INSTRUCTIONS The text between the lines BODY and ENDBODY is made of 1180 lines and 45448 bytes (not counting or ) In the following table this count is broken down by ASCII code; immediately following the code is the corresponding character. 30115 lowercase letters 1965 uppercase letters 767 digits 4 ASCII characters 9 5955 ASCII characters 32 14 ASCII characters 33 ! 23 ASCII characters 34 " 76 ASCII characters 35 # 1226 ASCII characters 36 $ 156 ASCII characters 37 % 103 ASCII characters 38 & 41 ASCII characters 39 ' 414 ASCII characters 40 ( 414 ASCII characters 41 ) 5 ASCII characters 42 * 16 ASCII characters 43 + 495 ASCII characters 44 , 150 ASCII characters 45 - 526 ASCII characters 46 . 5 ASCII characters 47 / 39 ASCII characters 58 : 1 ASCII characters 59 ; 5 ASCII characters 60 < 92 ASCII characters 61 = 2 ASCII characters 62 > 4 ASCII characters 64 @ 56 ASCII characters 91 [ 1548 ASCII characters 92 \ 55 ASCII characters 93 ] 108 ASCII characters 94 ^ 244 ASCII characters 95 _ 42 ASCII characters 96 ` 381 ASCII characters 123 { 9 ASCII characters 124 | 381 ASCII characters 125 } 11 ASCII characters 126 ~ BODY %%%%%% FORMAT %%%%%%%%%%%% \magnification\magstep1 \hoffset=0.5truecm \voffset=0.5truecm \hsize=15.8truecm \vsize=23.truecm \baselineskip=14pt plus0.1pt minus0.1pt \parindent=19pt \lineskip=4pt\lineskiplimit=0.1pt \parskip=0.1pt plus1pt %%%%%% Extra Fonts %%%%%%%%% \def\got#1{\bf#1} %% If eufm10 fonts are available, you will get a much more %% gothic effect for Lie algebras by replacing previous line %% with the following: %% \font\got=eufm10 scaled\magstep1 \def\G{\hbox{\got g}} \def\P{\hbox{\got p}} \def\V{\hbox{\got a}} %%%%%% SYMBOLS %%%%%%%%%%%%% \def\A{{\cal A}} \def\tildA{\tilde{\cal A}} \def\tildAnul{\tilde{\cal A}_0} \def\a{\alpha} \def\B{{\cal B}} \def\b{\beta} \def\C{{\cal C}} \def\Co{{\bf C}} \def\d{\delta} \def\D{\Delta} \def\E{{\cal E}} \def\f{\varphi} \def\g{\gamma} \def\H{{\cal H}} \def\K{{\cal K}} \def\L{\Lambda} \def\m{\mu} \def\Na{{\bf N}} \def\o{\omega} \def\O{{\cal O}} \def\p{\pi} \def\Q{\Omega} \def\r{\rho} \def\R{{\cal R}} \def\Re{{\bf R}} \def\s{\sigma} \def\t{\tau} \def\T{\Theta} \def\th{\theta} \def\U{{\cal U}} \def\W{{\cal W}} \def\x{\xi} \def\Z{{\cal Z}} \def\Ze{{\bf Z}} \def\ad{{\rm ad}} \def\Lor{{\cal L}_+^\uparrow} \def\Poi{{\cal P}_+^\uparrow} \def\Spin{\widetilde{\cal P}} \def\imply{\Rightarrow} \def\np{\par\noindent} \def\nu{\char'43} \def\lar{\longrightarrow} \def\ov{\overline} \def\mappazza{\buildrel{\widetilde U}\over\lar} \def\derteqzero{{d\over{dt}}\Bigr\vert_{t=0} \,\,} \def\quot#1#2{\hbox{$\matrix{#1\cr\vphantom{#2}}$ \hskip-.8cm\vbox{\vfill/\vfill}\hskip-12.4cm $\matrix{\vphantom{#1}\cr#2}$}} %%%%%%%% Macros for theorems %%%%%%%% \def\proof{\medskip\noindent{\bf Proof.}\quad} \def\proofof#1{\medskip\noindent{\bf Proof of #1.}\quad} \def\square{\hbox{$\sqcap\!\!\!\!\sqcup$}} \def\endproof{\hskip-4mm\hfill\vbox{\vskip3.5mm\square\vskip-3.5mm}\bigskip} \def\section #1\par{\vskip0pt plus.3\vsize\penalty-75 \vskip0pt plus -.3\vsize\bigskip\bigskip \noindent{\sectionfont #1}\nobreak\smallskip\noindent} \def\subs#1{\medskip\noindent{\it#1}\qquad} \def\claim#1#2\par{\vskip.1in\medbreak\noindent{\bf #1.} {\sl #2}\par \ifdim\lastskip<\medskipamount\removelastskip\penalty55\medskip\fi} \def\rmclaim#1#2\par{\vskip.1in\medbreak\noindent{\bf #1.} { #2}\par \ifdim\lastskip<\medskipamount\removelastskip\penalty55\medskip\fi} %%%%%%%% AUTO REF. %%%%%%%%%%%%% \font\sectionfont=cmbx10 scaled\magstep1 \newcount\REFcount \REFcount=1 \def\numref{\number\REFcount} \def\addref{\global\advance\REFcount by 1} \def\wdef#1#2{\expandafter\xdef\csname#1\endcsname{#2}} \def\wdch#1#2#3{\ifundef{#1#2}\wdef{#1#2}{#3} \else\write16{!!doubly defined#1,#2}\fi} \def\wval#1{\csname#1\endcsname} \def\ifundef#1{\expandafter\ifx\csname#1\endcsname\relax} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \def\ref(#1){\wdef{q#1}{yes}\ifundef{r#1}$\diamondsuit$#1 \write16{!!ref #1 was never defined!!}\else\wval{r#1}\fi} \def\inputreferences{ \def\REF(##1)##2\endREF{\wdch{r}{##1}{\numref}\addref} \REFERENCES} \def\references{ \def\REF(##1)##2\endREF{ \ifundef{q##1}\write16{!!ref. [##1] was never quoted!!}\fi \item{[\ref(##1)]}##2} \section{References}\par\REFERENCES} %%%%%%REFERENCES%%%%%%%% \def\REFERENCES{ \REF(BiWi1)Bisognano J., Wichmann E. , ``{\it On the duality condition for a Hermitian scalar field}", J. Math. Phys. {\bf 16} (1975), 985-1007. \endREF \REF(Borc1)Borchers H.J. , ``{\it The CPT theorem in two-dimensional theories of local observables}", Commun. Math. Phys. {\bf 143} (1992), 315. \endREF \REF(Brow1)Brown K.S., ``{\it Cohomology of groups}" Springer, New York Heidelberg Berlin, 1982. \endREF \REF(BGL)Brunetti R., Guido D., Longo R. , ``{\it Modular structure and duality in conformal quantum field theory}", Commun. Math. Phys., {\bf 156} (1993), 201-219. \endREF \REF(BuSu1)Buchholz, D. Summers, S.J. ``{\it An algebraic characterization of vacuum states in Minkowski space}", preprint DESY 92-119 \endREF \REF(DoLo1)Doplicher