cdot\|_{\hbox{\fiverm H}}$ on $\a_\C^*$ by 
1524:  
1525: $$\|\lambda\|_{\hbox{\fiverm H}}\:=\sup_{w\in {\cal W}/{\cal W}_0} |\lambda(wX_{\hbox{\fiverm H}})|\qquad (\lambda\in \a_\C^*) \ .$$ 
1526: Then [KS01b, Prop. 5.4] and its proof readily gives the following generalization: 
1527:  
1528: \Proposition 3.1.1. Let $\mu$ be a Borel measure on $\hat G_s$ with the property 
1529: $$(\forall 0\leq c<2)\qquad \int_{\hat G_s} e^{c\|\Im \lambda\|_{\hbox{\fiverm H}}}\ d\mu(\lambda)<\infty\ .\leqno(3.1.4)$$ 
1530: Then $\mu$ is the  Plancherel measure of an  
1531: invariant Hilbert space ${\cal H}(\mu)$ on $\Xi_H$. \qed 
1532:  
1533: \subheadline{3.2. The definition of the Hardy space} 
1534:  
1535: \noindent 
1536: We are now ready to give the definition of the Hardy space on $\Xi_H$. 
1537: Let us denote by $i\a_+^*$ an open Weyl chamber in  $i\a^*$. In the sequel we will 
1538: consider $i\a_+^*$ mainly as a subset of $\hat G_s$. Let 
1539: $\cal O$ be a neighborhood of $i\a^*$ such that $\sum_{w\in {\cal W}/ {\cal W}_0} z_{\hbox{\fiverm H}}^{-2w^{-1}\lambda}$ 
1540: %Change GO: Added a minus sign
1541: has a holomorphic square root 
1542: ${\bf z}_{\hfH}(\lambda)$ on $\cal O$. Define a holomorphic function $\cH$ on 
1543: $\cal O$ by 
1544: $$\cH (\lambda)=\c(\lambda)\cdot {\bf z}_{\hfH}(\lambda )$$
1545: and define a Borel measure $\mu$ on $i\a_+^*$ by 
1546: $$d\mu (\lambda )=\frac{d\lambda}{|\cH (\lambda )|^2} \leqno(3.2.1)$$
1547: where $d\lambda$ denotes the Lebesgue measure. Then we have:
1548: 
1549: 
1550: \Lemma 3.2.1. The measure $\mu$ satisfies the  condition (3.1.4); in particular
1551: $\mu$ is the Plancherel measure of an invariant Hilbert space ${\cal H}(\mu)$ on $\Xi_H$.
1552: 
1553: 
1554: \Proof. Recall the growth behaviour of the $\c$-function on the imaginary axis: There
1555: exists constants $C, N>0$ such that
1556: $$(\forall \lambda\in i\a^*)\qquad {1\over |\c(\lambda)|^2} \leq C (1+|\lambda|^N)\ . $$
1557: Moreover for $\lambda \in i\a^*$ one has 
1558: $z_{\hbox{\fiverm H}}^{\lambda}=e^{\lambda (iX_{\hbox{\fiverm H}})}>0$. 
1559: Hence
1560: $$\sum_{w\in {\cal W}/ {\cal W}_0} |z_{\hbox{\fiverm H}}^{-w^{-1}\lambda}|^2\geq
1561: e^{2\|{\rm Im}\lambda\|_{\hbox{\fiverm H}}}\. $$
1562: %Chane GO: I made several changes here in the proof.
1563: 
1564: Combining these two facts now yields that $\mu$ satisfies (3.1.4).\qed
1565: 
1566: Using Proposition 3.1.1 and Lemma 3.2.1 we now can give an adhoc-definition of the  Hardy 
1567: space on $\Xi_H$. 
1568:  
1569: \Definition 3.2.2. {\bf(Hardy space)} Let $G/H$ be a NCC symmetric space and $\Xi_H$ its associated domain 
1570: in $G_\C/ K_\C$. Then we define the {\it Hardy space} ${\cal H}^2(\Xi_H)$ on $\Xi_H$ by 
1571: $${\cal H}^2(\Xi_H)={\cal H}(\mu)$$ 
1572: with $\mu$ as in (3.2.1). \qed  
1573:  
1574:  
1575: Recall the Cauchy-Szeg\"o kernel ${\cal K}(z,w)$ of the invariant 
1576: Hilbert space ${\cal H}^2(\Xi_H)$ from Subsection 3.1. 
1577:  
1578: \Lemma 3.2.3. Let ${\cal K}$ be the Cauchy-Szeg\"o kernel of ${\cal H}^2(\Xi_H)$. 
1579: Then the limits 
1580: $$\Psi(z)=\lim_{t\nearrow 1} {\cal K}(z,a_tK_\C)\qquad (z\in \Xi_H)$$ 
1581: exist locally uniformly. In particular, $\Psi\:\Xi_H\to \C$ 
1582: is an $H$-invariant holomorphic function. 
1583:  
1584: \Proof. Fix $z\in \Xi_H$ and let $U\subeq \Xi_H$ be a compact 
1585: neighborhood of $z$. Choose $\eps>0$ small enough such that 
1586: $a_\eps U\subeq \Xi_H$. Then, by $G$-invariance, we have 
1587: $${\cal K}(z, a_tK_\C)={\cal K}(a_\eps z, a_{-\eps}a_tK_\C)= 
1588: {\cal K}(a_\eps z,a_{t-\eps}K_\C)$$ 
1589: for all $\eps <t<1$. The claim follows now, because 
1590: $]\eps, 1+\eps [\ni t\mapsto {\cal K}(a_\eps z,a_{t-\eps}K_\C)\in \C$ 
1591: is continuous, and hence 
1592: $\lim_{t\nearrow 1}{\cal K}(z, a_tK_\C)={\cal K}(a_{\eps} z, a_{1-\eps}K_\C)$ 
1593: exists and the convergence is uniform on compact subsets.\qed 
1594: %Also recall the inequality 
1595: %$|{\cal K}(z,w)|\leq \sqrt{{\cal K}(z,z)}\cdot\sqrt{{\cal K}(w,w)}$, valid 
1596: %for all positive definite kernels. 
1597: %Then we have 
1598: % 
1599: %$$\eqalign{\left|\Psi(z)\right|&=|\lim_{t\nearrow 1} {\cal K}(z, a_tK_\C)| 
1600: %=\left|{\cal K}(a_\eps z, a_{1-\eps}K_\C)\right|\cr 
1601: %&\leq \sqrt{{\cal K}(a_\eps z, a_\eps z)} \cdot \sqrt{{\cal K}(a_{1-\eps}K_\C, a_{1-\eps}K_\C )}\ .\cr} $$ 
1602: %Now the assertion follows from (3.1.2).\qed 
1603:  
1604: We refer to 
1605: $\Psi$ as the {\it Cauchy-Szeg\"o function} of ${\cal H}^2(\Xi_H)$. 
