Content-Type: multipart/mixed; boundary="-------------0011091627298" This is a multi-part message in MIME format. ---------------0011091627298 Content-Type: text/plain; name="00-444.comments" Content-Transfer-Encoding: 7bit Content-Disposition: attachment; filename="00-444.comments" AMS-Code: Primary 22E45, 22E70, 43A32, 44A20, 46L60, 47B49; Secondary 81R05, 81R30, 81T05. Dedicated to the memories of Irving E. Segal and Harish-Chandra. Appears in The Mathematical Legacy of Harish-Chandra: A Celebration of Representation Theory and Harmonic Analysis (R. Doran and V. Varadarajan, eds.), Proc. Sympos. Pure Math., vol. 68, American Mathematical Society, Providence, R.I., 2000, pp. 333--401. ---------------0011091627298 Content-Type: text/plain; name="00-444.keywords" Content-Transfer-Encoding: 7bit Content-Disposition: attachment; filename="00-444.keywords" Representations, Hilbert space, positivity, quantum theory, Segal-Bargmann transform, Lie groups, semisimple, Heisenberg group, reflection symmetry, spectrum, highest weight modules, discrete series, holomorphic series, Feynman-Kac formula, functional integral, isometric embedding, symmetric space, Cayley-type, complementary series, hyperbolic cone, ax+b group, duality ---------------0011091627298 Content-Type: application/x-tex; name="hcpaper4.tex" Content-Transfer-Encoding: 7bit Content-Disposition: inline; filename="hcpaper4.tex" % Equation numbers corrected to start with .1 in each section, % and cross-references checked, 1/13/2000 --BFT % Revised according to authors' marks 6/1/1999 --BFT %% This document created by Scientific Word (R) Version 3.0 \documentclass{amsproc} \usepackage{amsxtra} \usepackage{amsmath} \usepackage{graphicx} \usepackage{latexsym} \usepackage{amsfonts} \usepackage{amssymb} %TCIDATA{OutputFilter=latex2.dll} %TCIDATA{CSTFile=amsprtci.cst} %TCIDATA{LastRevised=Thu Apr 15 16:36:01 1999} %TCIDATA{} %TCIDATA{Language=American English} \theoremstyle{definition} \newtheorem{definition}{Definition}[section] \newtheorem{remark}[definition]{Remark} \newtheorem{example}[definition]{Example} \newtheorem{conjecture}[definition]{Conjecture} \theoremstyle{plain} \newtheorem{theorem}[definition]{Theorem} \newtheorem{lemma}[definition]{Lemma} \newtheorem{corollary}[definition]{Corollary} \newtheorem{proposition}[definition]{Proposition} \theoremstyle{remark} \newtheorem*{acknowledgements}{Acknowledgements} %\input inner \newbox\ipbox \newcommand{\ip}[2]{\left\langle #1\mathrel{\mathchoice {\setbox\ipbox=\hbox{$\displaystyle \left\langle\mathstrut #1#2\right\rangle$} \vrule height\ht\ipbox width0.25pt depth\dp\ipbox} {\setbox\ipbox=\hbox{$\textstyle \left\langle\mathstrut #1#2\right\rangle$} \vrule height\ht\ipbox width0.25pt depth\dp\ipbox} {\setbox\ipbox=\hbox{$\scriptstyle \left\langle\mathstrut #1#2\right\rangle$} \vrule height\ht\ipbox width0.25pt depth\dp\ipbox} {\setbox\ipbox=\hbox{$\scriptscriptstyle \left\langle\mathstrut #1#2\right\rangle$} \vrule height\ht\ipbox width0.25pt depth\dp\ipbox} } #2\right\rangle} %\input openone \def\openone%{\hbox{\upshape \small1\kern-3.3pt\normalsize1}} {\mathchoice {\hbox{\upshape \small1\kern-3.3pt\normalsize1}} {\hbox{\upshape \small1\kern-3.3pt\normalsize1}} {\hbox{\upshape \tiny1\kern-2.3pt\SMALL1}} {\hbox{\upshape \Tiny1\kern-2pt\tiny1}}} \newlength{\customskipamount} \setlength{\customskipamount}{5pt plus2.5pt minus1pt} \newlength{\customleftmargin} \settowidth{\customleftmargin}{PR2)} \addtolength{\customleftmargin}{1.1em} \addtolength{\customleftmargin}{\labelsep} \renewcommand{\theequation}{\thesection .\arabic{equation}} \renewcommand{\theenumi}{\roman{enumi}} \renewcommand{\labelenumi}{(\theenumi)} \setlength{\headheight}{1.2\headheight} \newlength{\qedskip} \newlength{\qedadjust} \input cyracc.def \font\eightcyr=wncyr8 \def\cyr{\eightcyr\cyracc} \renewcommand{\emptyset}{\varnothing} \def\F{\Phi} \def\G{\Gamma} \def\D{\Delta} \def\L{\Lambda} \def\go{\omega} \def\P{\Psi} \def\gS{\Sigma} \def\Ps{\Psi} \def\T{\Theta} \def\X{\Xi} \def\a{\alpha} \def\b{\beta} \def\d{\delta} \def\k{\kappa} \def\e{\epsilon} \def\f{\varphi} \def\g{\gamma} \def\h{\eta} \def\l{\lambda} \def\m{\mu} \def\n{\nu} \def\gO{\Omega} \def\p{\pi} \def\ps{\psi } \def\t{\tau} \def\r{\rho} \def\s{\sigma} \def\hooklongrightarrow{\lhook\joinrel\longrightarrow} \begin{document} \title[Unitary Representations and O-S Duality]{Unitary Representations and Osterwalder-Schrader Duality} \author{Palle E. T. Jorgensen} \address{Department of Mathematics\\ University of Iowa\\ Iowa City\\ IA 52242} \email{jorgen@math.uiowa.edu} \author{Gestur \'Olafsson} \address{Department of Mathematics\\ Louisiana State University\\ Baton Rouge\\ LA 70803} \email{olafsson@math.lsu.edu} \thanks{Both authors are supported in part by the U.S. National Science Foundation.} \thanks{The second author was supported by LEQSF grant (1996-99)-RD-A-12.} \dedicatory{Dedicated to the memory of Irving E. Segal} \begin{abstract} The notions of reflection, symmetry, and positivity from quantum field theory are shown to induce a duality operation for a general class of unitary representations of Lie groups. The semisimple Lie groups which have this $c$-duality are identified and they are placed in the context of Harish-Chandra's legacy for the unitary representations program. Our paper begins with a discussion of path space measures, which is the setting where reflection positivity (Osterwalder-Schrader) was first identified as a useful tool of analysis. \end{abstract}\maketitle \begin{center} \textit{Le plus court chemin entre deux v\'{e}rit\'{e}s dans}\linebreak \textit{le domaine r\'{e}el passe par le domaine complexe.\bigskip}% \linebreak \settowidth{\qedskip}{\textit{le domaine r{\'{e}}% el passe par le domaine complexe.}}\makebox[\qedskip]{\hfill\textsc {---Jacques Hadamard}}\vspace{24pt} \end{center} \section*{\label{Introduction}Introduction} In this paper, we present an idea which serves to unify the following six different developments: \begin{enumerate} \item \label{IntCat(1)}reflection positivity of quantum field theory, \item \label{IntCat(2)}the role of reflection positivity in functional integration, \item \label{IntCat(3)}the spectral theory of unitary one-parameter groups in Hilbert space, \item \label{IntCat(4)}an extension principle for operators in Hilbert space, \item \label{IntCat(5)}the Bargmann transform, and \item \label{IntCat(6)}reflection positivity and unitary highest weight modules for semisimple Lie groups. \end{enumerate} The emphasis is on (\ref{IntCat(6)}), but the common thread in our paper is the unity of the six areas, which, on the face of it, may perhaps appear to be unrelated. We also stress the interconnection between the six projects, and especially the impact on (\ref{IntCat(6)}) from (\ref{IntCat(1)}% )--(\ref{IntCat(5)}). Readers who may not be familiar with all six of the subjects (\ref{IntCat(1)}% )--(\ref{IntCat(6)}) may wish to consult the bibliography; for example \cite{GlJa87} is an excellent background reference on (\ref{IntCat(1)}% )--(\ref{IntCat(2)}), \cite{FOS83} is especially useful on (\ref{IntCat(2)}), and \cite{Arv84} covers the theory underlying (\ref{IntCat(3)}). Area (\ref{IntCat(4)}) is covered in \cite{Jor80} and \cite{JoMu80}, while \cite{BH94} and \cite{OO96} cover (\ref{IntCat(5)}). Background references on (\ref{IntCat(6)}) include \cite{JO97}, \cite{Jor86}, and \cite{Jor87}. Our main point is to show how the concepts of reflection, symmetry and positivity, which are central notions in quantum field theory, are related to a duality operation for certain unitary representations of semisimple Lie groups. On the group level this duality corresponds to the $c$-duality of causal symmetric spaces, a duality relating the ``compactly causal'' spaces to the ``non-compactly causal'' spaces. On the level of representations one starts with a unitary representation of a group $G$ with an involution $\tau$ corresponding to a non-compactly causal space, then uses an involution on the representation space satisfying a certain positivity condition on a subspace to produce a contraction representation of a semigroup $H\exp\left( C\right) $, where $H=G^{\tau}$ and $C$ is an $H$-invariant convex cone lying in the space of $\tau$-fixed points in the Lie algebra. Now a general result of L\"uscher and Mack and the co-authors can be used to produce a unitary representation of the $c$-dual group $G^{c}$ on the same space. (See \cite{HO95}, \cite{HiNe93}, \cite{LM75}, and \cite{O98} for these terms.) We aim to address several target audiences: workers in representation theory, mathematical physicists, and specialists in transform theory. This diversity has necessitated the inclusion of a bit more background material than would perhaps otherwise be called for: certain ideas are typically explained slightly differently in the context of mathematical physics from what is customary among specialists in one or more of the other areas. The symmetry group for classical mechanics is the \textit{Euclidean motion group} $E_{n}=SO(n)\times_{\mathrm{sp}}\mathbb{R}^{n}$, where the subscript ${}_{\mathrm{sp}}$ stands for semidirect product. Here the action of $(A,\mathbf{x})\in E_{n}$ on $\mathbb{R}^{n}$ is given by $(A,\mathbf{x}% )\cdot\mathbf{v}=A(\mathbf{v})+\mathbf{x}$, that is $SO(n)$ acts by rotations and $\mathbb{R}^{n}$ acts by translations. The connected symmetry group for the space-time of relativity is the \textit{Poincar\'{e} group} $P_{n}% =SO_{o}(n-1,1)\otimes_{sp}\mathbb{R}^{n}$. Let $x_{n}$ stand for the time coordinate, that is $t=x_{n}$. Those two symmetry groups of physics are related by transition to imaginary time, that is multiplying $x_{n}$ by $i$. This in particular changes the usual Euclidean form $(\mathbf{x}\mid \mathbf{y})=x_{1}y_{n}+\cdots+x_{n}y_{n}$ into the Lorentz form $q_{n-1,1}% (\mathbf{x},\mathbf{y})=x_{1}y_{1}+\cdots+x_{n-1}y_{n-1}-x_{n}y_{n}$, which is invariant under the group $SO_{o}(n-1,1)$. Those groups and the corresponding geometry can be related by the $c$-duality. Define an involution $\tau\colon E_{n}\rightarrow E_{n}$ by \[ \tau(A,\mathbf{x})=(I_{n-1,1}AI_{n-1,1},I_{n-1,1}\mathbf{x})\,,\qquad I_{n-1,1}=\left( \begin{array} [c]{cc}% I_{n-1} & 0\\ 0 & -1 \end{array} \right) \,. \] The differential $\tau\colon\frak{e}_{n}\rightarrow\frak{e}_{n}$ is given by the same formula, and $\tau$ is an involution on $\frak{{e}_{n}}$. Let \begin{align*} \frak{h}:= & \{X\in\frak{e}_{n}\mid\tau(X)=X\}\simeq\frak{s}\frak{o}% (n-1)\times_{sp}\mathbb{R}^{n-1}\\ \frak{q}:= & \{X\in\frak{e}_{n}\mid\tau(X)=-X\} \end{align*} The $c$-dual Lie algebra $\frak{e}^{c}$ is defined by: \[ \frak{e}_{n}^{c}:=\frak{h}\oplus i\frak{q}\,. \] A simple calculation shows that $\frak{e}_{n}\simeq\frak{p}_{n}$. Let $G=E_{n}$ and let $G^{c}$ denote the simply connected Lie group with Lie algebra $\frak{{e}^{c}}$. Then $G^{c}=\tilde{P}_{n}$, the universal covering group of $P_{n}$. Given that a physical system is determined by a \textit{unitary representation} $(\pi,\mathbf{H}(\pi))$ of the symmetry group, in our case $E_{n}$ and $P_{n}$, the problem is reduced to use ``analytic continuation'' to move unitary representations of $E_{n}$ to unitary representations of $P_{n}$ by passing over to imaginary time. This idea was used in the paper by J. Fr\"{o}hlich, K. Osterwalder, and E. Seiler in \cite{FOS83}, see also \cite{KlLa83}, to \textit{construct quantum field theoretical systems using Euclidean field theory}. In this paper we will give some general constructions and ideas related to this problem in the context of the applications (\ref{IntCat(1)})--(\ref{IntCat(5)}) mentioned above, and work out some simple examples. In \cite{Sch86} R. Schrader used this idea to construct, from a \textit{complementary series representation} of $SL(2n,\mathbb{C})$, a unitary representation of the group $SU(n,n)\times SU(n,n)$. In that paper the similarities to the \textit{Yang-Baxter} relation were also discussed, a theme that we will leave out in this exposition. What was missing in R.~Schrader's paper was the \textit{identification} of the resulting representations and a general procedure \textit{how} to construct those representations. We will see that we can do this for all simple Lie groups where the associated Riemannian symmetric space $G/K$ is a tube domain, and the that the duality works between complementary series representations and highest weight representations. In general this problem can be formulated in terms of $c$\textit{-duality} of Lie groups and the analytic continuation of unitary representations from one real form to another. (See \cite{HO95} for these terms.) The representations that show up in the case of semisimple groups are generalized principal series representations on the one side and highest weight representations on the other. The symmetric spaces are the causal symmetric spaces, and the duality is between non-compactly causal symmetric spaces and compactly causal symmetric spaces. The latter correspond bijectively to real forms of bounded symmetric domains. Therefore both the geometry and the representations are closely related to the work of Harish-Chandra on bounded symmetric domains and the holomorphic discrete series \cite{HCIV,HCV,HCVI}. But the ideas are also related to the work of I. Segal and S. Paneitz through the notion of causality and invariant cones, \cite{Segal,Paneitz81,Paneitz84}. A more complete exposure can be found in the joint paper by the coauthors: \textit{Unitary Representations of Lie Groups with Reflection Symmetry}, \cite{JO97}. There are other interesting and related questions, problems, and directions. In particular we would like to mention the analytic continuation of the $H$-invariant character of the highest weight representations, and the reproducing kernel of the Hardy space realization of this representation to a spherical distribution character (spherical function) of the generalized principal series representation that we started with. This connects the representations that show up in the duality on the level of distribution characters. For this we refer to \cite{O98, BK98}. The paper is organized as follows. The list also includes some sources for additional background references:\bigskip\newline \ref{Considerations}. Some Spectral Considerations Related to Reflection Positivity \cite{FOS83,OsSc73,OsSc75} \hrulefill\ \pageref{Considerations}\newline \ref{uopg}. Unitary One-Parameter Groups and Path Space Integrals \cite{Arv84,KlLa81} \hrulefill\ \pageref{uopg}\newline \ref{axbg}. The $(ax+b)$-Group \cite{JO97,Ped90} \hrulefill\ \pageref{axbg}\newline \ref{Hilbert}. The Hilbert Transform \cite{GF89,Seg98} \hrulefill\ \pageref{Hilbert}\newline \ref{RemConsiderationsMar.1}. One-Parameter Groups \hrulefill\ \pageref{RemConsiderationsMar.1}\newline \ref{Setting}. The General Setting \cite{JO97} \hrulefill\ \pageref{Setting}\newline \ref{Preliminaries}. Preliminaries \cite{GlJa87,Phil,Sto51,Mac} \hrulefill\ \pageref{Preliminaries}\newline \ref{Basic}. A Basic Lemma \cite{JO97} \hrulefill\ \pageref{Basic}\newline \ref{S-Hrep}. Holomorphic Representations \cite{HC65,HC70,Nee94,'OO88a,'OO88b} \hrulefill\ \pageref{S-Hrep}\newline \ref{LM}. The L\"uscher-Mack Theorem \cite{LM75} \hrulefill\ \pageref{LM}\newline \ref{SSS}. Bounded Symmetric Domains \cite{HiNe93,HO95,Hel62,He78} \hrulefill\ \pageref{SSS}\newline \ref{S-hwm}. Highest Weight Modules \cite{EHW83,Stanton,Ne94,Ne99,FK94,HO,Ja83,KNO98} \hrulefill\ \pageref{S-hwm}\newline \ref{S:Ex}. An Example: $SU(1,1)$ \cite{JO97} \hrulefill\ \pageref{S:Ex}\newline \ref{S:Sssp}. Reflection Symmetry for Semisimple Symmetric Spaces \cite{Jor86,Jor87,JO97,Sch86} \hrulefill\ \pageref{S:Sssp}\newline \ref{S:Bargmann}. The Segal-Bargmann Transform \cite{OO96} \hrulefill\ \pageref{S:Bargmann}\newline \ref{Diagonal}. The Heisenberg Group \cite{BH94,Jor88,JO97} \hrulefill\ \pageref{Diagonal}\newline \ref{axb}. The $(ax+b)$-Group Revisited \cite{JO97,LaPh,Hel64} \hrulefill\ \pageref{axb}\newline References \hrulefill\ \pageref{refpage} \section{\label{Considerations}Some Spectral Considerations Related to Reflection Positivity} \setcounter{equation}{0} \subsection{\label{uopg}Unitary One-Parameter Groups and Path Space Integrals} The term \emph{reflection positivity} is from quantum field theory (QFT) where it refers to a certain reflection in the time-variable. As we explained in the introduction this reflection is also the one which makes the analytic continuation between the (Newtonian) group of rigid motions and the Poincar\'{e} group of relativity. This is the approach to QFT of Osterwalder and Schrader \cite{OsSc73,OsSc75}. The approach implies a change of the inner product, and the new Hilbert space which carries a \emph{unitary} representation $\tilde{\pi}$ of the Poincar\'{e} group $P_{4}$ results from the corresponding ``old'' one by passing to a subspace where the positivity (the so-called Osterwalder-Schrader positivity) is satisfied. An energy operator may then be associated with this ``new'' unitary representation $\tilde{\pi}$ of $P_{4}$. This representation $\tilde{\pi}$ is ``physical'' in that the corresponding energy is positive. The basic connection between the two groups may further be understood from the corresponding quadratic forms on space-time $\left( x,t\right) $, $x=\left( x_{1},x_{2},x_{3}\right) $, $\left( x,t\right) \mapsto\left\| x\right\| ^{2}+t^{2}$ with $\left\| x\right\| ^{2}=x_{1}^{2}+x_{2}^{2}+x_{3}^{2}$. The analytic continuation $t\mapsto it$, $i=\sqrt{-1}$, turns this into the form $\left\| x\right\| ^{2}-t^{2}$ of relativity. This same philosophy may also be used in an analytic continuation argument connecting Feynman measure with the Wiener measure on path space. This is important since the Wiener measure seems to be the most efficient way of making precise the Feynman measure, which involves infinite ``renormalizations'' if given a literal interpretation. We refer to \cite{GlJa87} and \cite{Nel64} for more details on this point. For the convenience of the reader we include here a simple instance of reflection positivity for a path space measure which will be needed later: Let $\mathcal{D}=\mathcal{D}\left( \mathbb{R}\right) =C_{c}^{\infty}\left( \mathbb{R}\right) $ denote the usual test functions on $\mathbb{R}$, and the dual space $\mathcal{D}^{\prime}=\mathcal{D}^{\prime}\left( \mathbb{R}% \right) $ of distributions. (We use the notation $q\left( f\right) $, $q\in\mathcal{D}^{\prime}$, $f\in\mathcal{D}$.) Let $H_{0}=-\frac{1}% {2}\bigtriangleup+\frac{1}{2}q^{2}-\frac{1}{2}$ be the harmonic oscillator Hamiltonian, and form $\hat{H}=H_{0}-E_{0}$ picking $E_{0}$ such that $\hat {H}\geq0$ and $\hat{H}\Omega=0$ for a ground state vector $\Omega$. Then by \cite{GlJa87} there is a unique path-space measure $\phi_{0}$ on $\mathcal{D}^{\prime}$ such that \begin{align*} \int q\left( t\right) \,d\phi_{0}\left( q\right) & =0,\\ \int q\left( t_{1}\right) q\left( t_{2}\right) \,d\phi_{0}\left( q\right) & =\frac{1}{2}e^{-\left| t_{1}-t_{2}\right| }. \end{align*} One rigorous interpretation (see \cite{Nel64,Nel73}) is to view $\left( q\left( t\right) \right) $ here as a stochastic process, i.e., a family of random variables indexed by $t$. Further, for each $t>0$, and for each ``even + linear'' real potential $V\left( q\right) $, there is a unique measure $\mu_{t}$ on $\mathcal{D}^{\prime}$ such that \[ d\mu_{t}=Z_{t}^{-1}\exp\left( -\int_{-t/2}^{t/2}V\left( q\left( s\right) \right) \,ds\right) \,d\phi_{0}\left( q\right) \] with \[ Z_{t}=\int\exp\left( -\int_{-t/2}^{t/2}V\left( q\left( s\right) \right) \,ds\right) \,d\phi_{0}\left( q\right) . \] The consideration leading to the relation between the measures $d\phi_{0}$ and $d\mu_{t}$ is the Trotter approximation for the (analytically continued) semigroup,% \[ e^{-it\left( -\frac{1}{2}\bigtriangleup+V\right) }=\lim_{n\rightarrow\infty }\left( e^{\left( it/2n\right) \bigtriangleup}e^{-\left( it/n\right) V}\right) ^{n}. \] See \cite{Nel64} for more details on this point. When the operator $\left( \cdots\right) ^{n}$ on the right-hand side is computed, we find the integral kernel \[ K^{\left( n\right) }\left( x_{0},x_{n},t\right) =\frac{1}{N_{n}}\int e^{i\mathcal{S}\left( x_{0},\dots,x_{n},t\right) }\,dx_{1}\,\cdots\,dx_{n-1}% \] with% \[ N_{n}=\left( \frac{2\pi it}{n}\right) ^{3n/2}% \] and% \[ \mathcal{S}\left( x_{0},\dots,x_{n},t\right) =\frac{1}{2}\sum_{j=1}% ^{n}\left| x_{j-1}-x_{j}\right| ^{2}\left( \frac{t}{n}\right) ^{-1}% -\sum_{j=1}^{n}V\left( x_{j}\right) \left( \frac{t}{n}\right) . \] The heuristic motivation for $\phi_{0}$ and $\mu_{t}$ is then the ``action'' $\mathcal{S}$ approximating% \[ \mathcal{S}\left( q\right) =\frac{1}{2}\int_{0}^{t}\left( \dot{q}\left( s\right) \right) ^{2}\,ds-\int_{0}^{t}V\left( q\left( s\right) \right) \,ds \] via% \[ q\left( t_{j}\right) =x_{j},\qquad\bigtriangleup t_{j}=\frac{t}{n}, \] and% \[ \dot{q}\left( t_{j}\right) \sim\frac{x_{j}-x_{j-1}}{t/n}% \] If $-t/20,\,b\in\mathbb{R} $. We take $Q= \mathbb{R}$ and $H=\mathbb{R}^{+}$. Then $\tau$ is given by $\tau (a,b)\;=(a,-b)$. \textbf{5. The Heisenberg group:} Let $H_{n}$ be the $(2n+1)$-dimensional Heisenberg group. We write \[ H_{n}=\left\{ h(\mathbf{x},\mathbf{y},z)=\left( \begin{array} [c]{lll}% 1 & \mathbf{x}^{t} & z\\ 0 & I_{n-1} & \mathbf{y}\\ 0 & 0 & 1 \end{array} \right) \,\,\mathbf{x},\mathbf{y}\in\mathbb{R}^{n},\,z\in\mathbb{R}\right\} \simeq\mathbb{R}^{n}\times\mathbb{R}^{n}\times\mathbb{R}% \] with multiplication given by \[ h(\mathbf{x},\mathbf{y},z)h(\mathbf{x}^{\prime},\mathbf{y}^{\prime},z^{\prime })=h(\mathbf{x}+\mathbf{x}^{\prime},\mathbf{y}+\mathbf{y}^{\prime}% ,z+z^{\prime}+(\mathbf{x}\mid\mathbf{y}^{\prime}))\,. \] In particular $h(\mathbf{x},\mathbf{y},z)^{-1}=h(-\mathbf{x},-\mathbf{y}% ,(\mathbf{x}\mid\mathbf{y})-z)$. In this case we take $H=\{h(\mathbf{x}% ,0,0)\}\simeq\mathbb{R}^{n}$, which is abelian, and $Q=\{h(0,\mathbf{y},z)\in H_{n}\mid\mathbf{y}\in\mathbb{R}^{n},\,z\in\mathbb{R}\}\simeq\mathbb{R}^{n+1}% $. The involution is given by \[ \tau(h(\mathbf{x},\mathbf{y},z)):=h(\mathbf{x},-\mathbf{y},-z)\,. \] Starting from the pair $(G,\tau)$ and a unitary representation $(\pi ,\mathbf{H}(\pi))$ of $G$ we need a compactible involution on the Hilbert space $\mathbf{H}(\pi)$, that is a unitary linear map $J\colon\mathbf{H}% (\pi)\rightarrow\mathbf{H}(\pi)$ intertwining $\pi$ and $\pi\circ\tau$. Thus \[ J\pi(g)=\pi(\tau(g))J\qquad\forall g\in G\,. \] We will also need a semigroup $S$ such that $H\subset S$ or at least $H\subset\overline{S}$ where the bar denotes topological closure. We shall consider closed subspaces $\mathbf{K}_{0}\subset\mathbf{H}(\pi)$, where $\mathbf{H}(\pi)$ is the Hilbert space of $\pi$, such that $\mathbf{K}_{0}$ is invariant under $\pi(S^{o})$. Let $J\colon\mathbf{H}(\pi)\rightarrow \mathbf{H}(\pi)$ be a unitary intertwining operator for $\pi$ and $\pi \circ\tau$ such that $J^{2}=\operatorname*{id}$. We assume that $\mathbf{K}% _{0}$ may be chosen such that $\Vert v\Vert_{J}^{2}:=% %TCIMACRO{\TeXButton{ip}{\ip{v}{Jv}}}% %BeginExpansion \ip{v}{Jv}% %EndExpansion \geq0$ for all $v\in\mathbf{K}_{0}$. We will always assume our inner product conjugate linear in the first argument. We form, in the usual way, the Hilbert space $\mathbf{K}=\left( \mathbf{K}_{0}/\mathbf{N}\right) \sptilde$ by dividing out with $\mathbf{N}=\{v\in\mathbf{K}_{0}\mid% %TCIMACRO{\TeXButton{ip}{\ip{v}{Jv}}}% %BeginExpansion \ip{v}{Jv}% %EndExpansion =0\}$ and completing in the norm $\Vert\cdot\Vert_{J}$. (This is of course a variation of the Gelfand-Naimark-Segal (GNS) construction.) With the properties of $(G,\pi,\mathbf{H}(\pi),\mathbf{K})$ as stated, we show, using the L\"uscher-Mack theorem, that the simply connected Lie group $G^{c}$ with Lie algebra $\frak{g}^{c}=\frak{h}\oplus i\frak{q}$ carries a \textit{unitary} representation $\pi^{c}$ on $\mathbf{K}$ such that $\{\pi^{c}(h\exp(iY))\mid h\in H,Y\in C^{o}\}$ is obtained from $\pi$ by passing the corresponding operators $\pi(h\exp Y)$ to the quotient $\mathbf{K}_{0}/\mathbf{N}$. To see that this leads to a unitary representation $\pi_{c}$ of $G^{c}$ we use a basic result of L\"uscher and Mack \cite{LM75} and in a more general context one of Jorgensen \cite{Jor86}. In fact, when $Y\in C$, the selfadjoint operator $d\pi(Y)$ on $\mathbf{K}$ has spectrum contained in $(-\infty,0]$. As in Lemma \ref{MNoughtZero}, we show that in the case where $C$ extends to an $G^{c}$ invariant regular cone in $i\frak{g}^{c}=i\frak{h}\oplus\frak{q}$ and $\pi^{c}$ is injective, then each $\pi^{c}$ (as a unitary representation of $G^{c}$) must be a direct integral of highest-weight representations of $G^{c}$. The examples show that one can relax the condition in different ways, that is one can avoid using the L\"uscher-Mack theorem by instead constructing local representations and using only cones that are neither generating nor $H$-invariant. Assume now that $G$ is a semidirect product of $H$ and $N$ with $N$ normal and abelian. Define $\tau\colon G\rightarrow G$ by $\tau(hn)=hn^{-1}$. Let $\pi \in\hat{H}$ (the unitary dual) and extend $\pi$ to a unitary representation of $G$ by setting $\pi(hn)=\pi(h)$. In this case, $G^{c}$ is locally isomorphic to $G$, and $\pi$ gives rise to a unitary representation $\pi^{c}$ of $G^{c}$ by the formula $d\pi^{c}(X)=d\pi(X)$, $X\in\frak{h}$, and $d\pi^{c}% |_{i\frak{q}}=0$. A special case of this is the $3$-dimensional Heisenberg group, and the $(ax+b)$-group. In Sections \ref{Diagonal} and \ref{axb}, we show that, if we induce instead a character of the subgroup $N$ to $G$, then we have $\left( \mathbf{K}_{0}/\mathbf{N}\right) \sptilde=\{0\}$. Our approach to the general representation correspondence $\pi\mapsto\pi^{c} $ is related to the integrability problem for representations of Lie groups (see \cite{JoMo84}); but the present positivity viewpoint comes from Osterwalder-Schrader positivity; see \cite{OsSc73, OsSc75}. In addition the following other papers are relevant in this connection: \cite{FOS83,Jor86,Jor87,KlLa83,Pra89,Sch86}. \section{\label{Preliminaries}Preliminaries} \setcounter{equation}{0} The setting for the paper is a general Lie group $G$ with a nontrivial involutive automorphism $\tau$. \begin{definition} \label{ReflectionSymmetric}A unitary representation $\pi$ acting on a Hilbert space $\mathbf{H}(\pi)$ is said to be \textit{reflection symmetric} if there is a unitary operator $J\colon\mathbf{H}(\pi)\rightarrow\mathbf{H}(\pi)$ such that \begin{list}{}{\setlength{\leftmargin}{\customleftmargin} \setlength{\itemsep}{0.5\customskipamount} \setlength{\parsep}{0.5\customskipamount} \setlength{\topsep}{\customskipamount} \setlength{\parindent}{0pt}} \item[\hss\llap{\rm R1)}] ${\displaystyle J^2 = \operatorname*{id}}$; \item[\hss\llap{\rm R2)}] ${\displaystyle J\pi(g) = \pi(\tau(g))J\, , \quad g\in G}$. \end{list} \end{definition} If (R1) holds, then $\pi$ and $\pi\circ\tau$ are equivalent. Furthermore, generally from (R2) we have $J^{2}\pi(g) = \pi(g) J^{2}$. Thus, if $\pi$ is irreducible, then we can always renormalize $J$ such that (R1) holds. Let $H = G^{\tau}= \{g\in G\mid\tau(g) = g\}$ and let $\frak{h}$ be the Lie algebra of $H$. Then $\frak{h} = \{ X\in\frak{g}\mid\tau(X) = X\}$. Define $\frak{q} = \{Y\in\frak{g}\mid\tau(Y) = -Y\}$. Then $\frak{g} = \frak{h} \oplus\frak{q}$, $[\frak{h},\frak{q}]\subset\frak{q}$ and $[\frak{q},\frak{q}]\subset\frak{h}$. \begin{definition} \label{Hyperbolic}A closed convex cone $C\subset\frak{q}$ is \textit{hyperbolic} if $C^{o}\not =\emptyset$ and if $\operatorname{ad}X$ is semisimple with real eigenvalues for every $X\in C^{o}$. \end{definition} We will assume the following for $(G, \pi,\tau, J)$: \begin{list}{}% {\setlength{\leftmargin}{\customleftmargin} \setlength{\itemsep}{0.5\customskipamount} \setlength{\parsep}{0.5\customskipamount} \setlength{\topsep}{\customskipamount}} \item[\hss\llap{{\rm PR1)}}] $\pi $ is reflection symmetric with reflection $J$; \item[\hss\llap{{\rm PR2)}}] there is an $H$-invariant hyperbolic cone $C\subset\frak{q}$ such that $S(C) = H\exp C$ is a closed semigroup and $S(C)^o = H\exp C^o$ is diffeomorphic to $H\times C^o$; \item[\hss\llap{{\rm PR3)}}] there is a subspace ${0}\not= \mathbf{K}_0\subset\mathbf{H}(\pi )$ invariant under $S(C)$ satisfying the positivity condition \[ \ip{v}{v}_J:= \ip{v}{J(v)} \ge0,\quad\forall v\in\mathbf{K}_0\, .