S., Longo R., ``{\it Standard and split inclusions of von Neumann algebras}", Invent. Math. {\bf 73} (1984), 493-536. \endREF \REF(FGL)In preparation. \endREF \REF(FrGa1)Gabbiani F., Fr\"ohlich J., ``{\it Operator algebras and Conformal Field Theory}", pre-print ETH Z\"urich, 1992. \endREF \REF(GL1)Guido, D. Longo, R., ``{\it Relativistic invariance and charge conjugation in quantum field theory}", Commun. Math. Phys. {\bf 148} 521-551. \endREF \REF(Haag1)Haag R., {\it Local Quantum Physics}, Springer Verlag, Berlin Heidelberg 1992. \endREF \REF(HiLo1)Hislop P., Longo R., ``{\it Modular structure of the von~Neumann algebras associated with the free massless scalar field theory}", Commun. Math. Phys. {\bf 84} (1982), 84. \endREF \REF(McLa1)Mc Lane S. ``{\it Homology}", Springer, Berlin, 1963. \endREF \REF(Mich1) Michel L., ``{\it Sur les extensions centrales du groupe de Lorentz inhomog\`ene connexe}'', Nucl. Phys. 57 (1965) 356-385. \endREF \REF(Miln1)Milnor J., ``{\it Introduction to algebraic K-theory}", Ann. of Math Studies {\bf 72}, Princeton University Press, Princeton, 1971. \endREF \REF(Moor1)Moore C.C., ``{\it Group extensions and cohomology for locally compact groups. III}'', Trans. Amer. Math. Soc. {\bf 221} (1976) 1. \endREF \REF(Moor2)Moore C.C., ``{\it Group extensions and cohomology for locally compact groups. IV}'', Trans. Amer. Math. Soc. {\bf 221} (1976) 35. \endREF \REF(Simm) Simms D.J., ``{\it Lie groups and quantum mechanics}" Lecture Notes in Math., Vol. 52, Springer, Berlin, 1968. \endREF \REF(StraZs1)Str\u atil\u a S., Zsido L., {\it Lectures on von Neumann algebras}, Abacus press, England 1979. \endREF \REF(Wies1)Wiesbrock H.V., ``{\it A comment on a recent work of Borchers}", to appear in Lett. Math. Phys. \endREF \REF(Wies2)Wiesbrock H.V., ``{\it Conformal quantum field theory and half-sided modular inclusions of von~Neumann algebras}", to appear in Commun. Math. Phys. \endREF \REF(Yngv1)Yngvason J. ``{\it A note on essential duality}", preprint. % available via anonymous ftp in mp_arc@math.utexas.edu \endREF } \inputreferences \font\ftitle=cmbx10 scaled\magstep1 \font\fabs=cmbx10 scaled800 \vskip2truecm \centerline{\ftitle GROUP COHOMOLOGY, } \medskip\centerline{\ftitle MODULAR THEORY} \medskip\centerline{\ftitle AND SPACE-TIME SYMMETRIES} \bigskip\bigskip \centerline{R. Brunetti$^1$\footnote{$^*$}{ Supported in part by MURST and CNR-GNAFA.} \footnote{$^\bullet$}{Supported in part by INFN, sez. Napoli and by Fondazione ``A. della Riccia"}, D. Guido$^{2*}$ and R. Longo$^{23*}$} \footnote{}{E-mail\ brunetti@na.infn.it, guido@mat.utovrm.it, longo@mat.utovrm.it } \vskip1.truecm \item{$(^1)$} Department of Physics, Syracuse University \par 201 Physics Building, Syracuse, NY 13244-1130. \par \item{$(^2)$} Dipartimento di Matematica, Universit\`a di Roma ``Tor Vergata'' \par Via della Ricerca Scientifica, I--00133 Roma, Italia. \item{$(^3)$} Accademia Nazionale dei Lincei, \par via della Lungara 10, I--00165 Roma, Italia \vskip 3cm\noindent {\bf Abstract. } The Bisognano-Wichmann property on the geometric behavior of the modular group of the von Neumann algebras of local observables associated to wedge regions in Quantum Field Theory is shown to provide an intrinsic sufficient criterion for the existence of a covariant action of the (universal covering of) the Poincar\'e group. In particular this gives, together with our previous results, an intrinsic characterization of positive-energy conformal pre-cosheaves of von Neumann algebras. To this end we adapt to our use Moore theory of central extensions of locally compact groups by polish groups, selecting and making an analysis of a wider class of extensions with natural measurable properties and showing henceforth that the universal covering of the Poincar\'e group has only trivial central extensions (vanishing of the first and second order cohomology) within our class. \section{Introduction} \par In this paper we discuss the problem of characterizing the existence of a Poincar\'e covariant action for a net of local observable algebras in terms of net-intrinsic algebraic properties. In other words, given a state with a Reeh-Schlieder cyclicity property, we look for a condition that ensures it to be a relativistic vacuum. The work done by Bisognano and Wichmann [\ref(BiWi1)] during the seventies about the essential duality has shown the geometrical character of the Tomita-Takesaki modular operators associated with von Neumann algebras generated by Wightman fields with support based on wedge regions and the vacuum vector. While an algebraic version of the Bisognano-Wichmann theorem in the Poin\-ca\-r\'e covariant case is not yet available, there are general results in this direction given by the recent work of Borchers [\ref(Borc1)], see also [\ref(BuSu1),\ref(Wies1),\ref(Yngv1)] for related results. Based on the result of Borchers, there is a complete algebraic analysis in [\ref(BGL)] and [\ref(FrGa1)] showing that, for a conformally invariant theory, the modular groups associated