1606: As ${\cal K}$ is $G$-invariant, it follows that ${\cal K}$  
1607: can be reconstructed from $\Psi$.  
1608: Moreover, as $H_\C T(\Omega_{\hbox{\fiverm H}})K_\C/ K_\C$ 
1609: meets $\Xi_H$ in an open set, we conclude that $\Psi$ is 
1610: uniquely determined by its restriction to $T(\Omega_{\hbox{\fiverm H}})K_\C/ K_\C \subeq \Xi_H$. 
1611:  
1612: Using Theorem 2.5.2 (i) we finally obtain the spectral 
1613: resolution of $\Psi$. 
1614:  
1615:  
1616: \Theorem  3.2.4. For $a\in T(\Omega_{\hbox{\fiverm H}})$ we have 
1617:  
1618: $$\Psi(aK_\C)=\int_{i\a_+^*} \phi_\lambda(z_{\hbox{\fiverm H}}a) 
1619: \ {d\lambda\over |\cH(\lambda)|^2}, $$ 
1620: where the integrals on the right hand side converge uniformly 
1621: and absolutely on compact subsets of $T(\Omega_{\hbox{\fiverm H}})$.\qed 
1622:  
1623:  
1624: \ssk We now discuss the boundary value map $b\: {\cal H}^2(\Xi_H)\to 
1625: L^2(G/H)_{\rm mc}$. As usual, this boundary value map can be 
1626: nicely defined pointwise only on an appropriate dense subspace of ${\cal H}^2(\Xi_H)$. 
1627: Write ${\cal H}^2(\Xi_H)^\omega$ for the analytic  
1628: vectors of the left regular representation  
1629: $(L, {\cal H}^2(\Xi_H))$. Fix $f\in 
1630: {\cal H}^2(\Xi_H)^\omega$. Then for every  
1631: compact subset $C\subeq G$ there exists  
1632: an $0<\eps<1$ such that $L(a_{-\eps}g^{-1})f$ exists 
1633: for all $g\in G$. In particular, if $0<\eps\leq t<1$, then  
1634: $$f(ga_tK_\C)=f(ga_\eps a_{t-\eps} K_\C)= 
1635: [L(a_{-\eps}g^{-1})f)](a_{t-\eps}K_\C)\ .$$ 
1636: and so  
1637: $$\lim_{t\nearrow 1} f(ga_tK_\C)= 
1638: [L(a_{-\eps}g^{-1})f)](a_{1-\eps}K_\C)\ .$$ 
1639: It follows that we have a well defined  
1640: $G$-equivariant boundary value map: 
1641:  
1642: $$b^\omega\: {\cal H}^2(\Xi_H)^\omega\to C(G/H), 
1643: \ \  b^\omega(f)(gH)= 
1644: \lim_{t\nearrow 1} f(ga_tK_\C)\ \ .\leqno(3.2.2)$$ 
1645:  
1646: \msk Recall from (3.1.3) the isomorphism 
1647: $\Phi\: 
1648: \int_{i\a_+^*}^\oplus{\cal H}_\lambda\ d\mu(\lambda) 
1649: \to {\cal H}^2(\Xi_H)={\cal H}^2(\mu)$: 
1650: $$s=(s_\lambda)_\lambda\mapsto \left(xK_\C\mapsto \int_{\hat G_s} \la \pi_\lambda(x^{-1})s_\lambda, 
1651: v_{{\hbox{\fiverm K}, \scriptscriptstyle{\lambda}}}\ra \ d\mu(\lambda)\right)\, .$$ 
1652: It is 
1653: useful to have the corresponding formula for $b^\omega$ 
1654: on the space of 
1655: sections with values in 
1656: $\left(\int_{i\a_+^*}^\oplus{\cal H}_\lambda\ d\mu(\lambda)\right)^\omega$, i.e., on 
1657: the space of 
1658: analytic sections. 
1659: In this regard, it is better to replace 
1660: ${\cal H}^2(\Xi_H)^\omega$ by some smaller  
1661: but dense subspace ${\cal H}^2(\Xi_H)_0$.  
1662: In order to define  ${\cal H}^2(\Xi_H)_0$ 
1663: we have to introduce some terminology.  
1664: For a section $s=(s_\lambda)_\lambda\in  
1665: \int_{i\a_+^*}^\oplus{\cal H}_\lambda\  
1666: d\mu(\lambda)$ we define its support  
1667: by $\supp(s)=\oline{\{\lambda\in i\a_+^*\: s_\lambda 
1668: \neq 0\}}$. Futhermore we shall use  
1669: the identifications ${\cal H}_\lambda^\omega= 
1670: C^\omega(M\bs K)$ for $\lambda\in i\a_+^*$.  
1671: Recall that if $f\in{\cal H}^2(\Xi_H)^\omega$ 
1672: and $s=(s_\lambda)_\lambda=\Phi^{-1}(f)$, then  
1673: almost each stalk $s_\lambda$ is an analytic  
1674: vector, i.e. $s_\lambda\in C^\omega(M\bs K)$.  
1675: The subspace ${\cal H}^2(\Xi_H)_0$ is  
1676: then defined by 
1677: $${\cal H}^2(\Xi_H)_0= 
1678: \left\{ f\in {\cal H}^2(\Xi_H)^\omega\quad : \quad \eqalign{& 
1679: f \quad \hbox {is $K$-finite, }\cr 
1680: & s=(s_\lambda)_\lambda=\Phi^{-1}(f) 
1681: \quad \hbox{has compact support, }\cr 
1682: & s\: i\a_+^*\to C^\omega(M\bs K) 
1683: \quad \hbox{is weakly smooth.}}\right\}$$ 
1684: It is an easy verification that  
1685: ${\cal H}^2(\Xi_H)_0$ is a dense subspace 
1686: of ${\cal H}^2(\Xi_H)$. Write $b_0\: {\cal H}^2(\Xi_H)_0\to 
1687: C(G/H)$ for the restriction of $b^\omega$ to ${\cal H}^2 
1688: (\Xi_H)_0$.  
1689:  
1690: \par In the sequel we will often identify a function  
1691: $f\in {\cal H}^2(\Xi_H)_0$ with its corresponding  
1692: section $s=(s_\lambda)=\Phi^{-1}(f)$.  
1693: We then claim that  
1694:  
1695: $$b_0\: {\cal H}^2(\Xi_H)_0\to C(G/H), \ \ s=(s_\lambda) 
1696: \mapsto\left(gH\mapsto\int_{i\a_+^*} 
1697: \la \pi_\lambda(g^{-1})s_\lambda, v_{{\hbox{\fiverm H}, 
1698: \scriptscriptstyle{\lambda}}}\ra \ d\mu(\lambda)\right)\leqno(3.2.3)$$ 
1699: Notice that it is a priori not even clear that the  
1700: right hand side of (3.2.3) is 
1701: well defined. To establish (3.2.3) fix  
1702: $f\in {\cal H}^2(\Xi_H)_0$ and $g\in G$.  
