INSTRUCTIONS The text between the lines BODY and ENDBODY is made of 1306 lines and 54406 bytes (not counting or ) In the following table this count is broken down by ASCII code; immediately following the code is the corresponding character. 34992 lowercase letters 2897 uppercase letters 1206 digits 2 ASCII characters 9 6764 ASCII characters 32 18 ASCII characters 33 ! 15 ASCII characters 34 " 95 ASCII characters 35 # 1280 ASCII characters 36 $ 516 ASCII characters 37 % 31 ASCII characters 38 & 120 ASCII characters 39 ' 523 ASCII characters 40 ( 523 ASCII characters 41 ) 30 ASCII characters 42 * 20 ASCII characters 43 + 490 ASCII characters 44 , 253 ASCII characters 45 - 646 ASCII characters 46 . 15 ASCII characters 47 / 31 ASCII characters 58 : 4 ASCII characters 59 ; 7 ASCII characters 60 < 130 ASCII characters 61 = 7 ASCII characters 62 > 2 ASCII characters 64 @ 73 ASCII characters 91 [ 1911 ASCII characters 92 \ 73 ASCII characters 93 ] 130 ASCII characters 94 ^ 374 ASCII characters 95 _ 96 ASCII characters 96 ` 513 ASCII characters 123 { 20 ASCII characters 124 | 514 ASCII characters 125 } 85 ASCII characters 126 ~ BODY %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%% AN ALGEBRAIC SPIN AND STATISTICS THEOREM. I %%%%%%%%%% %%%%%% D. Guido - R. Longo, 15 pages, plain TeX %%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%% Special Fonts %%%%%%% %Number Sets % Blackboard bold \def\Co{{\bf C}} %\def\Co{{\Bbb C}} \def\Na{{\bf N}} %\def\Na{{\Bbb N}} \def\Re{{\bf R}} %\def\Re{{\Bbb R}} \def\To{{\bf T}} %\def\To{{\Bbb T}} \def\Ze{{\bf Z}} %\def\Ze{{\Bbb Z}} %%%%%% SYMBOLS %%%%%%%%%%%%% \def\A{{\cal A}} \def\a{\alpha} \def\B{{\cal B}} \def\b{\beta} \def\C{{\cal C}} \def\d{\delta} \def\D{\Delta} \def\cD{{\cal D}} \def\f{\varphi} \def\F{{\cal F}} \def\g{\gamma} \def\G{\Gamma} \def\bG{{\bf G}} \def\H{{\cal H}} \def\k{\kappa} \def\K{{\cal K}} \def\bK{{\bf K}} \def\l{\lambda} \def\L{\Lambda} \def\m{\mu} \def\n{\nu} \def\o{\omega} \def\O{{\cal O}} \def\p{\pi} \def\Q{\Omega} \def\r{\rho} \def\s{\sigma} \def\S{{\cal S}} \def\t{\tau} \def\T{\Theta} \def\th{\vartheta} \def\W{{\cal W}} \def\x{\xi} \def\Z{{\cal Z}} \def\Lp{{\cal L}_+} %Lorentz proper \def\Lpo{{\cal L}_+^\uparrow} %Lorentz proper orthochronous \def\Pp{{\cal P}_+} %Poincare' proper \def\Ppo{{\cal P}_+^\uparrow} %Poincare' proper orthochronous \def\Sp{\widetilde{\cal P}_+} %covering of Poincare' proper \def\Spo{\widetilde{\cal P}_+^\uparrow} %covering of Poincare' prop. ort. \def\imply{\Rightarrow} \def\Ind#1#2{{\rm Ind}_{#1}^{#2}} %%%%%%%% Macros for theorems %%%%%%%% \font\sectionfont=cmbx10 scaled\magstep1 \def\begProof{\medskip\noindent{\bf Proof.}\quad} \def\begProofof#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\titlea#1{\vskip0pt plus.3\vsize\penalty-75 \vskip0pt plus -.3\vsize\bigskip\bigskip \noindent{\sectionfont #1}\nobreak\smallskip\noindent} \def\titleb#1{\medskip\noindent{\it#1}\qquad} \def\claim#1#2{\vskip.1in\medbreak\noindent{\bf #1.} {\sl #2}\par \ifdim\lastskip<\medskipamount\removelastskip\penalty55\medskip\fi} \def\begProof{\medskip\noindent{\bf Proof.}\quad} \def\begProofof#1{\medskip\noindent{\bf Proof of #1.}\quad} \def\square{\hbox{$\sqcap\!\!\!\!\sqcup$}} \def\endProof{\hbox{\hskip-4mm\hfill\vbox{\vskip3.5mm\square\vskip-3.5mm}} \bigskip} \def\beglemma#1#2\endlemma{\claim{#1 Lemma}{#2}} \def\begdefinition#1#2\enddefinition{\claim{#1 Definition}{#2}} \def\begtheorem#1#2\endtheorem{\claim{#1 Theorem}{#2}} \def\begcorollary#1#2\endcorollary{\claim{#1 Corollary}{#2}} \def\begremark#1#2\endremark{\claim{#1 Remark}{#2}} \def\begproposition#1#2\endproposition{\claim{#1 Proposition}{#2}} \def\refno#1#2{\item{[#1]}{#2}} \def\begref#1#2{\titlea{#1}} \def\endref{} \newcount\FNOTcount \FNOTcount=1 \def\numfnot{\number\FNOTcount} \def\addfnot{\global\advance\FNOTcount by 1} \def\fonote#1{\footnote{$^\numfnot$}{#1}\addfnot} %%%%%%%% AUTO REF. %%%%%%%%%%%%% \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\rfr(#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{[\rfr(##1)]}##2} \refno{\rfr(##1)}##2} \begref{References}{99}\REFERENCES\endref} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%% TITLE %%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \font\ftitle=cmbx10 scaled\magstep1 \centerline{\ftitle AN ALGEBRAIC SPIN AND STATISTICS THEOREM. I} \bigskip \centerline{Daniele Guido$^1$\footnote{$^*$} { Supported in part by MURST and CNR-GNAFA.