with the von~Neumann algebras of double cones are associated with special one-parameter subgroups of the conformal group leaving invariant the given double cone (cf. [\ref(HiLo1)]). Applications of the Bisognano-Wichmann theorem to the analysis of the relations between the space-time symmetries and particle-antiparticle symmetry in quantum field theory are contained in [\ref(GL1)]. Here we show how to reconstruct the Poincar\'e group representation from the net-intrinsic property of {\it modular covariance} (assumption 2.1), i.e. the modular operators associated with the algebras of the wedge regions act geometrically as boosts. In our previous paper [\ref(BGL)] we showed in particular that, for a Poincar\'e covariant net, the modular unitaries associated with wedge regions should necessarily coincide with the boost unitaries in the given representation if the modular covariance and the split property are assumed. Here, we do not have any given unitary representation of the Poincar\'e group from the start. Instead we recover a canonical representation of the Poincar\'e group, which is unique if the split property holds, just by using the modular unitaries corresponding to all wedge algebras. This step forward makes two significant changes. The first concerns physics, since our result shows that the net itself contains all the space-time symmetry informations, and that the modular covariance property gives rise to a canonical representation even when (non-split case) it is not unique. The second concerns mathematics, providing a new application of the group cohomology describing central extensions of groups. As is well known, central extensions of groups appear naturally in physics, as shown for example by the Wigner theorems [\ref(Simm)] on the (anti-)unitary realization of the symmetries in quantum physics. The class of results in this context which is relevant for our purposes concerns the triviality of (suitable classes of) central extensions of the universal covering of the Poincar\'e group. One of the first results of this type is due to Michel [\ref(Mich1)], but his hypotheses seems unapplicable to our problem. Another result in this direction is a particular case of the theory of Moore [\ref(Moor1),\ref(Moor2)] on central extensions of locally compact groups via polish groups. The extensions described by Moore are continuous and open, and his theorem on the universal extension concerns extensions by closed subgroups of $\U(\H)$, the group of unitary operators on a separable Hilbert space $\H$. As a main step in our analysis, we study a class of extensions with natural measurable properties, which we call weak Lie central extensions. These extensions are given by a (not necessarily closed) subgroup $G$ of $\U(\H)$, and we assume that there are continuous one-parameter subgroups in $G$ which correspond to a set of Lie-algebra generators (see section 2 for the precise definition). We show that weak Lie central extensions of a simply connected perfect Lie group are trivial, and this applies to the universal covering of the Poincar\'e group. Our proof strongly relies on Moore theory. In fact, even though the continuity properties of the central extensions described by Moore are too restrictive for our purposes, we prove that his theorem on the structure of the first and second order (measurable) cohomology of a connected perfect Lie group (see [\ref(Moor2)] or theorem 1.3) can be generalized to the weak Lie central extensions (see theorem 1.7). The cases of the Poincar\'e group and of the conformal group find application in section 2. Finally we mention that our results, together with theorem 2.3 in our previous paper [\ref(BGL)], give a complete characterization of positive-energy conformal pre-cosheaves as those where the modular groups associated with double cones act geometrically as expected (see corollary 2.9). The plan of the paper is the following: in the first section we review some known facts about cohomology and extensions of groups, recall elements of Moore theory and provide a generalization of results needed in the following. In the second section we present our results about the generation of the unitary representation of the group of space-time symmetries by modular groups, in case of local algebras on a separable Hilbert space. The last section contains an outlook. \section 1. Group extensions and cohomology \par We begin by shortly reviewing basic elements of central extensions of groups (see [\ref(McLa1),\ref(Brow1)]). We first deal with {\it algebraic central extensions}, i.e. with short exact sequences of groups $$ 1\to A\buildrel i\over\to G\buildrel\pi\over\to P\to 1\eqno(1.1) $$ where $P$ is the group to extend and $A$ is a central subgroup of $G$. The pair $(G,\pi)$ determines the extension, and we often refer to it as a central extension of $P$ (by $A = ker(\pi)$). \par Two extensions $(G_1,\pi_1)$, $(G_2,\pi_2)$ are called {\it equivalent} if there is an isomorphism between $G_1$ and $G_2$ such that the following diagram