1703: As $f$ is an analytic  
1704: vector for the left regular representation  
1705: $(L, {\cal H}^2(\Xi_H))$ it follows that  
1706: there exists an $0<\eps<1$ such that $L(a_\eps g^{-1})f$  
1707: exists. Using standard procedures one deduces   
1708: that  
1709: $\pi_\lambda(a_\eps g^{-1})s_\lambda$ exists  
1710: for almost all $\lambda$. In particular  
1711: $s_\lambda\in {\cal H}_\lambda$  
1712: is analytic for almost all $\lambda$.  
1713:  Furthermore, 
1714: $L(a_\eps g^{-1})f$ corresponds to the section  
1715: $(\pi_\lambda(a_\eps g^{-1})s_\lambda)_\lambda$ and so 
1716: $$\|L(a_\eps g^{-1})f\|^2=\int_{i\a_+^*} 
1717: \|\pi_\lambda(a_\eps g^{-1})s_\lambda\|^2 \ d\mu(\lambda) 
1718: <\infty\ .\leqno(3.2.4)$$ 
1719:  
1720: With the  
1721: convention $\pi_\lambda(a_1) 
1722: v_{{\hbox{\fiverm K }, \scriptscriptstyle{\lambda}}} 
1723: =v_{{\hbox{\fiverm H }, \scriptscriptstyle{\lambda}}}$ 
1724: we then have for all $\eps\leq t\leq 1$ and almost all $\lambda$  
1725: the estimate 
1726: $$\eqalign{|\la \pi_\lambda(g^{-1})s_\lambda,  
1727: \pi_\lambda(a_t)v_{{\hbox{\fiverm K}, \scriptscriptstyle{\lambda}}}\ra| 
1728: &= |\la \pi_\lambda(a_\eps g^{-1})s_\lambda,  
1729: \pi_\lambda(a_{t-\eps}) v_{{\hbox{\fiverm K}, 
1730:  \scriptscriptstyle{\lambda}}}\ra|\cr  
1731: &\leq  \|\pi_\lambda(a_\eps g^{-1})s_\lambda\| 
1732: \cdot \|\pi_\lambda(a_{t-\eps}) v_{{\hbox{\fiverm K}, 
1733:  \scriptscriptstyle{\lambda}}}\|\cr  
1734: &\leq M\cdot   \|\pi_\lambda(a_\eps g^{-1})s_\lambda\|\cr}\leqno 
1735: (3.2.5)$$ 
1736: with $M=\sup_{\lambda\in \supp(s)\atop  
1737: \eps\leq t\leq 1}  
1738: \|\pi_\lambda(a_{t-\eps}) v_{{\hbox{\fiverm K}, 
1739:  \scriptscriptstyle{\lambda}}}\|<\infty$ as $\supp(s)$ is  
1740: compact.  
1741: \par Recall our notion of holomorphic extension from  
1742: Definition 2.2.1.  As almost  
1743: each stalk $s_\lambda$ is  
1744: an analytic vector in ${\cal H}_\lambda$, it follows  
1745: from  estimates (3.2.4-5) and the compactness  
1746: of $\supp (s)$ that  
1747: $$\eqalign{\int_{i\a_+^*} 
1748:  \la \pi_\lambda(g^{-1})s_\lambda,  
1749: v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}\ra \ d\mu(\lambda) 
1750: &= 
1751: \int_{\supp (s)} 
1752: \lim_{t\nearrow 1} \la \pi_\lambda(g^{-1})s_\lambda,  
1753: \pi(a_t)v_{{\hbox{\fiverm K}, \scriptscriptstyle{\lambda}}}\ra \ d\mu(\lambda)\cr 
1754: &= 
1755: \lim_{t\nearrow 1}  
1756: \int_{\supp(s)}\la \pi_\lambda(g^{-1})s_\lambda,  
1757: \pi_\lambda(a_t)v_{{\hbox{\fiverm K}, \scriptscriptstyle{\lambda}}}\ra \ d\mu(\lambda)\cr 
1758: &= b^\omega(f)(gH)\ .\cr}$$ 
1759: As $f$ and $g$ were arbitray, this completes the proof  
1760: of (3.2.3).  
1761:  
1762:  
1763:  
1764: \subheadline{3.3. The Plancherel Theorem for $L^2(G/H)_{\rm mc}$} 
1765:  
1766: \noindent 
1767: Before we can  show that $b_0$ has image in $L^2(G/H)_{\rm mc}$ and 
1768: extends to an isometric embedding, we need to recall some facts about the 
1769: most continuous spectrum  $L^2(G/H)_{\rm mc}$ of $L^2(G/H)$  (cf.\ [vdBS97a] and [D98]). 
1770: All the results collected below are proved in [vdBS97a] or might be considered as special 
1771: cases of [D98]. The crucial way where our assumption that $G/H$ 
1772: is NCC, $H=G^\tau$, and $G\subseteq G_\C$ with $G_\C$ simply connected, enters is the fact 
1773: that $Z_H(\a)=Z_K(\a)$ and $H=Z_H(\a )H_0$. 
1774:  
1775: \ssk Recall from Subsection 2.7 the mapping 
1776: $j(\lambda)\: \C^{\left| {\cal W}/ {\cal W}_0\right|}\to ({\cal H}_\lambda^{-\infty})^H$ 
1777: and the intertwining  matrix $A_H(\lambda, w\lambda)^*\: ({\cal H}_{w\lambda}^{-\infty})^H\to 
1778: ({\cal H}_\lambda^{-\infty})^H$ both defined for generic $\lambda\in \a_\C^*$, and all $w\in {\cal W}$. 
1779: For generic $\lambda$ we  define 
1780: $$j^0(\lambda)\:  \C^{\left| {\cal W}/ {\cal W}_0\right|}\to ({\cal H}_\lambda^{-\infty})^H$$ 
1781: by 
1782: $$j^0(\lambda)\:=[A_H(w_0\lambda, \lambda)^*]^{-1}\circ  j(w_0\lambda)\leqno(3.3.1)$$ 
1783: with $w_0\in {\cal W}$ the longest element. Then $j^0$ has no poles on $i\a^*$ 
1784: (cf.\ [vdBS97b, Th.\ 1]). 