\] \end{list} \begin{remark} \label{WhatToAssume}In (PR3) we can always assume that $\mathbf{K}_{0}$ is closed, as the invariance and the positivity pass over to the closure. In (PR2) it is only necessary to assume that $\mathbf{K}_{0}$ is invariant under $\exp C$, as one can always replace $\mathbf{K}_{0}$ by $\overline {\left\langle \pi(H)\mathbf{K}_{0}\right\rangle }$, the closed space generated by $\pi(H)\mathbf{K}_{0}$, which is $S(C)$-invariant, as $C$ is $H$-invariant. For the exact conditions on the cone for (PR2) to hold see the original paper by J. Lawson \cite{JL94}, or the monograph \cite[pp. 194 ff.]{HiNe93}. \end{remark} In some of the examples we will replace (PR2)and (PR3) by the following: weaker conditions \begin{list}{}{\setlength{\leftmargin}{\customleftmargin} \setlength{\itemsep}{0.5\customskipamount} \setlength{\parsep}{0.5\customskipamount} \setlength{\topsep}{\customskipamount}} \item[\hss\llap{\rm PR2$^{\prime}$)}] $C$ is (merely) some nontrivial cone in $\frak{q} $. \item[\hss\llap{\rm PR3$^{\prime}$)}] There is a subspace $0\not= \mathbf{K}_0\subset\mathbf{H} (\pi)$ invariant under $H$ and $\exp C$ satisfying the positivity condition from (PR3). \end{list} (See Section \ref{Diagonal} for further details.) Since the operators $\{\pi(h)\mid h\in H\}$ commute with $J$, they clearly pass to the quotient by \[ \mathbf{N}:=\{v\in\mathbf{K}_{0}\mid% %TCIMACRO{\TeXButton{ip}{\ip{v}{ Jv}}}% %BeginExpansion \ip{v}{ Jv}% %EndExpansion =0\} \] and implement unitary operators on $\mathbf{K}:=\left( \mathbf{K}% _{0}/\mathbf{N}\right) \sptilde$ relative to the inner product induced by \begin{equation}% %TCIMACRO{\TeXButton{ip}{\ip{u}{v}}}% %BeginExpansion \ip{u}{v}% %EndExpansion _{J}:=% %TCIMACRO{\TeXButton{ip}{\ip{u}{J(v)}}}% %BeginExpansion \ip{u}{J(v)}% %EndExpansion \,. \label{E:innerpr}% \end{equation} which will be denoted by the same symbol. Hence we shall be concerned with passing the operators $\{\pi(\exp Y)\mid Y\in C\}$ to the quotient $\mathbf{K}_{0}/\mathbf{N}$, and for this we need a basic Lemma. In general, when $\left( \mathbf{K}_{0},J\right) $ is given, satisfying the positivity axiom, then the corresponding composite quotient mapping \[ \mathbf{K}_{0}\longrightarrow\mathbf{K}_{0}/\mathbf{N}\hooklongrightarrow \left( \mathbf{K}_{0}/\mathbf{N}\right) \sptilde=:\mathbf{K}% \] is \emph{contractive} relative to the respective Hilbert norms. The resulting (contractive) mapping will be denoted $\beta$. An operator $\gamma$ on $\mathbf{H}$ which leaves $\mathbf{K}_{0}$ invariant is said to \emph{induce} the operator $\tilde{\gamma}$ on $\mathbf{K}$ if $\beta\circ\gamma =\tilde{\gamma}\circ\beta$ holds on $\mathbf{K}_{0}$. In general, an induced operation $\gamma\mapsto\tilde{\gamma}$ may not exist; and, if it does, $\tilde{\gamma}$ may fail to be bounded, even if $\gamma$ is bounded. This above-mentioned operator-theoretic formulation of reflection positivity has applications to the Feynman-Kac formula in mathematical physics, and there is a considerable literature on that subject, with work by E. Nelson \cite{Nel64,Nel73}, A. Klein and L.J. Landau \cite{Kle78,KlLa75,KlLa81}, B. Simon, and W.B. Arveson \cite{Arv84}. Since we shall not use path space measures here, we will omit those applications, and instead refer the reader to the survey paper \cite{Arv84} (lecture 4) by W.B. Arveson. In addition to mathematical physics, our motivation also derives from recent papers on non-commutative harmonic analysis which explore analytic continuation of the underlying representations; see, e.g., \cite{HOO91,BK98,Nee94,'O90a,O93,Ol82}. \section{\label{Basic}A Basic Lemma} \setcounter{equation}{0} \begin{lemma} \label{BasicLemma} \begin{enumerate} \item [\hss\llap{\rm1)}]Let $J$ be a period-$2$ unitary operator on a Hilbert space $\mathbf{H}$, and let $\mathbf{K}_{0}\subset\mathbf{H}$ be a closed subspace such that $% %TCIMACRO{\TeXButton{ip}{\ip{v}{J(v)}}}% %BeginExpansion \ip{v}{J(v)}% %EndExpansion \geq0$, $v\in\mathbf{K}_{0}$. Let $\gamma$ be an invertible operator on $\mathbf{H}$ such that $J\gamma=\gamma^{-1}J$ and which leaves $\mathbf{K}% _{0}$ invariant and has $(\gamma^{-1})^{\ast}\gamma$ bounded on $\mathbf{H}$. Then $\gamma$ induces a bounded operator $\tilde{\gamma}$ on $\mathbf{K}% =\left( \mathbf{K}_{0}/\mathbf{N}\right) \sptilde$, where $\mathbf{N}% =\{v\in\mathbf{K}_{0}\mid% %TCIMACRO{\TeXButton{ip}{\ip{v}{Jv}}}% %BeginExpansion \ip{v}{Jv}% %EndExpansion =0\}$, and the norm of $\tilde{\gamma}$ relative to the $J$-inner product in $\mathbf{K}$ satisfies \begin{equation} \Vert\tilde{\gamma}\Vert\leq\Vert(\gamma^{-1})^{\ast}\gamma\Vert_{sp}^{1/2}\,, \label{E:3.1}% \end{equation} where $\Vert\cdot\Vert_{sp}$ is the spectral radius. \item[\hss\llap{\rm2)}] If we have a semigroup $S$ of operators on $\mathbf{H}$ satisfying the conditions in \textup{(1)}, then \begin{equation} (\gamma_{1}\gamma_{2})\sptilde=\tilde{\gamma_{1}}\tilde{\gamma_{2}}% \,,\quad\gamma_{1},\gamma_{2}\in S\,. \label{E:3.2}% \end{equation} \end{enumerate} \end{lemma} \begin{proof} For $v\in\mathbf{K}_{0}$, $v\not =0$, we have \begin{align*} \Vert\gamma(v)\Vert_{J}^{2} & =% %TCIMACRO{\TeXButton{ip}{\ip{\gamma(v)}{J\gamma(v)}}}% %BeginExpansion \ip{\gamma(v)}{J\gamma(v)}% %EndExpansion \\ & =% %TCIMACRO{\TeXButton{ip}{\ip{\gamma(v)}{\gamma^{-1}J(v)}}}% %BeginExpansion \ip{\gamma(v)}{\gamma^{-1}J(v)}% %EndExpansion \\ & =% %TCIMACRO{\TeXButton{ip}{\ip{(\gamma^{-1})^{*}\gamma(v)}{J(v)}}}% %BeginExpansion \ip{(\gamma^{-1})^{*}\gamma(v)}{J(v)}% %EndExpansion \\ & =% %TCIMACRO{\TeXButton{ip}{\ip{(\gamma^{-1})^{*}\gamma(v)}{v}}}% %BeginExpansion \ip{(\gamma^{-1})^{*}\gamma(v)}{v}% %EndExpansion _{J}\\ & \leq\Vert(\gamma^{-1})^{\ast}\gamma(v)\Vert_{J}\Vert v\Vert_{J}\\ & \leq\Vert((\gamma^{-1})^{\ast}\gamma)^{2}(v)\Vert_{J}^{1/2}\Vert v\Vert _{J}^{1+1/2}\\ & \mkern4.5mu\vdots\\ & \leq\Vert((\gamma^{-1})^{\ast}\gamma)^{2^{n}}(v)\Vert_{J}^{1/2^{n}}\Vert v\Vert_{J}^{1+1/2+\cdots+1/2^{n}}\\ & \leq\left( \Vert((\gamma^{-1})^{\ast}\gamma)^{2^{n}}\Vert\Vert v\Vert\right) ^{1/2^{n}}\Vert v\Vert_{J}^{2}\,. \end{align*} Since $\displaystyle{\lim_{n\rightarrow\infty}\Vert((\gamma^{-1})^{\ast}% \gamma)^{2^{n}}\Vert^{1/2^{n}}=\Vert(\gamma^{-1})^{\ast}\gamma\Vert_{sp}}$, and ${\displaystyle \lim_{n\rightarrow\infty}\Vert v\Vert^{1/2^{n}}=1}$, the result follows. By this we get \[% %TCIMACRO{\TeXButton{ip}{\ip{\gamma(v)}{J\gamma(v)}}}% %BeginExpansion \ip{\gamma(v)}{J\gamma(v)}% %EndExpansion \leq\Vert(\gamma^{-1})^{\ast}\gamma\Vert_{sp}% %TCIMACRO{\TeXButton{ip}{\ip %{v}{J(v)}}}% %BeginExpansion \ip {v}{J(v)}% %EndExpansion \] which shows that $\gamma(\mathbf{N})\subset\mathbf{N}$, whence $\gamma$ passes to a bounded operator on the quotient $\mathbf{K}_{0}/\mathbf{N}$ and then also on $\mathbf{K}$ satisfying the estimate stated in (1). If both the operators in (\ref{E:3.2}) leave $\mathbf{N}$ invariant, so does $\gamma _{1}\gamma_{2}$ and the operator induced by $\gamma_{1}\gamma_{2}$ is $\tilde{\gamma_{1}}\tilde{\gamma_{2}}$ as stated. \end{proof} \begin{corollary} \label{GammaContraction}Let the notation be as above and assume that $\gamma$ is unitary on $\mathbf{H}$. Then the constant on the right in \textup{(\ref{E:3.1})} is one. Hence $\tilde{\gamma}$ is a contraction on $\mathbf{K}$. \end{corollary} To understand the assumptions on the space $\mathbf{K}_{0}$, that is positivity and invariance, we include the following which is based on an idea of R.S. Phillips \cite{Phil}. \begin{proposition} \label{P:3.3}Let $\mathbf{H}$ be a Hilbert space and let $J$ be a period-$2$ unitary operator on $\mathbf{H}$. Let $S$ be a commutative semigroup of unitary operators on $\mathbf{H}$ such that $S=S_{+}S_{-}$ with $S_{+}% =\{\gamma\in S\mid J\gamma=\gamma J\}$ and $S_{-}=\{\gamma\in S\mid J\gamma=\gamma^{-1}J\}$. Then $\mathbf{H}$ possesses a maximal positive and invariant subspace, that is a subspace $\mathbf{K}_{0}$ such that $% %TCIMACRO{\TeXButton{ip}{\ip{v}{J(v)}}}% %BeginExpansion \ip{v}{J(v)}% %EndExpansion \geq0$, $v\in\mathbf{K}_{0}$, and $\gamma\mathbf{K}_{0}\subset\mathbf{K}_{0}$, $\gamma\in S$. \end{proposition} \begin{proof} The basic idea is contained in \cite[pp. 386 ff.]{Phil}. \end{proof} \begin{remark} \label{RemBasicNew.1} A nice application is to the case $\mathbf{H}=L^{2}\left( X,m\right) $ where $X$ is a Stone space. There is a $m$-a.e.-defined automorphism $\theta\colon X\rightarrow X$ such that \[ J\left( f\right) =f\circ \theta ,\qquad f\in L^{2}\left( X,m\right) \] and $S$ is represented by multiplication operators on $L^{2}\left( X,m\right) $. By \cite{Phil} we know that there are clopen subsets $A,B\subset X$ such that with $M_{0}:=\left\{ x\in X\mid \theta \left( x\right) =x\right\} $ and $M_{1}=X\setminus M_{0}$ we have $A,B\subset M_{1}$, $A\cap B=\varnothing $, $A\cup B=M_{1}$ and $\theta \left( A\right) =B$. Let $\mathbf{K}_{0}:=L^{2}\left( M_{0}\cup A\right) $. Then $\mathbf{K}_{0}$ is a maximal positive and invariant subspace. \end{remark} \begin{lemma} \label{MNoughtZero}If $M_{0}\subset X$ is of measure zero, then the space $\mathbf{K}$ is trivial, that is $% %TCIMACRO{\TeXButton{ip}{\ip{f}{J(f)}}}% %BeginExpansion \ip{f}{J(f)}% %EndExpansion =0$ for all $f\in\mathbf{K}_{0}$. \end{lemma} \begin{remark} \label{AbelianSubspace}Assume that we have (PR1) and (PR2). By \cite{KrNe96} there is an abelian subspace $\frak{a}\subset\frak{q}$ such that $C^{o}=\operatorname{Ad}(H)(C^{o}\cap\frak{a})$. Let $S_{A}=\exp(C^{o}% \cap\frak{a})$. Then $S_{A}$ is an abelian semigroup, so one can use Proposition \ref{P:3.3} to construct a maximal positive and invariant subspace for $S_{A}$. But in general we can not expect this space to be invariant under $S$. \end{remark} We read off from the basic Lemma the following Proposition: \begin{proposition} \label{Contractive}Let $\pi$ be a unitary representation of the group $G$. Assume that $(\tau,J,C,\mathbf{K}_{0})$ satisfies the conditions \textup{(PR1), (PR2}$^{\prime}$\textup{),} and \textup{(PR3}$^{\prime}% $\textup{).} If $Y\in C$, then $\pi(\exp Y)$ induces a contractive selfadjoint operator $\tilde{\pi}(\exp Y)$ on $\mathbf{K}$. \end{proposition} \begin{proof} If $Y\in C$, then $\pi(\exp Y)\mathbf{K}_{0}\subset\mathbf{K}_{0}$, and $\pi(\exp Y)$ is unitary on $\mathbf{H}(\pi)$. Thus \begin{align*}% %TCIMACRO{\TeXButton{ip}{\ip{\pi(\exp Y)u}{J(v)}}}% %BeginExpansion \ip{\pi(\exp Y)u}{J(v)}% %EndExpansion & =% %TCIMACRO{\TeXButton{ip}{\ip{u}{\pi(\exp(-Y))J(v)}}}% %BeginExpansion \ip{u}{\pi(\exp(-Y))J(v)}% %EndExpansion \\ & =% %TCIMACRO{\TeXButton{ip}{\ip{u}{J(\pi(\exp Y)v)}}}% %BeginExpansion \ip{u}{J(\pi(\exp Y)v)}% %EndExpansion \,, \end{align*} proving that $\pi(\exp Y)$ is selfadjoint in the $J$-inner product. Since $\pi(\exp Y)$ is unitary on $\mathbf{H}(\pi)$ \[ \Vert\pi(\exp Y)\Vert=\Vert\pi(\exp Y)\Vert_{sp}=1\,, \] and the contractivity property follows. \end{proof} \begin{lemma} \label{HolomorphicRPlus}Let $\pi$ be a unitary representation of $G$ such that $(\tau,J,C,\mathbf{K}_{0})$ satisfies the conditions \textup{(PR1), (PR2}$^{\prime}$\textup{),} and \textup{(PR3}$^{\prime}$\textup{).} Then for $Y\in C$ there is a selfadjoint operator $d\tilde{\pi}(Y)$ in $\mathbf{K}% =\left( \mathbf{K}_{0}/\mathbf{N}\right) \sptilde$, with spectrum contained in $(-\infty,0]$, such that \[ \tilde{\pi}(\exp(tY))=e^{t\,d\tilde{\pi}(Y)},\quad t\in\mathbb{R}_{+}% \] is a contractive semigroup on $\mathbf{K}$. Furthermore the following hold: \begin{enumerate} \item [\hss\llap{\rm1)}]$t\mapsto e^{t\,d\tilde{\pi}(Y)}$ extends to a continuous map $z\mapsto e^{z\,d\tilde{\pi}(Y)}$ on $\{z\in\mathbb{C}% \mid\operatorname{Re}(z)\geq0\}$ holomorphic on the open right half-plane, and such that $e^{(z+w)\,d\tilde{\pi}(Y)}=e^{z\,d\tilde{\pi}(Y)}e^{w\,d\tilde{\pi }(Y)}$. \item[\hss\llap{\rm2)}] There exists a one-parameter group of unitary operators \[ \tilde{\pi}\left( \exp(itY)\right) :=e^{it\,d\tilde{\pi}(Y)},\quad t\in\mathbb{R}% \] on $\mathbf{K}$. \end{enumerate} \end{lemma} \begin{proof} The last statement follows by the spectral theorem. By construction $\{\tilde{\pi}(\exp(tY))\mid t\in\mathbb{R}_{+}\}$ is a semigroup of selfadjoint contractive operators on $\mathbf{K}$. The existence of the operators $d\tilde{\pi}(Y)$ as stated then follows from a general result in operator theory; see, e.g., \cite{Fr80} or \cite{KlLa81}. \end{proof} \begin{corollary} \label{Id}Let the situation be as in the last corollary. If $Y\in C\cap-C$ then $e^{t\,d\tilde{\pi}(Y)}=\operatorname*{id}$ for all $t\in\mathbb{R}_{+}$. In particular $d\tilde{\pi}(Y)=0$ for every $Y\in C\cap-C$. \end{corollary} \begin{proof} This follows as the spectrum of $d\tilde{\pi}(Y)$ and $d\tilde{\pi}(-Y)$ is contained in $(-\infty,0]$. \end{proof} We remark here that we have introduced the map $d\pi$ without using the space of \textit{smooth vectors} for the representation $\pi$. Let us recall that a vector $\mathbf{v}\in\mathbf{H}(\pi)$ is called \textit{smooth} if the map \[ \mathbb{R}\ni t\longmapsto\hat{\mathbf{v}}(t):=\pi(\exp tX)\mathbf{v}% \in\mathbf{H}(\pi) \] is smooth for all $X\in\frak{g}$. The vector is \textit{analytic} if the above map is analytic. We denote by $\mathbf{H}^{\infty}(\pi)$ the space of smooth vectors and by $\mathbf{H}^{\omega}(\pi)$ the space of analytic vectors. Both $\mathbf{H}^{\omega}(\pi)$ and $\mathbf{H}^{\infty}(\pi)$ are $G$-invariant dense subspaces of $\mathbf{H}% (\pi)$. We define a representation of $\frak{g}$ on $\mathbf{H}^{\infty}(\pi)$ by \[ d\pi(X)\mathbf{v}=\lim_{t\rightarrow0}\frac{\pi(\exp tX)\mathbf{v}-\mathbf{v}% }{t}\,. \] Recall that if $\pi$ is infinite-dimensional, then $d\pi$ is a representation of $\frak{g}$ by unbounded operator on $\mathbf{H} (\pi)$, but the analytic vectors and the $C^{\infty}$-vectors form dense domains for $d\pi$; see \cite{Nel59,Pou92,WaI72}. The operator $d\tilde{\pi}(X)$ in the above statements is an extension of the operator $d\tilde{\pi}(X)$ on the space of smooth vectors. This allows us to use the same notation for those two objects. We extend this representation to $\frak{g}_{{\mathbb{C}}}$ by complex linearity, $d\pi(X+iY)=d\pi (X)+i\,d\pi(Y)$, $X,Y\in\frak{g}$. Let $U(\frak{g})$ denote the universal enveloping algebra of $\frak{g}_{{\mathbb{C}}}$. Then $d\pi$ extends to a representation on $U(\frak{g})$, again denoted by $d\pi$. The space $\mathbf{H}^{\infty}(\pi)$ is a topological vector space in a natural way, cf.\ \cite{WaI72}. Furthermore $\mathbf{H}^{\infty}(\pi)$ is invariant under $G$ and $U(\frak{g})$. As $\pi(g)\pi(\exp(tX))\mathbf{v}=\pi(\exp (t\operatorname{Ad}(g)X))\pi(g)\mathbf{v}$, we get \[ \pi(g)\,d\pi(X)\mathbf{v}=d\pi(\operatorname{Ad}(g)X)\pi(g)\mathbf{v}, \qquad \mathbf{v}\in\mathbf{H}^{\infty}(\pi) \] for all $g\in G$ and all $X\in\frak{g}$. Define $Z^{\ast}=-\sigma(Z)$, $Z\in\frak{g}_{{\mathbb{C}}}$, where $\sigma$ is the conjugation $X+iY\mapsto X-iY$, $X,Y\in\frak{g}$. Then a simple calculation shows that for the densely defined operator $\pi(Z)$, $Z\in\frak{g}_{{\mathbb{C}}}$, we have $\pi(Z)^{\ast}=\pi(Z^{\ast})$ on $\mathbf{H}^{\infty}(\pi)$. When (R1--2) and (PR1--3) hold, and $Y\in C$, we showed in Lemma \ref{HolomorphicRPlus} that the operator $\tilde{\pi}\left( Y\right) $ is selfadjoint in $\mathbf{K}=\left( \mathbf{K}_{0}/\mathbf{N}\right) \sptilde$ with spectrum in $\left[ 0,\infty\right) $. Once $\tilde{\pi}$ is identified as a unitary representation of $G^{c}$, then $\tilde{\pi}\left( iY\right) $ is automatically a selfadjoint operator in $\mathbf{K}$ by \cite{NeSt59}, but semiboundedness of the corresponding spectrum of $\tilde{\pi}\left( Y\right) $ only holds for $Y\in C$. Yet if $\tilde{\pi}$ is obtained, as in Lemma \ref{HolomorphicRPlus}, from a unitary representation $\pi$ of $G$ acting on $\mathbf{H}$, then the spectrum of $\pi\left( Y\right) $ is contained in the purely imaginary axis $i\mathbb{R}$, and yet $\tilde{\pi}\left( Y\right) $ has spectrum in $\left[ 0,\infty\right) \subset\mathbb{R}$. The explanation is that the Hilbert spaces $\mathbf{H}$ and $\mathbf{K}$ are different for the two representations $\pi$ and $\tilde{\pi}$. \section{\label{S-Hrep}Holomorphic Representations} \setcounter{equation}{0} The unitary representations that show up in the duality are direct integrals of highest weight $G^{c}$-modules. Those representations can also be viewed as \textit{holomorphic representations} of a semigroup related to an extension of the $H$-invariant cone $C$ that we started with. We will therefore give a short overview over this theory, while referring to the forthcoming monograph \cite{Ne99} for more details. The theory of highest weight modules and holomorphic representations will always be related to the name of Harish-Chandra because of his fundamental work on bounded symmetric domains and the holomorphic discrete series, \cite{HCIV,HCV,HCVI}. Later Gelfand and Gindikin in \cite{GG} proposed a new approach for studying the Plancherel formula for semisimple Lie group $G$. Their idea was to consider functions in $\mathbf{L}^{2}(G)$ as the sum of boundary values of holomorphic functions defined on domains in $G_{{\mathbb{C}}}$. The first deep results in this direction are due to Ol'shanskii \cite{Ol82} and Stanton \cite{Stanton}, who realized the holomorphic discrete series of the group $G$ in a Hardy space of a local tube domain containing $G$ in the boundary. The generalization to noncompactly causal symmetric spaces was carried out in \cite{HO95,HOO91,'OO88a,'OO88b}. This program was carried out for solvable groups in \cite{HO} and for general groups in \cite{BK95,Nee94}. Let $G_{{\mathbb{C}}}$ be a complex Lie group with Lie algebra $\frak{g}% _{{\mathbb{C}}}$ and let $\frak{g}$ be a real form of $\frak{g}_{{\mathbb{C}}% }$. We assume for simplicity that $G$, the analytic subgroup of $G_{{\mathbb{C}}}$ with Lie algebra $\frak{g}$, is closed in $G_{{\mathbb{C}}% }$. Let $C$ be a regular $G$-invariant cone in $\frak{g}$ such that the set $S(C)=G\exp iC$ is a closed semigroup in $G_{{\mathbb{C}}}$. Moreover, we assume that the map \[ G\times C\ni(a,X)\longmapsto a\exp iX\in S(C) \] is a homeomorphism and even a diffeomorphism when restricted to $G\times C^{o}$. Finally, we assume that there exists a real automorphism $\sigma$ of $G_{{\mathbb{C}}}$ whose differential is the complex conjugation of $\frak{g}_{{\mathbb{C}}}$ with respect to $\frak{g}$, that is $\sigma (X+iY)=X-iY$ for $X,Y\in\frak{g}$. We notice that in this case $G_{{\mathbb{C}% }}/G$ is a symmetric space and that the corresponding subspace $\frak{q}$ is just $i\frak{g}$. Those hypotheses are also satisfied for Hermitian Lie groups and also for some solvable Lie groups; cf.\ \cite{HO}. Define% \begin{equation} W(\pi):=\{X\in\frak{g}\mid\forall u\in\mathbf{H}^{\infty}(\pi)\,:\,\ip {i\,d\pi(X)u}{u}\leq0\}\,. \label{Cpi}% \end{equation} Thus $W(\pi)$ is the set of elements of $\frak{g}$ for which $d\pi(iX)$ is \textit{negative}. The elements of $W(\pi)$ are called \textit{negative elements} for the representation $\pi$. \begin{lemma} \label{LemS-HrepNew.1}$W(\pi)$ is a closed $G$-invariant convex cone in $\frak{g}$. \end{lemma} \begin{definition} \label{DefS-HrepMar.1}Let $W$ be a $G$-invariant cone in $\frak{g}$. We denote the set of all unitary representations $\pi$ of $G$ with $W\subset W(\pi)$ by $\mathcal{A}(W)$. A unitary representation $\pi$ is called \textit{$W$% }\emph{-admissible} if $\pi\in\mathcal{A}(W)$. \end{definition} The representations in $\mathcal{A}(W)$ will be studied in detail in Section \ref{S-hwm} below. We show in Theorem \ref{AChw} that a $\rho\in \mathcal{A}(W)$ which is irreducible is in fact a \emph{highest weight representation}, and the corresponding $K^{c}$-weights are determined. The representations are then identified as discrete summands in $\mathbf{L}% ^{2}\left( G^{c}\right) $. Let $S$ be a semigroup with unit and let $\sharp\colon S\rightarrow S$ be a bijective involutive antihomomorphism, that is% \[ (ab)^{\sharp}=b^{\sharp}a^{\sharp}\qquad\text{and}\qquad\quad a^{\sharp\sharp }=a \] We call $\sharp$ an \textit{involution} on the semigroup $S$, and we call the pair $(S,\sharp)$ a \textit{semigroup with involution} or an \textit{involutive semigroup}. For us the important examples are the semigroups of the form $S(C)=H\exp C$ with $\gamma^{\sharp}=\tau(\gamma^{-1}% )$. Another class of examples consists of the \textit{contractive semigroups} on a Hilbert space $\mathbf{H}$. Let $S(\mathbf{H})=\{T\in B(\mathbf{H}% )\mid\Vert T\Vert\leq1\}$. Denote by $T^{\ast}$ the adjoint of $T$ with respect to the inner product on $\mathbf{H}$. Then $(S,\ast)$ is a semigroup with involution. \begin{definition} Let $(S,\sharp)$ be a topological semigroup with involution: then a semigroup homomorphism $\rho\colon S\rightarrow S(\mathbf{H})$ is called a contractive representation of $(S,\sharp)$ if $\rho(g^{\sharp})=\rho(g)^{\ast}$ and $\rho$ is continuous with respect to the weak operator topology of $S(\mathbf{H})$. A contractive representation is called irreducible\label{irr} if there is no closed nontrivial subspace of $\mathbf{H}$ invariant under $\rho(S)$. \end{definition} \begin{definition} \label{DefS-HrepNew.3}Let $\rho$ be a contractive representation of the semigroup $S(W)=G\exp iW\subset G_{{\mathbb{C}}}$. Then $\rho$ is holomorphic if the function $\rho\colon S(C)^{o}\rightarrow B(\mathbf{V})$ is holomorphic. \end{definition} The following lemma shows that, if a unitary representation of the group $G$ extends to a holomorphic representation of $S (C)$, then this extension is unique. \begin{lemma} \label{unique} If $f\colon S(W)\rightarrow S(\mathbf{H})$ is continuous and $f|_{S(W)^{o}}$ is holomorphic such that $f|_{G}=0$, then $f=0$. \end{lemma} To construct a holomorphic extension $\rho$ of a representation $\pi$ we have to assume that $\pi\in\mathcal{A}(W)$. Then for any $X\in W$, the operator $i\,d\pi(X)$ generates a self adjoint contraction semigroup which we denote by% \[ T_{X}(t)=e^{ti\,d\pi(X)}\,. \] For $s=g\exp iX\in S(C)$ we define% \begin{equation} \rho(s):=\rho(g)T_{X}(1) \label{extension}% \end{equation} \begin{theorem} \label{Th:extension} $\rho$ is a contractive and holomorphic representation of the semigroup $S(W)$. In particular, every representation $\pi\in \mathcal{A}(W)$ extends uniquely to a holomorphic representation of $S(W)$ which is uniquely determined by $\pi$. \end{theorem} We will usually denote the holomorphic extension of the representation $\pi$ by the same letter. For the converse of Theorem \ref{Th:extension}, we remark the following simple fact: Let $(S,\sharp)$ be a semigroup with involution and let $\rho$ be a contractive representation of $S$. Let \[ G(S):=\{s\in S\mid s^{\sharp}s=ss^{\sharp}=1\} \] Then $G(S)$ is a closed subgroup of $S$ and $\pi:=\rho|_{G(S)}$ is a unitary representation of $G(S)$. Obviously, \[ G\subset G(S(W))\,. \] Thus every holomorphic representation of $S(W)$ defines a unique unitary representation of $G$ by restriction. \begin{theorem} \label{ThmS-HrepNew.6}Let $\rho$ be a holomorphic representation of $S(W)$. Then $\rho|_{G}\in\mathcal{A}(W)$ and the $\rho$ agrees with the extension of $\rho|_{G}$ to $S(W)$. \end{theorem} Two representations $\rho$ and $\pi$ of the semigroup $S(W)$ are said to (be unitarily) equivalent if there exists a unitary isomorphism $U\colon \mathbf{H}(\rho)\rightarrow\mathbf{H}(\pi)$ such that \[ U\rho(s)=\pi(s)U\quad\forall s\in S(W) \] In particular, two contractive representations $\rho$ and $\pi$ of $S(W)$ are equivalent if and only if $\rho|_{G}$ and $\pi|_{G}$ are unitarily equivalent. We call a holomorphic contractive representation $\rho$ of $S(W)$ $W$-\textit{admissible} if $\rho|_{G}\in\mathcal{A}(W)$ and write $\rho \in\mathcal{A}(W)$. We denote by $\widehat{S(W)}$ the set of equivalence classes of irreducible holomorphic representations of $S(W)$. We say that a representation $\rho$ is bounded if $\Vert\rho(s)\Vert\leq 1$ for all $s\in S(W)$. Note that this depends only on the unitary equivalence class of $\rho$. We denote by $\widehat{S(W)}_{b}$ the subset in $\widehat{S(W)}$ of bounded representations. Let $\rho$ and $\pi$ be holomorphic representations of $S(W)$. Define a representation of $S(W)$ in $\mathbf{H}(\rho)\hat{\otimes}\mathbf{H}(\pi)$ by% \[ \lbrack\rho\otimes\pi](s):=\rho(s)\otimes\pi(s) \] Then $\rho\otimes\pi\in\mathcal{A}(W)$. We denote the representation $s\mapsto\operatorname*{id}$ by $\iota$. \begin{theorem} [Neeb, Ol'shanskii \cite{HO95,Nee94,Ne99}]\label{DirInt} Let $\rho$ be a holomorphic representation of $S(W)_{b}$. Then there exists a Borel measure $\mu$ on $\widehat{S(W)}$ supported on $\widehat{S(W)}_{b}$ and a direct integral of representations \[ \left( \int_{\widehat{S(W)}_{b}}^{\oplus}\rho_{\omega}\,d\mu(\omega ),\int_{\widehat{S(W)}_{b}}^{\oplus}\mathbf{H}(\omega)\,d\mu (\omega)\right) \] such that: \begin{enumerate} \item \label{DirInt(1)}The representation $\rho$ is equivalent to $\int_{\widehat{S(W)}_{b}}^{\oplus}\rho_{\omega}\,d\mu(\omega)$ . \item \label{DirInt(2)}There exists a subset $N\subset\widehat{S(W)}_{b}$ such that $\mu(N)=0$ and if $\omega\in\widehat{S(W)}_{b}\setminus N$, then $\rho_{\omega}$ is equivalent to $(\pi_{\omega}\otimes\iota,\mathbf{H}% (\omega)\hat{\otimes}\mathbf{L}(\omega))$, where $\pi_{\omega}\in\omega$ and $\mathbf{L}(\omega)$ is a Hilbert space. \item \label{DirInt(3)}If $\omega\in\widehat{S(W)}_{b}$ then set $n(\omega):=\dim\mathbf{L}(\omega)$. Then $n$ is a $\mu$-measurable function from $\widehat{S(W)}_{b}$ to the extended positive axis $[0,\infty]$ which is called the multiplicity function. \end{enumerate} \end{theorem} \begin{proof} See \cite{Ne99}, Theorem XI.6.13. \end{proof} \section{\label{LM}The L\"uscher-Mack Theorem} \setcounter{equation}{0} We use reference \cite{HiNe93} for the L\"uscher-Mack Theorem, but \cite{FOS83}, \cite{GoJo83}, \cite{Jor86}, \cite{Jor87}, \cite{JoMo84}, \cite{KlLa83}, \cite{LM75}, and \cite{Sch86} should also be mentioned in this connection. We have two ways of making the connection between the unitary representations of $G$ and those of $G^{c}$: one is based on the L\"uscher-Mack principle, and the other on the notion of local representations from Jorgensen's papers \cite{Jor86} and \cite{Jor87}. Let $\pi$, $C$, $\mathbf{H}(\pi)$, $J$ and $\mathbf{K}_{0}$ be as before. We have proved that the operators \[ \{\pi(h\exp(Y))\mid h\in H,Y\in C\} \] pass to the space $\mathbf{K}=\left( \mathbf{K}_{0}/\mathbf{N}\right) \sptilde$ such that $\tilde{\pi}(h)$ is unitary on $\mathbf{K}$, and $\tilde{\pi}(\exp Y)$ is contractive and selfadjoint on $\mathbf{K}$. As a result we arrive at selfadjoint operators $d\tilde{\pi}(Y)$ with spectrum in $(-\infty,0]$ such that for $Y\in C$, $\tilde{\pi}(\exp Y)=e^{d\tilde{\pi}% (Y)}$ on $\mathbf{K}$. As a consequence of that we notice that \[ t\longmapsto e^{t\,d\tilde{\pi}(Y)}% \] extends to a continuous map on $\{z\in{{\mathbb{C}}}\mid\operatorname{Re}% (z)\geq0\}$ holomorphic on the open right half plane $\{z\in{{\mathbb{C}}}% \mid\operatorname{Re}(z)>0\}$. Furthermore, \[ e^{(z+w)\,d\tilde{\pi}(Y)}=e^{z\,d\tilde{\pi}(Y)}e^{w\,d\tilde{\pi}(Y)}\,. \] As $\mathbf{K}$ is a unitary $H$-module we know that the $H$-analytic vectors $\mathbf{K}^{\omega}(H)$ are dense in $\mathbf{K}$. Thus $\mathbf{K}% _{oo}:=S(C^{o})\mathbf{K}^{\omega}(H)$ is dense in $\mathbf{K}$. We notice that for $u\in\mathbf{K}_{oo}$ and $X\in C^{o}$ the function $t\mapsto \tilde{\pi}(\exp tX)u$ extends to a holomorphic function on an open neighborhood of the right half-plane. This and the Campbell-Hausdorff formula are among the main tools used in proving the following Theorem of L\"uscher and Mack \cite{LM75}. We refer to \cite[p. 292]{HiNe93} for the proof. Our present use of Lie theory, cones, and semigroups will follow standard conventions (see, e.g., \cite{FHO93,Hel62,JL94,WaI72,Yos91}): the exponential mapping from the Lie algebra $\frak{g}$ to $G$ is denoted $\exp$, the adjoint representation of $\frak{g}$, $\operatorname{ad}$, and that of $G$ is denoted $\operatorname{Ad}$. \begin{theorem} [L\"uscher-Mack \cite{LM75}]\label{LuscherMack} Let $\rho$ be a strongly continuous contractive representation of $S(C)$ on the Hilbert space $\mathbf{H}$ such that $\rho(s)^{\ast}=\rho(\tau(s)^{-1})$. Let $G^{c}$ be the connected, simply connected Lie group with Lie algebra $\frak{g}^{c}% =\frak{h}\oplus i\frak{q}$. Then there exists a continuous unitary representation $\rho^{c}\colon G^{c}\rightarrow\mathrm{U}(\mathbf{H})$, extending $\rho$, such that for the differentiated representations $d\rho$ and $d\rho^{c}$ we have: \begin{enumerate} \item [\hss\llap{\rm1)}]\label{LuscherMack(1)}$d\rho^{c}(X)=d\rho (X)\,\quad\forall X\in\frak{h}$. \item[\hss\llap{\rm2)}] \label{LuscherMack(2)}$d\rho^{c}(iY)=i\,d\rho (Y)\,\quad\forall Y\in C$. \end{enumerate} \end{theorem} We apply this to our situation to get the following theorem: \begin{theorem} \label{PiCIrreducible}Assume that $(\pi,C,\mathbf{H},J)$ satisfies \textup{(PR1)--(PR3).} Then the following hold: \begin{enumerate} \item [\hss\llap{\rm1)}]\label{PiCIrreducible(1)}$S(C)$ acts via $s\mapsto\tilde{\pi}(s)$ by contractions on $\mathbf{K}$. \item[\hss\llap{\rm2)}] \label{PiCIrreducible(2)}Let $G^{c}$ be the simply connected Lie group with Lie algebra $\frak{g}^{c}$. Then there exists a unitary representation $\tilde{\pi}^{c}$ of $G^{c}$ such that $d\tilde{\pi }^{c}(X)=d\tilde{\pi}(X)$ for $X\in\frak{h}$ and $i\,d\tilde{\pi}% ^{c}(Y)=d\tilde{\pi}(iY)$ for $Y\in C$. \item[\hss\llap{\rm3)}] \label{PiCIrreducible(3)}The representation $\tilde{\pi}^{c}$ is irreducible if and only if $\tilde{\pi}$ is irreducible. \end{enumerate} \end{theorem} \begin{proof} (1) and (2) follow by the L\"uscher-Mack theorem and Proposition \ref{Contractive}, as the resulting representation of $S$ is obviously continuous. (3) Let $\mathbf{L}$ be a $G^{c}$-invariant subspace in $\mathbf{K}$. Then $\mathbf{L}$ is $\tilde{\pi}(H)$ invariant. Let $Y\in C^{o}$, $u\in \mathbf{L}^{\omega}$ and $v\in\mathbf{L}^{\perp}$. Define $f\colon \{z\in{{\mathbb{C}}}\mid\operatorname{Re}(z)\geq0\}\rightarrow{{\mathbb{C}}}$ by \[ f(z):=% %TCIMACRO{\TeXButton{ip}{\ip{v}{e^{zd\tilde{\pi}(Y)}u}}}% %BeginExpansion \ip{v}{e^{zd\tilde{\pi}(Y)}u}% %EndExpansion _{J}\,. \] Then $f$ is holomorphic in $\{z\in{{\mathbb{C}}}\mid\operatorname{Re}(z)>0\}$, and $f(it)=0$ for every (real) $t$. Thus $f$ is identically zero. In particular $f(t)=0$ for every $t>0$. Thus \[ 0=% %TCIMACRO{\TeXButton{ip}{\ip{v}{e^{td\tilde{\pi}(Y)}u}}}% %BeginExpansion \ip{v}{e^{td\tilde{\pi}(Y)}u}% %EndExpansion _{J}=% %TCIMACRO{\TeXButton{ip}{\ip{v}{\tilde{\pi}(\exp tY)u}}}% %BeginExpansion \ip{v}{\tilde{\pi}(\exp tY)u}% %EndExpansion _{J}\,. \] As $S^{o}=H\exp C^{o}$ it follows that $\tilde{\pi}(S^{o})(\mathbf{L}^{\omega })\subset(\mathbf{L}^{\perp})^{\perp}=\mathbf{L}$. By continuity we get $\tilde{\pi}(S)\mathbf{L}\subset\mathbf{L}$. Thus $\mathbf{K}$ is reducible as an $S$-module. The other direction follows in exactly the same way. \end{proof} We notice now that $-iC\subset W(\tilde{\pi}^{c})$. Thus $W(\tilde{\pi}^{c})$ is non-trivial and contain the $-\tau$-stable and $G$-invariant cone generated by $-iC$, i.e. $-i\cdot\overline{\operatorname{conv}\{\operatorname{Ad}% (G)C\}}\subset W(\tilde{\pi}^{c})$. But in general $W(\tilde{\pi}^{c})$ is neither generating nor pointed. It even does not have to be $-\tau$-invariant. In fact, the Lie algebra of the $(ax+b)$-group, and the Heisenberg group, do not have \emph{any} pointed, generating, invariant cones. \begin{lemma} \label{KernelPiC}$W(\tilde{\pi}^{c})\cap-W(\tilde{\pi}^{c})=\ker(\tilde{\pi }^{c})$. \end{lemma} \begin{proof} This is obvious from the spectral theorem. \end{proof} \begin{lemma} \label{IdealGC}$\frak{g}_{1}^{c}:=W(\tilde{\pi}^{c})-W(\tilde{\pi}^{c})$ is an ideal in $\frak{g}^{c}$. Furthermore, $[\frak{q},\frak{q}]\oplus i\frak{q}\subset\frak{g}_{1}^{c}$. \end{lemma} \begin{proof} Let $X\in\frak{g}^{c}$. Then, as $W(\tilde{\pi}^{c})$ is invariant by construction, we conclude that \[ e^{t\operatorname{ad}(X)}\left( W(\tilde{\pi}^{c})-W(\tilde{\pi}^{c})\right) \subset W(\tilde{\pi}^{c})-W(\tilde{\pi}^{c}),\,t\in\mathbb{R}\,. \] By differentiation at $t=0$, it follows that $[X,\frak{g}_{1}^{c}% ]\subset\frak{g}_{1}^{c}$. This shows that $\frak{g}_{1}^{c}$ is an ideal in $\frak{g}^{c}$. The last part follows as $C$ is generating (in $\frak{q}$). \end{proof} It is not clear if $\frak{g}_{1}^{c}$ is $\tau$-stable. To get a $\tau$-stable subalgebra one can replace $W(\tilde{\pi}^{c})$ by the cone generated by $-\operatorname{Ad}(G)C\subset W(\tilde{\pi}^{c})$ or by the maximal $G$- and $-\tau$-stable cone $W(\tilde{\pi}^{c})\cap(-\tau(W(\tilde{\pi}^{c})))$ in $W(\pi^{c})$. We have now the following important consequence of the Neeb-Ol'shanskii theorem: \begin{theorem} \label{HighestWeight}Let the analytic subgroup $G_{1}^{c}$ of $G^{c}$ corresponding to $\frak{g}_{1}^{c}$ be as described, and let $W(\tilde{\pi }^{c})$ be the corresponding module. Then $\tilde{\pi}^{c}|_{G_{1}^{c}}$ is a direct integral of irreducible representations in $\mathcal{A}(W)$. \end{theorem} \section{\label{SSS}Bounded Symmetric Domains} \setcounter{equation}{0} We have seen that the representations $\pi^{c}$ that we can produce using the duality are direct integrals of holomorphic representations of suitable subsemigroups of $G_{\mathbb{C}}^{c}$ (or a subgroup). Those on the other hand only exist if there is a $G^{c}$ invariant cone in $\frak{g}$. We will discuss the case of simple Lie group $G^{c}$ in some detail here. We refer to \cite{JO97}, chapter 2, and the references therein for proofs. In the duality $G\leftrightarrow G^{c}$ it will be the group $G^{c}$ that has holomorphic representations. Therefore we will start using the notation $G^{c}$ for Hermitian groups. \begin{theorem} [Kostant]\label{ThmKostant}Suppose that $\mathbf{V}$ is a finite-dimensional real vector space. Let $L$ be a connected reductive subgroup of $GL(\mathbf{V}% )$ acting irreducibly on $\mathbf{V}$. Let $G^{c}=L^{\prime}$ be the commutator subgroup of $L$. Further let $K^{c}$ be a maximal compact subgroup of $G^{c}$. Then the following properties are equivalent: \begin{enumerate} \item \label{ThmKostant(1)}There exists a regular $L$-invariant closed cone in $\mathbf{V}$. \item \label{ThmKostant(2)}The $G^{c}$-module $\mathbf{V}$ is spherical. \end{enumerate} \end{theorem} Let $C\subset\mathbf{V}$ be a regular $L$-invariant cone. Then $\mathbf{V}$ is spherical as a $G^{c}$-module. Let $K^{c}$ be a maximal compact subgroup of $G^{c}$. A $K^{c}$-invariant vector $u_{K^{c}}$ can be constructed in the following way: Let $u\in C^{o}$, the interior of $C$, be arbitrary. Define \[ u_{K^{c}}=\int_{K^{c}}k\cdot u\,dk\,. \] Then $u_{K^{c}}\in C^{o}$ is $K^{c}$-invariant. Suppose that the group $L$ acts on $\mathbf{V}$. Let $\operatorname{Cone}_{L}(\mathbf{V})$ denote the set of regular $L$-invariant cones in $\mathbf{V}$. \begin{theorem} [Vinberg]\label{ThmVinberg}Let $L$, $G^{c}$, and $\mathbf{V}$ be as in the theorem of Kostant. Then the following properties are equivalent: \begin{enumerate} \item \label{ThmVinberg(1)}$\operatorname{Cone}_{L}(\mathbf{V})\not =% \emptyset$; \item \label{ThmVinberg(2)}The $G^{c}$-module $\mathbf{V}$ is spherical; \item \label{ThmVinberg(3)}There exists a ray in $\mathbf{V}$ through $0$ which is invariant with respect to some minimal parabolic subgroup $P$ of $G^{c}$. \end{enumerate} If those conditions hold, every invariant pointed cone in $\mathbf{V}$ is regular. \end{theorem} For the next theorem, see \cite{Paneitz81,Paneitz84}. \begin{theorem} [Paneitz, Vinberg]\label{ThmPV} Let $G^{c}$ be a connected semisimple Lie group and $(\mathbf{V,\pi)}$ a real finite-dimensional irreducible $G^{c}$-module such that $\operatorname{Cone}% _{G^{c}}(\mathbf{V})\not =\emptyset$. Let $\theta$ be a Cartan involution on $G^{c}$. Choose an inner product on $\mathbf{V}$ such that $\pi(x)^{\ast}% =\pi(\theta(x)^{-1})$ for all $x\in G^{c}$. Then there exists a unique up to multiplication by $(-1)$ invariant cone $C_{\min}\in\operatorname{Cone}% _{G^{c}}(\mathbf{V})$ given by \[ C_{\min}=\operatorname{conv}(\pi(G^{c})u)\cup\left\{ 0\right\} =\overline{\operatorname{conv}\{\pi(G^{c})(\mathbb{R}^{+}v_{K^{c}})\}}\,, \] where $u$ is a highest weight vector, $v_{K^{c}}$ is a nonzero $K^{c}$-fixed vector unique up to scalar multiple, and $(u,v_{K^{c}})>0$. The unique \textup{(}up to multiplication by $(-1)$\textup{)} maximal cone is given by \[ C_{\max}=C_{\min}^{\ast}:=\left\{ w\in\mathbf{V}\mid\forall v\in C_{\min }\,:\,(w,v)\geq0\right\} \,. \] \end{theorem} Assume now that $G^{c}$ is a connected simple Lie group. Then $G^{c}$ acts on $\frak{g}^{c}$ by the adjoint action. Let $K^{c}\subset G^{c}$ be a maximal almost compact subgroup. Then by Kostant's Theorem we have $\operatorname{Cone}_{G^{c}}(\frak{g}^{c})\not =\emptyset$ if and only if there exists a $Z^{0}\in\frak{g}^{c}$ which is invariant under $\operatorname{Ad}(K^{c})$. Let $\frak{k}^{c}$ be the Lie algebra of $K^{c}$. Then $[\frak{k}^{c},Z^{0}]=0$. Hence $\mathbb{R}Z^{0}+\frak{k}^{c}% \not =\frak{g}^{c}$ is a Lie algebra containing $\frak{k}^{c}$. But $\frak{k}^{c}$ is maximal in $\frak{g}^{c}$. Hence $Z^{0}\in\frak{k}^{c}$. Similarly it follows that $\frak{z}_{\frak{g}^{c}}(Z^{0})=\frak{k}^{c}$. Finally the Theorem of Paneitz and Vinberg implies that $\operatorname{Cone}% _{G^{c}}(\frak{g}^{c})\not =\emptyset$ if and only if the center of $\frak{k}^{c}$ is one dimensional. In that case we can normalize the element $Z^{0}$ such that $\operatorname{ad}(Z^{0})$ has eigenvalues $0,i,-i$. Let $\frak{t}$ be a Cartan subalgebra of $\frak{g}^{c}$ containing $Z^{0}$. Then $\frak{k}^{c}\frak{\subset z}_{\frak{g}^{c}}(Z^{0})\subset\frak{k}^{c}$. Hence $\frak{t}$ is contained in $\frak{k}^{c}$. For $\alpha\in\frak{t}_{\mathbb{C}% }^{\ast}$ let \[ \frak{g}_{{\mathbb{C}\alpha}}^{c}:=\left\{ X\in\frak{g}_{\mathbb{C}}^{c}% \mid\forall Z\in\frak{t}_{\mathbb{C}}\,:\,[Z,X]=\alpha(Z)X\right\} \,. \] It is well known that $\dim\frak{g}_{{\mathbb{C}\alpha}}^{c}\leq1$ for all $\alpha\not =0$ and \[ \frak{g}_{\mathbb{C}}^{c}=\frak{t}_{\mathbb{C}}\oplus\bigoplus_{\alpha \in\Delta}\frak{g}_{{\mathbb{C}\alpha}}^{c}% \] where $\Delta=\left\{ \alpha\in\frak{t}_{\mathbb{C}}^{\ast}\setminus\left\{ 0\right\} \mid\frak{g}_{{\mathbb{C}\alpha}}^{c}\not =\left\{ 0\right\} \right\} $. We notice that $\alpha(\frak{t})\subset i\mathbb{R}$ for all $\alpha\in\Delta$ as $\frak{t}\subset\frak{k}$ and $a\,d(X)$ is skew-symmetric for all $X\in\frak{k}$. Let $\theta\colon\frak{g}^{c}\frak{\rightarrow }\frak{g}^{c}$ be the Cartan involution corresponding to $\frak{k}^{c}$. We do denote the corresponding involution $\theta\otimes1$ on $\frak{g}_{\mathbb{C}% }^{c}$ and the integrated involution on $G^{c}$ by the same letter. Then $\frak{g}^{c}=\frak{k}^{c}\oplus\frak{p}^{c}$ where $\frak{p}^{c}% \frak{=}\left\{ X\in\frak{g}^{c}\mid\theta(X)=-X\right\} $. Thus $\frak{g}_{\mathbb{C}}^{c}=\frak{k}_{\mathbb{C}}^{c}\oplus\frak{p}% _{\mathbb{C}}^{c}$. Let $\Delta_{c}=\left\{ \alpha\in\Delta\mid\alpha (Z^{0})=0\right\} $ and $\Delta_{p}=\left\{ \alpha\in\Delta\mid\alpha (Z^{0})=\pm i\right\} $. As $\frak{z}_{\frak{g}_{\mathbb{C}}}(Z^{0}% )=\frak{k}_{\mathbb{C}}$ we get% \[ \Delta_{c}=\left\{ \alpha\mid\frak{g}_{{\mathbb{C}\alpha}}^{c}\subset \frak{k}_{\mathbb{C}}^{c}\right\} \qquad\text{and}\qquad\Delta_{p}=\left\{ \alpha\mid\frak{g}_{{\mathbb{C}\alpha}}^{c}\subset\frak{p}_{\mathbb{C}}% ^{c}\right\} \,. \] Choose a positive system $\Delta^{+}$ in $\Delta$ such that $\Delta_{p}% ^{+}:=\left\{ \alpha\in\Delta_{p}\mid\alpha(Z^{0})=i\right\} \subset \Delta^{+}$. Then $\Delta^{+}=\Delta_{c}^{+}\cup\Delta_{p}^{+}$ and $\Delta_{c}^{+}$ is a positive system in $\Delta_{c}$. For $\Gamma \subset\Delta$ let $\frak{g}_{\mathbb{C}}^{c}(\Gamma):=\bigoplus_{\alpha \in\Gamma}\frak{g}_{\mathbb{C\alpha}}^{c}$. Then% \begin{align*} \frak{p}^{+}:= & \left\{ X\in\frak{g}_{\mathbb{C}}^{c}\mid\lbrack Z^{0},X]=iX\right\} =\frak{g}_{\mathbb{C}}^{c}(\Delta_{p}^{+})\,;\\ \frak{p}^{-}:= & \left\{ X\in\frak{g}_{\mathbb{C}}^{c}\mid\lbrack Z^{0},X]=-iX\right\} =\frak{g}_{\mathbb{C}}^{c}(-\Delta_{p}^{+})\,. \end{align*} Furthermore $\frak{p}^{+}$ and $\frak{p}^{-}$ are abelian subalgebras with $\frak{p}_{\mathbb{C}}=\frak{p}^{+}\oplus\frak{p}^{-}$. Let $P^{\pm}% :=\exp(\frak{p}^{\pm})$ and $K_{\mathbb{C}}^{c}=\exp(\frak{k}_{\mathbb{C}}% ^{c})$. Both $P^{+}$ and $P^{-}$ are simply connected closed abelian subgroups of $G_{\mathbb{C}}^{c}$. Hence $\exp\colon\frak{p}^{\pm}\rightarrow P^{\pm}$ is a diffeomorphism. Let% \begin{equation} \zeta=(\exp|_{\frak{p}^{+}})^{-1}\colon P^{+}\longrightarrow\frak{p}% ^{+}\,\text{.} \label{E:zeta}% \end{equation} The set $P^{+}K_{\mathbb{C}}^{c}P^{-}$ is open and dense in $G_{\mathbb{C}% }^{c}$, $G^{c}\subset P^{+}K_{\mathbb{C}}^{c}P^{-}$, $G^{c}K_{\mathbb{C}}% ^{c}P^{-}$ is open in $G_{\mathbb{C}}^{c}$, and $G^{c}\cap K_{\mathbb{C}}% ^{c}P^{-}=K^{c}$. Thus $G^{c}/K^{c}$ is holomorphically equivalent to an open submanifold $D$ of the complex flag manifold $X_{\mathbb{C}% }=G_{\mathbb{C}}^{c}/K_{\mathbb{C}}^{c}P^{-}$. Furthermore the map $pK_{\mathbb{C}}^{c}P^{-}\mapsto\zeta(p)$ induces a biholomorphic map---also denoted by $\zeta$---of $G^{c}/K^{c}$ onto a bounded symmetric domain $\Omega_{\mathbb{C}}\subset\frak{p}^{-}\simeq\mathbb{C}^{\dim(G^{c}/K^{c})}$. For $x\in P^{+}K_{\mathbb{C}}^{c}P^{-}$ we can write in a unique way% \begin{equation} x=p(x)k_{\mathbb{C}}(x)q(x) \label{E:P+KP-projections}% \end{equation} with $p(x)\in P^{+}$, $k_{\mathbb{C}}(x)\in K_{\mathbb{C}}^{c}$ and $q(x)\in P^{-}$. For $g\in G_{\mathbb{C}}^{c}$ and $Z\in\frak{p}^{+}$ we introduce the following notations when ever they make sense:% \begin{align*} g\cdot Z & =\zeta(p(g\exp(Z))\in\frak{p}^{+}\\ j(g,Z) & =k_{\mathbb{C}}(g\exp Z)\in K_{\mathbb{C}}^{c}\,. \end{align*} If $Z\in\Omega_{\mathbb{C}}$ and $g\in G^{c}$ then $g\cdot Z$ is defined and $g\cdot Z\in\mathbb{\Omega}_{\mathbb{C}}$. Furthermore $(g,Z)\mapsto g\cdot Z$ defines an action of $G^{c}$ on $\mathbb{\Omega}_{\mathbb{C}}$ such that $\zeta\colon G^{c}/K^{c}\rightarrow\Omega_{\mathbb{C}}$ is a $G^{c}$-map. The map $j$ is \textit{the universal automorphic factor} and it satisfies the following:% \begin{align} j(k,Z) & =k,\nonumber\\ j(p,Z) & =1,\nonumber\\ j(ab,Z) & =j(a,b\cdot Z)j(b,Z), \label{E:Multipl}% \end{align} if $k\in K_{\mathbb{C}}^{c}$, $Z\in\frak{p}^{+}$, $p\in P^{+}$, and $a,b\in G_{\mathbb{C}}^{c}$ are such that the expressions above are defined. Define% \[ S(\Omega_{\mathbb{C}}):=\left\{ \gamma\in G_{\mathbb{C}}^{c}\mid\gamma ^{-1}\cdot\Omega_{\mathbb{C}}\subset\Omega_{\mathbb{C}}\right\} \] and% \[ S(\Omega_{\mathbb{C}})^{o}:=\left\{ \gamma\in G_{\mathbb{C}}^{c}\mid \gamma^{-1}\cdot\overline{\Omega_{\mathbb{C}}}\subset\Omega_{\mathbb{C}% }\right\} \] where $\overline{\left\{ \Omega_{\mathbb{C}}\right\} }$ stands for the topological closure of $\Omega_{\mathbb{C}}$ in $\frak{p}^{+}$. Then $S(\Omega_{\mathbb{C}})$ is a closed semigroup in $\frak{g}_{\mathbb{C}}^{c}$ of the form% \[ S(\Omega_{\mathbb{C}})=G^{c}\exp(iC_{\max})\, \] where $C_{\max}$ is the maximal $G^{c}$-invariant cone in $\frak{g}% _{\mathbb{C}}^{c}$ containing $-Z^{0}$. Furthermore $S(\Omega_{\mathbb{C}% })^{o}$ is the topological interior of $S(\Omega_{\mathbb{C}})$ and% \[ S(\Omega_{\mathbb{C}})^{o}=S(C_{\max}^{o})=G^{c}\exp(iC_{\max}^{o})\,. \] We refer to \cite{HO95} or \cite{HiNe93} for all of this. \section{\label{S-hwm}Highest Weight Modules} \setcounter{equation}{0} Our notion of reflection for unitary representations leads to the class of representations in $\mathcal{A}(W)$ of Definition \ref{DefS-HrepMar.1}, and in the present section we analyze these representations more closely. The analysis is based in large part on \cite{HO95}, and involves results of (among others) M.~Davidson and R.~Fabec \cite{MDRF96}, K.-H. Neeb \cite{Nee94,Ne94,Ne99}, Harish-Chandra \cite{HC65,HC70}, Ol'shanskii \cite{Ol82}, R.J. Stanton \cite{Stanton}, Wallach \cite{WaI72}, and H.~Rossi and M.~Vergne \cite{VR76}. We have seen that the interesting representations are those in $\mathcal{A}% (W)$ where $W$ is an invariant cone in $\frak{g}^{c}$. It turns out that the irreducible representations in $\mathcal{A}(W)$ are \textit{highest weight representations. }A $(\frak{g}^{c},K^{c})$-module\label{gK-modul} is a complex vector space $\mathbf{V}$ such that \begin{enumerate} \item [1)]$\mathbf{V}$ is a $\frak{g}^{c}$-module. \item[2)] $\mathbf{V}$ carries a representation of $K^{c}$, and the span of $K^{c}\cdot v$ is finite-dimensional for every $v\in\mathbf{V}$. \item[3)] For $v\in\mathbf{V}$ and $X\in\frak{k}^{c}$ we have \[ X\cdot v=\lim_{t\rightarrow0}\frac{\exp(tX)\cdot v-v}{t}. \] \item[4)] For $Y\in\frak{g}^{c}$ and $k\in K^{c}$ the following holds for every $v\in\mathbf{V}$: \[ k\cdot(X\cdot v)=(\operatorname{Ad}(k)X)\cdot\lbrack k\cdot v]\,. \] \end{enumerate} Note that (3) makes sense, as $K^{c}\cdot v$ is contained in a finite dimensional vector space and this space contains a unique Hausdorff topology as a topological vector space. The $(\frak{g}^{c},K^{c})$-module is called \textit{admissible}\label{st-ad2} if the multiplicity of every irreducible representation of $K^{c}$ in $\mathbf{V}$ is finite. If $(\pi,\mathbf{V})$ is an irreducible unitary representation of $G^{c}$, then the space of $K^{c}% $-finite elements in $\mathbf{V}$, denoted by $\mathbf{V}_{K^{c}}$, is an admissible $(\frak{g}^{c},K^{c})$-module. Let $\frak{t}$ be a Cartan subalgebra of $\frak{k}^{c}$ and $\frak{g}^{c}$ as in the last section. \begin{definition} \label{DefNewS-hwm.1}Let $\mathbf{V}$ be a $(\frak{g}^{c},K^{c})$-module. Then $\mathbf{V}$ is a highest-weight module if there exists a nonzero element $v\in\mathbf{V}$ and a $\lambda\in\frak{t}_{\mathbb{C}}^{\ast}$ such that \begin{enumerate} \item [1)]\label{DefNewS-hwm.1(1)}$X\cdot v=\lambda(X)v$ for all $X\in \frak{t}$. \item[2)] \label{DefNewS-hwm.1(2)}There exists a positive system $\Delta^{+}$ in $\Delta$ such that $\frak{g}_{\mathbb{C}}^{c}(\Delta^{+})\cdot v=0$. \item[3)] \label{DefNewS-hwm.1(3)}$\mathbf{V}=U(\frak{g}^{c})\cdot v$. \end{enumerate} The element $v$ is called a primitive element of weight $\lambda$. \end{definition} Let $W\in\operatorname{Cone}_{G^{c}}(\frak{g}^{c})$ and $\pi\in\mathcal{A}(W)$ irreducible. We assume that $-Z^{o}\in W^{o}$. Then $\mathbf{V}_{K^{c}}$ is an irreducible admissible $(\frak{g}^{c},K^{c})$-module, and \[ \mathbf{V}_{K^{c}}=\bigoplus_{\lambda\in\frak{t}_{\mathbb{C}}^{\ast}% }\mathbf{V}_{K^{c}}(\lambda) \] where $\mathbf{V}_{K^{c}}(\lambda)=\mathbf{V}_{K^{c}}(\lambda,\frak{t}% _{\mathbb{C}})$. Let $v\in\mathbf{V}_{K^{c}}(\lambda)$ be nonzero. Let $\alpha\in\Delta_{p}^{+}$ and let $X\in\frak{p}_{\alpha}^{+}\setminus\{0\}$. Then \[ X^{k}\cdot v\in\mathbf{V}_{K^{c}}(\lambda+k\alpha). \] In particular, \[ -iZ^{0}\cdot(X^{k}\cdot v)=[-i\lambda(Z^{0})+k]v. \] This yields the following lemma. \begin{lemma} \label{LemNewS-hwm.2}Let the notation be as above. Then the following holds: \begin{enumerate} \item \label{LemNewS-hwm.2(1)}$-i\lambda(Z^{0})\leq0$. \item \label{LemNewS-hwm.2(2)}There exists a $\lambda$ such that $\frak{p}% ^{+}\cdot\mathbf{V}_{K^{c}}(\lambda)=\{0\}$. \end{enumerate} \end{lemma} \begin{lemma} \label{LemNewS-hwm.3}Let $\mathbf{W}^{\lambda}$ be the $K^{c}$-module generated by $\mathbf{V}_{K^{c}}(\lambda)$. Then $\mathbf{W}^{\lambda}$ is irreducible, $\mathbf{V}_{K^{c}}=U(\frak{p}^{-})\mathbf{W}^{\lambda}$ and the multiplicity of $\mathbf{W}^{\lambda}$ in $\mathbf{V}_{K^{c}}$ is one. \end{lemma} Let $\alpha\in\Delta_{p}^{+}$ then there exists a unique element $H_{\alpha }\in i\frak{t}\cap\lbrack\frak{g}_{\mathbb{C\alpha}},\frak{g}_{\mathbb{C}% \alpha}]$ such that $\alpha(H_{\alpha})=2$. Let $\mu$ be the highest weight of $\mathbf{W}^{\lambda}$ with respect to $\Delta_{c}^{+}$ and let $v^{\lambda}$ be a nonzero highest weight vector. Then $v^{\lambda}$ is a primitive element with respect to the positive system $\Delta_{c}^{+}\cup\Delta_{p}^{+}$. \begin{theorem} \label{AChw} Let $\rho\in\mathcal{A}(W)$ be irreducible. Then the corresponding $(\frak{g}^{c},K^{c})$-module is a highest-weight module and equals $U(\frak{p}^{-})\mathbf{W}^{\lambda}$. In particular, every weight of $\mathbf{V}_{K^{c}}$ is of the form \[ \nu-\sum_{\alpha\in\Delta(\frak{p}^{+},\frak{t}_{\mathbb{C}})}n_{\alpha}% \alpha\,. \] Furthermore, $\left\langle \nu,H_{\alpha}\right\rangle \leq0$ for all $\alpha\in\Delta_{p}^{+}$. \end{theorem} The $K^{c}$-representation $\pi^{\lambda}$ on $\mathbf{W}^{\lambda}$ is called the \emph{minimal}\textit{ }$K^{c}$-type of $\mathbf{V}$ and $\mathbf{V}% _{K^{c}}$. The multiplicity of $\pi^{\lambda}$ in $\mathbf{V}$ is one. We recall how to realize highest-weight modules in a space of holomorphic functions on $G^{c}/K^{c}$. We follow here the geometric construction by M. Davidson and R. Fabec \cite{MDRF96}. For a more general approach, see \cite{Ne94,Ne99}. To explain the method we start with the example $G^{c}=SU(1,1)=\left\{ \left( \begin{array} [c]{cc}% \alpha & \beta\\ \bar{\beta} & \bar{\alpha}% \end{array} \right) \biggm|\left| \alpha\right| ^{2}-\left| \beta\right| ^{2}=1\right\} $. We set $X=\left( \begin{array} [c]{cc}% 0 & 1\\ 0 & 0 \end{array} \right) $, $Y=\left( \begin{array} [c]{cc}% 0 & 0\\ 1 & 0 \end{array} \right) $ and $H:=H_{1}=\left( \begin{array} [c]{cc}% 1 & 0\\ 0 & -1 \end{array} \right) $. Then $Z^{0}=\frac{i}{2}H$ and $\frak{p}^{+}=\mathbb{C}X$, $\frak{k}_{\mathbb{C}}^{c}=\mathbb{C}H$ and $\frak{p}^{-}={\ \mathbb{C}}Y$. We use this to identify those spaces with $\mathbb{C}$. Let $\gamma=\left( \begin{array} [c]{cc}% a & b\\ c & d \end{array} \right) \in SL(2,\mathbb{C})\,$. Then% \begin{align*} \left( \begin{array} [c]{cc}% a & b\\ c & d \end{array} \right) & =\left( \begin{array} [c]{cc}% 1 & z\\ 0 & 1 \end{array} \right) \left( \begin{array} [c]{cc}% \gamma & 0\\ 0 & \gamma^{-1}% \end{array} \right) \left( \begin{array} [c]{cc}% 1 & 0\\ y & 1 \end{array} \right) \\ & =\left( \begin{array} [c]{cc}% \gamma+\gamma^{-1}z & \gamma^{-1}z\\ \gamma^{-1}y & \gamma^{-1}% \end{array} \right) \,. \end{align*} Hence $P^{+}K_{\mathbb{C}}^{c}P^{-}=\left\{ \left( \begin{array} [c]{cc}% a & b\\ c & d \end{array} \right) \biggm|d\not =0\right\} $ and if $x=\left( \begin{array} [c]{cc}% a & b\\ c & d \end{array} \right) \in P^{+}K_{\mathbb{C}}^{c}P^{-}$, then% \[ p(x)=b/d\,,\quad k_{\mathbb{C}}(x)=d^{-1}\quad\text{and}\quad q(x)=c/d\,. \] Thus \begin{align} \zeta(xK_{\mathbb{C}}^{c}P^{-}) & =b/d\\ x\cdot z & =\frac{az+b}{cz+d}\\ j(x,z) & =(cz+d)^{-1}\,. \label{E:jsu}% \end{align} To identify $\Omega_{\mathbb{C}}$ we notice that on $SU(1,1)$ we have $\zeta(x)=\beta/\bar{\alpha}$. Hence $G^{c}/K^{c}\simeq D=\{z\in\mathbb{C}% \mid|z|<1\}$. The finite-dimensional holomorphic representations of $K_{\mathbb{C}}^{c}$ are the characters \[ \chi_{n}(\exp ziH)=e^{inz}\,. \] In particular, $d\chi_{n}(Z^{0})=in/2$ or \[ -i\,d\chi_{n}(Z^{0})=\frac{n}{2}\,. \] Let $(\pi,\mathbf{V})$ be a unitary highest-weight representation of $SU(1,1)$ and assume that $(\pi,\mathbf{V})\in\mathcal{A}(W)$. Then $n\leq0$ by Lemma \ref{LemNewS-hwm.2} and Theorem \ref{AChw}. Let $\mathbf{V}(n)$ be the one-dimensional space of $\chi_{n}$-isotropic vectors. Then \[ \mathbf{V}_{K^{c}}=\bigoplus_{k\in\,\mathbb{N}}\mathbf{V}(n-2k), \] and the spaces $\mathbf{V}(m)$ and $\mathbf{V}(k)$ are orthogonal if $m\not =k$. Let $\sigma$ be the conjugation of $\frak{s}\frak{l}(2,\mathbb{C})$ with respect to $SU(1,1)$. Then $\sigma$ is given by% \[ \sigma\left( \left( \begin{array} [c]{cc}% a & b\\ c & d \end{array} \right) \right) =\left( \begin{array} [c]{cc}% -\bar{a} & \bar{c}\\ \bar{b} & -\bar{a}% \end{array} \right) \] so that $\sigma(X)=Y$. Since $\pi(T)^{\ast}=-\pi(\sigma(T))$ for all $T\in\frak{s}\frak{l}(2,\mathbb{C})$ we get% \[ \pi(Y)^{\ast}=\pi(-X). \] Finally, it follows from $[Y,X]=-H$ that for $v\in V(n)$: \begin{align*} \Vert\pi(Y)^{k}v\Vert^{2} & =% %TCIMACRO{\TeXButton{ip}{\ip{\pi(Y)^{k}v}{\pi(Y)^{k}v}}}% %BeginExpansion \ip{\pi(Y)^{k}v}{\pi(Y)^{k}v}% %EndExpansion \\ & =% %TCIMACRO{\TeXButton{ip}{\ip{\pi((-X)^{k}Y^{k})v}{v}}}% %BeginExpansion \ip{\pi((-X)^{k}Y^{k})v}{v}% %EndExpansion \end{align*} \begin{lemma} \label{LemNewS-hwm.5}Let the notation be as above. Then \begin{align*} \pi(-X)^{k}\pi(Y)^{k}v & =(-1)^{k}k!\frac{\Gamma(n+1)}{\Gamma(n-k+1)}\,v\\ & =(-n)_{k}v \end{align*} where $(a)_{k}=a(a+1)\cdots(a+k-1)$. \end{lemma} As \[ \sum_{k=0}^{\infty}(-n)_{k}\frac{|z^{2}|^{k}}{k!}=(1-|z|^{2})^{n}, \] (see\ \cite{GR80}) converges if and only if $|z|<1$, it follows that \[ q_{(zX)}v:=\sum_{k=0}^{\infty}\overline{z}^{k}\frac{Y^{k}v}{n!}% \] converges if and only if $zX\in{\Omega}_{\mathbb{C}}$. Let now $G^{c}$ be arbitrary. Let $\sigma\colon\frak{g}_{\mathbb{C}}% ^{c}\rightarrow\frak{g}_{\mathbb{C}}^{c}$ be the conjugation with respect to $\frak{g}^{c}$. We use the notation from earlier in this section. Using the usual $\frak{sl}(2,\mathbb{C})$ reduction, we get the following theorem. \begin{theorem} [Davidson-Fabec]\label{ThDF1} Let $T\in\frak{p}^{+}$. Define $q_{T}% \colon\mathbf{W}^{\lambda}\rightarrow\mathbf{V}$ by the formula \[ q_{T}v:=\sum_{n=0}^{\infty}\frac{\sigma(T)^{n}v}{n!}\,. \] \begin{enumerate} \item [1)]If $v\not =0$, then the series that defines $q_{T}$ converges in the Hilbert space $\mathbf{V}$ if and only if $T\in{\Omega}_{\mathbb{C}}$. \item[2)] Let $\pi_{\lambda}$ be the representation of $K^{c}$ on $\mathbf{W}^{\lambda}$. Let% \begin{equation} J_{\lambda}(g,Z):=\pi_{\lambda}(j(g,Z))\,. \label{E:jlambda}% \end{equation} Then \begin{equation} \pi(g)v=q_{g\cdot0}J_{\lambda}(g,0)^{\ast-1}v \label{E:pilambda}% \end{equation} for $g\in G^{c}$ and $v\in\mathbf{W}^{\lambda}$. \end{enumerate} \end{theorem} It follows that the span of the $q_{Z}\mathbf{W}^{\lambda}$ with $Z\in{\Omega }_{{\mathbb{C}}}$ is dense in $\mathbf{V}$, since $\mathbf{V}$ is assumed to be irreducible. Define $Q\colon{\Omega}_{\mathbb{C}}\times{\Omega}% _{\mathbb{C}}\rightarrow GL(\mathbf{W}^{\lambda})$ by% \[ Q(W,Z)=q_{W}^{\ast}q_{Z}\,. \] \label{QWZ} Then the following theorem holds. \begin{theorem} [Davidson-Fabec]\label{S:DF1}Let the notation be as above. Then the following hold: \begin{enumerate} \item \label{S:DF1(1)}$Q(W,Z)=J_{\lambda}(\exp(-\sigma(W)),Z)^{\ast-1}$. \item \label{S:DF1(2)}$Q(W,Z)$ is holomorphic in the first variable and antiholomorphic in the second variable. \item \label{S:DF1(3)}$% %TCIMACRO{\TeXButton{ip}{\ip{v}{Q(W,Z)u}}}% %BeginExpansion \ip{v}{Q(W,Z)u}% %EndExpansion =% %TCIMACRO{\TeXButton{ip}{\ip{q_{W}v}{q_{Z}u}}}% %BeginExpansion \ip{q_{W}v}{q_{Z}u}% %EndExpansion $ for all $u,v\in\mathbf{W}_{\lambda}$. \item \label{S:DF1(4)}$Q$ is a positive-definite reproducing kernel. \item \label{S:DF1(5)}$Q(g\cdot W,g\cdot Z)=J_{\lambda}(g,W)Q(W,Z)J_{\lambda }(g,Z)^{\ast}$. \end{enumerate} \end{theorem} For $Z\in{\Omega}_{\mathbb{C}}$ and $u\in\mathbf{W}^{\lambda}$, let $F_{Z,u}\colon{\Omega}_{\mathbb{C}}\rightarrow\mathbf{W}^{\lambda}$ be the holomorphic function \begin{equation} F_{Z,u}(W):=Q(W,Z)u \label{E:FQ}% \end{equation} and define \begin{equation}% %TCIMACRO{\TeXButton{ip}{\ip{F_{T,w}}{F_{Z,u}}}}% %BeginExpansion \ip{F_{T,w}}{F_{Z,u}}% %EndExpansion _{Q}:=% %TCIMACRO{\TeXButton{ip}{\ip{w}{Q(W,Z)u}}}% %BeginExpansion \ip{w}{Q(W,Z)u}% %EndExpansion . \label{E:Innpr}% \end{equation} Let $\mathbf{H}({\Omega}_{\mathbb{C}},\mathbf{W}^{\lambda})$ be the completion of the span of $\{F_{Z,u}\mid Z\in{\Omega}_{\mathbb{C}},u\in\mathbf{W}% ^{\lambda}\}$ with respect to this inner product. Then $\mathbf{H}({\Omega }_{\mathbb{C}},\mathbf{W}^{\lambda})$ is a Hilbert space consisting of $\mathbf{W}^{\lambda}$-valued holomorphic functions. Define a representation of $G^{c}$ in $\mathbf{H}({\Omega}_{\mathbb{C}},\mathbf{W}^{\lambda})$ by \begin{equation} (\rho(g)F)(W):=J_{\lambda}(g^{-1},W)^{-1}F(g^{-1}\cdot W)\,. \label{E:intrep}% \end{equation} Then $\rho$ is a unitary representation of $G^{c}$ in $\mathbf{H}({\Omega }_{\mathbb{C}},\mathbf{W}^{\lambda})$ called the \textit{geometric realization}\label{geometric} of $(\pi,\mathbf{V})$. \begin{theorem} [Davidson-Fabec]\label{ThmS-hwmNew.8}The map $q_{Z}v\mapsto F_{Z,v}$ extends to a unitary intertwining operator $U$ between $(\pi,\mathbf{V})$ and $(\rho,\mathbf{H}({\Omega}_{\mathbb{C}},\mathbf{W}^{\lambda}))$. It can be defined globally by% \[ \lbrack Uw](Z)=q_{Z}^{\ast}w,\quad w\in\mathbf{V},\,Z\in{\Omega}_{\mathbb{C}% }\,. \] \end{theorem} As the theorem stands, it gives a geometric realization for every unitary highest-weight module. What it does not do is give a natural analytic construction of the inner product on $\mathbf{H}({\Omega}_{\mathbb{C}% },\mathbf{W}^{\lambda})$. This is known only for some special representations, e.g., the \textit{holomorphic discrete series} of the group $G^{c}$ \cite{HCVI,DS,TI} or symmetric spaces of Hermitian type \cite{'OO88a,'OO88b}. At this point we will only discuss the holomorphic discrete series, which was constructed by Harish-Chandra in \cite{HCVI}, in particular Theorem 4 and Lemma 29. For that, let $\rho=\frac{1}{2}\sum_{\alpha\in\Delta^{+}}\alpha$ and let $\mu$ denote the highest weight of the representation of $K^{c}$ on $\mathbf{W}^{\lambda}$. For $f,g\in\mathbf{H}({\Omega}_{\mathbb{C}},\mathbf{W}^{\lambda})$, let $\mu$ be the $G^{c}$-invariant measure on $\mathbb{\Omega}_{\mathbb{C}}$ and \[% %TCIMACRO{\TeXButton{ip}{\ip{g}{f}}}% %BeginExpansion \ip{g}{f}% %EndExpansion _{\lambda}:=\int_{G^{c}/K^{c}}% %TCIMACRO{\TeXButton{ip}{\ip{g(Z)}{Q(Z,Z)^{-1}f(Z)}}}% %BeginExpansion \ip{g(Z)}{Q(Z,Z)^{-1}f(Z)}% %EndExpansion _{\mathbf{W}^{\lambda}}\,d\mu\,. \] \begin{theorem} [Harish-Chandra \cite{HC65,HC70}]\label{s-HC} Assume that \[ \left\langle \mu + \rho ,H_{\alpha}\right\rangle <0\text{\qquad for all }\alpha\in\Delta_{p}^{+}. \] Then $% %TCIMACRO{\TeXButton{ip}{\ip{g}{f}}}% %BeginExpansion \ip{g}{f}% %EndExpansion _{\lambda}$ is finite for $f,g\in\mathbf{H}({\Omega}_{\mathbb{C}}% ,\mathbf{W}^{\lambda})$ and there exists a positive constant $c_{\lambda}$ such that \[% %TCIMACRO{\TeXButton{ip}{\ip{g}{f}}}% %BeginExpansion \ip{g}{f}% %EndExpansion _{Q}=c_{\lambda}% %TCIMACRO{\TeXButton{ip}{\ip{g}{f}}}% %BeginExpansion \ip{g}{f}% %EndExpansion _{\lambda}\,. \] Moreover, $(\rho,\mathbf{H}({\Omega}_{\mathbb{C}},\mathbf{W}_{\lambda}))$ is unitarily equivalent to a discrete summand in $\mathbf{L}^{2}(G^{c})$. \end{theorem} The Theorem of Harish-Chandra relates some of the unitary highest weight modules to the discrete part of the Plancherel measure. It was shown by Ol'shanskii \cite{Ol82} and Stanton \cite{Stanton} that this ``holomorphic'' part of the discrete spectrum can be realized as a Hardy space of holomorphic functions on a local tube domain. Those results were generalized to symmetric spaces of \emph{Hermitian type} (or \emph{compactly causal symmetric spaces}) in a series of papers \cite{'OO88a,'OO88b,HOO91,OO96b,BO98} The last theorem shows in particular that the corresponding highest weight modules are unitary. It was shown by Wallach \cite{WaI72} and Rossi and Vergne \cite{VR76} that those are not all the unitary highest weight modules. The problem is to decide for which representations of $K^{c}$ the reproducing kernel $Q(Z,W)$ is positive definite. We refer to \cite{EHW83,Ja83} for the classification of unitary highest weight modules. We will from now on assume that the representation of $K^{c}$ is a character $\chi_{\lambda}$ where $\lambda\in i\frak{t}^{\ast}$ is trivial on $\frak{t}$ $\cap\lbrack \frak{k},\frak{k}]$. Choose a maximal set $\left\{ \gamma_{1},\dots ,\gamma_{r}\right\} $ of long strongly orthogonal roots in $\Delta_{p}^{+}$. This can be done by putting $r=\operatorname*{rank}(G^{c}/K^{c})$ and then choosing $\gamma_{r}$ to be a maximal root in $\Delta_{p}^{+}$, $\gamma_{r-1}$ maximal in $\{\gamma\in\Delta_{p}^{+}\mid\gamma\,\ $strongly orthogonal to $\gamma_{r}\}$, etc. Let $H_{j}:=H_{\gamma_{j}}$ and \begin{equation} \frak{a}=i\bigoplus _{j}\mathbb{R}H_{j}\subset\frak{t}. \label{eqfraktura} \end{equation} By the theorem of Moore (see\ \cite{He78}% ) we know that the roots in $\Delta_{p}$ restricted to $\frak{a}$, are given by $\pm\frac{1}{2}(\gamma_{i}+\gamma_{j})$, $1\leq i\leq j\leq r$ and possibly $\frac{1}{2}\gamma_{j}$. The root spaces for $\gamma_{j}$ are all one-dimensional and the root spaces $\frak{g}_{\pm\frac{1}{2}(\gamma _{i}+\gamma_{j})}$, $1\leq i0$ for $\left| t\right| <1$. The semigroup $S$ acts on $\mathbf{K}_{0}$ and by the L\"uscher-Mack Theorem \ref{LM} we get an highest weight module for $\widetilde{SU(1,1)}$, which is irreducible as we will see in a moment. We also know (see\ \cite{HO95}) that $S=H\exp C$ is a closed semigroup and that $\gamma I\subset I$, and actually $S$ is exactly the semigroup of elements in $SL(2,\mathbb{R})$ that act by contractions on $I$. Hence $S$ acts on $\mathbf{K}$. By a theorem of L\"uscher and Mack \cite{HiNe93,LM75}, the representation of $S$ on $\mathbf{K}$ extends to a representation of $G^{c}$, which in this case is the universal covering of $SU(1,1)$ that is locally isomorphic to $SL(2,\mathbb{R})$. According to Theorem \ref{HighestWeight} the resulting representation is a direct integral of highest weight representations. We notice that this defines a representation of $SL(2,\mathbb{R})$ if and only if certain integrality conditions hold; see \cite{JoMo84}. The question then arises to identify this direct integral and construct an explicit intertwining operator into the corresponding space of holomorphic functions on $D$. We notice first that the kernel $(y,x)\mapsto Q(y,x)=(1-yx)^{s-1}$ is the reproducing kernel of the irreducible highest weight representation given by% \begin{equation} \lbrack\rho_{s}(g)f](z)=(-cz+a)^{s-1}f\left( \frac{dz-b}{-cz+a}\right) \,,\quad g=\left( \begin{array} [c]{cc}% a & b\\ c & d \end{array} \right) \,. \label{eqEx.30}% \end{equation} In particular $Q(y,x)$ extends to a holomorphic function on $D\times\sigma(D)$ according to Remark \ref{R:oppo}. For $f\in\mathcal{C}_{c}^{\infty}(-1,1)$ define% \begin{equation} Uf(z):=\frac{1}{\pi}\int_{-1}^{1}f(u)(1-zu)^{s-1}\,du=\frac{1}{\pi}\int _{-1}^{1}f(u)Q(z,u)\,du\,. \label{eqEx.31}% \end{equation} By simple calculation, using (\ref{E:tautwist}), we get for $\gamma=\left( \begin{array} [c]{cc}% a & b\\ c & d \end{array} \right) \in S_{-}(\Omega)$:% \begin{align*} U(\pi_{s}(\gamma)f)(z) & =\frac{1}{\pi}\int_{-1}^{1}(d-bt)^{-s-1}f\left( \frac{-c+at}{at-c}\right) (1-zt)^{s-1}\,dt\\ & =\frac{1}{\pi}\int_{-1}^{1}f(u)(cu+d)^{s-1}\left( 1-z\frac{du+c}% {bu+a}\right) ^{s-1}\,du\\ & =\frac{1}{\pi}\int_{-1}^{1}f(u)\left( bu+a-dzu-cz\right) ^{s-1}\,du\\ & =\frac{1}{\pi}\int_{-1}^{1}f(u)\left( -cz+a-(dz-b)u\right) ^{s-1}\,du\\ & =(-cz+a)^{s-1}\frac{1}{\pi}\int_{-1}^{1}f(u)\left( 1-\frac{dz-b}% {-cz+a}\right) ^{s-1}\,du\\ & =\rho_{s}(\gamma)Uf(z)\,, \end{align*} where the respective representations are given by (\ref{eqEx.21}) and (\ref{eqEx.30}). Here the last equality follows from (\ref{E:jsu}), (\ref{E:jlambda}) and (\ref{E:intrep}). As $\rho_{s}$ is irreducible it follows that either $U$ is surjective or identically zero. Using that $Q(z,u)$ is the reproducing kernel for the representation $\rho_{s}$ we get for $f$ and $g$ with compact support:% \begin{align*}% %TCIMACRO{\TeXButton{ip}{\ip{Uf}{Ug}}}% %BeginExpansion \ip{Uf}{Ug}% %EndExpansion & =\frac{1}{\pi^{2}}\int_{-1}^{1}\int_{-1}^{1}\overline{f(u)}g(v)\,% %TCIMACRO{\TeXButton{ip}{\ip{Q(\cdot,u)}{Q(\cdot,v)}}}% %BeginExpansion \ip{Q(\cdot,u)}{Q(\cdot,v)}% %EndExpansion \,dv\,du\\ & =\frac{1}{\pi^{2}}\int_{-1}^{1}\int_{-1}^{1}\overline{f(u)}% g(v)Q(u,v)\,dv\,du\\ & =% %TCIMACRO{\TeXButton{ip}{\ip{f}{g}}}% %BeginExpansion \ip{f}{g}% %EndExpansion \,. \end{align*} It follows that $U$ is a unitary isomorphism. We can describe $U$ in a different way using the representation $\rho_{s}$ instead of the reproducing kernel. Let $\openone$ be the constant function $z\mapsto1$. Then% \begin{equation} \lbrack\rho_{s}(g)\openone](z)=J_{s}(g^{-1},z)=(-cz+a)^{s-1}\,. \label{eqEx.32}% \end{equation} We therefore get% \begin{align} \int_{H}f(h)\rho_{s}(h)\openone(z)\,dh & =\frac{1}{\pi}\int_{-\infty }^{\infty}\left[ \cosh(t)^{-s-1}f(\tanh(t))\right] (-\sinh(t)z+\cosh (t))^{s-1}\,dt\label{E:R*one}\\ & =\frac{1}{\pi}\int_{-1}^{1}f(u)(1-uz)^{s-1}\,du\nonumber\\ & =Uf(z)\,.\nonumber \end{align} We will meet the transform in (\ref{E:R*one}) again in the generalization of the Bargmann transform in Section \ref{S:Bargmann}. That shows that the Bargmann transform introduced in \cite{OO96} is closely related to the reflection positivity and the Osterwalder-Schrader duality. In summary, we have the representation $\pi_{s}$ from (\ref{eqEx.21}) acting on the Hilbert space $\widehat{\mathbf{H}_{+}(s)}$ of distributions obtained from completion with respect to% \[ \int_{-1}^{1}\int_{-1}^{1}\overline{f(x)}\,f(y)\left( 1-xy\right) ^{s-1}\,dx\,dy\,, \] and the unitarily equivalent representation $\rho_{s}$ from (\ref{eqEx.30}). The operator $U$ from (\ref{eqEx.31}) intertwines the two. Moreover $U$ passes to the distributions on $\left( -1,1\right) $, in the completion $\widehat{\mathbf{H}_{+}(s)}$, and we have% \begin{equation} U\left( \delta^{\left( n\right) }\right) =\frac{\left( s-1\right) \left( s-2\right) \cdots\left( s-n\right) }{\pi}\,z^{n},\label{eqEx.34}% \end{equation} where, for $n=0,1,2,\dots$, $\delta^{\left( n\right) }=\left( d/dx\right) ^{n}\delta$ are the derivatives of the Dirac ``function'', defined by \begin{equation} \left\langle f,\delta^{\left( n\right) }\right\rangle =\left\langle \left( -1\right)^{n}f^{\left( n\right) },\delta\right\rangle =\left( -1\right)^{n}f^{\left( n\right) }\left( 0\right) ,\label{eqExApr.35} \end{equation} where $f$ is a test function. Furthermore, $z^{n}$ are the monomials in the reproducing kernel Hilbert space $\mathbf{H}(s)$ corresponding to the complex kernel $\left( 1-\bar{z}w\right) ^{s-1}$. This Hilbert space consists of analytic functions $f\left( z\right) =\sum _{n=0}^{\infty}C_{n}z^{n}$, defined in $D=\left\{ z\in\mathbb{C}\mid\left| z\right| <1\right\} $, and satisfying% \[ \sum_{n=0}^{\infty}\left| C_{n}\right| ^{2}\frac{1}{\left| \binom{s-1}% {n}\right| }<\infty, \] where the $\binom{s-1}{n}$ refers to the (fractional) binomial coefficients. This sum also defines the norm in $\mathbf{H}(s)$. For every $w\in D$, the function $u_{w}\left( z\right) :=\left( 1-\bar{w}z\right) ^{s-1}$ is in $\mathbf{H}(s)$, and for the inner product, we have% \[% %TCIMACRO{\TeXButton{ip}{\ip{u_{w_{1}}}{u_{w_{2}}}}}% %BeginExpansion \ip{u_{w_{1}}}{u_{w_{2}}}% %EndExpansion _{\mathbf{H}(s)}=\left( 1-\bar{w}_{1}w_{2}\right) ^{s-1}\,. \] So $\mathbf{H}(s)$ is indeed a reproducing kernel Hilbert space, as it follows that the values $f\left( z\right) $, for $f\in\mathbf{H}(s)$ and $w\in D$, are given by the inner products% \[ f\left( w\right) =% %TCIMACRO{\TeXButton{ip}{\ip{u_{w}}{f}}}% %BeginExpansion \ip{u_{w}}{f}% %EndExpansion _{\mathbf{H}(s)}\,. \] Since $u_{w}\left( z\right) =\sum_{n=0}^{\infty}\binom{s-1}{n}\bar{w}% ^{n}z^{n}$, we conclude that the monomials $z^{n}$ form an orthogonal basis in $\mathbf{H}(s)$, and it follows from (\ref{eqEx.34}) that the distributions $\delta^{\left( n\right) }$, $n=0,1,2,\dots$, form an orthogonal basis in the Hilbert space $\widehat{\mathbf{H}_{+}(s)}$, and that% \[ \left\| \delta^{\left( n\right) }\right\| _{\widehat{\mathbf{H}_{+}(s)}% }^{2}= \frac{n!\left( 1-s\right) \left( 2-s\right) \cdots\left( n-s\right) } {\pi^{2\mathstrut}}. \] \section{\label{S:Sssp}Reflection Symmetry for Semisimple Symmetric Spaces} \setcounter{equation}{0} The main results in this section are Theorems \ref{PR} and \ref{S:Posref}. They are stated for non-compactly causal symmetric spaces, and the proofs are based on our Basic Lemma and the L\"uscher-Mack theorem. At the end of the section we show that results from Jorgensen's paper \cite{Jor86} lead to an extension of the scope of the two theorems. We now generalize the construction from the last section to a bigger class of semisimple symmetric pairs. We restrict ourself to the case of characters induced from a maximal parabolic subgroup, which leads to highest weight modules with one-dimensional lowest $K^{c}$-type. This is meant as a simplification and not as a limitation of our method. An additional source of inspiration for the present chapter is the following series of papers: \cite{Nel59,'O90a,'OO89a,'OO88a,'OO88b,OO96,JO97,KlLa83,OsSc73,Pra89,Sch86}. Assume that $G^{c}/K^{c}% =D\simeq\Omega_{\mathbb{C}}\subset\frak{p}^{+}$ is a bounded symmetric domain with $G^{c}$ simply connected and simple. Let $\theta^{c}$ be the Cartan involution on $G^{c}$ corresponding to $K^{c}$. Let $\sigma\colon D\rightarrow D$ be a conjugation, that is a non-trivial order two antiholomorphic map. Those involutions were classified in \cite{HJ75,HJ78}, see also \cite{HO95,'O90b,'O90a}. Then $\sigma$ defines an involution on the group $I_{o}(D)$, the connected component of holomorphic isometries of $D$, by% \[ \tau(f)(Z)=\sigma\left( f(\sigma(Z))\right) \,. \] But $I_{o}(D)$ is locally isomorphic to $G^{c}$, see \cite{He78}, Chapter VIII. Hence $\tau$ defines an involution on $G^{c}$ and $\frak{g}^{c}$. Let $H^{c}=G^{c\tau}$, and $\frak{h=}\left\{ X\in\frak{g}^{c}\mid\tau (X)=X\right\} $ and $\frak{q}^{c}=\left\{ X\in\frak{g}^{c}\mid \tau(X)=-X\right\} $. Then $\frak{g}^{c}=\frak{h}\oplus\frak{q}^{c}$. We define% \[ \frak{g}:=\frak{h}\oplus i\frak{q}^{c}% \] and $\frak{q=}\left\{ X\in\frak{g}\mid\tau(X)=-X\right\} =i\frak{q}^{c}$. Then $(\frak{g},\tau)$ is a symmetric pair. Let $G_{\mathbb{C}}$ be a simply connected Lie group with Lie algebra $\frak{g}_{\mathbb{C}}$ and let $G\subset G_{\mathbb{C}}$ be the connect Lie group with Lie algebra $\frak{g}$ . Then $\tau$ integrates to an involution on $G$. Let $H=G^{\tau}=\left\{ a\in G\mid\tau(a)=a\right\} $. Then $G/H$ is a symmetric space. The involution $\theta^{c}$ integrates to an involution on $G$ and $\theta :=\tau\theta^{c}$ is a Cartan involution on $G$ that commutes with $\tau$. Let $K$ be the corresponding maximal almost compact subgroup. Denote the corresponding Cartan decomposition as usually by $\frak{g}=\frak{k}% \oplus\frak{p}$. As $\tau$ is antiholomorphic it follows that $\tau(Z^{0})=-Z^{0}$, where $Z^{0}$ is a central element in $\frak{k}^{c}$ with eigenvalues $0,i,-i$% .\ Hence $G^{c}/H^{c}$ is a symmetric space of \emph{Hermitian type}, in the sense of \cite{'OO88a}. Those spaces are now usually called \emph{compactly causal symmetric spaces} because those are exactly the symmetric spaces such that $\frak{q}$ contains a regular $H$-invariant cone $C$ with $C^{o}% \cap\frak{k\not =\emptyset}$. The minimal cone is given by% \[ C_{\min}^{c}=\mathbb{R}^{+}\cdot\overline{\operatorname{conv}\left\{ \operatorname{Ad}(H)Z^{0}\right\} }\,. \] The dual spaces $G/H$ are exactly the \emph{non-compactly causal symmetric spaces}. Those are the symmetric spaces containing $H$-invariant regular cones $C$ such that $C^{o}\cap\frak{p}\not =\emptyset$. We use \cite{HO95} as a standard reference to the causal symmetric spaces. \begin{example} [Cayley type spaces]\label{ExaSsspNew.1}A special case of the above construction is when $G^{c}/K^{c}$ is a tube type domain. Let $\mathbf{c}$ be a Cayley transform from the bounded realization of $G^{c}/K^{c}$ to the unbounded realization. This can be done by choosing $\mathbf{c}% =\operatorname{Ad}(\exp(\frac{\pi i}{2}Y^{0}))$ where $\operatorname{ad}% (Y^{0})$ has eigenvalues $0,1,-1$. Then $\operatorname{Ad}(\mathbf{c}% )^{4}=\operatorname*{id}$ and $\tau=\operatorname{Ad}(\mathbf{c)}^{2}% (G^{c})=G^{c}$. Hence $\tau$ is an involution on $G^{c}$. It is also well known that $\tau(Z^{0})=-Z^{0}$. Hence $\tau$ defines a conjugation on $D$. The symmetric spaces $G/H$ are the symmetric spaces of Cayley type. We have $Y^{0}\in\frak{h}$ is central and $\frak{z}_{\frak{g}}(Y^{0})=\frak{h}$. Furthermore $\operatorname{Ad}(\mathbf{c})$ is an isomorphism $\frak{g}% ^{c}\simeq\frak{g}$. The spaces that we get from this construction are locally isomorphic to one of the following symmetric spaces, where we denote by the subscript $+$ the group of elements having positive determinant: $Sp(n,\mathbb{R})/GL(n,\mathbb{R})_{+}$, $SU(n,n)/GL(n,\mathbb{C})_{+}$, $SO^{\ast}(4n)/SU^{\ast}(2n)\mathbb{R}_{+}$, $SO(2,k)/SO(1,k-1)\mathbb{R}_{+}$ and $E_{7(-25)}/E_{6(-26)}\mathbb{R}_{+}$. \end{example} \begin{example} \label{ExaSsspNew.2}Assume that $H$ is a connected Lie group such that $H/K_{H}$, $K_{H}$ a maximal compact subgroup of $H$, is a bounded symmetric domain. Let $G^{c}=H\times H$ and $D=H/H_{K}\times\overline{H/H_{K}}$, where the bar denotes opposite complex structure. Let $\tau(d,c)=(c,d)$. Then $\tau$ is a conjugation with fixed-point set the diagonal. The corresponding involution on $G^{c}$ is $\tau(a,b)=(b,a)$. Thus $G^{c\tau}% =\operatorname*{diagonal}\simeq H$. Identify $G^{c\tau}$ with $H$. Then $G/H\ni(a,b)H\mapsto ab^{-1}\in H$ is an isomorphism. In this case $G$ is locally isomorphic to $H_{\mathbb{C}}$ and the involution $\tau$ on $\frak{g}$ is the conjugation with respect to the real form $\frak{h}\subset\frak{g}$. Let $H_{1}$ be the corresponding analytic subgroup. Then $H_{1}$ is locally isomorphic to $H$ and the symmetric space we are looking at is $G_{\mathbb{C}% }/H_{1}$. \end{example} We will need the following facts. Let $X^{0}=-iZ^{0}\in\frak{q}\cap\frak{p}$. Then $X^{0}$ is $H\cap K$-invariant,% \begin{equation} \frak{z}_{\frak{g}}(X^{0})=\frak{k}_{\mathbb{C}}^{c}\cap\frak{g=k}\cap \frak{h}\oplus\frak{p}\cap\frak{q}, \label{E:ZXo}% \end{equation} Let $\frak{n}:=\frak{p}^{+}\cap\frak{g}$, $\frak{\bar{n}}:=\frak{p}^{-}% \cap\frak{g}$, and $\frak{p}_{\max}:=(\frak{k}_{\mathbb{C}}^{c}\oplus \frak{p}^{+})\cap\frak{g}$. Then $\frak{p}_{\max}$ is a maximal parabolic subgroup of $\frak{g}$ of the form $\frak{p}_{\max}=\frak{m}\oplus \mathbb{R}X^{0}\oplus\frak{n}$, where $\frak{m}=\left\{ X\in\frak{k}% \cap\frak{h}\oplus\frak{p}\cap\frak{q}\mid B(X,X^{0})=0\right\} $, $B$ the Killing form on $\frak{g}$. We have $H\cap K=Z(X^{0})$. Let $A:=\exp (\frak{a})$, $N:=\exp(\frak{n})$, $\bar{N}:=\exp(\frak{\bar{n}})$. $M_{0}$ the analytic subgroup of $G$ corresponding to $\frak{m}$, and $M:=(H\cap K)M_{0}$. Then $M$ is a closed and $\tau$-stable subgroup of $G$, $M\cap A=\left\{ 1\right\} $, $MA=Z_{G}(A)$, and $P_{\max}=N_{G}(\frak{p}_{\max})=MAN$. We have \[ \frak{g}=\frak{h}\oplus\frak{p}_{\max}\,. \] Let $\Omega=\tau(\Omega_{\mathbb{C}})\cap\frak{g\subset\bar{n}}$. Then by \cite{FHO93}: \begin{lemma} \label{HPminOpenG}$HP_{\mathrm{min}}$ is open in $G$ and $HP_{\max}% =\exp(\Omega)P_{\max}\subset$ $\bar{N}P_{\mathrm{max}}$. \end{lemma} Let $\frak{a}=\mathbb{R}X^{0}$ and $A:=\exp(\frak{a})$. We need to fix the normalization of measures before we discuss the generalized principal series representations. Let the measure $da$ on $A$ be given by \[ \int_{A}f(a)\,da=\frac{1}{\sqrt{2\pi}}\int_{-\infty}^{\infty}\,f(a_{t}% )\,dt,\quad a_{t}=\exp2tX^{0}\,. \] Then Fourier inversion holds without any additional constants. We fix the Lebesgue measure $dX$ on $\bar{\frak{n}}$ such that, for $d\bar{n}=\exp(dX)$, we then have \[ \int_{\bar{\frak{n}}}a(\bar{n})^{-2\rho}\,d\bar{n}=1\,. \] Here $\rho(X)=\frac{1}{2}\operatorname*{tr}(${$\operatorname{ad}$% }$(X))|_{\frak{n}}$ as usual, and $a(g)\in A$, $g\in G$, is determined by $g\in KMa(g)N$. The Haar measure on compact groups will usually be normalized to have total measure one. The measure on $N$ is $\theta(d\bar{n})$. We fix a Haar measure $dm$ on $M$, and $dg$ on $G$ such that \[ \int_{G}f(g)\,dg=\int_{K}\int_{M}\int_{A}\int_{N}f(kman)a^{2\rho }\,dn\,da\,dm\,dk\,,\qquad f\in\mathcal{C}_{c}^{\infty}(G). \] Then we can normalize the Haar measure $dh$ on $H$ such that for $f\in\mathcal{C}_{c}^{\infty}(G)$, $\operatorname*{supp}(f)\subset HP_{\mathrm{max}}$, we have, see \cite{'O87}: \[ \int_{G}f(g)\,dg=\int_{H}\int_{M}\int_{A}\int_{N}f(hman)a^{2\rho }\,dn\,da\,dm\,dh\,. \] The invariant measure $d\dot{x}$ on $G/H$ is then given by \[ \int_{G}f(x)\,dx=\int_{G/H}\int_{H}f(xh)\,dh\,d\dot{x},\quad f\in \mathcal{C}_{c}(G) \] and similarly for $K/H\cap K$. \begin{lemma} \label{L:Int} Let the measures be normalized as above. Then the following hold: \begin{enumerate} \item [\hss\llap{\rm1)}]\label{L:Int(1)}Let $f\in\mathcal{C}_{c}(\bar{N}MAN)$. Then \[ \int_{G}f(g)\,dg=\int_{\bar{N}}\int_{M}\int_{A}\int_{N}f(\bar{n}man)a^{2\rho }\,d\bar{n}\,dm\,da\,dn\,. \] \item[\hss\llap{\rm2)}] \label{L:Int(2)}Let $f\in\mathcal{C}_{c}(\bar{N})$. For $y\in\bar{N}MAN$ write $y=\bar{n}(y)m_{\bar{N}}(y)a_{\bar{N}}(y)n_{\bar {N}}(y)$. Let $x\in G$. Then \[ \int_{\bar{N}}f(\bar{n}(x\bar{n}))a_{\bar{N}}(x\bar{n})^{-2\rho}\,d\bar {n}=\int_{\bar{N}}f(\bar{n})\,d\bar{n}\,. \] \item[\hss\llap{\rm3)}] \label{L:Int(3)}Write, for $g\in G$, $g=k(g)m(g)a(g)n(g)$ according to $G=KMAN$. Let $h\in\mathcal{C}(K/H\cap K)$. Then \[ \int_{K/H\cap K}h(\dot{k})\,d\dot{k}=\int_{\bar{N}}h(k(\bar{n})H\cap K)a(\bar{n})^{-2\rho}\,d\bar{n}\,. \] \item[\hss\llap{\rm4)}] \label{L:Int(4)}Let $h\in\mathcal{C}(K/H\cap K)$ and let $x\in G$. Then \[ \int_{K/H\cap K}f(k(xk)H\cap K)a(xk)^{-2\rho}\,d\dot{k}=\int_{K/H\cap K}% f(\dot{k})\,d\dot{k}% \] \item[\hss\llap{\rm5)}] \label{L:Int(5)}Assume that $\operatorname*{supp}% (f)\subset H/H\cap K\subset K/H\cap K$. Then \[ \int_{K/H\cap K}f(\dot{k})\,d\dot{k}=\int_{H/H\cap K}f(k(h)H\cap K)a(h)^{-2\rho}\,d\dot{h}\,. \] \item[\hss\llap{\rm6)}] \label{L:Int(6)}Let $f\in\mathcal{C}_{c}(\bar{N})$. Then \[ \int_{\bar{N}}f(\bar{n})\,d\bar{n}=\int_{H/H\cap K}f(\bar{n}(h))a_{\bar{N}% }(h)^{-2\rho}\,d\bar{n}\,. \] \item[\hss\llap{\rm7)}] \label{L:Int(7)}For $x\in HP_{\mathrm{max}}$ write $x=h(x)m_{H}(x)a_{H}(x)n_{H}(x)$ with $h(x)\in H$, $m_{H}(x)\in M$, $a_{H}(x)\in A$, and $n_{H}(x)\in N$. Let $f\in\mathcal{C}_{c}^{\infty }(H/H\cap K)$ and let $x\in G$ be such that $xHP_{\mathrm{max}}\subset HP_{\mathrm{max}}$. Then \[ \int_{H/H\cap K}f(h(xh)H\cap K)a_{H}(xh)^{-2\rho}\,d\dot{h}=\int_{H/H\cap K}f(\dot{h})\,d\dot{h}% \] \end{enumerate} \end{lemma} Identify $\frak{a}_{\mathbb{C}}^{\ast}$ with $\mathbb{C}$ by $\frak{a}% _{\mathbb{C}}^{\ast}\ni\nu\mapsto2\nu(X^{0})\in\mathbb{C}\,$. Then $\rho$ corresponds to $\dim\frak{n}$. For $\nu\in\frak{a}_{\mathbb{C}}^{\ast}$, let $\mathcal{C}^{\infty}(\nu)$ be the space of $\mathcal{C}^{\infty}$-functions $f\colon G\rightarrow\mathbb{C}$ such that, for $a_{t}=\exp t(2X^{0})$, we have \[ f(gma_{t}n)=e^{-(\nu+\rho)t}f(g)=a_{t}^{-(\nu+\rho)}f(g)\,. \] Define an inner product on $\mathcal{C}^{\infty}(\nu)$ by \[% %TCIMACRO{\TeXButton{ip}{\ip{f}{g}}}% %BeginExpansion \ip{f}{g}% %EndExpansion _{\nu}:=\int_{K}\,\overline{f(k)}g(k)\,dk=\int_{K/H\cap K}\,\overline {f(k)}g(k)\,d\dot{k}\,. \] Then $\mathcal{C}^{\infty}(\nu)$ becomes a pre-Hilbert space. We denote by $\mathbf{H}(\nu)$ the completion of $\mathcal{C}^{\infty}(\nu)$. Define $\pi(\nu)$ by \[ \lbrack\pi(\nu)(x)f](g):=f(x^{-1}g),\quad x,g\in G,\quad f\in\mathcal{C}% ^{\infty}(\nu)\,. \] Then $\pi(\nu)(x)$ is bounded, so it extends to a bounded operator on $\mathbf{H}(\nu)$, which we denote by the same symbol. Furthermore $\pi(\nu)$ is a continuous representation of $G$ which is unitary if and only if $\nu\in i\mathbb{R}$. By \cite{Pou92} we have $\mathbf{H}(\nu)^{\infty}=\mathcal{C}% ^{\infty}(\nu)$. We can realize $\mathbf{H}(\nu)$ as $\mathbf{L}^{2}(K/H\cap K)$ and as $\mathbf{L}^{2}(\bar{N},a(\bar{n})^{2\operatorname{Re}(\nu)}% \,d\bar{n})$ by restriction (see Lemma \ref{L:5.14}). In the first realization the representation $\pi(\nu)$ becomes \[ \lbrack\pi(\nu)(x)f](k)=a(x^{-1}k)^{-\nu-\rho}f(k(x^{-1}k)) \] and in the second \[ \lbrack\pi(\nu)(x)f](\bar{n})=a_{\bar{N}}(x^{-1}\bar{n})^{-\nu-\rho}f(\bar {n}(x^{-1}\bar{n}))\,. \] \begin{lemma} \label{L:invpar} The pairing \[ \mathbf{H}(\nu)\times\mathbf{H}(-\bar{\nu})\ni(f,g)\longmapsto% %TCIMACRO{\TeXButton{ip}{\ip{f}{g}}}% %BeginExpansion \ip{f}{g}% %EndExpansion _{\nu}:=\int_{K}\overline{f(k)}g(k)\,dk=\int_{K/H\cap K}\overline {f(k)}g(k)\,d\dot{k}% \] is $G$-invariant, i.e. \[% %TCIMACRO{\TeXButton{ip}{\ip{\pi(\nu)(x)f}{g}}}% %BeginExpansion \ip{\pi(\nu)(x)f}{g}% %EndExpansion _{\nu}=% %TCIMACRO{\TeXButton{ip}{\ip{f}{\pi(-\bar{\nu})(x^{-1})g}}}% %BeginExpansion \ip{f}{\pi(-\bar{\nu})(x^{-1})g}% %EndExpansion _{\nu}\,. \] \end{lemma} \begin{remark} \label{ImaginaryNu}We notice that if $\nu$ is purely imaginary, that is $-\bar{\nu}=\nu$, the above shows that $(\pi(\nu),\mathbf{H}(\nu))$ is then unitary. \end{remark} \begin{lemma} \label{L:5.14}{\ Let the notation be as above.} \begin{enumerate} \item [\hss\llap{\rm1)}]\label{L:5.14(1)}On $\bar{N}$ the invariant pairing $% %TCIMACRO{\TeXButton{ip}{\ip{\,\cdot\,}{\,\cdot\,}}}% %BeginExpansion \ip{\,\cdot\,}{\,\cdot\,}% %EndExpansion _{\nu}$ is given by \[% %TCIMACRO{\TeXButton{ip}{\ip{f}{g}}}% %BeginExpansion \ip{f}{g}% %EndExpansion _{\nu}=\int_{\bar{N}}\overline{f(\bar{n})}g(\bar{n})\,d\bar{n}\,,\quad f\in\mathbf{H}(\nu),\,g\in\mathbf{H}(-\bar{\nu})\,. \] \item[\hss\llap{\rm2)}] \label{L:5.14(2)}Let $\mathbf{H}_{H}(\nu)$ be the closure of $\{f\in\mathcal{C}^{\infty}(\nu)\mid\operatorname*{supp}(f)\subset HP_{\mathrm{max}}\}$. Then $\mathbf{H}_{H}(\nu)\ni f\mapsto f|_{H}% \in\mathbf{L}^{2}(H/H\cap K,a(h)^{2\operatorname{Re}(\nu)}\,d\dot{h})$ is an isometry. \item[\hss\llap{\rm3)}] \label{L:5.14(3)}Let $f\in\mathbf{H}(\nu)$, $g\in\mathbf{H}(-\bar{\nu})$ and assume that $\operatorname*{supp}(fg)\subset HP_{\mathrm{max}}$. Then \[% %TCIMACRO{\TeXButton{ip}{\ip{f}{g}}}% %BeginExpansion \ip{f}{g}% %EndExpansion _{\nu}=\int_{H/H\cap K}\overline{f(h)}g(h)\,d\dot{h}\,. \] \end{enumerate} \end{lemma} Let us assume, from now on, that there exists an element $w\in K$ such that {$\operatorname{Ad}$}$(w)(X^{0})=-X^{0}$. In particular such an element exists if $-1$ is in the Weyl group $W(\frak{a}_{q}):=N_{K}(\frak{a}_{q}% )/Z_{K}(\frak{a}_{q})$, where $\frak{a}_{q}\subset\frak{p}\cap\frak{q}$ is maximal abelian (and then maximal abelian in $\frak{p}$ and $\frak{q})$. This is always the case if $G/H$ is a Cayley type space because $G$ is then a Hermitian groups which implies that $\theta$ is an inner automorphism. The element $w$ does also exists if $G$ is a complexification of one of the groups $\frak{s}\frak{p}(n,\mathbb{R})$, $\frak{s}\frak{u}(n,n)$, $\frak{s}% \frak{o}^{\ast}(4n)$, $\frak{s}\frak{o}(2,k)$ and $\frak{e}_{7(-25)}$, see \cite{JO97}, Lemma 5.20. Let us work out more explicitly the details for the representations of the Cayley-type spaces in order to compare the existence of $(\rho_{\nu },\mathbf{K}_{\nu})$ to the existence of the complementary series, see\ Lemma \ref{CayleyConstant} below: For ``big'' $\nu$ define the intertwining operator $A(\nu)\colon\mathbf{H}% (\nu)\rightarrow\mathbf{H}(-\nu)$ by the converging integral% \[ \lbrack A(\nu)f](x):=\int_{N}f(gnw)\,dn\,, \] see \cite{KS80,Wal92}. The map $\nu\mapsto A(\nu)$ has an analytic continuation to a meromorphic function on $\frak{a}_{\mathbb{C}}^{\ast}$ intertwining $\pi(\nu)$ and $\pi(-\nu)$. Using Lemma \ref{L:invpar} we define a new invariant bilinear form on $\mathcal{C}_{c}^{\infty}(\nu)$ by \[% %TCIMACRO{\TeXButton{ip}{\ip{f}{g}}}% %BeginExpansion \ip{f}{g}% %EndExpansion _{\nu}:=% %TCIMACRO{\TeXButton{ip}{\ip{f}{A(\nu)g}}}% %BeginExpansion \ip{f}{A(\nu)g}% %EndExpansion _{\nu}. \] If there exists a (maximal) constant $R>0$ such that the invariant bilinear form $% %TCIMACRO{\TeXButton{ip}{\ip{\,\cdot\,}{\,\cdot\,}}}% %BeginExpansion \ip{\,\cdot\,}{\,\cdot\,}% %EndExpansion _{\nu}$ is positive definite for $|\nu|0$. Let \[ S(\Omega):=\{g\in G\mid gH\subset HP_{\mathrm{max}}\}=\left\{ g\in G\mid g\cdot\Omega\subset\Omega\right\} \,. \] Then $S(\Omega)$ is a closed semigroup invariant under $s\mapsto s^{\sharp }:=\tau(s)^{-1}$. It follows by construction that $S(\Omega)\subset HP_{\max}% $. We remark the following results: \begin{lemma} Let $C=C_{\mathrm{max}}$ be the maximal pointed generating cone in $\frak{q}$ containing $X^{0}$. Then the following hold: \begin{enumerate} \item Let $t>0$ and $Y\in\Omega$. Then $\exp tX^{0}\in S$ and $\exp tX^{0}\cdot Y=e^{-t}Y$. \item $S(\Omega)=H\exp C_{\mathrm{max}}$. \end{enumerate} \end{lemma} \begin{proof} (1) is a simple calculation. For (2) see \cite{HiNe93} and \cite{HO95}. \end{proof} \begin{corollary} \label{CorSsspNew.17}The semigroup $S(C_{\min})=H\exp(C_{\min})$ acts by contractions on $\Omega$. \end{corollary} \begin{lemma} \label{CSInvariant}Let $s\in S(\Omega)$ and $f\in\mathcal{C}_{c}^{\infty }(\Omega,\nu)$. Then $\pi(\nu)(s)f\in\mathcal{C}_{c}^{\infty}(\Omega,\nu)$, that is $\mathcal{C}_{c}^{\infty}(\Omega)$ is $S(\Omega)$-invariant. \end{lemma} \begin{proof} Let $f\in\mathcal{C}_{c}^{\infty}(\Omega)$ and $s\in S$. Then $\pi (\nu)(s)f(x)=f(s^{-1}x)\not =0$ only if $s^{-1}x\in\operatorname*{supp}% (f)\subset HP_{\mathrm{max}}$. Thus $\operatorname*{supp}(\pi(\nu)(s)f)\subset s\operatorname*{supp}(f)\subset sHP_{\mathrm{max}}\subset HP_{\mathrm{max}}$. \end{proof} We still assume that $G^{c}$ is simple. Let $(\rho_{\nu},\mathbf{K}_{\nu})$ be as above. Let $% %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion \in\mathbf{K}_{\nu}(\lambda-\rho)$ be the constant function $Z\mapsto1$. Then $\left\| %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion \right\| =1$. Let $H^{c}:=\left( G^{c}\right) ^{\tau}$. Then $H^{c}$ is connected. Let $\tilde{H}$ be the universal covering of $H^{c}$ and then also $H_{o}$. We notice that \[ H^{c}/H^{c}\cap K^{c}=H/H\cap K\,. \] Denote the restriction of $\rho_{\nu}$ to $H^{c}$ by $\rho_{\nu,H}$. We can lift $\rho_{\nu,H}$ to a representation of $\tilde{H}$ also denoted by $\rho_{\nu,H}$. We let $C=C_{\mathrm{min}}$ be the minimal $H$-invariant cone in $\frak{{q}}$ generated by $X^{0}$. We denote by $\tilde{C}=\tilde {C}_{\mathrm{min}}$ the minimal $G^{c}$-invariant cone in $i\frak{{g}^{c}}$ with $\tilde{C}\cap\frak{q}=\operatorname*{pr}{}_{\frak{q}}(\tilde{C})=C$, where $\operatorname*{pr}_{\frak{q}}\colon\frak{g}\rightarrow\frak{q}$ denotes the orthogonal projection (see\ \cite{HO95,'O90b}). As $L_{\mathrm{pos}}\leq0$ it follows that $\rho_{\lambda}$ extends to a holomorphic representation of the universal semigroup $\Gamma(G^{c},\tilde{C})$ corresponding to $G^{c}$ and $\tilde{C}$ (see \cite{HiNe93,KNO98}). Let $G_{1}^{c}$ be the analytic subgroup of $G_{\mathbb{C}}$ corresponding to the Lie algebra $\frak{g}^{c}$. Let $H_{1}$ be the analytic subgroup of $G_{1}^{c}$ corresponding to $\frak{h}$. Then---as we are assuming that $G\subset G_{\mathbb{C}}$---we have $H_{1}=H_{o}$. Let $\kappa\colon G^{c}\rightarrow G_{1}^{c}$ be the canonical projection and let $Z_{H}=\kappa^{-1}(Z_{G_{1}^{c}}\cap H_{o})$. Then $\rho_{\nu}$ is trivial on $Z_{H}$ as $\nu-\rho$ is trivial on $\exp([\frak{k}^{c},\frak{k}^{c}])\supset H^{c}\cap K^{c}$. Thus $\rho_{\nu}$ factors to $G^{c}/Z_{H}$, and to $\Gamma(G^{c},\tilde{C})/Z_{H}$. Notice that $(G^{c}/Z_{H})_{o}^{\tau}$ is isomorphic to $H_{o}$. Therefore we can view $H_{o}$ as subgroup of $G^{c}/Z_{H}$, and $S_{o}(C)=H_{o}\exp C$ as a subsemigroup of $\Gamma (G^{c},\tilde{C})/Z_{H}$. In particular $\tau_{\nu}(s)$ is defined for $s\in S_{o}(C)$. This allows us to write $\rho_{\nu}(h)$ or $\rho_{\nu,H}(h)$ for $h\in H_{o}$. Using (\ref{E:aNbarkC}) and Lemma \ref{LemNewS-hwm.2} we get \[ a_{\bar{N}}(h)^{\nu-\rho}=\left\langle %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion ,\rho_{\nu,H}(h)% %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion \right\rangle \,. \] In particular we get that $(h,k)\mapsto a_{\bar{N}}(h^{-1}k)^{\nu-\rho}$ is positive semidefinite if $\nu-\rho\le L_{\mathrm{pos}}$. Let us now consider the case $G=H_{\mathbb{C}}$ and $G^{c}=\tilde{H}% \times\tilde{H}$. Denote the constant $L_{\mathrm{pos}}$ for $\tilde{H}$ by $S_{\mathrm{pos}}$ and denote, for $\mu\leq S_{\mathrm{pos}}$, the representation with lowest $\tilde{H}\cap\tilde{K}$-type $\mu$ by $(\tau_{\mu },L_{\mu})$. Let $\bar{\tau}_{\mu}$ be the conjugate representation. Recall that we view $\tilde{H}$ as a subset of $G^{c}$ by the diagonal embedding \[ \tilde{H}\ni h\longmapsto(h,h)\in\Delta(G^{c}):=\{(x,x)\in G^{c}\mid x\in\tilde{H}\}\,. \] The center of $\frak{k}^{c}$ is two dimensional (over $\mathbb{R}$) and generated by $i(X^{0},X^{0})$ and $i(X^{0},-X^{0})$. We choose $Z^{0}% =i(X^{0},-X^{0})$. Then $\frak{p}^{+}=\frak{n}\times\bar{\frak{n}}$. Let $\openone$ again be a lowest weight vector of norm one. Denote the corresponding vector in the conjugate Hilbert space by $\overline{\openone}$. Then for $h\in\tilde{H}$:% \begin{align*}% %TCIMACRO{\TeXButton{ip}{\ip{\openone\otimes\overline{\openone}}{\tau_{\mu}\otimes\bar{\tau}_{\mu %}(h,h)\openone\otimes\overline{\openone}}}}% %BeginExpansion \ip{\openone\otimes\overline{\openone}}{\tau_{\mu}\otimes\bar{\tau}_{\mu }(h,h)\openone\otimes\overline{\openone}}% %EndExpansion & =% %TCIMACRO{\TeXButton{ip}{\ip{\openone}{\tau_{\lambda}(h)\openone}}}% %BeginExpansion \ip{\openone}{\tau_{\lambda}(h)\openone}% %EndExpansion \overline{% %TCIMACRO{\TeXButton{ip}{\ip{\openone}{\tau_{\lambda}(h)\openone}}}% %BeginExpansion \ip{\openone}{\tau_{\lambda}(h)\openone}% %EndExpansion }\\ & =\left| %TCIMACRO{\TeXButton{ip}{\ip{\openone}{\tau_{\lambda}(h)\openone}}}% %BeginExpansion \ip{\openone}{\tau_{\lambda}(h)\openone}% %EndExpansion \right| ^{2}\\ & =a_{\bar{N}}(h)^{2\mu}% \end{align*} Thus we define in this case $L_{\mathrm{pos}}:=2S_{\mathrm{pos}}$. As before we notice that $\tau_{\nu}\otimes\bar{\tau}_{\nu}(h,h)\openone\otimes\overline{\openone}$ is well defined on $H$. We now have a new proof that $% %TCIMACRO{\TeXButton{ip}{\ip{\,\cdot\,}{\,\cdot\,}}}% %BeginExpansion \ip{\,\cdot\,}{\,\cdot\,}% %EndExpansion $ is positive definite on $\Omega$. \begin{lemma} \label{ahaloN} For $\nu-\rho\leq L_{\mathrm{pos}}$ there exists an unitary irreducible highest weight representation $(\rho_{\nu},\mathbf{K}_{\nu})$ of $G^{c}$ and a lowest $K^{c}$-type vector $% %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion $ of norm one such that for every $h\in H$% \[ a_{\bar{N}}(h)^{\nu-\rho}=% %TCIMACRO{\TeXButton{ip}{\ip{\openone}{\tau_{\lambda}(h)\openone}}}% %BeginExpansion \ip{\openone}{\tau_{\lambda}(h)\openone}% %EndExpansion \,. \] Hence the kernel \[ (H\times H)\ni(h,k)\longmapsto a_{\bar{N}}(k^{-1}h)^{\nu-\rho}\in\mathbb{R}% \] is positive semidefinite. In particular $% %TCIMACRO{\TeXButton{ip}{\ip{\,\cdot\,}{\,\cdot\,}}}% %BeginExpansion \ip{\,\cdot\,}{\,\cdot\,}% %EndExpansion _{J}$ is positive semidefinite on $\mathbf{C}_{c}^{\infty}(\Omega,\nu)$ for $\nu-\rho\leq L_{\mathrm{pos}}$. \end{lemma} The Basic Lemma and the L\"uscher-Mack Theorem, together with the above, now imply the following Theorem: \begin{theorem} [Reflection Symmetry for Complementary Series]\label{PR} Let $G/H$% \linebreak be a non-compactly causal symmetric space such that there exists a $w\in K$ such that $\operatorname{Ad}(w)|_{\frak{a}}=-1$. Let $\pi(\nu)$ be a complementary series such that $\nu\leq L_{\mathrm{pos}}$. Let $C$ be the minimal $H$-invariant cone in $\frak{q}$ such that $S(C)$ is contained in the contraction semigroup of $HP_{\mathrm{max}}$ in $G/P_{\mathrm{max}}$. Let $\Omega$ be the bounded realization of $H/H\cap K$ in $\bar{\frak{n}}$. Let $J(f)(x):=f(\tau(x)w^{-1})$. Let $\mathbf{K}_{0}$ be the closure of $\mathcal{C}_{c}^{\infty}(\Omega,\nu)$ in $\mathbf{H}(\nu)$. Then the following hold: \begin{enumerate} \item [\hss\llap{\rm1)}]\label{PR(1)}$(G,\tau,\pi(\nu),C,J,\mathbf{K}_{0})$ satisfies the positivity conditions \textup{(PR1)--(PR2).} \item[\hss\llap{\rm2)}] \label{PR(2)}$\pi(\nu)$ defines a contractive representation $\tilde{\pi}(\nu)$ of $S(C)$ on $\mathbf{K}$ such that \[ \tilde{\pi}(\nu)(\gamma)^{\ast}=\tilde{\pi}(\nu)(\tau(\gamma)^{-1}). \] \item[\hss\llap{\rm3)}] \label{PR(3)}There exists a unitary representation $\tilde{\pi}^{c}$ of $G^{c}$ such that \begin{enumerate} \item [\hss\llap{\rm i)}]\label{PR(3)(1)}$d\tilde{\pi}(\nu)^{c}(X)=d\tilde {\pi}(\nu)(X)\,\quad\forall X\in\frak{h}$. \item[\hss\llap{\rm ii)}] \label{PR(3)(2)}$d\tilde{\pi}(\nu)^{c}% (iY)=i\,d\tilde{\pi}(\nu)(Y)\,\quad\forall Y\in C$. \end{enumerate} \end{enumerate} \end{theorem} We remark that this Theorem includes the results of R. Schrader for the complementary series of $SL(2n,\mathbb{C})$ \cite{Sch86}. In a moment we will show that actually $\tilde{\pi}(\nu)^{c}\simeq\rho_{\nu}$, where $\rho_{\nu}$ is the irreducible unitary highest weight representation of $G^{c}$ such that \[ a(h)^{\nu-\rho}=% %TCIMACRO{\TeXButton{ip}{\ip{\openone}{\rho_{\nu}(h)\openone}}}% %BeginExpansion \ip{\openone}{\rho_{\nu}(h)\openone}% %EndExpansion \] as before. {}From now on we assume that $\nu-\rho\leq L_{\mathrm{pos}}$. We notice that the inner product $% %TCIMACRO{\TeXButton{ip}{\ip{\,\cdot\,}{A(\nu)J(\,\cdot\,)}}}% %BeginExpansion \ip{\,\cdot\,}{A(\nu)J(\,\cdot\,)}% %EndExpansion $ makes sense independent of the existence of $w$. Let $\mathbf{K}_{0}$ be the completion of $\mathcal{C}_{c}^{\infty}(\Omega,\nu)$ in the norm $% %TCIMACRO{\TeXButton{ip}{\ip{\,\cdot\,}{A(\nu)J(\,\cdot\,)}}}% %BeginExpansion \ip{\,\cdot\,}{A(\nu)J(\,\cdot\,)}% %EndExpansion $. Let $\mathbf{N}$ be the space of vectors of zero length and let $\mathbf{K}$ be the completion of $\mathbf{K}_{0}/\mathbf{N}$ in the induced norm. First of all we have to show that $\pi(\nu)(\gamma)$ passes to a continuous operator $\tilde{\pi}(\nu)(\gamma)$ on $\mathbf{K}$ such that $\tilde{\pi}(\nu)(\gamma)^{\ast}=\tilde{\pi}(\nu)(\tau(\gamma)^{-1})$. For that we recall that \begin{equation} H/H\cap K=H_{o}/H_{o}\cap K=\Omega\label{eqSsspNew.14}% \end{equation} so we my replace the integration over $H$ in $% %TCIMACRO{\TeXButton{ip}{\ip{f}{A(\nu)Jf}}}% %BeginExpansion \ip{f}{A(\nu)Jf}% %EndExpansion _{\nu}$ with integration over $H_{o}$. Motivated by (\ref{E:R*one})\ we define for $f\in\mathcal{C}_{c}^{\infty}(\Omega,\nu)$ \begin{align} U(f)=\rho_{\nu}(f)% %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion & :=\int_{H_{o}}f(h)\rho_{\nu}(h)% %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion \,dh\,\label{eqSsspNew.15}\\ & =\int_{H_{o}}a_{\bar{N}}(h)^{-\nu-\rho}f(h\cdot0)J_{\nu}(h,\cdot)^{-1}\,dh. \label{E:U}% \end{align} \begin{lemma} \label{L:posref} Assume that $\nu-\rho\leq L_{\mathrm{pos}}$. Let $\rho_{\nu },\,\mathbf{K}_{\nu}$ and $% %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion $ be as specified in Lemma \textup{\ref{ahaloN}} and let $f,g\in \mathcal{C}_{c}^{\infty}(\Omega,\nu)$ and $s\in S(C)$. Then the following hold: \begin{enumerate} \item [\hss\llap{\rm1)}]\label{posref(1)}$% %TCIMACRO{\TeXButton{ip}{\ip{f}{[A(\nu)J](g)}}}% %BeginExpansion \ip{f}{[A(\nu)J](g)}% %EndExpansion _{\nu}=% %TCIMACRO{\TeXButton{ip}{\ip{Uf}{Ug}}}% %BeginExpansion \ip{Uf}{Ug}% %EndExpansion $. \item[\hss\llap{\rm2)}] \label{posref(2)}$U(\pi(\nu)(s)f)=\rho_{\nu}(s)U(f)$. \item[\hss\llap{\rm3)}] \label{posref(3)}$\pi_{\nu}(s)$ passes to a contractive operator $\frak{\pi(\nu)}(s)$ on $\mathbf{K}$ such that $\tilde{\pi}(\nu)(s)^{\ast}=\tilde{\pi}(\nu)(\tau(s)^{-1})$. \end{enumerate} \end{lemma} \begin{proof} (1) Let $f$ and $g$ be as above. Then \begin{align*}% %TCIMACRO{\TeXButton{ip}{\ip{f}{[A(\nu)J](g)}}}% %BeginExpansion \ip{f}{[A(\nu)J](g)}% %EndExpansion & =\int_{H_{o}/H_{o}\cap K}\int_{H_{o}/H_{o}\cap K}\overline{f(h)}% g(k)a_{\bar{N}}(h^{-1}k)^{\nu-\rho}\,dh\,dk\\ & =\int_{H_{o}/H_{o}\cap K}\int_{H_{o}/H_{o}\cap K}\overline{f(h)}g(k)% %TCIMACRO{\TeXButton{ip}{\ip{\openone}{\rho_{\nu}(h^{-1}k)\openone}}}% %BeginExpansion \ip{\openone}{\rho_{\nu}(h^{-1}k)\openone}% %EndExpansion \,dh\,dk\\ & =\int_{H_{o}/H_{o}\cap K}\int_{H_{o}/H_{o}\cap K}\overline{f(h)}g(k)% %TCIMACRO{\TeXButton{ip}{\ip{\rho_{\nu}(h)\openone}{\rho_{\nu}(k)\openone}}}% %BeginExpansion \ip{\rho_{\nu}(h)\openone}{\rho_{\nu}(k)\openone}% %EndExpansion \,dh\,dk\\ & =% %TCIMACRO{\TeXButton{ip}{\ip{Uf}{Ug}}}% %BeginExpansion \ip{Uf}{Ug}% %EndExpansion . \end{align*} This proves (1). (2) This follows from Lemma \ref{L:Int},7) and the following calculation: \begin{align*} U(\pi_{\nu}(s)f)% %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion & =\int f(s^{-1}h)\rho_{\nu}(h)% %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion \,dh\\ & =\int f(h(s^{-1}h))a_{H}(s^{-1}h)^{-(\nu+\rho)}\rho_{\nu}(h)% %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion \,dh\\ & =\int f(h(s^{-1}h))a_{H}(sh(s^{-1}))^{\nu-\rho}\rho_{\nu}(h)a_{H}% (s^{-1}h)^{-2\rho}% %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion \,dh\\ & =\int f(h)\rho_{\nu}(sh)% %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion \,dh\\ & =\rho_{\nu}(s)U(f)\,, \end{align*} where we have used that \[ \rho_{\nu}(sh)% %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion =a_{H}(sh)^{\nu-\rho}\rho_{\nu}(h(sh))% %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion \,. \] (3) By (1) and (2) we get: \begin{align*} \Vert\pi_{\nu}(s)f\Vert_{J}^{2} & =\Vert\rho_{\nu}(s)U(f)\Vert^{2}\\ & \leq\Vert U(f)\Vert^{2}\\ & =% %TCIMACRO{\TeXButton{ip}{\ip{f}{[A(\nu)J]f}}}% %BeginExpansion \ip{f}{[A(\nu)J]f}% %EndExpansion _{\nu}\qquad\left( =\Vert f\Vert_{J}^{2}\right) \,. \end{align*} Thus $\pi_{\nu}(s)$ passes to a contractive operator on $\mathbf{K}$. That $\tilde{\pi}_{\nu}(s)^{\ast}=\tilde{\pi}_{\nu}(\tau(s)^{-1})$ follows from Lemma \ref{L:AnuJ}. \end{proof} \begin{theorem} [Identification Theorem]\label{S:Posref} Assume that $G/H$ is non-compactly causal and that $\nu-\rho\leq L_{\mathrm{pos}}$. Let $\rho_{\nu}$, $\mathbf{K}_{\nu}$ and $% %TCIMACRO{\TeXButton{openone}{\openone}}% %BeginExpansion \openone %EndExpansion \in\mathbf{K}_{\nu}$ be as in Lemma \textup{\ref{ahaloN}.} Then the following hold: \begin{enumerate} \item [\hss\llap{\rm1)}]\label{S:Posref(1)}There exists a continuous contractive representation $\frak{\pi(\nu)}$ of $S_{o}(C)$ on $\mathbf{K}$ such that \[ \frak{\pi(\nu)}(s)^{\ast}=\frak{\pi(\nu)}(\tau(s)^{-1})\,,\quad\forall s\in S_{o}(C)\,. \] \item[\hss\llap{\rm2)}] \label{S:Posref(2)}There exists a unitary representation $\tilde{\pi}(\nu)^{c}$ of $G^{c}$ such that \begin{enumerate} \item [\hss\llap{\rm i)}]\label{S:Posref(2)(1)}$d\tilde{\pi}(\nu )^{c}(X)=d\tilde{\pi}(\nu)(X)\,\quad\forall X\in\frak{h}$. \item[\hss\llap{\rm ii)}] \label{S:Posref(2)(2)}$d\tilde{\pi}(\nu )^{c}(iY)=i\,d\tilde{\pi}(\nu)(Y)\,\quad\forall Y\in C$. \end{enumerate} \item[\hss\llap{\rm3)}] \label{S:Posref(3)}The map \[ \mathcal{C}_{c}^{\infty}(\Omega,\nu)\ni f\longmapsto U(f)\in\mathbf{K}_{\nu}% \] extends to an isometry $\mathbf{K}\simeq\mathbf{K}_{\nu}$ intertwining $\tilde{\pi}(\nu)^{c}$ and $\rho_{\nu}$. In particular $\tilde{\pi}(\nu)^{c}$ is irreducible and isomorphic to $\rho_{\nu}$. \end{enumerate} \end{theorem} \begin{proof} (1) follows from Lemma \ref{L:posref} as obviously $\tilde{\pi}(\nu )(sr)=\tilde{\pi}(\nu)(s)\tilde{\pi}_{\nu}(r)$. (2) This follows now from the Theorem of L\"uscher-Mack. (3) By Lemma \ref{L:posref} we know that $f\mapsto U(f)$ defines an isometric $S_{o}(C)$-intertwining operator. Let $f\in\mathcal{C}_{c}^{\infty}(\Omega ,\nu)$. Differentiation and the fact that $\rho_{\nu}$ is holomorphic gives \begin{enumerate} \item [\hss\llap{\rm i)}]$U(d\tilde{\pi}(\nu)^{c}(X)f)=d\rho_{\nu }(X)U(f)\,,\quad\forall X\in\frak{h}$. \item[\hss\llap{\rm ii)}] $U(i\,d\tilde{\pi}(\nu)^{c}(Y)f)=i\,d\rho_{\nu }(Y)U(f)\,,\quad\forall Y\in C$. \end{enumerate} But those are exactly the relations that define $\tilde{\pi}(\nu)^{c}$. The fact that $\frak{h}\oplus iC$ generates $\frak{g}^{c}$ implies that $f\mapsto U(f)$ induces an $\frak{g}^{c}$-intertwining operator intertwining $\tilde {\pi}(\nu)^{c}$ and $\rho_{\nu}$. As both are also representations of $G^{c}$, it follows that this is an isometric $G^{c}$-map. In particular as it is an isometry by Lemma \ref{L:posref}, part 1, it follows that the map $U$ is an isomorphism. This proves the theorem. %\hfill$\Box$ % % % % % \end{proof} We will now explain another view of the above results using local representations instead of the L\"uscher-Mack Theorem. Let $\frak{a}_{p}$ be a maximal abelian subspace of $\frak{p}$ containing $X^{0}$. Then $\frak{a}_{p}$ is contained in $\frak{q}$. Let $\Delta(\frak{g},\frak{a})$ be the set of roots of $\frak{a}$ in $\frak{g}$. We choose a positive system such that $\Delta_{+}=\left\{ \alpha\mid\alpha(X^{0})=1\right\} \subset\Delta ^{+}(\frak{g},\frak{a})$. Choose a maximal set of long strongly orthogonal roots $\gamma_{1},\dots,\gamma_{r}$, $r=rank(H/H\cap K)$. Choose $X_{j}% \in\frak{g}_{\gamma_{j}}$ such that with $X_{-j}=\tau(X_{j})$ we have $[X_{j},X_{-j}]=H_{j}:=H_{\gamma_{j}}$. Then by Theorem 5.1.8 in \cite{HO95} we have% \[ \Omega={\operatorname{Ad}}(H\cap K)\left\{ \sum_{j=1}^{r}t_{j}X_{-j}\biggm |-10$, let \[ B_{R}:={\operatorname{Ad}}(H\cap K)\left\{ \sum_{j=1}^{r}t_{j}X_{-j}\biggm |-R0$ be such that $B_{R}\subset U$. Then $\mathcal{C}_{c}^{\infty}(B_{R},\nu)\subset \mathcal{C}_{c}^{\infty}(U,\nu)$. Hence we can assume that $U=B_{R}$. Let $g\in\mathcal{C}_{c}^{\infty}(U,\nu)^{\perp}$ and let $f\in\mathcal{C}% _{c}^{\infty}(\Omega,\nu)$. We want to show that $% %TCIMACRO{\TeXButton{ip}{\ip{g}{f}}}% %BeginExpansion \ip{g}{f}% %EndExpansion _{J}=0$. Choose $0s_{0}$. Then $\pi(\nu)(a_{t})(f)\in\mathcal{C}_{c}^{\infty}(U,\nu)$ for every $t>s_{0}$. It follows that for $t>s_{0}$: \begin{align*} 0 & =% %TCIMACRO{\TeXButton{ip}{\ip{g}{\pi(\nu)(a_{t})f}}}% %BeginExpansion \ip{g}{\pi(\nu)(a_{t})f}% %EndExpansion _{J}\\ & =\int_{\Omega}\int_{\Omega}\overline{g(x)}\left[ \pi(\nu)(a_{t})f\right] (y) Q_{\sigma}(x,y) \,dx\,dy\\ & =e^{(\lambda+1)t}\int_{\Omega}\int_{\Omega}\overline{g(x)}f(e^{2t}% y) Q_{\sigma}(x,y) \,dx\,dy\\ & =e^{(\lambda-1)t}\int_{\Omega}\int_{\Omega}\overline{g(x)}f(y) Q_{\sigma}(x,e^{-2t}y) \,dx\,dy\,. \end{align*} By Lemma \ref{S:DF1} we know that $z\mapsto Q(zX,Y)$ is holomorphic on $D=\left\{ z\mid\left| z\right| <1\right\} $. As $g$ and $f$ both have compact support it follows by (\ref{eqApr.38}) that \[ F(z):=\int_{\Omega}\int_{\Omega}\overline{g(x)}f(y)Q(x,zy)\,dx\,dy \] is holomorphic on $D$. But $F(z)=0$ for $00$. By continuity $% %TCIMACRO{\TeXButton{ip}{\ip{g}{f}}}% %BeginExpansion \ip{g}{f}% %EndExpansion _{J}=0$. Thus $g=0$. \end{proof} Let us recall some basic facts from \cite{Jor86}. Let $\rho$ be a local homomorphism of a neighborhood $U$ of $e$ in $G$ into the space of linear operators on the Hilbert space $\mathbf{H}$ such that $\rho(g)$ is densely defined for $g\in U$. Furthermore $\rho|_{(U\cap H)}$ extends to a strongly continuous representation of $H$ in $\mathbf{H}$. $\rho$ is called a \emph{local representation} if there exists a dense subspace $\mathbf{D}% \subset\mathbf{H}$ such that the following hold: \begin{list}{}{\setlength {\leftmargin}{\customleftmargin} \setlength{\itemsep}{0.5\customskipamount} \setlength{\parsep}{0.5\customskipamount} \setlength{\topsep}{\customskipamount}} \item[\hss\llap{\rm LR1)}] $\forall g\in U$, ${\bf D} \subset{\bf D} (\rho (g))$, where ${\bf D} (\rho(g))$ is the domain of definition for $\rho(g)$. \item[\hss\llap{\rm LR2)}] If $g_1,g_2,g_1g_2\in U$ and $u\in{\bf D}$ then $\rho(g_2)u \in{\bf D} (\rho(g_1))$ and \[ \rho(g_1)[\rho(g_2)u] = \rho(g_1g_2)u\, . \] \item[\hss\llap{\rm LR3)}] Let $Y\in{\frak h}$ such that $\exp tY\in U$ for $0\le t\le1$. Then for every $u\in{\bf D}$ \[ \lim_{t\to0} \rho(\exp tY) u = u\, . \] \item[\hss\llap{\rm LR4)}] $\rho(Y){\bf D}\subset{\bf D}$ for every $Y\in {\frak h}$. \item[\hss\llap{\rm LR5)}] $\forall u\in{\bf D}\, \exists V_u$ an open $1$-neighborhod in $H$ such that $UV_u\subset U^2$ and $\rho(h)u \in{\bf D}$ for every $h\in V_u$. \item[\hss\llap{\rm LR6)}] For every $Y\in{\frak q}$ and every $u\in{\bf D}$ the function \[ h\longmapsto\rho(\exp(\operatorname{Ad}(h)Y))u \] is locally integrable on $\{h\in H\mid\exp(\operatorname{Ad}(h)Y)\in U\}$. \end{list} \cite{Jor86} now states that every local representation extends to a unitary representation of $G^{c}$. We now want to use Theorem \ref{Dense} to construct a local representation of $G$. For that let $00}$ is a semigroup, that is \begin{equation} H_{t}\ast H_{s}=H_{t+s}. \label{E:Heatsemigr}% \end{equation} This gives us the following form for $RR^{\ast}$ and $\left| R^{\ast}\right| =\sqrt{RR^{\ast}}$:% \begin{align} RR^{\ast}h(x) & =e^{-x^{2}/2}R^{\ast}h(x)\nonumber\\ & =H_{1}\ast h(x)\,. \label{E:RR*classical}% \end{align} Thus% \begin{align} \sqrt{RR^{\ast}}(h)(x) & =H_{1/2}\ast h(x)\label{E:SqrRR*classical}\\ & =2^{n/2}\int h(y)e^{-(x-y)^{2}}\,dy\,.\nonumber \end{align} {}From this we derive the following expression for the unitary part of the polar decomposition of $R^{\ast}=B\sqrt{RR^{\ast}}$ of $R^{\ast}$:% \begin{align*} Bh(z) & =R^{\ast}\left| R^{\ast}\right| ^{-1}h(z)\\ & =e^{z^{2}/2}H_{1/2}\ast h(z)\\ & =2^{n/2}\int h(x)e^{-x^{2}+2zx-z^{2}/2}\,dy\\ & =2^{n/2}e^{-z^{2}/2}\int h(x)e^{-x^{2}+2xz}\,dx \end{align*} which is, up to a scaling factor, the usual Bargmann transform, see \cite{GF89}, p.~40. In particular it follows directly from our construction that the Bargmann transform is a unitary isomorphism. This example shows that we can recover the Bargmann transform from the \emph{restriction principle}, see \cite{OO96}, that is we have \begin{itemize} \item Manifolds $M\subset M_{\mathbb{C}}$ where $M_{\mathbb{C}}$ is a ``complexification'' of $M$; \item Groups $H\subset G$ such that $H$ acts on $M$ and $G$ acts on $M_{\mathbb{C}}$; \item Measures $\mu$ and $\lambda$ on $M$ respectively $M_{\mathbb{C}}$ with the measure $\mu$ being $H$-invariant; \item A reproducing Hilbert space $\mathbf{F}(M_{\mathbb{C}})\subset \mathbf{L}^{2}(M_{\mathbb{C}},\lambda)\cap\mathcal{O}(M_{\mathbb{C}})$ with a representation $\pi_{G}$ of $G$ given by% \[ \lbrack\pi_{G}(g)F](z)=m(g^{-1},z)^{-1}F(g^{-1}\cdot z) \] where $m$ is a ``multiplier''. \item A function $\chi$ such that the ``restriction''\ map $R(F)(x):=\chi (x)F(x)$, $x\in M$, from $\mathbf{F}(M_{\mathbb{C}})\rightarrow\mathbf{L}% ^{2}(M,\mu)$ is a closed (or continuous) $H$-intertwining operator with dense image. \end{itemize} Denote the reproducing kernel of $\mathbf{F}(M_{\mathbb{C}})$ by $K(z,w)=K_{w}(z)$. Then $K_{w}\in\mathbf{F}(M_{\mathbb{C}})$ and $K(z,w)=\overline{K(w,z)}$. The map $R^{\ast}\colon L^{2}(M,\mu)\rightarrow \mathbf{F}(M_{\mathbb{C}})$ has the form% \begin{align} R^{\ast}h(w) & =% %TCIMACRO{\TeXButton{ip}{\ip{K_w}{R^*h}}}% %BeginExpansion \ip{K_w}{R^*h}% %EndExpansion \nonumber\\ & =% %TCIMACRO{\TeXButton{ip}{\ip{RK_w}{h}}}% %BeginExpansion \ip{RK_w}{h}% %EndExpansion \nonumber\\ & =\int_{M}h(m)\overline{\chi(m)}K(w,m)\,d\mu(m)\,. \label{E:R*}% \end{align} Hence% \begin{equation} RR^{\ast}h(x)=\int_{M}h(m)\chi(x)\overline{\chi(m)}\,K(x,m)\,d\mu(m)\,. \label{E:RestrPrinc}% \end{equation} Write $R^{\ast}=B\left| R^{\ast}\right| $ for the polar decomposition of $R^{\ast}$. \begin{definition} The unitary isomorphism $B\colon\mathbf{L}^{2}(M,\mu)\rightarrow \mathbf{F}(M_{\mathbb{C}})$ is called the (generalized) Bargmann transform. \end{definition} The natural setting that we are looking at now is $H/H\cap K\simeq \Omega\subset G^{c}/K^{c}=\Omega_{\mathbb{C}}$ and one of the highest weight modules as a generalization of the Fock spaces with the representation $[\rho_{\nu}(g)F](Z)=J_{\lambda}(g^{-1},Z)^{-1}F(g^{-1}\cdot Z)$, see (\ref{E:intrep}). Here $\lambda$ corresponds to $\nu-\rho$ in our previous sections. Define% \begin{equation} RF(x)=J_{\lambda}(x,0)^{-1}F(x\cdot0),\quad x\in H\mathbb{\,}. \label{E:DefRbd}% \end{equation} Using the multiplier relation $J_{\lambda}(h^{-1}x,0)=J_{\lambda}% (h^{-1},x\cdot0)J_{\lambda}(x,0)$ which follows from (\ref{E:Multipl}) \ we get:% \begin{align*} R\left[ \rho_{\lambda}(h)F\right] (x) & =J_{\lambda}(x,0)^{-1}\left[ \rho_{\lambda}(h)F\right] (x\cdot0)\\ & =J_{\lambda}(x,0)^{-1}J_{\lambda}(h^{-1},x\cdot0)^{-1}F(h^{-1}\cdot (x\cdot0))\\ & =J_{\lambda}(h^{-1}x,0)^{-1}F((h^{-1}x)\cdot0)\\ & =[L(h)RF](x) \end{align*} where $L$ stands for the left regular representation of $H$ on $H/H\cap K$. Hence $R$ is an intertwining operator. Let $\frak{a}_{p}$, $\Delta$ and $\Delta^{+}$ be as in the end of Section \ref{S:Sssp}. By \cite{'OO88a,'OO88b} the following is known: \begin{theorem} Suppose that $% \left\langle \lambda+\rho,H_{\alpha}\right\rangle <0$ for all $\alpha\in\Delta_{+}$. Then $\rho_{\lambda}$ is equivalent to a discrete summand in $\mathbf{L}^{2}(G^{c}/H)$, that is there exists an injective $G^{c}$-map $T\colon\mathbf{H}(\rho_{\lambda})\rightarrow \mathbf{L}^{2}(G^{c}/H)$. \end{theorem} We call the resulting discrete part of $\mathbf{L}^{2}(G^{c}/H)$ the \emph{holomorphic discrete series} of $G^{c}/H$. We have (see \cite{OO96}): \begin{theorem} Assume that $\rho_{\lambda}$ is a holomorphic discrete series representation of $G^{c}/H$ then the following holds: \begin{enumerate} \item The restriction map is injective, closed and with dense image. \item The generalized Bargmann transform $B\colon\mathbf{L}^{2}(H/H\cap K)\rightarrow\mathbf{H}(\rho_{\lambda})$ is a $H$-isomorphism. \end{enumerate} \end{theorem} Denote as before the reproducing kernel of $\mathbf{H}(\rho_{\lambda})$ by $Q(W,Z)$. Then by (\ref{E:R*}) and Theorem \ref{S:DF1}:\ \begin{align} R^{\ast}f(Z) & =\int_{H/H\cap K}f(h\cdot0)\cdot\overline{J_{\lambda }(h,0)^{-1}Q(h\cdot0,Z)}\,dh\nonumber\\ & =\int_{H/H\cap K}f(h\cdot0)J_{\lambda}(h^{-1},Z)\,dh\,. \label{E:R*general}% \end{align} If $Z=x\cdot0$, $x\in H$, then $J_{\lambda}(h^{-1},Z)=J_{\lambda}% (h^{-1}x,0)J(x,0)^{-1}$. Let $\Psi_{\lambda}(h):=J_{\lambda}(h,0)$ and view $f(h\cdot0)$ as a right $H\cap K$-invariant function on $H$. Then $RR^{\ast}$ is the convolution operator% \[ RR^{\ast}f(x)=\int_{H}f(h)\Psi_{\lambda}(h^{-1}x)\,dh\,. \] The problem that we face here is, that $\lambda\mapsto\Psi_{\lambda}\ast$ is \emph{not} a semigroup of operators. Hence it becomes harder to find the square root of $RR^{\ast}$, but it can still be shown, that it is a convolution operator. The final remark in this section is the following connection between our map $U$ that comes from the reflection positivity and $R^{\ast}$. \begin{theorem} Let $f\in\mathcal{C}_{c}^{\infty}(\Omega,\nu)$. Define a function $F$ on $H/H\cap K$ by $F(h)=a_{\bar{N}}(h)^{-\nu-\rho}f(h\cdot_{\operatorname*{opp}% }0)$. Then% \[ Uf(Z)=R^{\ast}(F)(Z)\,. \] \end{theorem} \begin{proof} This follows from (\ref{E:R*general}) and (\ref{E:U}) \end{proof} \section{\label{Diagonal}The Heisenberg Group} \setcounter{equation}{0} A special case of the setup in Definition \ref{ReflectionSymmetric} above arises as follows: Let the group $G$, and $\tau\in\operatorname{Aut}_{2}(G)$ be as described there. Let $\mathbf{H}_{\pm}$ be two given complex Hilbert spaces, and $\pi_{\pm}\in\operatorname{Rep}(G,\mathbf{H}_{\pm})$ a pair of unitary representations. Suppose $T\colon\mathbf{H}_{-}\rightarrow \mathbf{H}_{+}$ is a unitary operator such that $T\pi_{-}=\left( \pi_{+}% \circ\tau\right) T$, or equivalently, \begin{equation} T\pi_{-}(g)f_{-}=\pi_{+}\left( \tau(g)\right) Tf_{-} \label{TPiMinus}% \end{equation} for all $g\in G$, and all $f_{-}\in\mathbf{H}_{-}$. Form the direct sum $\mathbf{H}:=\mathbf{H}_{+}\oplus\mathbf{H}_{-}$ with inner product \begin{equation}% %TCIMACRO{\TeXButton{ip}{\ip{f_{+}\oplus f_{-}}{f^{\prime}_{+}\oplus f^{\prime %}_{-}}}}% %BeginExpansion \ip{f_{+}\oplus f_{-}}{f^{\prime}_{+}\oplus f^{\prime }_{-}}% %EndExpansion :=% %TCIMACRO{\TeXButton{ip}{\ip{f_{+}}{f^{\prime}_{+}}}}% %BeginExpansion \ip{f_{+}}{f^{\prime}_{+}}% %EndExpansion _{+}+% %TCIMACRO{\TeXButton{ip}{\ip{f_{-}}{f^{\prime}_{-}}}}% %BeginExpansion \ip{f_{-}}{f^{\prime}_{-}}% %EndExpansion _{-} \label{IPPlusMinus}% \end{equation} where the $\pm$ subscripts are put in to refer to the respective Hilbert spaces $\mathbf{H}_{\pm}$, and we may form $\pi:=\pi_{+}\oplus\pi_{-}$ as a unitary representation on $\mathbf{H}=\mathbf{H}_{+}\oplus\mathbf{H}_{-}$ by \[ \pi(g)\left( f_{+}\oplus f_{-}\right) =\pi_{+}(g)f_{+}\oplus\pi_{-}% (g)f_{-}\,,\quad g\in G,\;f_{\pm}\in\mathbf{H}_{\pm}\,. \] Setting \begin{equation} J:=\left( \begin{matrix} 0 & T\\ T^{\ast} & 0 \end{matrix} \right) \,, \label{JMatrix}% \end{equation} that is $J\left( f_{+}\oplus f_{-}\right) =\left( Tf_{-}\right) \oplus\left( T^{\ast}f_{+}\right) $, it is then clear that properties (1)--(2) from Definition \ref{ReflectionSymmetric} will be satisfied for the pair $(J,\pi)$. Formula (\ref{TPiMinus}) may be recovered by writing out the relation \begin{equation} J\pi=\left( \pi\circ\tau\right) J \label{JPiRelation}% \end{equation} in matrix form, specifically \[ \left( \begin{matrix} 0 & T\\ T^{\ast} & 0 \end{matrix} \right) \left( \begin{matrix} \pi_{+}(g) & 0\\ 0 & \pi_{-}(g) \end{matrix} \right) =\left( \begin{matrix} \pi_{+}(\tau(g)) & 0\\ 0 & \pi_{-}(\tau(g)) \end{matrix} \right) \left( \begin{matrix} 0 & T\\ T^{\ast} & 0 \end{matrix} \right) \,. \] If, conversely, (\ref{JPiRelation}) is assumed for some unitary period-$2$ operator $J$ on $\mathbf{H}=\mathbf{H}_{+}\oplus\mathbf{H}_{-}$, and, if the two representations $\pi_{+}$ and $\pi_{-}$ are \textit{disjoint,} in the sense that no irreducible in one occurs in the other (or, equivalently, there is no nonzero intertwiner between them), then, in fact, (\ref{TPiMinus}) will follow from (\ref{JPiRelation}). The diagonal terms in (\ref{JMatrix}) will be zero if (\ref{JPiRelation}) holds. This last implication is an application of Schur's lemma. \begin{lemma} \label{Operator}Let $0\neq\mathbf{K}_{0}$ be a closed linear subspace of $\mathbf{H}=\mathbf{H}_{+}\oplus\mathbf{H}_{-}$ satisfying the positivity condition \textup{(PR3)} in Definition \textup{\ref{Hyperbolic},} that is \begin{equation}% %TCIMACRO{\TeXButton{ip}{\ip{v}{Jv}}}% %BeginExpansion \ip{v}{Jv}% %EndExpansion \geq0\,,\quad\forall v\in\mathbf{K}_{0} \label{vJvPositivity}% \end{equation} where \begin{equation} J=\left( \begin{matrix} 0 & T\\ T^{\ast} & 0 \end{matrix} \right) \label{JMatrixBis}% \end{equation} is given from a fixed unitary isomorphism $T\colon\mathbf{H}_{-}% \rightarrow\mathbf{H}_{+}$ as in \textup{(\ref{TPiMinus}).} For $v=f_{+}\oplus f_{-}\in\mathbf{H}=\mathbf{H}_{+}\oplus\mathbf{H}_{-}$, set $P_{+}v:=f_{+}$. The closure of the subspace $P_{+}\mathbf{K}_{0}$ in $\mathbf{H}_{+}$ will be denoted $\overline{P_{+}\mathbf{K}_{0}}$. Then the subspace \[ \mathbf{G}=\left\{ \left. \left( \begin{matrix} f_{+}\\ f_{-}% \end{matrix} \right) \in\mathbf{K}_{0}\,\right| \,f_{-}\in T^{\ast}\left( \overline {P_{+}\mathbf{K}_{0}}\right) \right\} \] is the graph of a closed linear operator $M$ with domain \begin{equation} \mathbf{D}=\left\{ f_{+}\in\mathbf{H}_{+}\,\left| \,\exists f_{-}\in T^{\ast}\left( \overline{P_{+}\mathbf{K}_{0}}\right) \mathop{{\rm s.t.}% }\left( \begin{matrix} f_{+}\\ f_{-}% \end{matrix} \right) \in\mathbf{K}_{0}\right. \right\} \,; \label{MDomain}% \end{equation} and, moreover, the product operator $L:=TM$ is dissipative on this domain, that is \begin{equation}% %TCIMACRO{\TeXButton{ip}{\ip{Lf_{+}}{f_{+}}}}% %BeginExpansion \ip{Lf_{+}}{f_{+}}% %EndExpansion _{+}+% %TCIMACRO{\TeXButton{ip}{\ip{f_{+}}{Lf_{+}}}}% %BeginExpansion \ip{f_{+}}{Lf_{+}}% %EndExpansion _{+}\geq0 \label{LDissipative}% \end{equation} holds for all $f_{+}\in\mathbf{D}$. \end{lemma} \begin{proof} The details will only be sketched here, but the reader is referred to \cite{Sto51} and \cite{Jor80} for definitions and background literature. An important argument in the proof is the verification that, if a column vector of the form $\left( \begin{matrix} 0\\ f_{-}% \end{matrix} \right) $ is in $\mathbf{G}$, then $f_{-}$ must necessarily be zero in $\mathbf{H}_{-}$. But using positivity, we have \begin{equation} \left| %TCIMACRO{\TeXButton{ip}{\ip{u}{Jv}}}% %BeginExpansion \ip{u}{Jv}% %EndExpansion \right| ^{2}\leq% %TCIMACRO{\TeXButton{ip}{\ip{u}{Ju}}}% %BeginExpansion \ip{u}{Ju}% %EndExpansion% %TCIMACRO{\TeXButton{ip}{\ip{v}{Jv}}}% %BeginExpansion \ip{v}{Jv}% %EndExpansion \,,\quad\forall u,v\in\mathbf{K}_{0}\,. \label{uJvBound}% \end{equation} Using this on the vectors $u=\left( \begin{matrix} 0\\ f_{-}% \end{matrix} \right) $ and $v=\left( \begin{matrix} k_{+}\\ k_{-}% \end{matrix} \right) \in\mathbf{K}_{0}$, we get \[% %TCIMACRO{\TeXButton{ip}{\ip{\left( %\begin{matrix} %0 \\ f_{-} %\end{matrix} %\right) }{\left( %\begin{matrix} %Tk_{-} \\ T^{*}k_{+} %\end{matrix} %\right) }}}% %BeginExpansion \ip{\left( \begin{matrix} 0 \\ f_{-} \end{matrix} \right) }{\left( \begin{matrix} Tk_{-} \\ T^{*}k_{+} \end{matrix} \right) }% %EndExpansion =% %TCIMACRO{\TeXButton{ip}{\ip{f_{-}}{T^{*}k_{+}}}}% %BeginExpansion \ip{f_{-}}{T^{*}k_{+}}% %EndExpansion =0\,,\quad\forall k_{+}=P_{+}v\,. \] But, since $f_{-}$ is also in $T^{\ast}\left( \overline{P_{+}\mathbf{K}_{0}% }\right) $, we conclude that $f_{-}=0$, proving that $\mathbf{G}$ is the graph of an operator $M$ as specified. The dissipativity of the operator $L=TM$ is just a restatement of \textup{(PR3)}. \end{proof} The above result involves only the operator-theoretic information implied by the data in Definition \ref{Hyperbolic}, and, in the next lemma, we introduce the representations: \begin{lemma} \label{Normal}Let the representations $\pi_{\pm}$ and the intertwiner $T$ be given as specified before. Let $H=G^{\tau}$; and suppose we have a cone $C\subset\frak{q}$ as specified in \textup{(PR1), (PR2}$^{\prime}$\textup{)} and \textup{(PR3}$^{\prime}$\textup{).} Assume further that\begin{list}% {}{\setlength{\leftmargin}{\customleftmargin} \setlength{\itemsep}{0.5\customskipamount} \setlength{\parsep}{0.5\customskipamount} \setlength{\topsep}{\customskipamount}} \item[\hss\llap{\rm PR4)}] $\mathbf{D}$ is dense in $\mathbf{H}_{+}$; \item[\hss\llap{\rm PR5)}] The commutant of $\pi(H)$ is abelian. \end{list} Then $L=TM$ is normal. \end{lemma} \begin{proof} Since $T$ is a unitary isomorphism $\mathbf{H}_{-}\rightarrow\mathbf{H}_{+}$ we may make an identification and reduce the proof to the case where $\mathbf{H}_{+}=\mathbf{H}_{-}$ and $T$ is the identity operator. We then have \[ \pi_{-}=T^{-1}\left( \pi_{+}\circ\tau\right) T=\pi_{+}\circ\tau\,; \] and if $h\in H$, then \[ \pi_{-}(h)=\pi_{+}\left( \tau(h)\right) =\pi_{+}(h)\,; \] while, if $\tau(g)=g^{-1}$, then \[ \pi_{-}(g)=\pi_{+}\left( \tau(g)\right) =\pi_{+}\left( g^{-1}\right) \,. \] Using only the $H$ part from (PR2${}^{\prime}$), we conclude that $\mathbf{K}_{0}$ is invariant under $\pi_{+}\oplus\pi_{+}(H)$. If the projection $P_{\mathbf{K}_{0}}$ of $\mathbf{H}_{+}\oplus\mathbf{H}_{+}$ onto $\mathbf{K}_{0}$ is written as an operator matrix $\left( \begin{matrix} P_{11} & P_{12}\\ P_{21} & P_{22}% \end{matrix} \right) $ with entries representing operators in $\mathbf{H}_{+}$, and satisfying \begin{align*} P_{11}^{\ast} & =P_{11}\,,\\ P_{22}^{\ast} & =P_{22}\,,\\ P_{12}^{\ast} & =P_{21}\,,\\ P_{ij} & =P_{i1}P_{1j}+P_{i2}P_{2j}\,, \end{align*} then it follows that \begin{equation} P_{ij}\pi_{+}(h)=\pi_{+}(h)P_{ij}\quad\forall i,j=1,2,\;\forall h\in H\,, \label{PijPiPlus}% \end{equation} which puts each of the four operators $P_{ij}$ in the commutant $\pi _{+}(H)^{\prime}$ from (PR5). Using (PR4), we then conclude that $L$ is a dissipative operator with $\mathbf{D}$ as dense domain, and that $\mathbf{K}_{0}$ is the graph of this operator. Using (PR5), and a theorem of Stone \cite{Sto51}, we finally conclude that $L$ is a normal operator, that is it can be represented as a multiplication operator with dense domain $\mathbf{D}$ in $\mathbf{H}_{+}$. \end{proof} We shall consider two cases below (the Heisenberg group, and the $(ax+b)$-group) when conditions (PR4)--(PR5) can be verified from the context of the representations. Suppose $G$ has two abelian subgroups $H$, $N$, and the second $N$ also a normal subgroup, such that $G=HN$ is a product representation in the sense of Mackey \cite{Mac}. Define $\tau\in \operatorname{Aut}_{2}(G)$ by setting \begin{equation} \tau(h)=h\,,\quad\forall h\in H\,,\text{ and }\tau(n)=n^{-1}\,,\quad\forall n\in N\,. \label{TauDef}% \end{equation} The Heisenberg group is a copy of $\mathbb{R}^{3}$ represented as matrices $\vphantom{\left( \begin{matrix} 1 & a & c \\ 0 & 1 & b \\ 0 & 0 & 1\end{matrix}\right) _X}\left( \begin{matrix} 1 & a & c\\ 0 & 1 & b\\ 0 & 0 & 1 \end{matrix} \right) $, or equivalently vectors $(a,b,c)\in\mathbb{R}^{3}$. Setting $H=\left\{ (a,0,0)\mid a\in\mathbb{R}\right\} $ and \begin{equation} N=\left\{ (0,b,c)\mid b,c\in\mathbb{R}\right\} \,, \label{NHeisenberg}% \end{equation} we arrive at one example. The $(ax+b)$-group is a copy of $\mathbb{R}^{2}$ represented as matrices $\left( \begin{matrix} a & b\\ 0 & 1 \end{matrix} \right) $, $a=e^{s}$, $b\in\mathbb{R}$, $s\in\mathbb{R}$. Here we may take $H=\left\{ \left. \left( \begin{matrix} a & 0\\ 0 & 1 \end{matrix} \right) \,\right| \,a\in\mathbb{R}_{+}\right\} $ and \begin{equation} N=\left\{ \left. \left( \begin{matrix} 1 & b\\ 0 & 1 \end{matrix} \right) \,\right| \,b\in\mathbb{R}\right\} \,, \label{Naxb}% \end{equation} and we have a second example of the Mackey factorization. Generally, if $G=HN$ is specified as described, we use the representations of $G$ which are induced from one-dimensional representations of $N$. If $G$ is the Heisenberg group, or the $(ax+b)$-group, we get all the infinite-dimensional irreducible representations of $G$ by this induction (up to unitary equivalence, of course). For the Heisenberg group, the representations are indexed by $\hslash\in\mathbb{R}\setminus\{0\}$, $\hslash$ denoting Planck's constant. The representation $\pi_{\hslash}$ may be given in $\mathbf{H}=\mathbf{L}% ^{2}(\mathbb{R})$ by \begin{equation} \pi_{\hslash}(a,b,c)f(x)=e^{i\hslash(c+bx)}f(x+a)\,,\quad\forall f\in\mathbf{L}^{2}(\mathbb{R}),\;(a,b,c)\in G\,. \label{PiHBar}% \end{equation} The Stone-von Neumann uniqueness theorem asserts that every unitary representation $\pi$ of $G$ satisfying \[ \pi(0,0,c)=e^{i\hslash c}I_{\mathbf{H}(\pi)}\quad(\hslash\neq0) \] is unitarily equivalent to a direct sum of copies of the representation $\pi_{\hslash}$ in (\ref{PiHBar}). The $(ax+b)$-group (in the form $\left\{ \left. \left( \begin{matrix} e^{s} & b\\ 0 & 1 \end{matrix} \right) \,\right| \,s,b\in\mathbb{R}\right\} $) has only two inequivalent unitary irreducible representations, and they may also be given in the same Hilbert space $\mathbf{L}^{2}(\mathbb{R})$ by \begin{equation} \pi_{\pm}\left( \begin{matrix} e^{s} & b\\ 0 & 1 \end{matrix} \right) f(x)=e^{\pm ie^{x}b}f(x+s)\,,\quad\forall f\in\mathbf{L}% ^{2}(\mathbb{R})\,. \label{PiPlusMinusaxb}% \end{equation} There are many references for these standard facts from representation theory; see, e.g., \cite{Jor88}. \begin{lemma} \label{Abel}Let the group $G$ have the form $G=HN$ for locally compact abelian subgroups $H,N$, with $N$ normal, and $H\cap N=\{e\}$. Let $\chi$ be a one-dimensional unitary representation of $N$, and let $\pi =\operatorname*{ind}{}_{N}^{G}(\chi)$ be the corresponding induced representation. Then the commutant of $\left\{ \pi(H)\mid h\in H\right\} $ is an abelian von Neumann algebra: in other words, condition \textup{(PR5)} in Lemma \textup{\ref{Normal}} is satisfied. \end{lemma} \begin{proof} See, e.g., \cite{Jor88}. \end{proof} In the rest of the present section, we will treat the case of the \emph{Heisenberg group,} and the $(ax+b)$\emph{-group} will be the subject of the next section. For both groups we get pairs of unitary representations $\pi_{\pm}$ arising from some $\tau\in\operatorname{Aut}_{2}(G)$ and described as in (\ref{JPiRelation}) above. But when the two representations $\pi_{+}$ and $\pi_{-}=\pi_{+}\circ\tau$ are irreducible and disjoint, we will show that there are no spaces $\mathbf{K}_{0}$ satisfying (PR1), (PR2$^{\prime}$), and (PR3) such that $\mathbf{K}=\left( \mathbf{K}_{0}/\mathbf{N}\right) \sptilde$ is nontrivial. Here (PR2) is replaced by \begin{list}{}% {\setlength{\leftmargin}{\customleftmargin} \setlength{\itemsep}{0.5\customskipamount} \setlength{\parsep}{0.5\customskipamount} \setlength{\topsep}{\customskipamount}} \item[\hss\llap{\rm PR2$^{\prime}$)}] $C$ is a nontrivial cone in $\frak {q} $. \end{list} Since for both groups, and common to all the representations, we noted that the Hilbert space $\mathbf{H}_{+}$ may be taken as $\mathbf{L}^{2}% (\mathbb{R})$, we can have $J$ from (\ref{JMatrixBis}) represented in the form $J=\left( \begin{matrix} 0 & I\\ I & 0 \end{matrix} \right) $. Then the $J$-inner product on $\mathbf{H}_{+}\oplus\mathbf{H}% _{-}=\mathbf{L}^{2}(\mathbb{R})\oplus\mathbf{L}^{2}(\mathbb{R})\simeq \mathbf{L}^{2}(\mathbb{R},\mathbb{C}^{2})$ may be brought into the form \begin{equation}% %TCIMACRO{\TeXButton{ip}{\ip{\left( \begin{matrix} f_{+} \\ f_{-}% %\end{matrix}\right) }{\left( \begin{matrix} f_{+} \\ f_{-} \end{matrix}% %\right) }}}% %BeginExpansion \ip{\left( \begin{matrix} f_{+} \\ f_{-}\end{matrix}\right) }{\left ( \begin{matrix} f_{+} \\ f_{-} \end{matrix}\right) }% %EndExpansion _{J}=2\operatorname{Re}% %TCIMACRO{\TeXButton{ip}{\ip{f_{+}}{f_{-}}}}% %BeginExpansion \ip{f_{+}}{f_{-}}% %EndExpansion =2\int_{-\infty}^{\infty}\operatorname{Re}\left( \overline{f_{+}(x)}% f_{-}(x)\right) \,dx\,. \label{JInnerProduct}% \end{equation} For the two examples, we introduce \[ N_{+}=\left\{ (0,b,c)\mid b,c\in\mathbb{R}_{+}\right\} \] where $N$ is defined in (\ref{NHeisenberg}), but $N_{+}$ is not $H$-invariant. Alternatively, set \[ N_{+}=\left\{ \left. \left( \begin{matrix} 1 & b\\ 0 & 1 \end{matrix} \right) \,\right| \,b\in\mathbb{R}_{+}\right\} \] for the alternative case where $N$ is defined from (\ref{Naxb}), and note that this $N_{+}$ is $H$-invariant. In fact there are the following $4$ invariant cones in $\frak{q}$: \begin{align*} C_{1}^{+} & =\left\{ (0,0,t)\mid t\geq0\right\} \\ C_{1}^{-} & =\left\{ (0,0,t)\mid t\leq0\right\} \\ C_{2}^{+} & =\left\{ (0,x,y)\mid x\in\mathbb{R},\;y\geq0\right\} \\ C_{2}^{-} & =\left\{ (0,x,y)\mid x\in\mathbb{R},\;y\leq0\right\} \end{align*} Let $\pi$ denote one of the representations of $G=HN$ from the discussion above (see formulas (\ref{PiHBar}) and (\ref{PiPlusMinusaxb})) and let $\mathcal{D}$ be a closed subspace of $\mathbf{H}=\mathbf{L}^{2}(\mathbb{R})$ which is assumed invariant under $\pi(HN_{+})$. Then it follows that the two spaces \begin{align} \mathcal{D}_{\infty} & :=\bigvee\left\{ \pi(n)\mathcal{D}\mid n\in N\right\} \label{DInfinity}\\ \mathcal{D}_{-\infty} & :=\bigwedge\left\{ \pi(n)\mathcal{D}\mid n\in N\right\} \label{DMinusInfinity}% \end{align} are invariant under $\pi(G)$, where the symbols $\bigvee$ and $\bigwedge$ are used for the usual lattice operations on closed subspaces in $\mathbf{H}$. We leave the easy verification to the reader, but the issue is resumed in the next section. If $P_{\infty}$, resp., $P_{-\infty}$, denotes the projection of $\mathbf{H}$ onto $\mathcal{D}_{\infty}$, resp., $\mathcal{D}_{-\infty}$, then we assert that both projections $P_{\pm\infty}$ are in the commutant of $\pi(G)$. So, if $\pi$ is irreducible, then each $P_{\infty}$, or $P_{-\infty }$, must be $0$ or $I$. Since $\mathcal{D}_{-\infty}\subset\mathcal{D}% \subset\mathcal{D}_{\infty}$ from the assumption, it follows that $P_{\infty }=I$ if $\mathcal{D}\neq\{0\}$. \begin{lemma} \label{Heisenberg}Let $G$ be the Heisenberg group, and let the notation be as described above. Let $\pi_{+}$ be one of the representations $\pi_{\hslash}$ and let $\pi_{-}$ be the corresponding $\pi_{-\hslash}$ representation. Let $0\neq\mathbf{K}_{0}\subset\mathbf{L}^{2}(\mathbb{R})\oplus\mathbf{L}% ^{2}(\mathbb{R})$ be a closed subspace which is invariant under $\left( \pi_{+}\oplus\pi_{-}\right) \left( HN_{+}\right) $. Then it follows that there are only the following possibilities for $\overline{P_{+}\mathbf{K}_{0}% }$: $\{0\}$, $\mathbf{L}^{2}(\mathbb{R})$, or $A\mathcal{H}_{+}$ where $\mathcal{H}_{+}$ denotes the Hardy space in $\mathbf{L}^{2}(\mathbb{R})$ consisting of functions $f$ with Fourier transform $\hat{f}$ supported in the half-line $\left[ 0,\infty\right) $, and where $A\in\mathbf{L}^{\infty }(\mathbb{R})$ is such that $\left| A(x)\right| =1\mathop{{\rm a.e.}}% x\in\mathbb{R}$. For the space $\overline{P_{-}\mathbf{K}_{0}}$, there are the possibilities: $\{0\}$, $\mathbf{L}^{2}(\mathbb{R})$, and $A\mathcal{H}_{-}$, where $A$ is a \textup{(}possibly different\/\textup{)} unitary $\mathbf{L}% ^{\infty}$-function, and $\mathcal{H}_{-}$ denotes the negative Hardy space. \end{lemma} \begin{proof} Immediate from the discussion, and the Beurling-Lax theorem classifying the closed subspaces in $\mathbf{L}^{2}(\mathbb{R})$ which are invariant under the multiplication operators, $f(x)\mapsto e^{iax}f(x)$, $a\in\mathbb{R}_{+}$. We refer to \cite{LaPh}, or \cite{Hel64}, for a review of the Beurling-Lax theorem. \end{proof} \begin{corollary} \label{Positive}Let $\pi_{\pm}$ be the representations of the Heisenberg group, and suppose that the subspace $\mathbf{K}_{0}$ from Lemma \textup{\ref{Heisenberg}} is chosen such that \textup{(PR1)--(PR3)} in Definition \textup{\ref{Hyperbolic}} hold. Then $\left( \mathbf{K}% _{0}/\mathbf{N}\right) \sptilde=\{0\}$. \end{corollary} \begin{proof} Suppose there are unitary functions $A_{\pm}\in\mathbf{L}^{\infty}% (\mathbb{R})$ such that $\overline{P_{\pm}\mathbf{K}_{0}}=A_{\pm}% \mathcal{H}_{\pm}$. Then this would violate the Schwarz-estimate (\ref{uJvBound}), and therefore condition (PR3). Using irreducibility of $\pi_{+}=\pi_{\hslash}$ and of $\pi_{-}=\pi_{+}\circ\tau=\pi_{-\hslash}$, we may reduce to considering the cases when one of the spaces $\overline{P_{\pm }\mathbf{K}_{0}}$ is $\mathbf{L}^{2}(\mathbb{R})$. By Lemma \ref{Normal}, we are then back to the case when $\mathbf{K}_{0}$ or $\mathbf{K}_{0}^{-1}$ is the graph of a densely defined normal and dissipative operator $L$, or $L^{-1}$, respectively. We will consider $L$ only. The other case goes the same way. Since \begin{equation} \left( \pi_{+}\oplus\pi_{-}\right) \left( 0,b,0\right) \left( f_{+}\oplus f_{-}\right) (x)=e^{i\hslash bx}f_{+}(x)\oplus e^{-i\hslash bx}f_{-}(x) \label{Pibf}% \end{equation} it follows that $L$ must anti-commute with the multiplication operator $ix$ on $\mathbf{L}^{2}(\mathbb{R})$. For deriving this, we used assumption (PR3) at this point. We also showed in Lemma \ref{Normal} that $L$ must act as a multiplication operator on the Fourier-transform side. But the anti-commutativity is inconsistent with a known structure theorem in \cite{Ped90}, specifically Corollary 3.3 in that paper. Hence there are unitary functions $A_{\pm}$ in $\mathbf{L}^{\infty}(\mathbb{R})$ such that $\overline{P_{\pm}\mathbf{K}_{0}}=A_{\pm}\mathcal{H}_{\pm}$. But this possibility is inconsistent with positivity in the form $\operatorname{Re}% %TCIMACRO{\TeXButton{ip}{\ip{f_{+}}{f_{-}}}}% %BeginExpansion \ip{f_{+}}{f_{-}}% %EndExpansion \geq0\,,\quad\forall(f_{+},f_{-})\in\mathbf{K}_{0}$ (see (\ref{JInnerProduct}% )) if $\left( \mathbf{K}_{0}/\mathbf{N}\right) \sptilde\neq\{0\}$. To see this, note that $\mathbf{K}_{0}$ is invariant under the unitary operators (\ref{Pibf}) for $b\in\mathbb{R}_{+}$. The argument from Lemma \ref{Heisenberg}, now applied to $\pi_{+}\oplus\pi_{-}$, shows that the two subspaces \[ \mathbf{K}_{0}^{\infty}:=\bigvee_{b\in\mathbb{R}}\left( \pi_{+}\oplus\pi _{-}\right) \left( 0,b,0\right) \mathbf{K}_{0}% \] and \[ \mathbf{K}_{0}^{-\infty}:=\bigwedge_{b\in\mathbb{R}}\left( \pi_{+}\oplus \pi_{-}\right) \left( 0,b,0\right) \mathbf{K}_{0}% \] are both invariant under the whole group $\left( \pi_{+}\oplus\pi_{-}\right) (G)$. But the commutant of this is $2$-dimensional: the only projections in the commutant are represented as one of the following, \[ \left( \begin{matrix} 0 & 0\\ 0 & 0 \end{matrix} \right) ,\;\left( \begin{matrix} I & 0\\ 0 & 0 \end{matrix} \right) ,\;\left( \begin{matrix} 0 & 0\\ 0 & I \end{matrix} \right) ,\text{\quad or\quad}\left( \begin{matrix} I & 0\\ 0 & I \end{matrix} \right) \,, \] relative to the decomposition $\mathbf{L}^{2}(\mathbb{R})\oplus\mathbf{L}% ^{2}(\mathbb{R})$ of $\pi_{+}\oplus\pi_{-}$. The above analysis of the anti-commutator rules out the cases $\left( \begin{matrix} I & 0\\ 0 & 0 \end{matrix} \right) $ and $\left( \begin{matrix} 0 & 0\\ 0 & I \end{matrix} \right) $, and if $\left( \mathbf{K}_{0}/\mathbf{N}\right) \sptilde \neq\{0\}$, we are left with the cases $\mathbf{K}_{0}^{\infty}=\{0\}$ and $\mathbf{K}_{0}^{\infty}=\mathbf{L}^{2}(\mathbb{R})\oplus\mathbf{L}% ^{2}(\mathbb{R})$. Recall, generally $\mathbf{K}_{0}^{-\infty}\subset \mathbf{K}_{0}\subset\mathbf{K}_{0}^{\infty}$, as a starting point for the analysis. A final application of the Beurling-Lax theorem (as in \cite{LaPh}; see also \cite{DyMc70}) to (\ref{Pibf}) then shows that there must be a pair of unitary functions $A_{\pm}$ in $\mathbf{L}^{\infty}(\mathbb{R})$ such that \begin{equation} \mathbf{K}_{0}=A_{+}\mathcal{H}_{+}\oplus A_{-}\mathcal{H}_{-} \label{KAAHH}% \end{equation} where $\mathcal{H}_{\pm}$ are the two Hardy spaces given by having $\hat{f}$ supported in $\left[ 0,\infty\right) $, respectively, $\left( -\infty,0\right] $. The argument is now completed by noting that (\ref{KAAHH}) is inconsistent with the positivity of $\mathbf{K}_{0}$ in (\ref{vJvPositivity}); that is, we clearly do not have $% %TCIMACRO{\TeXButton{ip}{\ip{\left( %\begin{matrix} %A_{+}h_{+} \\ A_{-}h_{-} %\end{matrix} %\right) }{J\left( %\begin{matrix} %A_{+}h_{+} \\ A_{-}h_{-} %\end{matrix} %\right) }}}% %BeginExpansion \ip{\left( \begin{matrix} A_{+}h_{+} \\ A_{-}h_{-} \end{matrix} \right) }{J\left( \begin{matrix} A_{+}h_{+} \\ A_{-}h_{-} \end{matrix} \right) }% %EndExpansion =2\operatorname{Re}% %TCIMACRO{\TeXButton{ip}{\ip{A_{+}h_{+}}{A_{-}h_{-}}}}% %BeginExpansion \ip{A_{+}h_{+}}{A_{-}h_{-}}% %EndExpansion $ semidefinite, for all $h_{+}\in\mathcal{H}_{+}$ and all $h_{-}\in \mathcal{H}_{-}$. This concludes the proof of the Corollary. \end{proof} \begin{remark} \label{UncorrelatedClosedSubspace}At the end of the above proof of Corollary \ref{Positive}, we arrived at the conclusion (\ref{KAAHH}) for the subspace $\mathbf{K}_{0}$ under consideration. Motivated by this, we define a closed subspace $\mathbf{K}_{0}$ in a direct sum Hilbert space $\mathbf{H}_{+}% \oplus\mathbf{H}_{-}$ to be \emph{uncorrelated} if there are closed subspaces $\mathbf{D}_{\pm}\subset\mathbf{H}_{\pm}$ in the respective summands such that \begin{equation} \mathbf{K}_{0}=\mathbf{D}_{+}\oplus\mathbf{D}_{-} \label{KDD}% \end{equation} Contained in the corollary is then the assertion that every semigroup-invariant $\mathbf{K}_{0}$ in $\mathbf{L}^{2}(\mathbb{R}% )\oplus\mathbf{L}^{2}(\mathbb{R})$ is uncorrelated, where the semigroup here is the subsemigroup $S$ in the Heisenberg group $G$ given by \begin{equation} S=\left\{ (a,b,c)\mid b\in\mathbb{R}_{+},\;a,c\in\mathbb{R}\right\} \,, \label{SinG}% \end{equation} and the parameterization is the one from (\ref{NHeisenberg}). We also had the representation $\pi$ in the form $\pi_{+}\oplus\pi_{-}$ where the respective summand representations $\pi_{\pm}$ of $G$ are given by (\ref{PiHBar}) relative to a pair $(\hslash,-\hslash)$, $\hslash\in\mathbb{R}\setminus\{0\}$ some fixed value of Planck's constant. In particular, it is assumed in Corollary \ref{Positive} that each representation $\pi_{\pm}$ is \emph{irreducible.} But for proving that some given semigroup-invariant $\mathbf{K}_{0}$ must be uncorrelated, this last condition can be relaxed considerably; and this turns out to be relevant for applications to Lax-Phillips scattering theory for the wave equation with obstacle scattering \cite{LaPh}. In that context, the spaces $\mathbf{D}_{\pm}$ will be outgoing, respectively, incoming subspaces; and the wave equation translates backwards, respectively forwards, according to the unitary one-parameter groups $\pi _{-}(0,b,0)$, respectively, $\pi_{+}(0,b,0)$, with $b\in\mathbb{R}$ representing the time-variable $t$ for the unitary time-evolution one-parameter group which solves the wave equation under consideration. The unitary-equivalence identity (\ref{JPiRelation}) stated before Lemma \ref{Operator} then implies equivalence of the wave-dynamics before, and after, the obstacle scattering. \end{remark} Before stating our next result, we call attention to the $(2n+1)$-dimensional Heisenberg group $G_{n}$ in the form $\mathbb{R}^{2n+1}=\mathbb{R}^{n}% \times\mathbb{R}^{n}\times\mathbb{R}$, in parameter form: $a,b\in \mathbb{R}^{n}$, $c\in\mathbb{R}$, and product rule \[ (a,b,c)\cdot(a^{\prime},b^{\prime},c^{\prime})=(a+a^{\prime},b+b^{\prime },c+c^{\prime}+a\cdot b^{\prime}) \] where $a+a^{\prime}=(a_{1}+a_{1}^{\prime},\dots,a_{n}+a_{n}^{\prime})$ and $a\cdot b^{\prime}=\sum_{j=1}^{n}a_{j}b_{j}^{\prime}$. For every (fixed) $b\in\mathbb{R}^{n}\setminus\{0\}$, we then have a subsemigroup \begin{equation} S(b)=\left\{ (a,\beta b,c)\mid\beta\in\mathbb{R}_{+},\;a\in\mathbb{R}% ^{n},\;c\in\mathbb{R}\right\} \,; \label{SubsemigroupS}% \end{equation} and we show in the next result that it is enough to have invariance under such a semigroup in $G_{n}$, just for a single direction, defined from some fixed $b\in\mathbb{R}^{n}\setminus\{0\}$. \begin{theorem} \label{KNoughtUncorrelated}Let $\pi_{\pm}$ be unitary representations of the Heisenberg group $G$ on respective Hilbert spaces $\mathbf{H}_{\pm}$, and let $T\colon\mathbf{H}_{-}\rightarrow\mathbf{H}_{+}$ be a unitary isomorphism which intertwines $\pi_{-}$ and $\pi_{+}\circ\tau$ as in \textup{(\ref{TPiMinus})} where \begin{equation} \tau(a,b,c)=(a,-b,-c)\,,\quad\forall(a,b,c)\in G\simeq\mathbb{R}^{2n+1}\,. \label{TauInHeisenberg}% \end{equation} Suppose there is $\hslash\in\mathbb{R}\setminus\{0\}$ such that \begin{equation} \pi_{+}(0,0,c)=e^{i\hslash c}I_{\mathbf{H}_{+}}\,. \label{PiPlusInHeisenberg}% \end{equation} If $\mathbf{K}_{0}\subset\mathbf{H}_{+}\oplus\mathbf{H}_{-}$ is a closed subspace which is invariant under \[ \left\{ (\pi_{+}\oplus\pi_{-})(a,\beta b,c)\mid a\in\mathbb{R}^{n},\;\beta \in\mathbb{R}_{+},\;c\in\mathbb{R}\right\} \] from \textup{(\ref{SubsemigroupS}),} $b\in\mathbb{R}^{n}\setminus\{0\}$, then we conclude that $\mathbf{K}_{0}$ must automatically be uncorrelated. \end{theorem} \begin{proof} The group-law in the Heisenberg group yields the following commutator rule: \[ (a,0,0)(0,b,0)(-a,0,0)=(0,b,a\cdot b) \] for all $a,b\in\mathbb{R}^{n}$. We now apply $\pi=\pi_{+}\oplus\pi_{-}$ to this, and evaluate on a general vector $f_{+}\oplus f_{-}\in\mathbf{K}% _{0}\subset\mathbf{H}_{+}\oplus\mathbf{H}_{-}$: abbreviating $\pi(a)$ for $\pi(a,0,0)$, and $\pi(b)$ for $\pi(0,b,0)$, we get \[ \pi(a)\pi(\beta b)\pi(-a)(f_{+}\oplus f_{-})=e^{i\hslash\beta a\cdot b}\pi _{+}(\beta b)f_{+}\oplus e^{-i\hslash\beta a\cdot b}\pi_{-}(\beta b)f_{-}% \in\mathbf{K}_{0}% \] valid for all $a\in\mathbb{R}^{n}$, $\beta\in\mathbb{R}_{+}$. Note, in (\ref{PiPlusInHeisenberg}), we are assuming that $\pi_{+}$ takes on some specific value $e^{i\hslash c}$ on the one-dimensional center. Since $\pi_{-}$ is unitarily equivalent to $\pi_{+}\circ\tau$ by assumption (see (\ref{PiPlusInHeisenberg})), we conclude that \[ \pi_{-}(0,0,c)=e^{-i\hslash c}I_{\mathbf{H}_{-}}\,,\quad\forall c\in \mathbb{R}\,. \] The argument really only needs that the two representations $\pi_{\pm}$ define \emph{different} characters on the center. (Clearly $\hslash\neq-\hslash$ since $\hslash\neq0$.) Multiplying through first with $e^{-i\hslash\beta a\cdot b}$, and integrating the resulting term \[ \pi_{+}(\beta b)f_{+}\oplus e^{-i2\hslash\beta a\cdot b}\pi_{-}(\beta b)f_{-}\in\mathbf{K}_{0}% \] in the $a$-variable, we get $\pi_{+}(\beta b)f_{+}\oplus0\in\mathbf{K}_{0}$. The last conclusion is just using that $\mathbf{K}_{0}$ is a closed subspace. But we can do the same with the term \[ e^{i2\hslash\beta a\cdot b}\pi_{+}(\beta b)f_{+}\oplus\pi_{-}(\beta b)f_{-}% \in\mathbf{K}_{0}\,, \] and we arrive at $0\oplus\pi_{-}(\beta b)f_{-}\in\mathbf{K}_{0}$. Finally letting $\beta\rightarrow0_{+}$, and using strong continuity, we get $f_{+}\oplus0$ and $0\oplus f_{-}$ both in $\mathbf{K}_{0}$. Recalling that $f_{\pm}$ are general vectors in $P_{\pm}\mathbf{K}_{0}$, we conclude that $P_{+}\mathbf{K}_{0}\oplus P_{-}\mathbf{K}_{0}\subset\mathbf{K}_{0}$, and therefore $\overline{P_{+}\mathbf{K}_{0}}\oplus\overline{P_{-}\mathbf{K}_{0}% }\subset\mathbf{K}_{0}$. Since the converse inclusion is obvious, we arrive at (\ref{KDD}) with $\mathbf{D}_{\pm}=\overline{P_{\pm}\mathbf{K}_{0}}$. \end{proof} The next result shows among other things that there are representations $\pi$ of the Heisenberg group $G_{n}$ (for each $n$) such that the reflected representation $\pi^{c}$ of $G_{n}^{c}\simeq G_{n}$ (see Theorem \ref{PiCIrreducible}) acts on a nonzero Hilbert space $\mathbf{H}^{c}=\left( \mathbf{K}_{0}/\mathbf{N}\right) \sptilde$. However, because of Lemma \ref{KernelPiC}, $\pi^{c}\left( G_{n}^{c}\right) $ will automatically be an \emph{abelian} group of operators on $\mathbf{H}^{c}$. To see this, note that the proof of Theorem \ref{MDomain} shows that $\pi^{c}$ must act as the identity operator on $\mathbf{H}^{c}$ when restricted to the one-dimensional center in $G_{n}^{c}\simeq G_{n}$. It will be convenient for us to read off this result from a more general context: we shall consider a general Lie group $G$, and we fix a right-invariant Haar measure on $G$. A distribution $F$ on the Lie group $G$ will be said to be positive definite (PD) if \begin{equation} \int_{G}\int_{G}F(uv^{-1})\overline{f(u)}f(v)\,du\,dv\geq0 \tag*{(PD)}% \end{equation} for all $f\in C_{c}^{\infty}(G)$; and we say that $F$ is PD on some open subset $\Omega\subset G$ if this holds for all $f\in C_{c}^{\infty}(\Omega)$. The interpretation of the expression in (PD) is in the sense of distributions. But presently measurable functions $F$ will serve as the prime examples. We say that the distribution is reflection-positive (RP) on $\Omega$ ((RP$_{\Omega}$) for emphasis) if, for some period-$2$ automorphism $\tau$ of $G$, we have \begin{equation} F\circ\tau=F \label{FTau}% \end{equation} and \begin{equation} \int_{G}\int_{G}F(\tau(u)v^{-1})\overline{f(u)}f(v)\,du\,dv\geq0 \tag*{(RP$ _\Omega$)}% \end{equation} for all $f\in C_{c}^{\infty}(\Omega)$. We say that $x\in G$ is (RP$_{\Omega}$)-contractive if (RP$_{\Omega}$) holds, and for all $f\in C_{c}^{\infty}(\Omega)$% \begin{align*} 0 & \leq\int_{G}\int_{G}F(\tau(u)v^{-1})\overline{f(ux)}f(vx)\,du\,dv\\ & \leq\int_{G}\int_{G}F(\tau(u)v^{-1})\overline{f(u)}f(v)\,du\,dv\,. \end{align*} Note that, since \[ \int_{G}\int_{G}F(\tau(u)v^{-1})\overline{f(ux)}f(vx)\,du\,dv=\int_{G}\int _{G}F(\tau(u)\tau(x)^{-1}xv^{-1})\overline{f(u)}f(v)\,du\,dv\, \] it follows that every $x\in H$ is contractive; in fact, isometric. If instead we have $\tau(x)=x^{-1}$, then contractivity amounts to the estimate \[ \int_{G}\int_{G}F(\tau(u)x^{2}v^{-1})\overline{f(u)}f(v)\,du\,dv\leq\int _{G}\int_{G}F(\tau(u)v^{-1})\overline{f(u)}f(v)\,du\,dv\,\, \] for all $f\in C_{c}^{\infty}(\Omega)$. Using the Basic Lemma one can also show that $x$ acts by contractions. The following result is useful, but an easy consequence of the definitions and standard techniques for positive definite distributions; see for example \cite{Jor88,Pra89}. \begin{theorem} \label{PiExtends}Let $F$ be a distribution on a Lie group $G$ with a period-$2$ automorphism $\tau$, and suppose $F$ is $\tau$-invariant, \textup{(PD)} holds on $G$, and \textup{(RP}$_{\Omega}$\textup{)} holds on some open, and semigroup-invariant, subset $\Omega$ in $G$. Then define \[ (\pi(u)f)(v):=f(vu)\,,\quad\forall u,v\in G,\;\forall f\in C_{c}^{\infty }(G)\,; \] and \[ Jf:=f\circ\tau\,. \] Let $\mathbf{H}(F)$ be the Hilbert space obtained from the GNS construction, applied to \textup{(PD),} with inner product on $C_{c}^{\infty}(G)$ given by \[% %TCIMACRO{\TeXButton{ip}{\ip{f}{g}}}% %BeginExpansion \ip{f}{g}% %EndExpansion :=\int_{G}\int_{G}F(uv^{-1})\overline{f(u)}g(v)\,du\,dv\,. \] Then $\pi$ extends to a unitary representation of $G$ on $\mathbf{H}(F)$, and $J$ to a unitary operator, such that \[ J\pi=(\pi\circ\tau)J\,. \] If \textup{(RP}$_{\Omega}$\textup{)} further holds, as described, then $\pi$ induces \textup{(}via Theorem \textup{\ref{PiCIrreducible})} a unitary representation $\pi^{c}$ of $G^{c}$ acting on the new Hilbert space $\mathbf{H}^{c}$ obtained from completing in the new inner product from \textup{(RP}$_{\Omega}$\textup{),} and dividing out with the corresponding kernel. \end{theorem} The simplest example of a function $F$ on the Heisenberg group $G_{n}$ satisfying (PD), but not (RP$_{\Omega}$), for nontrivial $\Omega$'s, may be obtained from the Green's function for the sub-Laplacian on $G_{n}$; see \cite[p. 599]{Ste93} for details. If complex coordinates are introduced in $G_{n}$, the formula for $F$ takes the following simple form: let $z\in\mathbb{C}^{n}$, $c\in\mathbb{R}$, and define \[ F(z,c)=\frac{1}{\left( \left| z\right| ^{4}+c^{2}\right) ^{n}}\,. \] Then we adapt the product in $G_{n}$ to the modified definition as follows: \[ (z,c)\cdot(z^{\prime},c^{\prime})=(z+z^{\prime},c+c^{\prime}+\left\langle z,z^{\prime}\right\rangle )\quad\forall z,z^{\prime}\in\mathbb{C}% ^{n},\;\forall c,c^{\prime}\in\mathbb{R}\,, \] where $\left\langle z,z^{\prime}\right\rangle $ is the symplectic form \[ \left\langle z,z^{\prime}\right\rangle :=2\operatorname{Im}(z\cdot\bar {z}^{\prime})\,. \] The period-$2$ automorphism $\tau$ on $G_{n}$ we take as \[ \tau(z,c)=(\bar{z},-c) \] with $\bar{z}$ denoting complex conjugation $(z_{1},\dots,z_{n})\mapsto (\bar{z}_{1},\dots,\bar{z}_{n})$. The simplest example where both (PD) and (RP$_{\Omega}$) hold on the Heisenberg group $G_{n}$ is the following: \begin{example} \label{SubLaplacian}Let $\zeta=(\zeta_{1},\dots\zeta_{n})\in\mathbb{C}^{n}$, $\xi_{j}=\operatorname{Re}\zeta_{j}$, $\eta_{j}=\operatorname{Im}\zeta_{j}$, $j=1,\dots,n$. Define \[ F(z,c)=\int_{\mathbb{R}^{2n}}\frac{e^{i\operatorname{Re}(z\cdot\bar{\zeta})}% }{\prod_{j=1}^{n}(\left| \zeta_{j}\right| ^{2}+1)}\,d\xi_{1}\cdots\,d\xi _{n}\,d\eta_{1}\cdots\,d\eta_{n}\,. \] Let $\Omega:=\{(z,c)\in G_{n}\mid z=(z_{j})_{j=1}^{n},\;\operatorname{Im}% z_{j}>0\}$. Then (PD) holds on $G_{n}$, and (RP$_{\Omega}$) holds, referring to this $\Omega$. Since the expression for $F(z,c)$ factors, the problem reduces to the $(n=1)$ special case. There we have \[ F(z,c)=\int_{\mathbb{R}^{2}}\frac{e^{i(x\xi+y\eta)}}{\xi^{2}+\eta^{2}+1}% \,d\xi\,d\eta\,; \] and if $f\in C_{c}^{\infty}(\Omega)$ with $\Omega=\left\{ (z,c)\mid y>0\right\} $, then \begin{multline*} \int_{G_{1}}\int_{G_{1}}F(\tau(u)v^{-1})\overline{f(u)}f(v)\,du\,dv\\ =\int_{\mathbb{R}^{8}}\frac{e^{i(x-x^{\prime})\xi}\,e^{-i(y+y^{\prime})\eta}% }{\xi^{2}+\eta^{2}+1}\overline{f(x+iy,c)}f(x^{\prime}+iy^{\prime},c^{\prime })\,d\xi\,d\eta\,dx\,dy\,dc\,dx^{\prime}\,dy^{\prime}\,dc^{\prime}\,. \end{multline*} Let $\tilde{f}$ denote the Fourier transform in the $x$-variable, keeping the last two variables $(y,c)$ separate. Then the integral transforms as follows: \[ \int_{\mathbb{R}^{5}}\frac{e^{-(y+y^{\prime})\sqrt{1+\xi^{2}}}}{\sqrt {1+\xi^{2}}}\overline{\tilde{f}(\xi,y,c)}\tilde{f}(\xi,y^{\prime},c^{\prime })\,d\xi\,dy\,dy^{\prime}\,dc\,dc^{\prime}\,. \] Introducing the Laplace transform in the middle variable $y$, we then get (since $f$ is supported in $y>0$) \[ \int_{0}^{\infty}e^{-y\sqrt{1+\xi^{2}}}\tilde{f}(\xi,y,c)\,dy=\tilde {f}_{\lambda}(\xi,\sqrt{1+\xi^{2}},c)\,; \] the combined integral reduces further: \[ \int_{\mathbb{R}}\,\left| \vphantom{\int\tilde{f}_\lambda(\xi,\sqrt{1+\xi ^2},c)\,dc}\smash{\int_{\mathbb{R}} \tilde{f}_\lambda(\xi,\sqrt{1+\xi ^2},c)\,dc}\right| ^{2}\frac{d\xi}{1+\xi^{2}}% \] which is clearly positive; and we have demonstrated that (RP$_{\Omega}$) holds. It is immediate that $F$ is $\tau$-invariant (see (\ref{FTau})), and also that it satisfies (PD) on $G_{n}$. \end{example} \section{\label{axb}The $(ax+b)$-Group Revisited} \setcounter{equation}{0} We showed that in general we get a unitary representation $\pi^{c}$ of the group $G^{c}$ from an old one $\pi$ of $G$, provided $\pi$ satisfies the assumptions of reflection positivity. The construction as we saw uses a certain cone $C$ and a semigroup $H\exp C$, which are part of the axiom system. What results is a new class of unitary representations $\pi^{c}$ satisfying a certain spectrum condition (semi-bounded spectrum). But, for the simplest non-trivial group $G$, this semi-boundedness turns out \textit{not} to be satisfied in the general case. Nonetheless, we still have a reflection construction getting us from unitary representations $\pi$ of the $(ax+b)$-group, such that $\pi\circ\tau\simeq\pi$ (unitary equivalence), to associated unitary representations $\pi^{c}$ of the same group. The (up to conjugation) unique non-trivial period-$2$ automorphism $\tau$ of $G$, where $G$ is the $(ax+b)$-group, is given by \[ \tau(a,b)=(a,-b)\,. \] Recall that the $G$ may be identified with the matrix-group \[ \left\{ \left. \left( \begin{matrix} a & b\\ 0 & 1 \end{matrix} \right) \,\right| \,a>0,b\in\mathbb{R}\right\} \] and $(a,b)$ corresponds to the matrix $\left( \begin{matrix} a & b\\ 0 & 1 \end{matrix} \right) $. In this realization the Lie algebra of $G$ has the basis \[ X=\left( \begin{matrix} 1 & 0\\ 0 & 0 \end{matrix} \right) \quad\mbox{and}\quad Y=\left( \begin{matrix} 0 & 1\\ 0 & 0 \end{matrix} \right) \,. \] We have $\exp(tX)=(e^{t},0)$ and $\exp(sY)=(1,s)$. Hence $\tau(X)=X$ and $\tau(Y)=-Y$. Thus $\frak{h}=\mathbb{R}X$ and $\frak{q}=\mathbb{R}Y$. We notice the commutator relation $[X,Y]=Y$. The possible $H$-invariant cones in $\frak{q}$ are $\pm\{tY\mid t\geq0\}$. It is known from Mackey's theory that $G$ has two inequivalent, unitary, irreducible, infinite-dimensional representations $\pi_{\pm}$, and it is immediate that we have the unitary equivalence (see details below):% \begin{equation} \pi_{+}\circ\tau\simeq\pi_{-}\,. \label{E:5.1}% \end{equation} Hence, if we set $\pi:=\pi_{+}\oplus\pi_{-}$, then $\pi\circ\tau\simeq\pi$, so we have the setup for the general theory. We show that $\pi$ may be realized on $\mathbf{L}^{2}(\mathbb{R})\oplus\mathbf{L}^{2}(\mathbb{R})\simeq \mathbf{L}^{2}(\mathbb{R},\mathbb{C}^{2})$, and we find and classify the invariant positive subspaces $\mathbf{K}_{0}\subset\mathbf{L}^{2}% (\mathbb{R},\mathbb{C}^{2})$. To understand the interesting cases for the $(ax+b)$-group $G$, we need to relax the invariance condition: We shall \textit{not} assume invariance of $\mathbf{K}_{0}$ under the semigroup $\{\pi(1,b)\mid b\geq0\}$, but only under the infinitesimal unbounded generator $\pi(Y)$. With this, we still get the correspondence $\pi\mapsto \pi_{\mathbf{K}_{0}}^{c}$ as described above. We use the above notation. We know from Mackey's theory \cite{Mac} that there are two inequivalent irreducible infinite-dimensional representations of $G$, and we shall need them in the following alternative formulations: Let $\mathcal{L}_{\pm}$ denote the respective Hilbert space $\mathbf{L}% ^{2}(\mathbb{R}_{\pm})$ with the multiplicative invariant measure $d\mu_{\pm }=dp/|p|$, $p\in\mathbb{R}_{\pm}$. Then the formula% \begin{equation} f\longmapsto e^{ipb}f(pa) \label{E:5.4}% \end{equation} for functions $f$ on $\mathbb{R}$ restricts to two unitary irreducible representations, denoted by $\pi_{\pm}$ of $G$ on the respective spaces $\mathcal{L}_{\pm}$. Let $Q(f)(p):=f(-p)$ denote the canonical mapping from $\mathcal{L}_{+}$ to $\mathcal{L}_{-}$, or equivalently from $\mathcal{L}_{-}$ to $\mathcal{L}_{+}$. Then we have for $g\in G$ (cf.\ (\ref{E:5.1})):% \begin{equation} Q\pi_{+}(g)=\pi_{-}(\tau(g))Q \label{E:5.5}% \end{equation} For the representation $\pi:=\pi_{+}\oplus\pi_{-}$ on $\mathbf{H}% :=\mathcal{L}_{+}\oplus\mathcal{L}_{-}$ we therefore have% \begin{equation} J\pi(g)=\pi(\tau(g))J,\quad g\in G\,, \label{E:5.6}% \end{equation} where $J$ is the unitary involutive operator on $\mathbf{H}$ given by% \begin{equation} J=\left( \begin{matrix} 0 & Q\\ Q & 0 \end{matrix} \right) \,. \label{E:5.7}% \end{equation} Instead of the above $p$-realization of $\pi$ we will mainly use the following $x$-formalism. The map $t\mapsto\pm e^{t}$ defines an isomorphism $L_{\pm }\colon\mathcal{L}_{\pm}\rightarrow\mathbf{L}^{2}(\mathbb{R})$, where we use the additive Haar measure $dx$ on $\mathbb{R}$. For $g=(e^{s},b)\in G$ and $f\in\mathbf{L}^{2}(\mathbb{R})$, set% \begin{equation} (\pi_{\pm}(g)f)(x):=e^{\pm ie^{x}b}f(x+s),\quad x\in\mathbb{R}\,. \label{E:5.8}% \end{equation} A simple calculation shows that $L_{\pm}$ intertwines the old and new construction of $\pi_{\pm}$, excusing our abuse of notation. In this realization $Q$ becomes simply the identity operator $Q(f)(x)=f(x)$. The involution $J\colon\mathbf{L}^{2}(\mathbb{R},\mathbb{C}^{2})$ is now simply given by \begin{equation} J(f_{0},f_{1})=(f_{1},f_{0}) \label{eqaxbNew.7}% \end{equation} or $J=\left( \begin{matrix} 0 & 1\\ 1 & 0 \end{matrix} \right) $. In this formulation the operator \begin{equation} L:=\pi_{\pm}(\Delta_{H}-\Delta_{Q})=\pi_{\pm}(X^{2}-Y^{2}) \label{E:5.9}% \end{equation} takes the form \begin{equation} L=\left( \frac{d\,}{dx}\right) ^{2}\,+\,e^{2x}\,, \label{E:5.10}% \end{equation} but it is on $\mathbf{L}^{2}(\mathbb{R})$ and $-\infty0\right\} $ will have $\mathbf{K}=\left( \mathbf{K}_{0}/\mathbf{N}\right) \sptilde$ equal to $0$. \end{proposition} \begin{lemma} \label{QField}Let $Q$ be the projection in $\mathbf{L}^{2}(\mathbb{R}% )\oplus\mathbf{L}^{2}(\mathbb{R})$ onto a translation-invariant $J$-positive subspace. Then $Q$ is represented by a measurable field of $2\times2$ complex matrices $\mathbb{R}\ni\xi\mapsto\left( Q_{ij}(\xi)\right) _{ij=1}^{2}$ such that $\left| Q_{12}(\xi)\right| ^{2}=Q_{11}(\xi)Q_{22}(\xi)$ a.e. on $\mathbb{R}$, and $Q_{12}(\xi)+Q_{21}(\xi)\geq0$ a.e.; and conversely. \end{lemma} \begin{corollary} \label{QMatrix} These relations imply the following for the matrix $Q$: \begin{enumerate} \item [\hss\llap{\rm1)}]If $Q_{12}(\xi)=0$ then we have the three possibilities: \begin{align*} Q(\xi) & =0\,,\\ Q(\xi) & =\left( \begin{matrix} 1 & 0\\ 0 & 0 \end{matrix} \right) \,,\mbox{ and }\\ Q(\xi) & =\left( \begin{matrix} 0 & 0\\ 0 & 1 \end{matrix} \right) \,. \end{align*} In all those cases, we have $Q(\xi)JQ(\xi)=0$. \item[\hss\llap{\rm2)}] If $Q_{12}(\xi)\neq0$, then $00$, then: \begin{align*} \pi_{+}(b)h & =Q_{11}\pi_{+}(b)h+Q_{12}\pi_{+}(-b)k\,,\\ \pi_{+}(-b)k & =Q_{21}\pi_{+}(b)h+Q_{22}\pi_{+}(-b)k\,, \end{align*} valid for any $\left( \begin{matrix} h\\ k \end{matrix} \right) \in\mathbf{K}_{0}$, and $b\in\mathbb{R}_{+}$. So it follows from Lemma \ref{BasicLemma} again that $\mathcal{D}$ is invariant under $\left\{ \pi_{+}(b)\mid b>0\right\} $, and also under the whole semigroup $\left\{ \pi_{+}(g)\mid g\in S\right\} $ where $\pi_{+}$ is now denoting the corresponding irreducible representation of $G$ on $\mathbf{L}^{2}% (\mathbb{R})$. Let \begin{align} \mathcal{D}_{\infty} & :=\bigvee_{b\in\mathbb{R}}\pi_{+}(b)\mathcal{D}% \,,\label{DInfinitybis}\\ \mathcal{D}_{-\infty} & :=\bigwedge_{b\in\mathbb{R}}\pi_{+}(b)\mathcal{D}\,, \label{DMinusInfinitybis}% \end{align} where $\bigvee$ and $\bigwedge$ denote the lattice operations on closed subspaces in $\mathbf{L}^{2}(\mathbb{R})$, and \[ \left( \pi_{+}(b)f\right) (x)=e^{ibe^{x}}f(x)\,,\quad f\in\mathbf{L}% ^{2}(\mathbb{R}),\;b,x\in\mathbb{R}. \] We may now apply the Lax-Phillips argument to the spaces $\mathcal{D}% _{\pm\infty}$. If $\left( \mathbf{K}_{0}/\mathbf{N}\right) \sptilde$ should be $\neq\{0\}$, then $\mathcal{D}=\{0\}$ by the argument. Since we are assuming $\left( \mathbf{K}_{0}/\mathbf{N}\right) \sptilde \neq\{0\}$, we get $\mathcal{D}=\{0\}$, and as a consequence the following operator graph representation for $\mathbf{K}_{0}$: $\left( \mathbf{K}% _{0}/\mathbf{N}\right) \sptilde=\beta\left( G(L)\right) $ where $G(L)$ is the graph of a closed operator $L$ in $\mathbf{L}^{2}(\mathbb{R})$. Specifically, this means that the linear mapping $\mathbf{K}_{0}/\mathbf{N}% \ni\left( \begin{matrix} h\\ k \end{matrix} \right) +\mathbf{N}\mapsto h$ is well-defined as a linear closed operator. This in turn means that $\mathbf{K}_{0}$ may be represented as the graph of a closable operator in $\mathbf{L}^{2}(\mathbb{R})$ as discussed in the first part of the proof. Hence such a representation could have been assumed at the outset. \begin{remark} \label{BorchersCMP92}In a recent paper on local quantum field theory \cite{Bor92}, Borchers considers in his Theorem II.9 a representation $\pi$ of the $(ax+b)$-group $G$ on a Hilbert space $\mathbf{H}$ such that there is a \emph{conjugate linear} $J$ (that is a period-$2$ antiunitary) such that $J\pi J=\pi\circ\tau$ where $\tau$ is the period-$2$ automorphism of $G$ given by $\tau(a,b):=(a,-b)$. In Borchers's example, the one-parameter subgroup $b\mapsto\pi(1,b)$ has semibounded spectrum, and there is a unit-vector $v_{0}\in\mathbf{H}$ such that $\pi(1,b)v_{0}=v_{0}$, $\forall b\in\mathbb{R}% $. The vector $v_{0}$ is cyclic and separating for a von Neumann algebra $M$ such that $\pi(1,b)M\pi(1,-b)\subset M$, $\forall b\in\mathbb{R}_{+}$. Let $a=e^{t}$, $t\in\mathbb{R}$. Then, in Borchers's construction, the other one-parameter subgroup $t\mapsto\pi(e^{t},0)$ is the modular group $\Delta^{it}$ associated with the cyclic and separating vector $v_{0}$ (from Tomita-Takesaki theory \cite[vol. I]{BrRo}). Finally, $J$ is the corresponding modular conjugation satisfying $JMJ=M^{\prime}$ when $M^{\prime}$ is the commutant of $M$. \end{remark} \begin{acknowledgements} The authors would like to thank Tom Branson, Paul Muhly, Bent \O rsted and Steen Pedersen for helpful discussions, and the referee for constructive suggestions. We would like to thank Brian Treadway for his excellent typesetting of the final version of the paper in \LaTeX. Both authors were also supported in part by the National Science Foundation. \end{acknowledgements} \begin{thebibliography}{999} \label{refpage} \bibitem{Arv84}W.B.~Arveson, \emph{Ten Lectures on Operator Algebras}, CBMS Regional Conf. Ser. in Math., vol.~55, American Mathematical Society, Providence, Rhode Island, 1984 (reprinted 1988), see ch.~4: ``Non-selfadjoint operator algebras and the Feynman-Kac formula''. \bibitem{BO98}F.~Betten, and G.~{\'{O}}lafsson, \emph{Causal compactification and Hardy spaces for spaces of Hermitian type}, preprint, 1998, http://math.lsu.edu/\raisebox{-4pt}{$\widetilde{\hbox to0.5cm{}}$}olafsson. \bibitem{ALMMT96}A.~Ashtekar, J.~Lewandowki, D.~Marolf, J.~Mourao, and T.~Thiemann, \emph{Coherent state transforms for spaces of connections}, J.~Funct. Anal. \textbf{135} (1996), 519--551. \bibitem{Bor92}H.-J. Borchers, \emph{The {CPT}-theorem in two-dimensional theories of local observables}, Comm. Math. Phys. \textbf{143} (1992), 315--332. \bibitem{BrRo}O.~Bratteli and D.W. Robinson, \emph{{O}perator {A}lgebras and {Q}uantum {S}tatistical {M}echanics}, vol.~I, 2nd ed., Springer-Verlag, New York--Berlin, 1987 (1st ed. 1979). \bibitem{MDRF96}M.G. Davidson and R.C. Fabec, \emph{Geometric realizations for highest weight representations}, Representation Theory and Harmonic Analysis (Cincinnati, OH, 1994) (T.~Ton-That, K.I. Gross, D.~St. P. Richards and P.J. Sally, Jr., eds.), Contemporary Mathematics, vol.~191, American Mathematical Society, Providence, Rhode Island, 1995, pp.~13--31. \bibitem{DS}M.G. Davidson and R.J. Stanke, \emph{Ladder representation norms for hermitian symmetric groups}, to appear in Journal of Lie Theory. \bibitem{DyMc70}H.~Dym and H.P. McKean, \emph{Fourier Series and Integrals}, Probability and Mathematical Statistics, vol.~14, Academic Press, New York, 1972. \bibitem{E83}T.J. Enright, \emph{Unitary representations for two real forms of a semisimple Lie algebra: A theory of comparison}, Lie Group Representations, I (College Park, Maryland, 1982/1983), Lecture Notes in Mathematics, vol.~1024, Springer, Berlin--New York, 1983, pp.~1--29. \bibitem{EHW83}T.J. Enright, R.~Howe, and N.~Wallach, \emph{A classification of unitary highest weight modules}, Representation Theory of Reductive Groups (Park City, Utah, 1982) (P. C. Trombi, ed.), Progress in Mathematics, vol.~40, Birkh\"{a}user, Boston, 1983, pp.~97--143. \bibitem{FHO93}J.~Faraut, J.~Hilgert, and G.~\'{O}lafsson, \emph{Spherical functions on ordered symmetric spaces}, Ann. Inst. Fourier (Grenoble) \textbf{44} (1994), 927--965. \bibitem{FK94}J.~Faraut and A.~Kor\'{a}nyi, \emph{Analysis on Symmetric Cones}, Oxford Mathematical Monographs, Clarendon Press, Oxford, 1994. \bibitem{GF89}G.B. Folland, \emph{Harmonic Analysis in Phase Space}, Annals of Mathematics Studies, vol.~122, Princeton University Press, Princeton, New Jersey, 1989. \bibitem{Fr80}J.~Fr\"{o}hlich, \emph{Unbounded, symmetric semigroups on a separable Hilbert space are essentially selfadjoint}, Adv. in Appl. Math. \textbf{1} (1980), 237--256. \bibitem{FOS83}J.~Fr\"{o}hlich, K.~Osterwalder, and E.~Seiler, \emph{On virtual representations of symmetric spaces and their analytic continuation}, Ann. of Math. (2) \textbf{118} (1983), 461--489. \bibitem{GG}I.M. Gel'fand and S.G. Gindikin, \emph{Complex manifolds whose skeletons are semisimple real Lie groups, and analytic discrete series of representations}, Functional Anal. Appl. \textbf{11} (1977), 258--265, translated from {\eightcyr\cyracc Funktsional. Anal. i Prilozhen.} \textbf{11} (1977), 19--27 (Russian). \bibitem{GlJa87}Glimm, J., and Jaffe, A., \emph{{Q}uantum {P}hysics: {A} {F}unctional {I}ntegral {P}oint of {V}iew}, 2nd ed., Springer-Verlag, New York--Berlin, 1987. \bibitem{GoJo83}F.M. Goodman and P.E.T. Jorgensen, \emph{Lie algebras of unbounded derivations}, J.~Funct. Anal. \textbf{52} (1983), 369--384. \bibitem{GR80}I.S. Gradshteyn and I.M. Ryzhik, \emph{Table of Integrals, Series, and Products}, Academic Press, New York--London--Toronto, 1980. \bibitem{BH94}B.C. Hall, \emph{The Segal-Bargmann ``coherent-state'' transform for compact Lie groups}, J.~Funct. Anal. \textbf{122} (1994), 103--151. \bibitem{HCIV}Harish-Chandra, \emph{Representations of semisimple Lie groups}, IV, Amer. J. Math. \textbf{77} (1955), 743--777. \bibitem{HCV}Harish-Chandra, \emph{Representations of semisimple Lie groups}, V, Amer. J. Math. \textbf{78} (1956), 1--41. \bibitem{HCVI}Harish-Chandra, \emph{Representations of semisimple Lie groups}, VI, Amer. J. Math. \textbf{78} (1956), 564--628. \bibitem{HC65}Harish-Chandra, \emph{Discrete series for semisimple Lie groups}, I, Acta Math. \textbf{113} (1965), 241--318; II, Acta Math. \textbf{116} (1966), 1--111. \bibitem{HC70}Harish-Chandra, \emph{Harmonic analysis on semisimple Lie groups}, Bull. Amer. Math. Soc. \textbf{76} (1970), 529--551. \bibitem{Hel62}S.~Helgason, \emph{Differential Geometry and Symmetric Spaces}, Pure and Applied Mathematics, vol.~XII, Academic Press, New York--London, 1962. \bibitem{He78}S.~Helgason, \emph{Differential Geometry, Lie Groups and Symmetric Spaces}, Pure and Applied Mathematics, vol.~80, Academic Press, New York--London, 1978. \bibitem{He84}S.~Helgason, \emph{Groups and Geometric Analysis: Integral Geometry, Invariant Differential Operators and Spherical Functions}, Pure and Applied Mathematics, vol.~113, Academic Press, New York--London, 1984. \bibitem{Hel64}H.~Helson, \emph{Lectures on Invariant Subspaces}, Academic Press, New York--London, 1964. \bibitem{OH94}O.~Hijab, \emph{Hermite functions on compact Lie groups}, I, J.~Funct. Anal. \textbf{125} (1994) 480--492. \bibitem{HiNe93}J.~Hilgert and K.-H. Neeb, \emph{Lie Semigroups and Their Applications}, Lecture Notes in Math., vol.~1552, Springer-Verlag, Berlin, 1993. \bibitem{HO}J.~Hilgert and G.~\'{O}lafsson, \emph{Analytic extensions of representations, the solvable case}, Japan. J. Math. (N.S.) \textbf{18} (1992) 213--289. \bibitem{HO95}J.~Hilgert and G.~\'{O}lafsson, \emph{Causal Symmetric Spaces: Geometry and Harmonic Analysis}, Perspectives in Mathematics, vol.~18, Academic Press, Inc., San Diego, California, 1997. \bibitem{HOO91}J.~Hilgert, G.~\'{O}lafsson, and B.~\O rsted, \emph{Hardy spaces on affine symmetric spaces}, J.~Reine Angew. Math. \textbf{415} (1991), 189--218. \bibitem{TI} T.~Inoue, \emph{Unitary representations and kernel functions associated with boundaries of a bounded symmetric domain}, Hiroshima Math. J. \textbf{10} (1980), 75--140. \bibitem{HJ75}H.A. Jaffee, \emph{Real forms of Hermitian symmetric spaces}, Bull. Amer. Math. Soc. \textbf{81} (1975), 456--458. \bibitem{HJ78}H.A. Jaffee, \emph{Anti-holomorphic automorphisms of the exceptional symmetric domains}, J.~Differential Geom. \textbf{13} (1978), 79--86. \bibitem{Ja83}H.P. Jakobsen, \emph{Hermitian symmetric spaces and their unitary highest weight modules}, J.~Funct. Anal. \textbf{52} (1983) 385--412. \bibitem{Jor75}P.E.T. Jorgensen, \emph{Representations of differential operators on a Lie group}, J.~Funct. Anal. \textbf{20} (1975), 105--135. \bibitem{Jor80}P.E.T. Jorgensen, \emph{Unbounded operators: Perturbations and commutativity problems}, J.~Funct. Anal. \textbf{39} (1980), 281--307. \bibitem{Jor86}P.E.T. Jorgensen, \emph{Analytic continuation of local representations of Lie groups}, Pacific J. Math. \textbf{125} (1986), 397--408. \bibitem{Jor87}P.E.T. Jorgensen, \emph{Analytic continuation of local representations of symmetric spaces}, J.~Funct. Anal. \textbf{70} (1987), 304--322. \bibitem{Jor88}P.E.T. Jorgensen, \emph{{O}perators and {R}epresentation {T}heory: {C}anonical {M}odels for {A}lgebras of {O}perators {A}rising in {Q}uantum {M}echanics}, North-Holland Mathematics Studies, vol.~147, Notas de Matem\'atica, vol.~120, North-Holland, Amsterdam--New York, 1988. \bibitem{JoMo84}P.E.T. Jorgensen and R.T. Moore, \emph{Operator Commutation Relations}, Mathematics and Its Applications, D.~Reidel Publishing Co., Dordrecht--Boston, 1984. \bibitem{JoMu80}P.E.T. Jorgensen and P.S. Muhly, \emph{Selfadjoint extensions satisfying the {W}eyl operator commutation relations}, J.~Analyse Math. \textbf{37} (1980), 46--99. \bibitem{JO97}P.E.T. Jorgensen and G.~\'{O}lafsson, \emph{Unitary representations of Lie groups with reflection symmetry}, J.~Funct. Anal. \textbf{158} (1998), 26--88. \bibitem{Kle78}A.~Klein, \emph{The semigroup characterization of {O}sterwalder-{S}chrader path spaces and the construction of {E}uclidean fields}, J.~Funct. Anal. \textbf{27} (1978), 277--291. \bibitem{KlLa75}A.~Klein and L.J. Landau, \emph{Singular perturbations of positivity preserving semigroups via path space techniques}, J.~Funct. Anal. \textbf{20} (1975), 44--82. \bibitem{KlLa81}A.~Klein and L.J. Landau, \emph{Construction of a unique selfadjoint generator for a symmetric local semigroup}, J.~Funct. Anal. \textbf{44} (1981), 121--137. \bibitem{KlLa83}A.~Klein and L.J. Landau, \emph{From the Euclidean group to the Poincar\'{e} group via Osterwalder-Schrader positivity}, Comm. Math. Phys. \textbf{87} (1982/83), 469--484. \bibitem{KS80}A.W. Knapp and E.M. Stein, \emph{Intertwining operators for semisimple groups}, II, Invent. Math. \textbf{60} (1980), 9--84. \bibitem{KW65}A.~Kor\'{a}nyi and J.A. Wolf, \emph{Realization of Hermitian symmetric spaces as generalized half-planes}, Ann. of Math. (2) \textbf{81} (1965), 265--288. \bibitem{BK95}B.~Kr\"{o}tz, \emph{Plancherel--Formel f\"{u}r Hardy--R\"{a}% ume}, Diplomarbeit, Darm\-stadt, 1995. \bibitem{BK98}B.~Kr\"{o}tz, \emph{Formal dimension for semisimple symmetric spaces}, preprint 1998, submitted. \bibitem{KrNe96}B.~Kr\"{o}tz and K.-H. Neeb, \emph{On hyperbolic cones and mixed symmetric spaces}, J. Lie Theory \textbf{6} (1996), 69--146. \bibitem{KNO98}B.~Kr\"{o}tz, K.-H. Neeb, and G.~\'{O}lafsson, \emph{Spherical representations and mixed symmetric spaces}, Represent. Theory \textbf{1} (1997), 424--461. \bibitem{KNO99}B.~Kr\"{o}tz, K.-H. Neeb, and G.~\'{O}lafsson, \emph{Spherical Functions on Mixed Symmetric Spaces}, preprint 1998, submitted. \bibitem{JL94}J.D. Lawson, \emph{Polar and Ol'shanskii decompositions}, J.~Reine Angew. Math. \textbf{448} (1994), 191--219. \bibitem{LaPh}P.D. Lax and R.S. Phillips, \emph{Scattering theory}, 2nd ed., Academic Press, New York, 1989 (1st ed. 1967). \bibitem{LM75}M.~L\"uscher and G.~Mack, \emph{Global conformal invariance in quantum field theory}, Comm. Math. Phys. \textbf{41} (1975), 203--234. \bibitem{Mac}G.W. Mackey, \emph{Induced Representations of Groups and Quantum Mechanics}, W.A. Benjamin, New York, 1968. \bibitem{Ma82}T.~Matsuki, \emph{Orbits on affine symmetric spaces under the action of parabolic subgroups}, Hiroshima Math.~J. \textbf{12} (1982), 307--320. \bibitem{Nee94}K.-H. Neeb, \emph{Holomorphic representation theory, \textup{{II}}}, Acta Math. \textbf{173} (1994), 103--133. \bibitem{Ne94}K.-H. Neeb, \emph{Realization of general unitary highest weight representations}, to appear in Forum Math. \bibitem{Ne99}K.-H. Neeb, \emph{Holomorphy and Convexity in Lie Theory}, Expositions in Mathematics, de Gruyter, Berlin, 1999. \bibitem{Nel59}E.~Nelson, \emph{Analytic vectors}, Ann. of Math. (2) \textbf{70} (1959), 572--615. \bibitem{Nel64}E.~Nelson, \emph{Feynman integrals and the {S}chr{\"{o}}dinger equation}, J.~Math. Phys. \textbf{5} (1964), 332--343. \bibitem{Nel73}E.~Nelson, \emph{Construction of quantum fields from {M}arkoff fields}, J.~Funct. Anal. \textbf{12} (1973), 97--112. \bibitem{NeSt59}E.~Nelson and W.F. Stinespring, \emph{Representation of elliptic operators in an enveloping algebra}, Amer. J. Math. \textbf{81} (1959), 547--560. \bibitem{'O87}G.~\'{O}lafsson, \emph{Fourier and Poisson transformation associated to a semisimple symmetric space}, Invent. Math. \textbf{90} (1987), 605--629. \bibitem{'O90b}G.~\'{O}lafsson, \emph{Causal symmetric spaces}, Mathematica Gottingensis \textbf{15} (1990). \bibitem{'O90a}G.~\'{O}lafsson, \emph{Symmetric spaces of Hermitian type}, Differential Geom. Appl. \textbf{1} (1991), 195--233. \bibitem{O93}G.~\'{O}lafsson, \emph{Spherical functions on ordered symmetric space; A Harish-Chandra type formula}, in preparation. \bibitem{O98}G.~\'{O}lafsson, \emph{Open problems in harmonic analysis on causal symmetric spaces}, Positivity in Lie Theory: Open Problems (J. Hilgert, J. D. Lawson, K-H. Neeb, and E. B. Vinberg, eds.), Walter de Gruyter, Berlin--New York, 1998. \bibitem{'OO88a}G.~\'{O}lafsson and B.~\O rsted, \emph{The holomorphic discrete series for affine symmetric spaces}, I, J.~Funct. Anal. \textbf{81} (1988), 126--159. \bibitem{'OO88b}G.~\'{O}lafsson and B.~\O rsted, \emph{The holomorphic discrete series of an affine symmetric space and representations with reproducing kernels}, Trans. Amer. Math. Soc. \textbf{326} (1991), 385--405. \bibitem{'OO89a}G.~\'{O}lafsson and B.~\O rsted, \emph{Is there an orbit method for affine symmetric spaces?} The Orbit Method in Representation Theory (M. Duflo, N.V. Pedersen, and M. Vergne, eds.), Birkh\"{a}user, Boston, 1990. \bibitem{OO96}G.~\'{O}lafsson and B.~\O rsted, \emph{Generalizations of the Bargmann transform}, Lie theory and its applications in physics (Clausthal, 1995) (H.-D. Doebner, V. K. Dobrev and J. Hilgert, eds.), World Scientific, River Edge, New Jersey, 1996. \bibitem{OO96b}G.~\'{O}lafsson and B.~\O rsted, \emph{Causal compactification and Hardy spaces}, Report No.~19, 1995/96, Institut Mittag-Leffler, to appear in Trans. Amer. Math. Soc. \bibitem{Ol82}G.I. Ol'shanskii, \emph{Invariant cones in Lie algebras, Lie semigroups, and the holomorphic discrete series}, Functional Anal. Appl. \textbf{15} (1982), 275--285, translated from {\eightcyr\cyracc Funktsional. Anal. i Prilozhen.} \textbf{15} (1981), 53--66 (Russian). \bibitem{OZ95}B.~\O rsted and G.K. Zhang, \emph{Generalized principal series representations and tube domains}, Duke Math.~J. \textbf{78} (1995), 335--357. \bibitem{OsS1}T.~Oshima and J.~Sekiguchi, \emph{Eigenspaces of invariant differential operators on an affine symmetric space}, Invent. Math. \textbf{57} (1980), 1--81. \bibitem{OsS2}T.~Oshima and J.~Sekiguchi, \emph{The restricted root system of a semisimple symmetric pair}, Group representations and systems of differential equations (Tokyo, 1982) (K. Okamoto, ed.), Advanced Studies in Pure Mathematics, vol.~4, North-Holland, Amsterdam--New York, 1984, pp.~433--497. \bibitem{OsSc73}K.~Osterwalder and R.~Schrader, \emph{Axioms for Euclidean Green's functions}, Comm. Math. Phys. \textbf{31} (1973), 83--112. \bibitem{OsSc75}K.~Osterwalder and R.~Schrader, \emph{Axioms for Euclidean Green's functions}, II, Comm. Math. Phys. \textbf{42} (1975), 281--305. \bibitem{Paneitz81}S.~Paneitz, \emph{Invariant convex cones and causality in semisimple Lie algebras and groups}, J.~Funct. Anal. \textbf{43} (1981), 313--359. \bibitem{Paneitz84}S.~Paneitz, \emph{Determination of invariant convex cones in simple Lie algebras}, Ark. Mat. \textbf{21} (1983), 217--228. \bibitem{Ped90}S.~Pedersen, \emph{Anticommuting selfadjoint operators}, J.~Funct. Anal. \textbf{89} (1990), 428--443. \bibitem{Phil}R.S. Phillips, \emph{The extension of dual subspaces invariant under an algebra}, Proc. Internat. Sympos. Linear Spaces (Jerusalem, 1960), Jerusalem Academic Press, Jerusalem, and Pergamon, Oxford, 1961, pp.~366--398. \bibitem{Pou92}N.S. Poulsen, \emph{On $C^{\infty}$-vectors and intertwining bilinear forms for representations of Lie groups}, J.~Funct. Anal. \textbf{9} (1992), 87--120. \bibitem{Pra89}H.~Prado, University of Iowa thesis, 1989. \bibitem{ReSi75}M.~Reed and B.~Simon, \emph{Methods of Modern Mathematical Physics} II: \emph{Fourier Analysis, Self-Adjointness}, Academic Press, New York--London, 1975. \bibitem{S92}S.~Sahi, \emph{Explicit Hilbert spaces for certain unipotent representations}, Invent. Math. \textbf{110} (1992), 409--418. \bibitem{S93}S.~Sahi, \emph{Unitary representations on the Shilov boundary of a symmetric tube domain}, Representation theory of groups and algebras (J.~Adams, R.~Herb, S.~Kudla, J.-S. Li, R.~Lipsman and J.~Rosenberg, eds.), Contemporary Mathematics, vol.~145, American Mathematical Society, Providence, Rhode Island, 1993, pp.~275--286 \bibitem{Sch86}R.~Schrader, \emph{Reflection positivity for the complementary series of $SL(2n,\boldsymbol{C})$}, Publ. Res. Inst. Math. Sci. \textbf{22} (1986), 119--141. \bibitem{Segal}I.E. Segal, \emph{Mathematical Cosmology and Extragalactic Astronomy}, Pure and Applied Mathematics, vol.~68, Academic Press, New York--London, 1976. \bibitem{Seg98}I.E. Segal, \emph{Real spinor fields and the electroweak interaction}, J.~Funct. Anal. \textbf{154} (1998), 542--558. \bibitem{Stanton}R.J. Stanton, \emph{Analytic extension of the holomorphic discrete series}, Amer. J. Math. \textbf{108} (1986), 1411--1424. \bibitem{Ste93}E.M. Stein, \emph{Harmonic Analysis: Real-variable Methods, Orthogonality, and Oscillatory Integrals}, Princeton Mathematical Series, vol.~43, Princeton University Press, Princeton, New Jersey, 1993. \bibitem{Sto51}M.H. Stone, \emph{On unbounded operators in Hilbert space}, J.~Indian Math. Soc. (N.S.) \textbf{15} (1951), 155--192. \bibitem{SzN74}B.~Sz\H{o}kefalvi-Nagy, \emph{{U}nitary {D}ilations of {H}ilbert {S}pace {O}perators and {R}elated {T}opics}, CBMS Regional Conf. Ser. in Math., vol.~19, American Mathematical Society, Providence, Rhode Island, 1974. \bibitem{VR76}M.~Vergne and H.~Rossi, \emph{Analytic continuation of the holomorphic discrete series of a semi-simple Lie group}, Acta Math. \textbf{136} (1976), 1--59. \bibitem{WaI72}G.~Warner, \emph{Harmonic Analysis on Semisimple Lie Groups}, I, Die Grundlehren der mathematischen Wissenschaften, Band 188, Springer-Verlag, New York--Heidelberg, 1972. \bibitem{NW79}N.R. Wallach, \emph{The analytic continuation of the discrete series}, I, Trans. Amer. Math. Soc. \textbf{251} (1979), 1--17; II, 19--37. \bibitem{Wal92}N.R. Wallach, \emph{Real Reductive Groups}, II, Pure and Applied Mathematics, 132-II, Academic Press, Inc., Boston, 1992. \bibitem{Yos91}K.~Yosida, \emph{Lectures on Differential and Integral Equations}, Dover Publications, Inc., New York, translated from the Japanese, reprint of the 1960 translation, Pure and Applied Mathematics, vol.~X, Interscience Publishers, New York--London, 1960. \end{thebibliography} \end{document} ---------------0011091627298--