commutes: $$ \matrix{ & & & & G_1 & & & & \cr & & &\nearrow& &\searrow& & & \cr 1&\to&A& &\updownarrow & &P&\to&1\cr & & &\searrow& &\nearrow& & & \cr & & & & G_2 & & & & \cr} \eqno(1.2) $$ An extension of $P$ via $A$ is {\it trivial} if it is equivalent to the direct product $(A\times P,\pi)$ where $\pi$ is the projection onto $P$, i.e. sequence (1.1) splits. A {\it section} of an extension $(G,\pi)$, i.e. a map $s:P\to G$ such that $\pi\cdot s=id_P$, determines an identification of $G$ with $A\times P$ as sets given by $g\to(s(\pi(g))^{-1}g,\pi(g))$. Then, the multiplication rule on $A\times P$ is given by $$ (a,p)\cdot(b,q)=(ab\o(pq)^{-1},pq),\eqno(1.3) $$ where $\o(p,q) = s(q)s(pq)^{-1}s(p)$ satisfies the 2-{\it cocycle} condition $\d_2\o=0$, where $\d_2$ is defined in $(1.4)$. Conversely a 2-cocycle gives an associative multiplication on $A\times P$, thus defining an extension. Two cocycles $\o_1$, $\o_2$ give rise to equivalent extensions {\it iff} there is a group automorphism of $A\times P$ which is compatible with diagram (1.2) and intertwines the products given by $\o_1$ and $\o_2$. This means that there is a map $\f:P\to A$ such that the morphism between $(A\times P,\o_1)$ and $(A\times P,\o_2)$ is given by $(a,p)\mapsto(a\f(p),p)$, and $\o_1=\o_2\cdot\d_1\f$, with $\d_1$ defined in (1.4), i.e. $\o_2$ differs by $\o_1$ of a coboundary, showing that equivalence classes of extensions are indeed cohomology classes. \par The $n$-{\it cochains} $C^n(P,A)$ of $P$ with values in $A$ are maps from $P^n$ to $A$, and the $n$-th {\it coboundary map} $\d_n:C^n(P,A)\to C^{n+1}(P,A)$ is given by $$ \eqalign{ \d_n f(p_1,\dots,p_{n+1}) &=f(p_2,\dots,p_{n+1})-f(p_1p_2,p_3,\dots,p_{n+1})\cr &+f(p_1,p_2p_3,\dots,p_{n+1})+\dots\cr &+(-1)^nf(p_1,\dots,p_np_{n+1})-(-1)^nf(p_1,\dots,p_n)} \eqno(1.4) $$ and verifies $\d_{n+1}\d_n=1$. \par The range of $\d_{n-1}$ is denoted by $B^n(P,A)$, the kernel of $\d_n$ by $Z^n(P,A)$, and their quotient by $H^n(P,A)$. Since $C^n(P,A)$ is an abelian group with respect to the pointwise multiplication, $H^n(P,A)$ is a group too. \par The above discussion shows that equivalence classes of central extensions of $P$ via $A$ are in $1-1$ correspondence with elements of $H^2(P,A)$, and therefore form a group. Note that the cocycle equation for 1-cochains means exactly that $Z^1(P,A)\equiv H^1(P,A)$ is the group of homomorphisms from $P$ to $A$, therefore the vanishing of the first cohomology group for all abelian $A$ is equivalent to the fact that $P$ coincides with its commutator $[P,P]$, i.e. $P$ is {\it perfect}. \par The groups $G$ for which all central extensions split in a unique way, namely $H^n(P,A) =0, n=1,2$ for all A as above, are called {\it algebraically simply connected} (see [\ref(Miln1),\ref(Brow1)]). \par A {\it universal central extension} $(E,\s)$ of $P$ is a central extension such that for each central extension $(G,\pi)$ of $P$ there is a homomorphism $U:E\to G$ such that the following diagram commutes: $$ \def\normalbaselines{\baselineskip20pt \lineskip3pt \lineskiplimit3pt } \def\mapse#1{\searrow\hskip-.3cm {}^#1} \def\mapright#1{\smash{ \mathop{\longrightarrow}\limits^{#1}}} \def\mapdown#1{\Big\downarrow \rlap{$\vcenter{\hbox{\hskip-.1cm $\scriptstyle#1\/$}}$}} \matrix{ E &\mapright U&G\cr &\mapse{\s}&\mapdown{\pi}\cr &&P\cr}\ .\eqno(1.5) $$ If a universal central extension exists, it is unique (up to equivalence) and the extensions of $P$ via $A$ may be easily described. Indeed consider the following commutative diagram: $$ \def\normalbaselines{\baselineskip20pt \lineskip3pt \lineskiplimit3pt } \def\mapright#1{\smash{ \mathop{\longrightarrow}\limits^{#1}}} \def\mapdownu#1{\Big\downarrow \rlap{$\vcenter{\hbox{$\scriptstyle#1$}}$}} \def\mapdownd#1{\Big\Vert \rlap{$\vcenter{\hbox{$\scriptstyle#1$}}$}} \matrix{ &1&\mapright{}&A&\mapright{}&E\times A&\mapright{id_E\cdot1_A}&E&\mapright{}&1&\cr &&&\mapdownd{}&&\mapdownu{U\cdot id_A}&&\mapdownu\sigma &&&\cr &1&\mapright{}&A&\mapright{}&G&\mapright\p &P&\mapright{}&1&.\cr}\eqno(1.6) $$ By a classical diagram chasing argument, it is easy to see that $U\cdot id_A$ is surjective, and that the restriction of the projection $id_E\cdot1_A$ to $\ker(U\cdot id_A)$ is an isomorphism onto $S:=\ker(\s)$. Moreover the embedding of $S$ in $E\times A$ via the preceding isomorphism has the form $id_S\times\psi$, where $\psi$ is a morphism from $S$ to $A$. As a consequence, there is an isomorphism $$ \quot{E\times A}S \rightarrow G\eqno(1.7) $$ where $S$ is embedded into $E\times A$ as described, therefore the extension $(G,\pi)$ is described in terms of this embedding, i.e. in terms of the morphism $\psi$, which implies that $H^1(S,A)\simeq H^2(P,A)$. \par A central extension $(E,\s)$ of $P$ where $E$ is algebraically simply connected is the universal central extension of $P$. Indeed fix an extension $(G,\pi)$ of $P$, choose a section $s$ of $\pi$ and consider the 2-cocycle $\o$ defined in equation (1.3). Then the map $$ \matrix{ \tilde\o&:E\times E&\to&A\cr &(g,h)&\mapsto&\o(\s(g)\s(h))}\eqno(1.8) $$ is indeed a cocycle in $Z^2(E,A)$. Since $H^2(E,A)$ is trivial, there is a 1-cochain $\f$ s.t. $\d_1\f=\tilde\o$. Then, setting $$ U(g)=s(\s(g))\f(g)^{-1},\qquad g\in E\eqno(1.9) $$ we easily get that $U$ is a homomorphism from $E$ to $G$ and makes diagram (1.5) commutative. We note that since $\s$ is surjective and $E$ is perfect, necessarily $P$ is perfect too. \par If $P$ is a topological group, extensions with given topological properties are of interest. A theory in this sense has been developed by Moore in a series of papers [\ref(Moor1),\ref(Moor2)]. We recall that a topological group $A$ is called {\it polish} if its topology may be obtained by a separable complete metric (that can be chosen compatible with the uniformities of $A$). Moore considers {\it topological central extensions} of a locally compact group $P$ via a polish group $A$, i.e. exact sequences (1.1) such that $i$ is a homeomorphism into its image and $\pi$ is continuous and open. He denotes by $Ext(P,A)$ the equivalence classes of topological extensions of $P$ via $A$, where the isomorphism in diagram (1.2) is asked to be a homeomorphism. In analogy with the algebraic case, topological cohomology groups are to be defined in such a way that the identifications of $Ext(P,A)$ with $H^2_{top}(P,A)$, and of $H^1_{top}(P,A)$ with the continuous homomorphisms from $P$ to $A$ hold. Moore proves this to be the case if cochains are Borel measurable. With this hypothesis, the previously mentioned identifications hold, cohomology groups are polish groups and all usual functorial properties are satisfied. From now on we consider topological cohomology groups only, therefore we drop the subscript $_{top}$. \par Let us call {\it unitary group} a subgroup of $\U(\H)$, the group of the unitary operators on a separable Hilbert space $\H$ equipped with the weak topology. It turns out that closed (i.e. complete) unitary groups are polish. Concerning universal extensions, Moore deals with {\it unitary topological extensions}, namely $A$ in (1.1) is a closed unitary group. We shall say that $E$ is $\U$-{\it simply connected} if all central unitary extensions split in a unique way, that is to say $H^1(E,A)$ and $H^2(E,A)$ vanish for all closed abelian unitary groups $A$. \claim{1.1 Theorem [\ref(Moor2)]} Let the locally compact group $P$ have a universal topological central extension $(E,\s)$ with $E$ a polish $\U$-simply connected group. Then $(E,\s)$ is a universal central extension for the class of all unitary central extensions, namely for every topological central extension $(G,\pi)$ via a closed unitary group $A$ there exists a continuous map $U:E\to G$ s.t. the diagram (1.5) commutes. As a consequence, the isomorphism (1.7) is a homeomorphism and $H^1(S,A)$ is isomorphic to $H^2(P,A)$, where $S:=\ker(\s)$. \par According to theorem 1.1, we call the $\U$-covering of $P$ the (unique) central extension of $P$ via a simply connected group (if it exists). We need however a generalization of theorem 2 to a larger class of extensions. \claim{1.2 Theorem} Let $P$ be a locally compact group which admits a $\U$-covering $(E,\s)$. Let $(G,\pi)$ be an algebraic central extension of $P$ with $G$ a unitary group and $A:=\ker(\pi)$ a topological subroup of $G$, closed as unitary group. If $\pi$ has a Borel section $s$, then $\pi$ splits on $E$, i.e there exists a continuous map $U:E\to G$ s.t. diagram (1.5) commutes. \np As a consequence, the group $\quot{E\times A}S$ is polish and isomorphism (1.7) is continuous. \np If, in addition, the map $\pi$ is continuous, then $G$ is a closed unitary group and the extension $(G,\pi)$ is a topological extension in the sense of definition 1.1. \par The last statement of the theorem shows that the difference between the extensions in theorems 1.1 and 1.2 lies in the continuity of the map $\pi$. \proof Since $s$ is a measurable section, the 2-cocycle defined in (1.3) is measurable too, and it is an element of $Z^2(P,A)$. Then we define $\tilde\o$ as in (1.8), and since $\s$ is continuous, $\tilde\o$ turns out to be an element of $Z^2(E,A)$. Since $A$ is a closed unitary group and $E$ is $\U$-simply connected, $\tilde\o$ is indeed a coboundary, i.e. there exists $\f\in C^1(E,A)$ s.t. $\tilde\o=\d_1\f$. Then $U$ defined by equation (1.9) is a homomorphism from $E$ to $G$, it is measurable by construction, and makes diagram (1.5) commute. By proposition $5(a)$ in [\ref(Moor1)], $U$ is indeed continuous. \np Now we come back to diagram (1.6). Surjectivity of $U\cdot id_A$ and the existence of the isomorphism $j:\ker(U\cdot id_A)\to S$ follow by algebraic reasons. Then we observe that $j$ is simply the restriction to $\ker(U\cdot id_A)$ of the projection $id_E\cdot1_A$, which is continuous and open, therefore $j$ is a homeomorphism. Then $\quot{E\times A}S$ is a polish group and the isomorphism (1.7) is continuous. \np If the map $\pi$ is continuous, then $$ g\in G\to\left(\s^{-1}\circ\pi(g)\ ,\ \left[U\circ\s^{-1}\circ\pi(g)\right]^{-1}\ g\right)\in E\times