1785: Denote by $(\e_w)_{w\in {\cal W}/ {\cal W}_0}$ the canonical basis of the Hilbert space 
1786: $\C^{\left| {\cal W}/ {\cal W}_0\right|}$. For $w\in {\cal W}/{\cal W}_0$ define 
1787: $$\eta_{\lambda,w}^0\:=j^0(\lambda) \e_w\in ({\cal H}_\lambda^{-\infty})^H\ .$$ 
1788: Define a Hilbert space structure on $\Hom (\C^{\left| {\cal W}/ {\cal W}_0\right|}, {\cal H}_\lambda)$ 
1789: using the identification 
1790: $$\Hom (\C^{\left| {\cal W}/ {\cal W}_0\right|}, {\cal H}_\lambda)\simeq {\cal H}_\lambda\otimes 
1791: [\C^{\left| {\cal W}/ {\cal W}_0\right|}]^*\ .$$ 
1792: \par Write ${\cal S}(G/H)$ for the Schwartz space on $G/H$ 
1793: and $p_{\rm mc}\: L^2(G/H)\to L^2(G/H)_{\rm mc}$ for  
1794: the orthogonal projection on the most continuous spectrum.  
1795: Set ${\cal S}_{\rm mc}(G/H)=p_{\rm mc}({\cal S}(G/H))$.  
1796: Then for functions $f\in {\cal S}_{\rm mc}(G/H)$  
1797: the Fourier transform is defined by  
1798: $${\cal F}(f)=\left(\pi_\lambda(f)j^0(\lambda)\right )_\lambda \ .\leqno(3.3.2) $$ 
1799: By [D98, Th.\ 3] or [vdBS97a,Cor. 18.2 and Prop. 18.3], 
1800: ${\cal F}$ extends to a $G$-equivariant unitary  
1801: isomorphism  
1802: $${\cal F}\: L^2(G/H)_{\rm mc} \to \int_{i\a_+^*}^\oplus 
1803: \Hom (\C^{\left| {\cal W}/ {\cal W}_0\right|}, {\cal H}_\lambda) \ d\lambda\ . \leqno(3.3.3)$$ 
1804: In particular, we have (using suitable normalization of measures) that 
1805: $$\|f\|^2= \int_{i\a_+^*} 
1806: \|{\cal F}(f)(\lambda)\|^2 \ d\lambda \leqno(3.3.4) $$ 
1807: for all $f\in {\cal S}_{\rm mc}(G/H)$.  
1808: \par Next we wish to describe ${\cal F}^{-1}$.  
1809: Let $(\e_w^*)_{w\in {\cal W}/ {\cal W}_0}$ be the 
1810: dual basis of $(\e_w)_{w\in {\cal W}/ {\cal W}_0}$. Then a section $s$ of 
1811: $\int_{i\a_+^*}^\oplus 
1812: \Hom (\C^{\left| {\cal W}/ {\cal W}_0\right|}, {\cal H}_\lambda) \ d\lambda$ can be written as 
1813: $s=(\sum_{w\in {\cal W}/ {\cal W}_0} s_{\lambda, w}\otimes \e_w^*)_\lambda$ 
1814: with $s_{\lambda, w}\in {\cal H}_\lambda$ for all $\lambda\in i\a_+^*$ and 
1815: $w\in {\cal W}/ {\cal W}_0$. Recall that if  
1816: $s=(\sum_{w\in {\cal W}/ {\cal W}_0} s_{\lambda, w}\otimes \e_w^*)_\lambda$ is a smooth vector, then  
1817: $s_{\lambda, w}$ is a smooth vector in ${\cal H}_\lambda$ 
1818: for almost all $\lambda$.  
1819: In the sequel we will use the identification  
1820: ${\cal H}_\lambda^\infty=C^\infty(M\bs K)$. Define a 
1821: subspace of $\left(\int_{i\a_+^*}^\oplus 
1822: \Hom (\C^{\left| {\cal W}/ {\cal W}_0\right|}, {\cal H}_\lambda) \ d\lambda\right)^\infty$ 
1823: by 
1824: $$ 
1825: %\eqalign{ 
1826: {\cal H}_0=\left\{ s%=(\sum_{w\in {\cal W}/ {\cal W}_0} s_{\lambda, w}\otimes \e_w^*)_\lambda & 
1827: \in \left(\int_{i\a_+^*}^\oplus 
1828: \Hom (\C^{\left| {\cal W}/ {\cal W}_0\right|}, 
1829: {\cal H}_\lambda) \ d\lambda\ \right)^\infty\quad :\quad 
1830: \eqalign{& s \ \hbox{is $K$-finite,}\  
1831: \supp (s) \ \hbox{is compact,}\cr 
1832: & s\: i\a_+^*\to \Hom(\C^{\left| {\cal W}/ {\cal W}_0\right|}, 
1833:  C^\infty(M\bs K)) \cr 
1834:  & \hbox{is weakly smooth}. \cr}\right\}$$ 
1835: It is not hard to see that ${\cal H}_0$ is  
1836: a dense subspace in $\int_{i\a_+^*}^\oplus 
1837: \Hom (\C^{\left| {\cal W}/ {\cal W}_0\right|},  
1838: {\cal H}_\lambda) \ d\lambda$.  
1839: Then for an element $s\in {\cal H}_0$ the inverse  
1840: Fourier-transform is given by [D98, Th.\ 3] 
1841:  
1842: $${\cal F}^{-1}(s)(gH)=\int_{i\a_+^*}\sum_{w\in {\cal W}/ {\cal W}_0} \la \pi_\lambda(g^{-1}) s_{\lambda, w}, 
1843: \eta_{\lambda, w}^0\ra \ d\lambda\ .\leqno(3.3.5)$$ 
1844: Moreover [D98, Th.\ 3] implies that 
1845: $${\cal F}^{-1}({\cal H}_0)\subeq  
1846: {\cal S}_{\rm mc}(G/H)\ ;  
1847: \leqno(3.3.6)$$ 
1848: in particular ${\cal F}({\cal F}^{-1}(s))$ is  
1849: given by the formula (3.3.2) for $s\in {\cal H}_0$.  
1850:   
1851: \Remark 3.3.1. We have normalized the invariant 
1852: measure on $G/H$ and the measure $d\lambda$ on ${\frak a}^*$ so that (3.3.3) and 
1853: (3.3.4) holds without any additional constants. This is possible, because we 
1854: are only working with the principial series of representations 
1855: and the most continuous part of the spectrum. In general, one 
1856: has to take into account the order of several Weyl groups. 
1857: We refer to Theorem 31 and Remark 32 in [vdB00] for general 
1858: discussion on the normalization of measures.\qed 
1859:  
1860: \subheadline{3.4. Isometry of the boundary value mapping} 
1861:  
1862: \noindent 
1863: In this subsection we  complete our 
1864: discussion of the boundary value mapping begun in Subsection 3.2. 