} and Roberto Longo$^{1,2*}$} \footnote{}{E-mail:\ guido@mat.utovrm.it, longo@mat.utovrm.it } \bigskip \item{$(^1)$} Dipartimento di Matematica, Universit\`a di Roma ``Tor Vergata'' \par via della Ricerca Scientifica, I--00133 Roma, Italia. \item{$(^2)$} Centro Linceo Interdisciplinare, \par via della Lungara 10, I--00165 Roma, Italia \bigskip\bigskip \noindent{\bf Abstract.} A spin-statistics theorem and a PCT theorem are obtained in the context of the superselection sectors in Quantum Field Theory on a 4-dimensional space-time. Our main assumption is the requirement that the modular groups of the von Neumann algebras of local observables associated with wedge regions act geometrically as pure Lorentz transformations. Such a property, satisfied by the local algebras generated by Wightman fields because of the Bisognano-Wichmann theorem, is regarded as a natural primitive assumption. %%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%% REFERENCES %%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%% \def\REFERENCES{ \REF(AHKT1)Araki H., Haag R., Kastler D., Takesaki M., ``{\it Extensions of KMS states and chemical potential}'', Commun. Math. Phys. {\bf 53} (1977), 97-134. \endREF \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(BiWi2)Bisognano J., Wichmann E., ``{\it On the duality condition for quantum fields}", J. Math. Phys. {\bf 17} (1976), 303-321. \endREF \REF(BjDr1)Bjorken J.D., Drell S.D., ``{\it Relativistic Quantum Fields}'', McGraw-Hill, New York 1965. \endREF \REF(Borc2)Borchers H.J., ``{\it Local rings and the connection between spin and statistics}", Commun. Math. Phys. {\bf 1} (1965), 281-307 \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(Borc3)Borchers H.J., ``{\it On the converse of the Reeh-Schlieder theorem}'', Commun. Math. Phys. {\bf 10} (1968), 269-273. \endREF \REF(BGL1)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(BGL2)Brunetti R., Guido D., Longo R., ``{\it Group cohomology, modular theory and space-time symmetries}", to appear in Rev. Math. Phys. \endREF \REF(BuEp1)Buchholz D., Epstein H., ``{\it Spin and statistics of quantum topological charges}'', Fizika {\bf 3} (1985), 329-343. \endREF \REF(BuFr1)Buchholz D., Fredenhagen K., ``{\it Locality and structure of particle states}'', Commun. Math. Phys. {\bf 84} (1982) 1-54. \endREF \REF(BuSu1)Buchholz D., Summers S.J., ``{\it An algebraic characterization of vacuum states in Minkowski space}", Commun. Math. Phys. {\bf 155} (1993), 442-458. \endREF \REF(Burgoyne)Burgoyne N., ``{\it On the connection of spin with statistics}'', Nuovo Cimento {\bf 8} (1958), 807. \endREF \REF(Dell'Antonio)Dell'Antonio G.F., ``{\it On the connection of spin with statistics}'', Ann. Phys. {\bf 16} (1961), 153. \endREF \REF(DHR1)Doplicher S., Haag R., Roberts J.E., ``{\it Local observables and particle statistics I}'', Commun. Math. Phys. {\bf 23} (1971), 199-230. \endREF \REF(DHR2)Doplicher S., Haag R., Roberts J.E., ``{\it Local observables and particle statistics II}'', Commun. Math. Phys. {\bf 35} (1974), 49-85. \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(DoRo)Doplicher S., Roberts J.E., ``{\it Why there is a field algebra with a compact gauge group describing the superselection structure in particle physics}'', Commun. Math. Phys. {\bf 131} (1990), 51-107. \endREF \REF(Dyson)Dyson F.J., ``{\it On the connection of weak local commutativity and regularity of Wightman functions}'', Phys. Rev. {\bf 110} (1958), 579. \endREF \REF(Fie)Fierz M., ``{\it \"Uber die relativische Theorie kr\"aftfreier Teilchen mit beliebigem spin}'', Helv. Phys. Acta {\bf 12} (1939), 3. \endREF \REF(Epst1)Epstein H., ``{\it CTP invariance in a theory of local observables}'', J. Math. Phys. {\bf 8} (1967), 750. \endREF \REF(GL1)Guido D., Longo R., ``{\it Relativistic invariance and charge conjugation in quantum field theory}'', Commun. Math. Phys. {\bf 148} (1992), 521-551. \endREF \REF(GL3)Guido D., Longo R., ``{\it An algebraic Spin and Statistics theorem. II}'' in preparation. \endREF \REF(Haag1)Haag R., {\it Local Quantum Physics}, Springer Verlag, Berlin Heidelberg 1992. \endREF \REF(HHW)Haag R., Hugenoltz N.M., Winnink M., ``{\it On the equilibrium states in quantum statistical mechanics}'', Commun. Math. Phys. {\bf 5} (1967), 215.\endREF \REF(HaKa1)Haag R., Kastler D., ``{\it An algebraic approach to Quantum Field Theory}'', J. Math. Phys. {\bf 5} (1964), 848-861. \endREF \REF(Jost1)Jost R., ``{\it The general theory of Quantized Fields}'' Amer. Math Soc., Providence RI 1965. \endREF \REF(Jos57)Jost R., ``{\it Eine Bemerkung zu CTP Theorem}'', Elv. Phys. Acta {\bf 30} (1957), 409. \endREF \REF(Kay)Kay B.S., ``{\it The free field on the wedge and uniqueness of KMS one-particle structure}'' preprint Feb. 1983. \endREF \REF(Kuck1)Kuckert B., ``{\it PCT und Kovarianz als modulare Strukturen}'', Diploma Thesis, Hamburg, in preparation. \endREF \REF(Lipsman)Lipsman R.L., ``{\it Group representations}'' Lect. Notes in Math. {\bf 388}, Springer Verlag, New York--Heidelberg--Berlin 1974. \endREF \REF(Land1)Landau L., ``{\it On local functions of fields}'' Commun. Math. Phys. {\it 39} (1974), 49-62. \endREF \REF(Long1)Longo R., ``{\it Index of subfactors and statistics of quantum fields. }'', $I$ Commun. Math. Phys. {\bf 126} (1989), 217-247; $II$ {\bf 130} (1990), 285-309. \endREF \REF(Lud54)L\"uders G., ``{\it Vertaugschungsrelationen zwischen verschiedenen Feldern}'', Z. Naturforsch. {\bf 13a} (1958), 254. \endREF \REF(LudersAndZumino)L\"uders G., Zumino B., ``{\it Connection between spin and statistics}'', Phys. Rev. {\bf 110} (1958), 1450. \endREF \REF(Pau)Pauli W., ``{\it On the connection between spin and statistics}'', Phys. Rev. {\bf 58} (1940), 716. \endREF \REF(Pau55)Pauli W., ``{\it Exclusion principle, Lorentz group and reflection of space time and charge}'', in Niels Bohr and the Development of Physics, W. Pauli (ed.) Pergamon Press, New York, 1955. \endREF \REF(Schwinger)Schwinger J., ``{\it On the theory of quantized fields I}'', Physics Reviews {\bf 82} (1951), 914. \endREF \REF(Sewell)Sewell G.L., ``{\it Relativity of temperature and Hawking effect}'', Phys. Lett. {\bf79A} (1980), 23. \endREF \REF(StZs1)Str\u atil\u a S., Zsido L., {\it Lectures on von~Neumann algebras}, Abacus press, England 1979. \endREF \REF(Streater)Streater R.F., ``{\it Local fields with the wrong connection between Spin and Statistics}'', Commun. Math. Phys. {\bf 5} (1967), 88-96. \endREF \REF(StWi1)Streater R.F., Wightman A.S., ``{\it PCT, Spin and Statistics, and all that}'', Benjamin, Reading (MA) 1964. \endREF \REF(Take1)Takesaki M., ``{\it Tomita theory of modular Hilbert algebras }'', Lect. Notes in Math. {\bf 128}, Springer Verlag, New York--Heidelberg--Berlin 1970. \endREF \REF(Wigh)Wightman A.S., ``{\it Quantum field theory in terms of vacuum expectation values}'', Phys. Rev. {\bf 101}, (1956) 860. \endREF \REF(Wies1)Wiesbrock H.V., ``{\it A comment on a recent work of Borchers}", Lett. Math. Phys. {\bf 25}, (1992), 157-159. \endREF \REF(Yngv1)Yngvason J., ``{\it A note on essential duality}'', preprint. \endREF} \inputreferences \titlea{Introduction} In this paper we shall reconsider from an intrinsic point of view two well known fashinating theorems in Quantum Field Theory: the PCT theorem and the Spin and Statistics theorem. Both of them have a long history, see [\rfr(Jost1),\rfr(StWi1)]. The spin and statistics theorem first appeared in the context of free fields in the work of Fierz [\rfr(Fie)] and Pauli [\rfr(Pau)]: one cannot second quantize particles with integer spin by anticommuting fields, i.e. fields obeying Fermi statistics, nor particles with half-integer spin by local fields, i.e. fields obeying Bose statistics. The PCT theorem originated in [\rfr(Lud54)] as a relation between the existence of the space-inversion symmetry P and the existence of the product of the charge and the time-inversion symmetry CT. Pauli proved in [\rfr(Pau55)] that PCT is always a symmetry of Lorentz invariant field equations. It was a success of the Wightman axiomatic approach [\rfr(Wigh)] to extabilish model independent results: the connection between spin and statistics was obtained by Burgoyne [\rfr(Burgoyne)], see also [\rfr(LudersAndZumino),\rfr(Dell'Antonio)], and a PCT theorem by Jost [\rfr(Jos57)], see also [\rfr(Dyson), \rfr(Schwinger)], both relying on the general holomorphic properties of the $n$-point functions. A spin and statistics theorem in the algebraic approach [\rfr(HaKa1)], see also [\rfr(Borc2)], was later given by Epstein [\rfr(Epst1)] and has a version for (Doplicher-Haag-Roberts) DHR superselection sectors [\rfr(DHR2)] and for more general topological charges [\rfr(BuEp1)]. All these approaches heavily rely on arguments of analytic continuation, whose nature give some mysterious effectiveness to the results. Moreover they make use of certain detailed structures, either because they deal with Wightman tempered distributions or because they treat the case of finite mass degeneracy, where a supeselection sector has to contain only finitely many particles of the same type and all of them are assumed to have strictly positive mass. The approach to Quantum Field Theory by local observable algebras [\rfr(HaKa1)] suggests however that a PCT symmetry and spin-statistics correspondence should be intrinsically associated with the net of local algebras and manifest itself as the consequence of the locality principle\fonote{The spin-statistics relation depends on sharp locality. In the second quantization of Bose particles by anti-commuting fields, microscopic causality is still asymptotically present and its violation is sizeable only at distances comparable to the Compton wave lenght [\rfr(BjDr1)]. This is perhaps an indication that the spin-statistics relation might be different in contexts like Quantum Gravity where a sharp causality principle does not occur.}. >From the mathematical point of view the spin-statistics correspondence is a relation between two quantities of different nature, the univalence and the statistics phase, and one is led to tie up these concepts on a general ground, somehow in the spirit of an index theorem. We shall extablish both a PCT and a spin-statistics theorem in the following general context. Let $\O\to\A(\O)$ be a net of von~Neumann algebras on a Hilbert space $\H$, i.e. an inclusion preserving association between regions $\O$ in the four-dimensional Minkowski space and von Neumann algebras of local observables, that we assume here to be irreducible. We make the following assumptions. {\it Locality.} If $\O_1$ and $\O_2$ are space-like separated regions, then $\A(\O_1)$ and $\A(\O_2)$ commute elementwise. {\it Modular covariance.} There is a vector $\Q\in\H$, the vacuum vector, cyclic for the algebras $\A(W)$ associated with all wedge region $W$ in the Minkowski space, such that $$ \D_W^{it}\A(\O)\D_W^{-it}=\A(\L_W(t)\O)\ ,\qquad t\in\Re $$ where $\O$ is any region, $\D$ is the Tomita-Takesaki modular operator [\rfr(Take1),\rfr(StZs1)] associated with $(\A(W),\Q)$ and $\L_W$ is the one-parameter rescaled group of pure Lorentz transformations preserving $W$. {\it Reeh-Schlieder property.} The vacuum vector $\Q\in\H$ is also cyclic for the algebras $\A(\S)$ associated with all space-like cones $\S$. Locality is the well-known expression of Einstein causality and we do not dwell on it. The Reeh Schlieder property is known to hold for the vacuum vector in a Poincar\`e covariant theory as a consequence of the positivity of the energy and the weak additivity assumption for the local algebras [\rfr(Borc3)]. Modular covariance needs however further comments. Postponing for a while the justification for such assumption, we recall that this entails the net to be covariant with respect to the universal covering $\Spo$ of the Poincar\'e group $\Ppo$, with positive energy [\rfr(BGL2)]. Indeed we shall prove here that it is actually covariant with respect to $\Ppo$ as a special case of our general Spin and Statistics theorem and taking to completion our previous work. Therefore modular covariance is a way to intrinsically encode the Poincar\'e covariance property in the net structure, providing a canonical representation of the Poincar\`e group $\Ppo$ (cf. also [\rfr(BuSu1)]). Let now $\r$ be a superselection sector of $\A$ in the sense of Doplicher-Haag-Roberts [\rfr(DHR1)] or more generally of Buchholz-Fredenhagen [\rfr(BuFr1)]. An index-statistics relation [\rfr(Long1)] shows that $$ {\rm Ind}(\r)=d(\r)^2 $$ where ${\rm Ind}(\r)$ is the Jones index of $\r$ and $d(\r)$ is the DHR statistical dimension, namely $$ d(\r)=|\l_\r|^{-1} $$ where the statistical parameter $\l_\r\in\Re$ classifies the statistics in $3+1$ space-time dimensions [\rfr(DHR1)]. Therefore the index is an intrinsic quantity that determines the statistics up to the Fermi-Bose alternative, i.e. the sign of $\l_\r$. On the other hand, the Poincar\'e representation in the vacuum sector being fixed by modular covariance, the representation of $\Spo$ associated with a covariant irreducible sector $\r$ is uniquely determined, therefore the univalence (integer or half-integer spin alternative) is intrinsically associated with $\r$. Since $\r$ is automatically $\Spo$-covariant if $d(\r)<\infty$ (assuming a regularity property for the net [\rfr(GL1)]), it is natural to expect a general algebraic Spin and Statistics theorem connecting these two intrinsic quantities for any sector with finite statistics. Our result in this respect will in fact show that on these general grounds $$ {\rm sign}(\l_\r)=U_\r(2\pi) $$ where $U_\r$ is the representation of $\Spo$ in the sector $\r$ and $U_\r(2\pi)$ denotes the corresponding rotation by $2\pi$. Modular covariance also implies that the anti-unitary involution $\T$, definable by the modular theory according to the Bisognano-Wichmann prescription [\rfr(BiWi2)], implements a complete space-time reflection. As shown in [\rfr(GL1)], this entails that $\T$ intertwines a sector with its conjugate. We therefore obtain a PCT symmetry. We come now back to the origin of the modular covariance property. Its main justification certainly comes from the Bisognano-Wichmann theorem [\rfr(BiWi1),\rfr(BiWi2)] to the effect that this property holds if the local algebras are constructed from Wightman fields. An algebraic version of the Bisognano-Wichmann theorem does not exist yet, except in the case of conformal theories where it holds in full generality [\rfr(BGL1)]. However a theorem of Borchers [\rfr(Borc1)] shows part of the geometric properties of the modular group for wedge regions to be always present and in particular every $1+1$ dimensional Poincar\'e covariant net satisfying essential duality has the modular covariance property. At the present time no counter-example to modular covariance is known to exist within Poincar\`e covariant theories (see however [\rfr(Land1),\rfr(Yngv1)]). There are nevertheless counter-examples to the spin-statistics theorem [\rfr(Streater)]: these are constructed by infinite multiplicity fields where the Poincar\'e group representation is not unique. It turns out that the wrong connection between spin and statistics depends on the wrong choice of the Poincar\`e group representation, while our canonical choice for the latter has the desired properties. We remark that an intrinsic way to eliminate pathological examples of the above kind comes by requiring the split property [\rfr(DoLo1)]; this indeed implies the uniqueness of the Poincar\`e group representation [\rfr(BGL1)] and we propose it as a natural candidate for a derivation of the modular covariance property by first principles. On the physical side modular covariance manifests an interesting analogy with the Unruh effect and with the Hawking black hole thermal radiation, as first noticed by Sewell [\rfr(Sewell)]. We sketch the essential ideas, see also [\rfr(Haag1)]. As is known the modular group of a von Neumann algebra with respect to a given state is characterized by the Kubo-Martin-Schwinger condition [\rfr(Take1)] and, on the other hand, KMS condition is peculiar of thermal equilibrium states in Statistical Mechanics [\rfr(HHW)]. By the Bisognano-Wichmann theorem the boosts satify the KMS condition with respect to the vacuum, as automorphisms of the von Neumann algebra of the corresponding wedge $W$ and, on the other hand, the orbits of the boosts are the trajectories of a uniformly accelerated motion for which the "Rindler universe" $W$ is a natural horizon; the equivalence principle in Relativity Theory then allows an interpretation of the thermal outcome as a gravitational effect. On this basis Haag has long proposed to derive the Bisognano-Wichmann theorem. The role of the modular covariance assumption may be also understood by its consequences. Among other things, it implies the positivity of the energy for the constructed Poincar\'e group representation [\rfr(Wies1),\rfr(BGL2)]. As is known the positivity of the energy is lost on a curved space-time, and the modular covariance seems to be the appropriate substitute in this case. Moreover, as already mentioned, it gives rise to the KMS condition, namely an analytic continuation property. It turns out that this analytic aspect of the modular covariance assumption encorporates all the holomorphic properties present in Quantum Field Theory. But, as a matter of facts, the modular group is an algebraic object, a manifestation of the $^*$-operation, thus providing us with an algebraic approach to our problems. We pass now to a description of the methods of our work. This paper relies on the modular theory of Tomita and Takesaki and on an analysis with the unitary representations of ${\rm SL}(2,\Re)$. We shall find a key relation arising from the comparison of the modular groups of different algebras, and we shall regard it as an identity concerning operators in the space of a representation of ${\rm SL}(2,\Re)$, because of the well known fact that the 2+1 dimensional Lorentz group is isomorphic to ${\rm SL}(2,\Re)/\{1,-1\}$. Section~1 contains the proof of this identity by the Mackey machine of the induced representations (e.g. [\rfr(Lipsman)]) and a free field verification. In section~2 the PCT theorem and the Spin and Statistics relation are proven in the context of the field algebras,\fonote{We have recently been informed by Kuckert of an independent complementary analysis based on assumptions of weak geometric type for the modular conjugation [\rfr(Kuck1)].} where the formalism is close to the classical formulation. Then, in section~3, we obtain our result in the context of local observables. This is done by rephrasing the statements in terms of the Doplicher-Roberts field algebra [\rfr(DoRo)]. This last step has certain pedagogical advantages, but has to be avoided in order to extend our work to more general settings where the field algebra does not exists. In a forthcoming paper [\rfr(GL3)] we shall indeed provide a more intrinsic approach in terms of local observables only, that will cover low dimensional and conformal theories in particular. The general picture will be clarified by examples. \titlea{1. An identity for operators associated with representations of ${\rm SL}(2,\Re)$} Let us consider two one-parameter subgroups of ${\rm SL}(2,\Re)$ $$ \mu(t)\equiv\left (\matrix{\cosh\pi t&-\sinh\pi t\cr-\sinh\pi t&\cosh\pi t\cr}\right)\ ,\quad \nu(t)\equiv\left (\matrix{e^{-\pi t}&0\cr0&e^{\pi t}\cr}\right)\ . $$ If $U$ is a unitary representation of ${\rm SL}(2,\Re)$ on a Hilbert space $\H$ we look at the corresponding selfadjoint infinitesimal generators $$ H = H_U = i{d\over dt}U(\mu(t))|_{t=0} $$ $$ K = K_U = i{d\over dt}U(\nu(t))|_{t=0}. $$ We shall denote by $\bG$ the group ${\rm PSL}(2,\Re)$ given by the quotient of ${\rm SL}(2,\Re)$ with its center $\{-1,1\}$ and by $\widetilde\bG$ the universal covering of $\bG$, which is of course the universal covering of ${\rm SL}(2,\Re)$ too. We take the same definition for $H$ and $K$ in the case of a unitary representation $U$ of the universal covering group $\widetilde\bG$. We will consider the following property for a representation $U$ of $\widetilde\bG$: $$ \eqalign{ T_t&\equiv e^{{ 1\over 2}K}e^{itH}e^{-{1\over 2}K}\subset e^{-itH}\cr {\rm and}\quad T_t&\hbox{\rm is densely defined}\quad\forall t\in\Re\cr} \eqno(1.1) $$ where the symbol $\subset$ denotes the extension of operators. Property~(1.1) refers to a representation $U$, but we omit the symbol $U$ when no confusion arises. \begtheorem{1.1} Property~(1.1) holds for all the unitary representations of $\widetilde\bG$. \endtheorem In order to prove this theorem, we first observe that it is enough to check Property~(1.1) on dense sets of vectors, not necessarily on the full domain of $T_t$. \beglemma{1.2} Let us assume that, for each real $t$, there is a dense subset ${\cal D}_t$ of the domain of $T_t$ such that $T_t|_{{\cal D}_t}\subset e^{-itH}$. Then Property~(1.1) holds for the given representation. \endlemma \begProof Note first that the matrix $\left(\matrix{0&1\cr-1&0\cr}\right )\in {\rm SL}(2,\Re)$ conjugates $\mu(t)$ with $\mu(-t)$ and $\nu(t)$ with $\nu(-t)$ therefore the assumption of the lemma remains true if we replace $K$ with $-K$ and $H$ with $-H$, in particular $e^{ -{1\over 2}K}e^{itH}e^{{1\over 2}K}$ is densely defined. Now $$ T_t^*\supset e^{ -{1\over 2}K}e^{-itH}e^{{1\over 2}K} $$ therefore $T_t^*$ is densely defined and $T_t$ is closable. Since $T_t\xi = e^{-itH}\xi$ for all $\xi$ in a dense set and $e^{-itH}$ is bounded, the equality $T_t\xi=e^{-itH}\xi$ must hold for all $\xi$ in the domain of $T_t$. \endProof \begcorollary{1.3}Let $U_1,U_2$ and $U$ be unitary representations of $\widetilde\bG$. \medskip \item{$(a)$} Property~(1.1) holds for $U_1\otimes U_2$ iff it holds for both $U_1$ and $U_2$. \item{$(b)$}If $U = \int^{\oplus}U(\lambda)dm(\lambda)$ is a direct integral decomposition of $U$, then Property~(1.1) holds for $U$ iff it holds for $U(\l)$, for $m$-almost all $\lambda$. \medskip \endcorollary \begProof Part $(b)$ and the implication $\Leftarrow$ of part $(a)$ are immediate by Lemma~1.2 since one can check Property~(1.1) on natural dense sets. If Property~(1.1) holds for $U_1\otimes U_2$ then it holds for $U_1$ and $U_2$ up to a constant, namely, considering for example the representation $U_1$, there exists a one dimensional character $z(\cdot)$ of $\Re$ such that $$ T_t\subset z(t)e^{-itH} $$ and $T_t$ is densely defined. Of course $z(\cdot)$ remains unchanged if we replace $\mu$ and $\nu$ by a pair of conjugate one-parameter subgroups. As in the proof of Lemma~1.2 we may thus replace $\mu(t)$ by $\mu(-t)$ and thus $z(t)$ by $z(-t)$, hence $z(t)=z(-t)=1$ and the proof is complete. \endProof We need now to verify Property~(1.1) in some specific representation. To this end recall that [\rfr(StWi1)], if $W_i$ is the wedge in the $3$-dimensional space-time along the axis $x_i$, $i=1,2$, and $\L_i(t)$ is the associated one-parameter group of pure Lorentz transformations (see section~2), there is an isomorphism of ${\rm PSL}(2,\Re)$ with the $2+1$-dimensional Lorentz group $\Lpo(3)$ determined by $$ \m(t)\to\L_1(t)\quad{\rm and}\quad \n(t)\to\L_2(t) . $$ Accordingly, we shall identify $\bG$ with a subgroup of the 2+1-dimensional Poincar\'e group $\Ppo(3)$. \beglemma{1.4} Let $V\equiv V_{m,0}$ be the positive energy representation of $\Ppo(3)$ of spin 0 and mass $m>0$. Then Property~(1.1) holds for the restriction $U\equiv V|_{\bG}$ of $V$ to $\bG$. \endlemma \begProof As is known, $V$ extends to a (anti-)representation of the proper Poincar\'e group $\Pp(3)$, namely there exists a anti-unitary involution $\Theta$, on the same Hilbert space, that commutes with $U$ and implements the change of sign on the translation operators in any space-time direction. By the one-particle version of the Bisognano Wichmann theorem (which follows of course from the Bisognano-Wichmann theorem in the free field setting, see [\rfr(Kay)] for a short direct verification of this special case) we may identify the rescaled boost transformations with the modular group of the real Hilbert subspace of the one-particle Hilbert space associated to the corresponding wedge region. Then Property~(1.1) holds because it is equivalent to the commutativity of $\Theta$ with the boosts (see Proposition~2.6). \endProof \begremark{1.5} The proof of Lemma~1.4 makes use of a one-particle version of the Bisognano-Wichmann theorem; as we mentioned this can be proved directly by mimicking the proof of the Bisognano-Wichmann in this special case. Since this quick verification requires an analytic continuation argument that does not fit with the spirit of this paper, we sketch here an algebraic derivation of Lemma~1.4. To begin with note that Lemma~1.4 would hold if $V$ were the irreducible positive-energy massless representation of $\Ppo(3)$ with helicity 0. Indeed in this case $V$ extends to a representation of the conformal group and the algebraic argument in [\rfr(BGL1)] applies. Now $V\otimes V$ has a direct integral decomposition into irreducible representations where massive representations occur and thus Lemma~1.2 implies Lemma~1.4 for some $m>0$. However the representation $U\equiv V|_{\bG}$ in Lemma~1.4 does not depend on $m>0$ up to unitary equivalence by the following proposition, hence Lemma~1.4 holds for all $m>0$. \endremark \begproposition{1.6} The representation $U\equiv V|_{\bG}$ in Lemma~1.4 is equivalent to the quasi-regular representation of $\bG$ corresponding to the rotation subgroup $\bK\equiv\left\{\left(\matrix{\cos\theta &\sin\theta\cr-\sin\theta&\cos\theta\cr}\right), 0\leq\theta<2\pi\right\}/\{1,-1\}$, namely $U$ is the representation of $\bG$ induced by the identity representation of $\bK$. \endproposition \begProof The $m>0$ hyperboloid $H_m=\{\vec{x}\in\Re^3/x_0^2-x_1^2- x_2^2 = m^2,\, x_0>0\}$ is a homogeneous space for $\bG$ whose stability subgroup at the point $(m,0,0)$ is $\bK$. $U$ is the corresponding representation on $L^2(H_m,\mu_m)$, with $\mu_m$ the Lorentz invariant measure on $H_m$, and this is by definition the quasi-regular represention with respect to $\bK$. \endProof \begcorollary{1.7} With $U\equiv V|_{\bG}$ the representation in Lemma~1.4, $U\otimes U$ is equivalent to an infinite multiple of the regular representation $\lambda$ of $\bG$ $$ U\otimes U = \infty\cdot\lambda. $$ In particular Property~(1.1) holds for $\lambda$. \endcorollary \begProof The first statement is a consequence of the Mackey tensor product theorem for induced representations, see [\rfr(Lipsman), Theorem~2 and Example~5]. Property~(1.1) then holds for $\lambda$ because of Lemma~1.2.\endProof Here is an alternative verification of Property~(1.1) for $\lambda$. By Proposition~1.6, taking tensor products and making use of Corollary~1.3, we check that Property~(1.1) is valid for the irreducible representation $V_{m,s}$ of $\Ppo(3)$ of any mass $m>0$ and any integral spin $s$. Now $U_{m,s}\equiv V_{m,s}|_\bG$ is the induced representation $$ U_{m,s}=\Ind{\bK}{\bG}(\chi_s) $$ where $\chi_s$ is character of $\bK\simeq\To$ associated with the integer $s$. By inducing at stages one has $$ \lambda=\Ind{\{1\}}{\bG}(id)=\Ind{\bK}{\bG}(\lambda_\bK) $$ where $\lambda_\bK=\Ind{\{1\}}{\bK}(id)$ is the regular representation of \bK, hence $$ \lambda =\Ind{\bK}{\bG}(\lambda_\bK) = \Ind{\bK}{\bG}\bigoplus_{s=-\infty}^{\infty}\chi_s = \bigoplus_{s=-\infty}^{\infty}\Ind{\bK}{\bG}(\chi_s) = \bigoplus_{s=-\infty}^{\infty}U_{m,s} $$ and the statement follows by Corollary~1.4. \begProofof{of Theorem~1.1} Suppose first that $U$ is a representation of $\bG$. By the absorbing property of the regular representation $\lambda$, $U\otimes\lambda$ is equivalent to a multiple of $\lambda$, therefore Property~(1.1) holds for $U$ because of Lemma~1.2 and Corollary~1.7. If now $U$ is a more general representation of $\widetilde\bG$, then $U$ detemines a projective representation of $\bG$, namely $U(l(\cdot))$ with $l$ a Borel section for the quotient map of $\widetilde\bG$ modulo $\bG$. The tensor product $U\otimes\overline U$ of $U$ with the conjugate representation is a true representation of $\bG$, hence Property~(1.1) holds for $U\otimes\overline U$. Then, by Lemma~1.2, Property~(1.1) holds for $U$. \endProof \titlea{2. PCT, Spin and Statistics on the field algebras} In this section we consider a pre-cosheaf $\O\to\F(\O)$ of von~Neumann algebras acting on a Hilbert space $\H$, where $\O$ is any open region of the $4$-dimensional Minkowski space $M$. We assume the following properties: \medskip \item{$(1)$} {\it Reeh-Schlieder property} for space-like cones: there is a vector $\Q\in\H$ ({\it vacuum}) which is cyclic for the algebras associated with all space-like cones. \item{$(2)$} {\it Normal commutation relations}: there is a vacuum-preserving self-adjoint unitary $\G$ (statistics operator) that implements an automorphism on every local von Neumaann algebra and the normal commutation relations between Bose and Fermi fields hold, i.e. setting $$ \F_{\pm}(\O):=\{A\in\F(\O):\G A\G=\pm A\} $$ we have that if $\O_1$ and $\O_2$ are space-like separated then $\F_+(\O_1)$ commutes with $\F(\O_2)$ and $\F_-(\O_1)$ anticommutes with $\F_-(\O_2)$. \item{$(3)$} {\it Modular covariance property} with respect to the vacuum vector $\Q$ (cf. Definition~2.3). \medskip \begproposition{2.1}{\rm(Twisted Locality)} Let $Z$ be the unitary operator defined by $$ Z={I+i\G\over 1+i}.\eqno(2.1) $$ Then $$ Z\F(\O)Z^*\subset\F(\O')'.\eqno(2.2) $$ \endproposition \begProof A direct computation shows that $$ \eqalign{ ZBZ^*=B,\qquad&B\in\F_+\cr ZFZ^*=i\G F\quad&F\in\F_-\cr}.\eqno(2.3) $$ Hence, if $\O_1$ and $\O_2$ are space-like separated and $F_j\in\F_-(\O_j)$, $j=1,2$ we have $$ [ZF_1Z^*,F_2]=i\G(F_1F_2+F_2F_1)=0 $$ and the thesis holds. \endProof We recall that a {\it wedge} region is any Poincar\'e transformed of the region $W_1:=\{\vec{x}\in\Re^n:|x_0|