A $$ gives a map from the subsets of $G$ to the subsets of $E\times A$ for which the preimage of an open set is open. On the other hand, composing such a map with the projection on $\quot{E\times A}S$, we get a map of points which is the inverse of the isomorphism (1.7). Such isomorphism is therefore a homeomorphism, and all other properties follow. \endproof Note that, in the topological setting, the only property that $[P,P]$ is dense in $P$ is needed for the existence of a universal central extension. A remarkable result of Moore shows that a perfect connected group $P$ admits the $\U$-covering. Moreover, if $P$ is a Lie group, central extensions may be completely described in terms of the Lie algebra cohomology and the fundamental group $\pi_1(P)$. \par We recall that when $\P$ is a Lie algebra, a central extension of $\P$ via a vector space $\V$ is an exact sequence $0\to\V\to\G\buildrel\pi\over\to\P\to0$ where $\V$ is a central Lie subalgebra of $\G$ and $\pi$ is a Lie algebra homomorphism. Extensions are still described in terms of 2-cocycles, and equivalence classes of extensions in terms of 2-cohomology classes. We denote by $H^n(\P,\V)$ the cohomology groups of $\P$ with respect to $\V$. \par \claim{1.3 Theorem [\ref(Moor2)]} The $\U$-covering $(E,\s)$ of a perfect connected Lie group $P$ is a perfect Lie group, and $$ \ker(\s)\simeq\pi_1(P)\times H^2(\P,\Re) $$ where $\P$ is the Lie algebra of $P$. If $H^2(\P,\Re)=0$, $E$ coincides with the universal covering of $P$. \par In the case of Lie groups, the existence of a measurable section in theorem 1.2 may be replaced by a natural condition. \claim{1.4 Definition} Let $(G,\pi)$ be an algebraic extension of the Lie group $P$ where $G$ is a unitary group. We say that $\pi$ is a weak Lie extension if there exists a set $\{L_1\dots L_n\}$ of generators of $\P$ as a Lie algebra ad a corresponding set $\{V_1(t),\dots,V_n(t)\}$ of strongly continuous 1-parameter groups in $G$ such that $$ \pi(V_i(t))=\exp(tL_i),\qquad i=1,\dots,n. $$ where we have denoted by $\P$ the Lie algebra of $P$. \par \claim{1.5 Lemma} Let $(G,\pi)$ be a central weak Lie extension of the Lie group $P$. Then there is a finite set of strongly continuous 1-parameter groups in $G$ such that the generators of their images in $P$ are a basis for $\P$ as a vector space. \par \proof Let us consider the map $v$ which associates to a given set of one parameter groups in $P$ the (finite dimensional) vector space spanned by their generators. \np We observe that given $A,B\in\P$, the commutator $[A,B]$ belongs to $$ X:=v(\{t\to\exp(sA)\ \exp(tB)\ \exp(-sA):s\in\Re\}). $$ Indeed, $X$ is spanned, by definition, by the range of the Lie algebra valued function $s\to\exp(sA)\ B\ \exp(-sA)$, whose derivative in 0 is $[A,B]$. \np As a consequence, by finite dimensionality, we may find a finite set $\{s_i^j\}$ such that the vector space $$ v\circ\pi\left\{t\to V_{h_1}(s_1^j)\cdot\dots\cdot V_{h_p}(s_p^j) \cdot V_{h_0}(t)\cdot V_{h_p}(s_p^j) \cdot\dots\cdot V_{h_1}(s_1^j):j=1,2,\dots\right\} $$ contains the element $[L_{h_1},[\dots[L_{h_p},L_{h_0}]\ ]$, where the $V_i$ are described in definition 1.4. Then the thesis easily follows. \endproof \claim{1.6 Proposition} Let $(G,\pi)$ be a central weak Lie extension of the connected Lie group $P$, with $G$ a unitary group. Then there exists a Borel measurable section $S:P\to G$ of the extension $\pi$. \par \proof According to the previous lemma, we may suppose that $\{L_1,\dots L_n\}$ is a basis for $\P$ as a vector space. \np The map $\a :\Re^n\to P$ given by $ \a (t_1,\dots t_n)=\prod_{i=1}^n \exp(t_iL_i) $ has non trivial Jacobian at the origin because $\{L_1\dots L_n\}$ is a basis. Therefore there there exist two open sets, $U\subset\Re^n$ and $V\equiv \a (U)\subset P$, where $U$ is a neighborhood of zero in $\Re^n$ and, consequently, $V$ is a neighborhood of the identity element in $P$, such that the map $\a \,:\, U\to V$ is a diffeomorphism. \np Now we define the (strongly) continuous map $\b:\Re^n\to G$ given by $\b(t_1,\dots t_n)=\prod_{i=1}^n V_i(t_i)$, and observe that the map $S_0:=\b\ \a^{-1}:V\to G$ is a (strongly) continuous section of $\pi$ on $V$, i.e. $\pi\cdot S_0\equiv id_V$. \np Since $V$ is an open neighborhood of the origin and $P$ is a connected group, then $P$ is algebraically generated by $V$, i.e., each element of $P$ can be written as a finite product of elements in $V$. \np Now we consider the open covering $\bigcup_{g\in P} gV= P.$ Since $P$ is $\sigma$-compact, we may extract a countable sub-covering, $$ P =\bigcup_{k\in\Na}g_k V. $$ Finally, define the measurable partition of $ P $ given by, $$ \eqalign{ A_1 &= g_1 V\cr A_{n+1}&=\, g_{n+1} V\cap\left(\bigcup_{k=1}^{n} g_k V\right )^c \quad n\in\Na.\cr } $$ \np For each $n\in\Na$, we may write $g_n =v_{n,1} \cdots v_{n,m_n}$, $v_{n,k}\in V$, hence, since $A_n \subset g_n V$, any element of $A_n$ may be written as $ v_{n,1} \cdots v_{n,m_n} v$, when $v$ varies in $V$. Then, we define $$ \matrix{ S_n :&\quad A_n &\longrightarrow&G&\cr &&&&\cr &v_{n,1} \cdots v_{n,m_n} v &\longmapsto&S_0 (v_{n,1}) \cdots S_0 (v_{n,m_n})S_0 (v)&\quad.