1865:  
1866:  
1867:  
1868: \Theorem 3.4.1. {\rm \bf(Isometry of the boundary value mapping)}
1869: The boundary value mapping, initially defined 
1870: by 
1871: $$b_0\: {\cal H}^2(\Xi_H)_0\to C(G/H), \ \ b_0(f)(gH)=\lim_{t\nearrow 1} 
1872: f(ga_tK_\C)$$ 
1873: (cf.\ {\rm (3.2.2-3)}) extends to a $G$-equivariant isometric embedding 
1874: $$b\:  {\cal H}^2(\Xi_H)\to L^2(G/H)_{\rm mc}\ .$$ 
1875:  
1876: \Proof. For each $\lambda\in i\a_+^*$ define a vector 
1877: $\b(\lambda)\in (\C^{\left |{\cal W} 
1878: / {\cal W}_0\right|})^*$ by 
1879: $$\b(\lambda)= \c(w_0\lambda)
1880: \sum_{w\in {\cal W}/ {\cal W}_0} z_{\hbox{\fiverm H}}^{-w^{-1}(w_0\lambda+\rho)}\e_w^*\ .$$
1881: Notice, that for $\lambda\in i\a^*$ we have
1882: $$|z_{\hbox{\fiverm H}}^{-2w^{-1}\lambda}|=
1883: z_{\hbox{\fiverm H}}^{-w^{-1}(\lambda+\rho)}\overline{z_{\hbox{\fiverm H}}^{-w^{-1}(\lambda+\rho)}}
1884: =z_{\hbox{\fiverm H}}^{-w^{-1}(\lambda+\rho)}z_{\hbox{\fiverm H}}^{w^{-1}(\bar{\lambda}+\rho)}\, .$$
1885: Therefore,
1886: employing the Maass-Selberg relation for $\c(\lambda)$ we obtain
1887: $$\|\b(\lambda)\|^2=|\c(\lambda)|^2 \cdot \sum_{w\in {\cal W}/ {\cal W}_0}
1888: |z_{\hbox{\fiverm H}}^{-w^{-1}\lambda}|^2 =|\c_{G/H}(\lambda)|^2\ .$$
1889: In particular we see that we have an  
1890: $G$-equivariant isometric embedding of  
1891: direct integrals  
1892: $$\iota\: \int_{i\a_+^*}^\oplus {\cal H}_\lambda\ d\mu(\lambda) 
1893: \to  \int_{i\a_+^*}^\oplus \Hom(\C^{\left |{\cal W} 
1894: / {\cal W}_0\right|}, {\cal H}_\lambda)\ d\lambda,  
1895: \ \  s=(s_\lambda)_\lambda\mapsto (s_\lambda\otimes \b(\lambda))_\lambda\ .$$ 
1896: {}From the definition of the spaces ${\cal H}^2(\Xi_H)_0$ 
1897: and ${\cal H}_0$ it is then clear that
1898: $$\iota \left(\Phi^{-1}({\cal H}^2(\Xi_H)_0)\right)\subeq 
1899: {\cal H}_0\ .\leqno(3.4.1)$$ 
1900: 
1901: \par As
1902: $$\iota (s)_\lambda = \c (w_0\lambda)\sum_{w\in {\cal W}/{\cal W}_0}
1903: z_{\hbox{\fiverm H}}^{-w^{-1}(w_0\lambda+\rho)}s_\lambda\otimes \e_w^*\ , $$
1904: we get by (3.2.3), (3.3.5),  Theorem
1905: 2.6.1 and Theorem 2.7.1 that
1906: $$\eqalign{[{\cal F}^{-1}(\iota (s))](gH) &=\int_{i\a_+^*}
1907: \sum_{w\in {\cal W}/{\cal W}_0}
1908: \c (w_0\lambda)
1909: z_{\hbox{\fiverm H}}^{-w^{-1}(w_0\lambda+\rho)}
1910: \la \pi_\lambda (g^{-1})s_\lambda, \eta^0_{\lambda, w}\ra \ d\lambda\cr
1911: &=\int_{i\a_+^*}
1912: \sum_{w\in {\cal W}/{\cal W}_0}
1913: \la \pi_\lambda (g^{-1})s_\lambda, \c (w_0\bar{\lambda})
1914: z_{\hbox{\fiverm H}}^{w^{-1}(w_0\bar{\lambda}+\rho)}\eta^0_{\lambda, w}\ra \ d\lambda\cr
1915: &=\int_{i\a_+^*}
1916: \la \pi_\lambda (g^{-1})s_\lambda, v_{{\hbox{\fiverm H}, \scriptscriptstyle{\lambda}}}\ra \ d\lambda
1917: \cr
1918: &=b_0(s)(gH) }
1919: $$
1920: 
1921: 
1922: \par From this and (3.3.6) it follows that
1923: $b_0(f)\in S_{\rm mc}(G/H)$; in particular we
1924: have
1925: $b_0(f)\in L^2(G/H)_{\rm mc}$. Finally,
1926: $$\|b_0(f)\|_{L^2(G/H)_{\rm mc}}=
1927: \|{\cal F}^{-1} (\iota (s))\|=\|\iota(s)\|=\|s\|=
1928: \|\Phi(s)\|=\|f\|$$ 
1929: as ${\cal F}$, $\iota$ and $\Phi$ are isometric.  
1930: This completes the proof of the theorem.\qed  
1931:  
1932:  
1933:  
1934: \Remark 3.4.2. The domain $\Xi_H$ is maximal in the sense that 
1935: generic functions in ${\cal H}^2(\Xi_H)$ do not extend holomorphically 
1936: over $\Xi_H$.\qed 
1937:  
1938:  
1939:  
1940:   
1941: \subheadline{3.5. Concluding remarks and the example of $G=\Sl(2,\R)$} 
1942:  
1943: \noindent 
1944: In [GK\'O01] we defined a Hardy space ${\cal H}^2(\Xi)$ on $\Xi$ for the  
1945: cases where $\Xi=\Xi_H$.  Let us briefly summarize its construction in order  
1946: to put it into perspective with the results in this section. 
1947:  
1948: \par Geometrically the situation $\Xi=\Xi_H$ is equivalent to the fact  
1949: that $\Xi$ is homogeneous 
1950: for a bigger Hermitian group $S\supeq G$ (cf.\ [KS01b]). More precisely, if $U<S$ denotes an appropriate  
1951: maximal compact subgroup with $K\subeq U$ then $\Xi$ is $G$-biholomorphic  
1952: to the Hermitian symmetric space $S/U$. For example if $G$ is Hermitian, then  
1953: $S=G\times G$ and $\Xi\simeq G/K\times \oline{G/K}$.  
1954:  
1955: \par The assumption $\Xi=\Xi_H$ thus allows us to identify $\Xi$ with a  bounded  
1956: symmetric domain ${\cal D}\simeq S/U$. Within this identification one shows that  
1957: $\partial_d\Xi\simeq G/H$ becomes a Zariski-open subset in the Shilov boundary  
1958: $\partial_s{\cal D}$ of ${\cal D}$.  
1959:  
1960:  
1961: \par The identification of $\Xi$ with ${\cal D}$ was used in [GK\'O01] in a crucial way:  
1962: One can transfer the action of an appropriate compression-semigroup $\Gamma\supeq G$ on ${\cal D}$ 
1963: to $\Xi$ and use this to give a definition of a Hardy space as follows: 
1964:  
1965: $${\cal H}^2(\Xi)=\{ f\in {\cal O}(\Xi)\: \|f\|^2=\sup_{\gamma\in \Int\Gamma} 
1966: \int_{G/H} |f(\gamma gz_{\hbox{\fiverm H}})|^2\ dgH<\infty\}\ .\leqno(3.5.1)$$ 
1967: In [GK\'O01] we have shown -- with entirely different methods -- that the Hardy space defined as  
1968: in (3.5.1) has the following properties: 
1969: \msk 
1970: \item{(3.5.2)} ${\cal H}^2(\Xi)$ is a Hilbert space of holomorphic functions 
1971: featuring an isometric boundary value mapping $b\:{\cal H}^2(\Xi)\into L^2(G/H)_{\rm mc}$.  
1972: Moreover, $\im b$ is a multiplicity one subspace of {\it full spectrum}.  
1973: \item{(3.5.3)} The Hardy space ${\cal H}^2(\Xi)$ is $G$-isometric to  
1974: $L^2(G/K)$ through a transform of Segal-Barg\-mann type.   
1975: \item{(3.5.4)} ${\cal H}^2(\Xi)$ is $G$-isometric to the classical Hardy space  
1976: ${\cal H}^2({\cal D})$ through an explicitely given mapping.  
1977: \msk 
1978: In particular for $\Xi=\Xi_H$ it follows from Theorem 3.4.1 and  
1979: (3.5.2) that the definition of (3.5.1) coincides 
1980: with our spectral definition of the Hardy space in Definition 3.2.2.  For the  
1981: cases where $\Xi\neq \Xi_H$ there is no apparent semigroup action on $\Xi_H$ and a definition  
1982: of ${\cal H}^2(\Xi_H)$ in the flavour of (3.5.1) seems presently not possible.  
1983:  
1984: \par Notice that (3.5.3) implies that the Plancherel measure of ${\cal H}^2(\Xi)$ has  
1985: support equal to $i\a_+^*$. However, in [GK\'O01] we could not determine this  
1986: measure explicitely. With the new approach given in this section this difficulty  
1987: is already taken care of with the definition of the Hardy space.  
1988: \par The explicit isomorphism of ${\cal H}^2(\Xi)$ with the classical  
1989: Hardy space ${\cal H}^2({\cal D})$ allows us to find also a nice closed  
1990: expression for the Cauchy-Szeg\"o function $\Psi$ (cf.\ [GK\'O01, Th.\ 5.7 and  
1991: Ex.\  5.10]). Combining this closed expression with the spectral resolution  
1992: of $\Psi$ in Theorem 3.2.4 one obtains interesting identities for  
1993: (generalized) hypergeometric functions. For example for $G=\Sl(2,\R)$ one obtains 
1994: the following formula: 
1995: $${1-\tanh^2 t\over 1+\tanh^2 t} ={\pi\over 2}\int_0^\infty F\left({1\over 4}+i{\lambda\over 4},  
1996: {1\over 4}-i{\lambda\over 4}, 1; -\sinh^2\left(2t+i{\pi\over 2}\right)\right) 
1997: \cdot\left|{\Gamma\left({i\lambda+1\over2}\right)\over \Gamma\left({i\lambda\over2}\right)} \right|^2 
1998: \ {d\lambda\over \cosh {\pi\over 2}\lambda}$$ 
1999: for all $t\in \R+i]-{\pi\over 4}, {\pi\over 4}[$. Here $F$ denotes the Gau\3 hypergeometric function.   
2000:  
2001:  
2002: \sectionheadline{A. Appendix: Analytic vectors for representations} 
2003:  
2004: In this appendix  we will summarize  some facts on analytic vectors  
2005: for representations. None of the results collected below is new,  
2006: however some of them might be hard to find explicitely in the  
2007: literature. In order to keep the exposition short, we will omit  
2008: proofs and often do not make the most general assumptions.  
2009: A more detailed account containing complete proofs can be found  
2010: in the forthcoming survey [K\'O03].   
2011:  
2012: \subheadline{A.1. Definition and topology of analytic vectors} 
2013:  
2014: Throughout this appendix $G$ will denote a connected unimodular Lie group  
2015: with $G\subeq G_\C$.  
2016:  
2017: \par Let $E$ be a complex Banach space and  
2018: $\Gl(E)$ the group of continuous invertible operators on $E$.  
2019: By a (Banach) representation of $(\pi, E)$ of $G$ we will understand  
2020: a group homomorphism $\pi\: G\to \Gl(E)$ such that  
2021: for all $v\in E$ the orbit mapping  
2022: $$\gamma_v\: G\to E, \ \ g\mapsto \pi(g)v$$ 
2023: is continuous.  
2024:  
2025: \par A vector $v\in E$ is called {\it analytic} if $\gamma_v$ is an  
2026: analytic $E$-valued map or, equivalently, if there exists  
2027: an open neighborhood $U$ of $\1$ in $G_\C$ and a $G$-equivariant  
2028: holomorphic mapping  
2029: $$\gamma_{v,U}\: GU\to E$$ 
2030: such that $\gamma_{v,U}(\1)=v$. In particular, $\gamma_{v, U}\res_G=\gamma_v$.  
2031:  
2032: \par The vector space of all analytic vectors  for $(\pi, E)$ is denoted  
2033: by $E^\omega$. We recall a fundamental result of Nelson which  
2034: states that $E^\omega$ is dense in $E$.  
2035:  
2036: \ssk  Next we are going to recall the definition of the topology  
2037: on $E^\omega$.  
2038: \par For a complex manifold 
2039: $M$ let us denote by ${\cal O}(M,E)$ the space of all $E$-valued  
2040: holomorphic mappings on $E$. Topologically we consider ${\cal O}(M,E)$  
2041: as a Fr\'echet space with the topology of compact convergence. 
2042:  
2043: \par  For any open neighborhood $U$  
2044: of $\1$ in $G_\C$ we write $E_U$ for the subspace  
2045: of $E^\omega$ for which $\gamma_{v,U}$ exists. Then we obtain a linear  
2046: embedding  
2047:  
2048: $$\eta_U\: E_U\to {\cal O}(GU, E), \ \ v\mapsto \gamma_{v,U}\ .$$ 
2049: The image of $\eta_U$ is closed and hence ${\cal O}(GU,E)$ induces  
2050: a Fr\'echet topology on $E_U$. Notice that for $U_1\subeq U_2$  
2051: we obtain a continuous embedding $E_{U_2}\to E_{U_1}$ via restriction.  