\cr } $$ As before, $S_n$ is a strongly continuous section of $\pi$ on $A_n$. Therefore it is easy to understand that the map $S=\sum_{n\in\Na}\chi_{A_n}\,\,S_n$, where $\chi_B$ is the characteristic function of $B$, is the desired measurable section. \endproof Making use of theorems 1.2 and 1.3 and proposition 1.6, we can prove the following theorem on weak Lie extensions. We remark that no condition on the closure of $A$ is needed. \claim{1.7 Theorem} Let $P$ be a perfect connected Lie group. Then all central weak Lie extensions $(G,\pi)$ of $P$, with $G$ a unitary group, split on the $\U$-covering $(E,\s)$, i.e. there exist a continuous homomorphism $U:E\to G$ such that diagram (1.5) commutes. \par \proof Let $\tilde G$ be the group algebraically generated by $G$ and $\ov A$, where $\ov A$ is the weak closure of $A$ in $\U(H)$, and extend $\pi$ to a morphism $\tilde\pi$ in such a way that $\tilde\pi|_{\ov A}=0$. Then, applying proposition~1.6 and theorem~1.2 to the extension $(\tilde G,\tilde\pi)$ we get a continuous map $U:E\to\tilde G$ such that $\tilde\pi\cdot U=\s$. \np By theorem 1.3 $E$ is perfect, therefore $U(E)=U([E,E])\subseteq[\tilde G,\tilde G]$. Since $\ov A$ commutes with $G$ and $\tilde G$, equality $[\tilde G,\tilde G]\equiv[G,G]$ holds. As a consequence, $Rg(U)\subseteq G$, and the proof is concluded. \endproof We recall that the (4-dimensional) Poincar\'e group $\Poi$ is perfect and its Lie algebra $\wp$ satisfies $H^2(\wp,\Re)=0$. Therefore, next corollary immediately follows. \claim{1.8 Corollary} Let $(G,\pi)$ be a central weak Lie extension of the Poincar\'e group $\Poi$ where $G$ is a unitary group. Then there exists a strongly continuous unitary representation $U$ of the universal covering $\Spin$ of $\Poi$ such that $\pi\cdot U=\s$, where $\s$ is the covering map, and there is a continuous isomorphism $$ \quot{\Spin\times A}{\Ze_2}\to G $$ where $\Ze_2$ is a suitable order two central subgroup of $\Spin\times A$. \par \proof The corollary easily follows by theorems 1.2, 1.3 and 1.7, the previous observations on the Lie algebra $\wp$ and the fact that the kernel of the covering map from $\Spin$ to $\Poi$ is $\Ze_2$. \endproof \section 2. Modular covariance and the reconstruction of space-time symmetries. \par In this section we study how the unitary representation of the group of the space-time symmetries of a Quantum Field Theory [\ref(Haag1)] may be generated by the modular unitaries associated with the algebras of a suitable class of regions. Local Quantum Theories are described by a local pre-cosheaf of von~Neumann algebras (see [\ref(GL1)]), i.e by a map $$ \A:\O\to\A(\O),\qquad \O\in\K $$ where $\K$ is the family of the double cones in the Minkowski space $M$, such that $$ \eqalign{ \O_1\subset\O_2&\imply\A(\O_1)\subset\A(\O_2) \hskip1.5cm {\rm(isotony)}\cr \A(\O)&\subset\A(\O')'\hskip3.cm {\rm(locality)}.\cr} $$ \par Local algebras are supposed to act on a {\it separable} Hilbert space $\H$, and $\A_0$ denotes the $C^*$-algebra of quasi-local observables generated by the local algebras. There is a vector $\Q$ (vacuum) which is cyclic for all of them. \par The pre-cosheaf is extended by additivity to general open regions of $M$. \par As a consequence of locality, $\Q$ turns out to be cyclic and separating for the algebras associated with all non empty open regions whose complement has non-empty interior (Reeh-Schlieder property). \par We recall that a {\it wedge} region is any Poincar\'e transformed of the region $W_1:=\{x\in\Re^n:|x_0|2$. Then there exists a vacuum preserving, positive energy, $\A$-covariant, unitary representation $U$ of $\Spin$ which is canonically determined by the property $$U(\tilde\L_W (t))=\D_W^{it},\qquad W\in\W$$ \par In order to prove Theorem $2.3$ we need some results about the structure of $\Poi$ (Lemma $2.5$), and about the group generated by modular operators associated with wedge regions. \claim{2.4 Proposition} The generators of the boost transformations generate the Lie algebra of the Poincar\'e group. As a consequence, the boosts algebraically generate $\Poi$. \par \proof The first statement is well known, and the second follows by the connectedness of $\Poi$. \endproof \claim{2.5 Lemma} Let $\A$ be a local pre-cosheaf on $M$ satisfying assumption $2.1$, and consider the set $$ H=\{g\in\Poi:\A(g\O)=\A(\O)\quad\forall\O\in\K\} $$ Then, H is a normal subgroup of $\Poi$ and either \item{$(a)$} $H=\{1\}$ \np or \item{$(b)$} for each $\O\in\K$, $\A(\O)$ is equal to $\A_0$ and is a maximal abelian subalgebra of $B(\H)$. \par \proof $H$ is clearly a group, and we prove that it is indeed normal in $\Poi$. In fact, by modular covariance, $\forall g\in H$ we get $$ \A(\L_W (t) g\L_W (-t)\O)=Ad\D_W^{it}\cdot\A(g\L_W (-t)\O)= Ad\D_W^{it}\cdot\A(\L_W (-t)\O)= \A(\O) $$ therefore, since the boosts generate $\Poi$ (proposition~2.4), $H$ is normal. \np Now we prove that each non trivial normal subgroup $K$ of $\Poi$ contains the translation subgroup $T=\{T(a):a\in\Re^n\}$. Let $1\not=g\in K$. If $g$ is not a translation then, by normality, $$ K\ni g T(a) g^{-1} T(-a)=T(g(a))T(-a)\in T $$ i.e. $K\cap T$ contains at least a non trivial element $g$. Again by normality, $K\cap T$ is globally invariant under the action of the Lorentz group. Since the (minimal) Lorentz-invariant subsets of $\Re^n$ are the future light cone, its surface, the past light cone, its surface, and the complement of the preceding ones, and since any of these generates $\Re^n$ as a group, then $K\supset T$. In particular, we proved that if $H$ is non trivial it contains $T$. In this case, for each $\O\in\K$ we can find a translation $T(a)$ such that $T(a)\O$ is spatially separated by $\O$, and therefore, by locality, $\A(\O)$ is abelian. Then it is easy to see that all local algebras coincide, and they are maximal abelian sub-algebras [\ref(StraZs1)] because the vacuum is cy\-clic. \endproof \rmclaim{Remark} If alternative $(b)$ of Lemma 2.5 holds, then Theorem 2.3 is true by taking the trivial representation of the universal covering of the Poincar\'e group. In the rest of the paper we shall discuss alternative $(a)$, i.e. we shall suppose $H=\{1\}$. \par \claim{2.6 Lemma} Let $G$ be the subgroup of $\U(\H)$ which is algebraically generated by the modular groups of the algebras $\A(W)$, $W\in\W$. Then $G$ is an algebraic central extension of the Poincar\'e group $\Poi$. \par \proof First we prove that the map $\pi$ which satisfies $$ \p (\D^{it}_{W})=\L_{W}(t)\qquad W\in\W,\quad t\in\Re\eqno(2.1) $$ extends to a well defined surjective group homomorphism $\p :G\to\Poi$. \np Indeed let $\D_1^{it_1}\cdot\dots\cdot \D_n^{it_n}=1$ be a non trivial identity in $G$, where $\D_i$ is the modular operator of $\A(W_i)$ and $\{W_i , i=1,\dots,n\}$ is any collection of $n$ wedges in $\cal W$. Then $$ \A(\O)= \D_1^{it_1}\cdot\dots\cdot\D_n^{it_n}\A(\O) \D_n^{-it_n}\cdot\dots\cdot\D_1^{-it_1}= \A(\p (\D_1^{it_1}\cdot\dots\cdot\D_n^{it_n})\O), \quad\forall\O\in\K. $$ From the Lemma 2.5 we get $\p(\D_1^{it_1}\cdot\dots\cdot \D_n^{it_n})=1$. As a consequence $\p$ is a well-defined homomorphism. Since the boosts algebraically generate $\Poi$ (proposition 2.4), $\pi$ is surjective. \np Now we observe that if $U\in\ker(\p)$, then $\ad U(\A(W))=\A(\p(U)W)=\A(W)$ for each wedge $W$, therefore $U$ commutes with the modular group of any wedge algebra [\ref(StraZs1)], and hence with any element in $G$. As a consequence, the exact sequence $1\lar\ker(\p)\lar G\lar\Poi\lar 1$ gives the announced central extension. \endproof \proofof{Theorem 2.3} Equation (2.1) and proposition 2.4 implies that the extension $\pi$ described in lemma~2.6 is a weak Lie extension. Since dim$(M)>2$, the Poincar\'e group is perfect and its Lie algebra has trivial second cohomology. Therefore, by lemma 2.6 and corollary 1.8, we get a strongly continuous unitary representation $U$ of $\Spin$. Setting $U_W(t):=U(\tilde\L_W(t))$, $W\in\W$, $t\in\Re$, the unitary operator $$ z(t)=\D_W^{it}U_W(-t) $$ implements internal symmetries by modular covariance, therefore is in the center of $G$. Let us denote by $\th_g$, $g\in\Poi$, the action of $\Poi$ on the central extension $G$. It follows that $\th_g(\D_W^{it})=\D_{gW}^{it}$, $\th_g(U_W(t))=U_{gW}(t)$. Since $\th_g$ acts trivially on the center of $G$ and $\Poi$ acts transitively on the family $\W$, then $z(t)$ does not depend on $W\in\W$. Moreover $$ z(s)z(t)=\D_W^{it}z(s)U_W(-t)=z(s+t)\qquad s,t\in\Re $$ i.e. $z(t)$ is a one-parameter group. Finally, since $\D_{W'}^{it}=\D_W^{-it}$ by essential duality (theorem 2.2), and $U_{W'}(t)=U_W(-t)$ because $\tilde\L_W (t)$ and $\tilde\L_{W'}({-t})$ are the unique lifting of $\L_W (t)=\L_{W'}({-t})$, we have $z(t)=z(-t)$, i.e. $z(t)\equiv I$. \np In order to check the positivity of the energy-momentum, we observe that the generator of any time-like translation is a convex combination of generators of light-like translations. Moreover each generator of a light-like translation gives rise to a one-parameter semigroup of endomorphisms of a suitable wedge. Therefore the result follows by proposition 2.7. \endproof The following proposition is a partial converse to Borchers theorem, in a slightly different form it appeared also in Wiesbrock paper [\ref(Wies1)], with the difference that we do not assume the commutation relations of the modular conjugation with the one-parameter group. \claim{2.7 Proposition} (Converse of Borchers theorem) Let $\R$ be a von Neumann algebra standard with respect to the vector $\Q$. If $U(a)$ is a one-parameter semigroup ($a\in \Re_+ $) of endomorphysms of $\R$ such that \item{$(i)$} $U(a)\Q=\Q$ \item{$(ii)$} $ad\D_{\R}^{it} U(a)=U(e^{-2\pi t} a)$ \np then $U(a)$ has positive generator. \par We give here an easy proof of the preceding Theorem. \proof From Lemma II.$3$ in [\ref(Borc1)] we know that the operator valued function $t\to U(e^{-2\pi t} a)$ can be analitically extended on the strip $\{ z\in\Co ; -1/2