2052: Thus  
2053: $$E^\omega=\lim_{U\to \{1\}} E_U=\bigcup_U E_U $$  
2054: and we can equip $E^\omega$ with the inductive limit topology, i.e. 
2055: the finest topology on $E^\omega$ for which all inclusion  
2056: mappings $E_U\to E^\omega$ become continuous. Notice that this  
2057: turns $E^\omega$ into a locally convex topological vector space.   
2058:  
2059: \par By $E^{-\omega}$ we will denote the antidual of $E^\omega$, i.e. 
2060: the space of all antilinear continuous functionals on $E^\omega$.  
2061: The space $E^{-\omega}$ is referred to as the space of {\it hyperfunction  
2062: vectors} of the representation $(\pi, E)$. We equip  
2063: $E^{-\omega}$ with the topology of bounded convergence.  
2064:  
2065:  
2066: \subheadline{A.2. Analytic vectors for $L^1(G/H)$} 
2067:  
2068: \par Let $H<G$ be a closed subgroup such that $G/H$  
2069: carries a $G$-invariant measure. We write $L^1(G/H)$ for the corresponding  
2070: Banach space of integrable functions and $(L, L^1(G/H))$ for the left regular representation  
2071: of $G$ on $L^1(G/H)$, i.e.,  
2072: $$(L(g)f)(xH)=f(g^{-1}xH)\qquad (g,x\in G, f\in L^1(G/H))\ .$$ 
2073:  
2074: Further it is convenient to assume that $G/H\subeq G_\C/ H_\C$.  
2075: Then we have the following characterization of the analytic 
2076: vectors:  
2077:  
2078: \Proposition A.2.1. Let $U$ be an open neighborhood of $\1$ in $G_\C$.  
2079: Then $f\in L^1(G/H)_U$ if and only if  
2080: there exists a holomorphic function $\tilde f$ on the open set  
2081: $$U^{-1} GH_\C/ H_\C \subeq G_\C/ H_\C$$  
2082: with the following properties: 
2083: \item{(1)} $\tilde f\res_{G/H}=f$.  
2084: \item{(2)} For all $x\in U$ the map  
2085: $$\tilde f_x\: G/H\to\C, \ \ gH\mapsto \tilde f(x^{-1}gH)$$ 
2086: belongs to $L^1(G/H)$.  
2087: \item{(3)} For all compact subsets $U^c\subeq U$ we have  
2088: $$\sup_{x\in U^c} \|\tilde f_x\|<\infty\ .$$ \qed  
2089:  
2090: There are two types  of homogeneous spaces  $G/H$ which will 
2091: be of particular interest for us. The first is when $H={\1}$. Then $L^1(G)^\omega$ 
2092: denotes the analytic for the left regular representation of $G$ on $L^1(G)$.  
2093: The second case is for $G=H\times H$ and $H<G$ the diagonal subgroup. In this  
2094: case $G/H\simeq H$ and $L$ becomes left-right regular representation of $H\times H$  
2095: on $H$. Here we shall write $L^1(H)^{\omega, \omega}$ for the analytic vectors.  
2096:  
2097:  
2098:  
2099: \subheadline{A.3. Averaging properties} 
2100:  
2101: Recall that the average map  
2102: $$C_c(G)\to C_c(G/H), \ \ f\mapsto f^H; f^H(xH)=\int_H f(xh)\ dh$$ 
2103: is contiunuous and onto. Further, this map extends to a surjective contraction of Banach spaces  
2104: $L^1(G)\to L^1(G/H)$. We will show that the  averaging operator  
2105: maps analytic vectors into analytic vectors. 
2106:  
2107: \par A standard application of the Bergman estimate gives: 
2108:  
2109: \Lemma A.3.1. Let $U\subeq G_\C$ be an open neighborhood  
2110: of $\1$. Then for any pair of compact subsets 
2111: $U_1, U_2\subeq U$ with $U_1\subeq \Int U_2$ there  
2112: exists a constant $C>0$ such that for all $f\in L^1(G)_U$ we have that  
2113: $$(\forall x\in U_1^{-1}G) \qquad \int_H |\tilde f(xh)|\ dh\leq  
2114: C \sup_{x\in U_2}\|\tilde f_x\|,  $$ 
2115: where $\tilde f$ denotes the extension of $f$ to a holomorphic  
2116: function on $U^{-1}G$ (cf.\ Proposition A.2.1).\qed  
2117:  
2118: Combining Lemma A.3.1 with Proposition A.2.1 we obtain:  
2119:  
2120: \Proposition A.3.2.  Let $U\subeq G_\C$ be an open neighborhood  
2121: of $\1$.  Then for every  
2122: $f\in L^1(G)_U$ and $g\in G$ the integral 
2123: $f^H(g)=\int_H f(gh)\ dh$ converges absolutely and $f^H\in L^1(G/H)_U$.  
2124: In particular, there is a well defined mapping  
2125: $$L^1(G)^\omega\to L^1(G/H)^\omega, \ \ f\mapsto f^H\ .$$\qed  
2126:  
2127:  
2128:  
2129: \subheadline{A.4. Mollifying properties} 
2130:  
2131: In this section $E={\cal H}$ will be a Hilbert space and $(\pi, {\cal H})$ a unitary 
2132: representation of $G$.  
2133: For $f\in L^1(G)$ one defines a continuous operator $\pi(f)\: {\cal H}\to {\cal H}$ by  
2134:  
2135: $$\pi(f)v=\int_G f(g) \pi(g)v\ dg\qquad (v\in {\cal H})\ .$$ 
2136: Notice that this defines a $*$-representation of the Banach algebra 
2137: $L^1(G)$, i.e. we have $\pi(f*g)=\pi(f)\pi(g)$ and $\pi(f)^*=\pi(f^*)$ with  
2138: $f^*(x)=\oline {f(x^{-1}) }$.  
2139:  
2140: \par Recall that $L^1(G)^{\omega, \omega}$ denotes the analytic  
2141: vectors for the left-right regular representation of $G\times G$ on  
2142: $L^1(G)$. It is easy to see  
2143: that $L^1(G)^{\omega,\omega}$ is $*$-closed subalgebra of $L^1(G)$.  
2144:  
2145: \par Let $f\in L^1(G)^{\omega,\omega}$. It follows readily  
2146: from Proposition A.2.1 and the definition of analytic vectors 
2147: that $\pi^\omega(f)$ maps ${\cal H}$ continuously into ${\cal H}^\omega$.  
2148: In particular the restriction $\pi^\omega(f)\:=\pi(f)\res_{{\cal H}^\omega}$ 
2149: gives rise to a continuos operator $\pi^\omega(f)\: {\cal H}^\omega\to {\cal H}^\omega$. 
2150: Hence we have an algebra representation: 
2151:  
2152: $$\pi^\omega\: L^1(G)^{\omega,\omega}\to \End({\cal H}^\omega), \ \ f\mapsto \pi^\omega(f)\ .$$  
2153: The corresponding dual representation is given by  
2154:  
2155: $$\pi^{-\omega}\: L^1(G)^{\omega,\omega}\to \End({\cal H}^{-\omega}); \ \   
2156: \pi^{-\omega}(f)\lambda=\lambda\circ \pi^\omega(f^*)\ .$$ 
2157: Another application of Proposition A.2.1 then gives us  
2158: the mollifying property:  
2159:  
2160: \Proposition  A.4.1. Let $(\pi, {\cal H})$ be a unitary  
2161: representation of a unimodular Lie group $G$. Then we have for all  
2162: $f\in L^1(G)^{\omega,\omega}$ that   
2163: $$\pi^{-\omega}(f){\cal H}^{-\omega}\subeq {\cal H}^\omega\ .$$\qed  
2164:  
2165: \nin {\bf Note:} For $f\in L^1(G)^{\omega, \omega}$, it is often convenient to write $\pi(f)$  
2166: instead of $\pi^{-\omega}(f)$. We will use this convention throughout Section 2 in the main text.  
2167:  
2168:  
2169:  
2170: \def\entries{ 
2171:  
2172:  
2173: \[AG90 Akhiezer, D.\ N., and S.\ G.\ Gindikin, {\it On Stein 
2174: extensions of 
2175: real symmetric spaces}, 
2176: Math.\ Ann.\ {\bf 286}, 1--12, 1990 
2177:  
2178: \[vdB88 van den Ban, E., {\it The principal series for a reductive symmetric space I, 
2179: $H$-fixed distribution vextors}, Ann. sci. \'Ec. Norm. Sup. {\bf 4}, {\bf 21} 
2180: (1988), 359--412 
2181:  
2182: \[vdB00 ---, {\it The Plancherel theorem for a reductive symmetric space}, Lectures 
2183: for the European School of Group Theory. August 14--26,  
2184: 2000, SDU-Odense University. 
2185: http://www.math.uu.nl/people/ban/publ.html 
2186:  
2187: \[vdBD88 van den Ban, E., and P. Delorme, {\it Quelques propri\'et\'es des repr\'esentations  
2188: sph\'eriques pour les espaces sym\'etriques r\'eductifs}, J. Funct. Anal. {\bf 80} (1988),  
2189: 284--307 
2190:  
2191: \[vdBS97a van den Ban, E., and H.\ Schlichtkrull, {\it The most 
2192: continuous part of the Plan\-che\-rel decomposition for a reductive 
2193: symmetric space}, Ann. of Math. {\bf (2) 145} (1997), no. {\bf 2}, 
2194: 267--364 
2195:  
2196: \[vdBS97b ---, {\it Fourier transform on a semisimple symmetric space},  Invent. Math. {\bf 130} (1997), no. {\bf 3}, 
2197: 517--574 
2198:  
2199: \[BD92 Brylinski, J.-L., and P. Delorme, {\it Vecteurs distributions $H$-invariants pour les s\'eries 
2200: principales g\'en\'eralis\'ees d'espaces sym\'etriques r\'eductifs et 
2201: prolongement m\'eromorphe d'int\'egralesd'Eisenstein}, Invent. Math. {\bf 109} (1992), no. {\bf 3}, 619--664 
2202:  
2203: \[D98  Delorme, P., {\it Formule de Plancherel pour les espaces 
2204: sym\'etriques r\'eductifs},  Ann. of Math. {\bf (2) 147} 
2205: (1998), no. {\bf 2}, 417--452 
2206:  
2207: \[FT99 Faraut, J., and E.  G.  F.  Thomas, {\it Invariant Hilbert spaces of 
2208: holomorphic functions}, J.  Lie Theory {\bf 9} (1999), no.  {\bf 2}, 383--402 
2209:  
2210:  
2211: \[GK02a Gindikin, S., and B.\ Kr\"otz, {\it Complex crowns of Riemannian 
2212: symmetric spaces and non-compactly causal symmetric spaces}, 
2213: Trans. Amer. Math. Soc. {\bf 354} (2002), no. {\bf 8}, 
2214: 3299--3327 
2215:  
2216: \[GK02b ---, {\it Invariant Stein domains in Stein symmetric spaces 
2217: and a non-linear complex convexity theorem}, IMRN {\bf 18} (2002), 959--971 
2218:  
2219: \[GK\'O083 Gindikin, S., B.\ Kr\"otz and G.\ \'Olafsson, {\it Hardy spaces for 
2220: non-compactly causal symmetric spaces and the most continuous spectrum}, Math. Ann. {\bf 327} (2003), 25--66
2221:  
2222: \[H84 Helgason, S., ``Groups and Geometric Analysis'', Academic Press, 1984  
2223:  
2224: \[H\'O96 Hilgert, J.\ and 
2225: G.\ \'Olafsson, ``Causal Symmetric Spaces, Geometry and 
2226: Harmonic Analysis,'' Acad. Press, 1996 
2227:  
2228: \[K99 Kr\"otz, B., {\it The Plancherel theorem for biinvariant Hilbert 
2229: spaces}, Publ.  Res.  Inst.  Math.  Sci.  {\bf 35} (1999), no.  {\bf 1}, 
2230: 91--122 
2231:  
2232: \[K\'O03 Kr\"otz, B., and G. \'Olafsson, {\it Analytic vectors for representations -- a  
2233: survey}, in preparation 
2234:  
2235: \[KS01a Kr\"otz, B., and R.J. Stanton, {\it Holomorphic extension of 
2236: representations: (I) 
2237: automorphic functions}, Annals of Mathematics, to appear 
2238:  
2239: \[KS01b Kr\"otz, B., and R.J. Stanton, {\it Holomorphic extensions of 
2240: representations: (II) geometry and harmonic analysis}, preprint 
2241:  
2242: \[M79 Matsuki, T., {\it The orbits of affine symmetric spaces under the action of minimal parabolic 
2243: subgroups}, J. Math. Soc. Japan {\bf 31} (1979), 331--357 
2244:  
2245: \[\'O87 \'Olafsson, G., {\it Fourier and Poisson transformation associated to 
2246: semisimple symmetric space}, Invent Math. {\bf 90} (1987), 605--629 
2247:  
2248:  
2249:  
2250: } 
2251:  
2252:  
2253: {\sectionheadline{\bf References} 
2254: \frenchspacing 
2255: \entries\par} 
2256: \dlastpage  
2257: \bye 
2258:  
2259: