arXiv:1704.01336v1 [math.RT] 5 Apr 2017 AAHCNE ULCTOS OUE** VOLUME PUBLICATIONS, CENTER BANACH 2010 h ae si nlfr n ovrino twl epublished be will it of version no and form final in in is modular paper half-sided The observables, local algebra, Neumann phrases and words Key nintr prtr naHletspace Hilbert a on operators antiunitary ihatuiayrpeettoso h utpiaiegr multiplicative the of representations antiunitary subspaces representations with standard group in modula antiunitary objects lo and of modular of Araki–Haag–Kastler) perspective theory of the the sense between from the borderline in the (QFT) at Theory concepts von key of of the vectors theory of separating modular cyclic the from in conjugations and symmetry, modular PCT a or inversion time Abstract. n oua nescin fsadr usae correspon subspaces standard Aff( of intersections modular and r fe brbnlsoe ree ymti pcs eexp We spaces. symmetric ordered o over geometry bundles the fiber and often representations are c group subspaces Lie standard algebra of antiunitary configurations Neumann of that von is of point context main classical localizati Our more modular of the context translation and the the Longo, in on arise lies they notes as these subspaces of develope emphasis and The Wiesbrock context. and Borchers of work breaking ground **************************************** R ahmtc ujc Classification Subject Mathematics ,adteepoietebscbidn lcsfragnrlt general a for blocks building basic the provide these and ), OIHAAEYO SCIENCES OF ACADEMY POLISH NTTT FMATHEMATICS OF INSTITUTE nintr ersnain fLegop aevle nth in values take groups Lie of representations Antiunitary ASAA201* WARSZAWA NINTR REPRESENTATIONS ANTIUNITARY eateto ahmtc,LusaaSaeUniversity State Louisiana mathematics, of Department eatetMteai,Universit¨at Erlangen-N¨urnberg Mathematik, Department nintr ersnain oua prtr oua c modular operator, modular representation, antiunitary : aesrse1,908Elne,Germany Erlangen, 91058 11, Cauerstrasse N OUA THEORY MODULAR AND ao og,L 00,USA 70803, LA Rouge, Baton -al [email protected] E-mail: ALHRANNEEB KARL-HERMANN -al [email protected] E-mail: GESTUR H V rmr 24;Scnay8R5 81T05. 81R05, Secondary 22E45; Primary : nqatmpyis nintr prtr implement operators antiunitary physics, quantum In . H ⊆ [1] OLAFSSON ´ hc ntr r noet-n correspondence one-to-one in are turn in which oup ndvlpdb rnti ud and Guido Brunetti, by developed on lso unu edtheory field quantum clusion eeacnrlpiti oencode to is point central a Here . emn lers esre some survey We algebras. Neumann oatuiayrpeettosof representations antiunitary to d ewe ofiuain fstandard of configurations between R nb tde rmteperspective the from studied be an h orsodn pcs which spaces, corresponding the f nvrosdrcin nteQFT the in directions various in d × a bevbe QatmField (Quantum observables cal ihcci eaaigvectors. separating cyclic with s er tre nte9swt the with 90s the in started heory afsddmdlrinclusions modular Half-sided . c hsprpciet provide to perspective this ect elsewhere. prtragba hyaieas arise they algebras operator hoyo prtralgebras operator of theory r ru fuiayand unitary of group e nuain von onjugation, 2 K.-H. NEEB AND G. OLAFSSON´ new and systematic insight into the much richer configurations of nets of local observables in QFT.

Contents 1. Introduction...... 3 2. Antiunitaryrepresentations ...... 7 2.1. Involutivegrouppairs ...... 8 2.2. Extendingunitaryrepresentations ...... 10 2.3. One-parametergroups ...... 15 2.4. Somelow-dimensionalgroups ...... 18 3. Modularobjectsandstandardsubspaces...... 22 3.1. Standardsubspaces...... 22 3.2. Symplecticaspectsofstandardsubspaces ...... 24 3.3. Orthogonalrealone-parametergroups ...... 25 3.4. Half-sided modular inclusions of standard subspaces ...... 26 3.5. Half-sided modular intersections ...... 31 4. Aglimpseofmodulartheory ...... 34 4.1. TheTomita–TakesakiTheorem ...... 34 4.2. Translating between standard subspaces and von Neumann algebras . . . 36 4.3. Borcherstriples...... 39 4.4. Modulargeometry ...... 43 5. Nets of standard subspaces and von Neumann algebras ...... 46 5.1. Netsofstandardsubspaces ...... 47 5.2. NetsofvonNeumannalgebras ...... 52 6. Second quantization and modular localization ...... 55 6.1. BosonicFockspace...... 55 6.2. FermionicFockspace...... 57 6.3. From antiunitary representations to local nets ...... 60 7. Perspectives...... 61 7.1. TheVirasorogroup...... 61 7.2. EuclideanJordanalgebras...... 61 7.3. Hermitiangroups...... 61 7.4. Analyticextension ...... 62 7.5. Geometricstandardsubspaces...... 62 7.6. DualpairsintheHeisenberggroup ...... 63 7.7. A representation theoretic perspective on modular localization ...... 64 A.Appendices ...... 64 A.1. AlemmaonvonNeumannalgebras ...... 64 A.2.Twolemmasongroups...... 65 A.3.Symmetricspaces...... 65 ANTIUNITARY REPRESENTATIONS 3

1. Introduction. One of the core ideas of quantum theory is that the states of a quan- tum system correspond to one-dimensional subspaces of a complex , i.e., H the elements [v] = Cv, v = 0, of its projective space P( ). This set carries a geometric 6 H structure defined by the transition probability v, w 2 τ([v], [w]) := |h i| [0, 1] v, v w, w ∈ h ih i between two states [v] and [w], where d([v], [w]) = arccos τ([v], [w]) [0,π/2] is the ∈ corresponding Riemannian metric (the Fubini–Study metric),p turning it into a Riemann– Hilbert manifold. Wigner’s Theorem characterizes the automorphisms of (P( ), τ), resp., H the for the metric, as those bijections induced on P( ) by elements of the H antiunitary group AU( ) of all linear and antilinear surjective isometries of ([Ba64]). H H Accordingly, we have an isomorphism Aut(P( ), τ) = AU( )/T1 =: PAU( ) H ∼ H H of Aut(P( ), τ) with the projective antiunitary group PAU( ). So any action of a group H H G by symmetries of a quantum system leads to a homomorphism π : G PAU( ) ♯ → H and further to a homomorphism of the pullback extension G = π∗ AU( ) of G by the H circle group T to AU( ), i.e., an antiunitary representation. More precisely, for a pair H (G, G ), where G G is a subgroup of index 2, a homomorphism U : G AU( ) is 1 1 ⊆ → H called an antiunitary representation of (G, G ) if U is antiunitary for g G . If G is 1 g 6∈ 1 a topological group with two connected components, then we obtain a canonical group pair by G1 := G0 (the identity component). In this case an antiunitary representation of G is a continuous homomorphism U : G AU( ) mapping G G into antiunitary → H \ 0 operators. In the mathematical literature on representations, antiunitary operators have never been in the focus, whereas in quantum physics one is forced to consider antiunitary op- erators to implement a time-reversal symmetry ([Wig59]). If the dynamics of a quantum itH system is described by a unitary one-parameter group Ut = e , where the Hamiltonian H is unbounded and bounded from below, then a unitary time reversal operator would T lead to the relation H = H, which is incompatible with H being bounded from be- T T − low. This problem is overcome by implementing time reversal by an antiunitary operator because it imposes no restrictions on the spectrum of the Hamiltonian. In particular, the PCT Theorem in Quantum Field Theory (QFT) which concerns the implementation of a symmetry reversing parity (P), charge (C) and time (T), leads to an extension of a uni- Rd ⋊ R tary representation of the Poincar´egroup P (d)+↑ = SO1,d 1( )↑ to an antiunitary d − representation of the larger group P (d)+ = R ⋊ SO1,d 1(R) ([Ha96, Thm. II.5.1.4]). ∼ − In the modular theory of operator algebras one studies pairs ( , Ω) consisting of M a von Neumann algebra B( ) and a cyclic separating unit vector Ω . Then M ⊆ H ∈ H S(MΩ) := M ∗Ω for M defines an unbounded antilinear involution, and the polar ∈ M 1/2 decomposition of its closure S = J∆ leads to a positive selfadjoint operator ∆ = S∗S, 1 an antiunitary involution J satisfying the modular relation J∆J = ∆− , and αt(M) := it it ∆ M∆− defines automorphisms of (see [BR87] and 4.1). In particular, we are M § naturally led to antiunitary symmetries. We say that (∆,J) is a pairs of modular objects 4 K.-H. NEEB AND G. OLAFSSON´ if J is a conjugation and ∆ a positive selfadjoint operator satisfying the modular relation. To connect this with QFT, we recall the notion of a Haag–Kastler net of C∗-sub- algebras ( ) of a C∗-algebra , associated to (bounded) regions in d-dimensional A O A O Minkowski space. The algebra ( ) is interpreted as observables that can be measured in A O the “laboratory” . Accordingly, one requires isotony, i.e., that implies ( ) O O1 ⊆ O2 A O1 ⊆ ( ) and that the ( ) generate . Causality enters by the locality assumption that A O2 A O A ( ) and ( ) commute if and are space-like separated, i.e., cannot correspond A O1 A O2 O1 O2 with each other (cf. Example 5.4). Finally one assumes an action σ : P (d)↑ Aut( ) + → A of the connected Poincar´egroup such that σ ( ( )) = (g ). Every Poincar´einvariant g A O A O state ω of the algebra now leads by the GNS construction to a covariant representation A (π , , Ω) of , and hence to a net ( ) := π ( ( ))′′ of von Neumann algebras on ω Hω A M O ω A O ω. Whenever Ω is cyclic and separating for ( ), we obtain modular objects (∆ ,J ). H M O O O This connection between the Araki–Haag–Kastler theory of local observables and modular theory leads naturally to antiunitary group representations (cf. Section 5). The starting point for the recent development that led to fruitful applications of mod- ular theory in QFT was the Bisognano–Wichmann Theorem, asserting that, the modular it it automorphisms α (M) = ∆ M∆− corresponding to the algebra (W ) of observables t M corresponding to a wedge domain W in Minkowski space (cf. Definition 4.11) are im- plemented by the unitary action of a one-parameter group of Lorentz boosts preserving W ([BW76]). This geometric implementation of modular automorphisms in terms of Poincar´etransformations was an important first step in a rich development based on the work of Borchers and Wiesbrock in the 1990s [Bo92, Bo95, Bo97, Wi92, Wi93, Wi93c]. They managed to distill the abstract essence from the Bisognano–Wichmann Theorem which led to a better understanding of the basic configurations of von Neumann al- gebras in terms of half-sided modular inclusions and modular intersections. This im- mediately led to very tight connections between the geometry of homogeneous spaces and modular theory [BGL93]. In his survey [Bo00], Borchers described how these con- cepts have revolutionized quantum field theory. Subsequent developments can be found in [Tr97, Sch97, Ar99, BGL02, Lo08, JM17]; for the approach to Quantum Gravity based on Non-commutative Geometry and Tomita–Takesaki Theory, see in particular [BCL10]. A key insight that simplifies matters considerably is that modular objects (∆,J) associated to a pair ( , Ω) of a von Neumann algebra and a cyclic separating vector M M Ω are completely determined by the real subspace

V := hΩ, where h = M : M ∗ = M . M M M { ∈M } It satisfies V iV = 0 and V + iV is dense in . Closed real subspaces V M ∩ M { } M M H ⊆ H with these two properties are called standard. Every standard subspace V determines by the polar decomposition of the closed operator S defined on V +iV by S (x+iy)= x iy V V − a pair (∆V ,JV ) of modular objects and, conversely, any such pair (∆,J) determines a standard subspace as the fixed point space of J∆1/2 (see Section 3). We refer to [Lo08] for an excellent survey on this correspondence. In QFT, standard subspaces provide the basis for the technique of modular localization, developed by Brunetti, Guido and Longo in [BGL02]. For some applications we refer to [Sch97, MSY06, Sch06, LW11, Ta12, LL14, Mo17]. ANTIUNITARY REPRESENTATIONS 5

From the perspective of antiunitary representations, standard subspaces V with mod- ular objects (∆,J) are in one-to-one correspondence with antiunitary representations

it/2π U : R× AU( ) by U 1 = J and Uet = ∆− (1) → H − (Proposition 3.2). Accordingly, antiunitary representations (U, ) of the affine group H Aff(R) = R⋊R× correspond to one-parameter families of standard subspaces (Vx)x R, ∼ ∈ where Vx corresponds to the affine stabilizer group of x. Borchers’ key insight was that the positive energy condition on the representation of the translation group is intimately related to inclusions of these subspaces. More precisely, U = eitP satisfies P 0 if (t,1) ≥ and only if U V V holds for all t 0 ( 3.4). This leads to Borchers pairs (V,U) (t,1) 0 ⊆ 0 ≥ § of a standard subspace V and a unitary one-parameter group (Ut)t R, a concept that ∈ is equivalent to the so-called half-sided modular inclusions V V of pairs of standard 1 ⊆ 2 subspaces, which was condensed from the corresponding concept of a half-sided modular inclusion of von Neumann algebras ( 3.4,4.2). §§ The main objective of this article is to describe certain structures arising in QFT, such as nets of von Neumann algebras and standard subspaces, from the perspective of antiunitary group representations. Since any standard subspace V corresponds to a representation of R× and inclusions of standard subspaces correspond to antiunitary positive energy representations of Aff(R), it is very likely that a better understanding of antiunitary representations and corresponding families of standard subspaces provides new insight into the geometric structures underlying QFT. This article is written from a mathematical perspective and we are rather brief on the concrete physical aspects mentioned in 5.2. We tried to describe the mathematical side of the theory as clearly as § possible to make it easier for mathematicians to understand the relevant aspects without going to much into physics. For more details of the physical side, we recommend [BDFS00, BGL02, Lo08, LL14]. In particular, the programs outlined by Borchers and Wiesbrock, see f.i., [Bo97, Bo00], [Wi93c], leave much potential for an analysis from the representation theoretic perspective. The structure of this paper is as follows. In Section 2 we discuss antiunitary represen- tations of group pairs (G, G1) and criteria for a unitary representation of G1 to extend to an antiunitary representation of G. An interesting simplifying feature is that, when- ever antiunitary extensions exist, they are unique up to equivalence (Theorem 2.11). We show that irreducible unitary representations of G1 fall into three types (real, complex and quaternionic) with respect to their extendability behavior to antiunitary representa- tions of G. We also take a closer look at antiunitary representations of one-dimensional Lie groups ( 2.3). Here R× plays a central role because its antiunitary representations § encode modular objects (∆,J) as in (1). We conclude Section 2 with a discussion of an- tiunitary representations of the affine group Aff(R), the projective group PGL2(R) and the 3-dimensional Heisenberg group Heis(R2). Section 3 is devoted to various aspects of standard subspaces as a geometric counter- part of antiunitary representations of R×. In particular, we discuss how the embedding V can be obtained from the orthogonal one-parameter group ∆it on V ( 3.3), and ⊆ H |V § in 3.4 we discuss half-sided modular inclusions of standard subspaces and how they are § 6 K.-H. NEEB AND G. OLAFSSON´ related to antiunitary representations of Aff(R), P (2)+ and PGL2(R). In Section 4 we first recall some of the key features of Tomita–Takesaki Theory. 4.2 § is of key importance because it is devoted to the translation between pairs ( , Ω) of von M Neumann algebras with cyclic separating vectors and standard subspaces V . We have already seen how to obtain a standard subspace V = hΩ from ( , Ω). Conversely, M M M one can use Second Quantization (see Section 6 for details) to associate to each stan- dard subspace V pairs ( (V ), Ω), where (V ) is a von Neumann algebra on ⊆ H R± R± the (bosonic/fermionic) Fock space ( ). This method has been invented and studied F± H thoroughly by Araki and Woods in the 1960s and 1970s in the context of free bosonic quantum fields ([Ar64, Ar99, AW63, AW68]); some of the corresponding fermionic re- sults are more recent (cf. [EO73], [BJL02]) and other statistics (anyons) are discussed in [Sch97]. A central point is that these correspondences permit to translate between results on configurations of standard subspaces and configurations of von Neumann algebras with common cyclic vectors. We explain this in detail for half-sided modular inclusions and Borchers pairs ( 4.2 and 4.3) but we expect it to go much deeper. Keeping in mind that §§ standard subspaces are in one-to-one correspondence with antiunitary representations of R× and half-sided modular inclusions with antiunitary positive energy representations of Aff(R), we expect that many interesting results on von Neumann algebras can be obtained from a better understanding of antiunitary representations of Lie group pairs (G, G ) and configurations of homomorphisms γ : (R×, R×) (G, G ). The construction 1 + → 1 of free fields by second quantization associates to an antiunitary representation (U, ) H of G on the one-particle spaces , resp., to the corresponding standard subspaces V , a H γ net of Neumann algebras on Fock space. However, there is also a converse aspect which is probably more important, namely that the passage from pairs ( , Ω) to the standard M subspaces V is not restricted to free fields and can be used to attach geometric structure M to nets of von Neumann algebras, all encoded in the subgroup of AU( ) generated by H all operators ∆it and J . To substantiate this remark, we discuss in Section 5 several M aspects of netsM of standard subspaces and von Neumann algebras as they arise in QFT. In particular, we consider nets of standard subspaces (Vℓ)ℓ L arising from antiunitary ∈ representations (U, ), which leads to the covariance relation U V = V for g G , H g ℓ g.ℓ ∈ 1 and one expects geometric information to be encoded in the G-action on the index set L. A common feature of the natural examples is that L has a fibration over a symmetric space that corresponds to the projection (∆ ,J ) J , forgetting the modular operator. ℓ ℓ 7→ ℓ For details we refer to the discussion of several examples in Section 5. Typical index sets g g g 1 L arise as conjugation orbits γ , (γ∨) : g G Hom(R×, G), where γ (t)= gγ(t)g− 1 { ∈ }⊆ and γ∨(t) = γ(t− ). In this picture, the above projection simply corresponds to the evaluation map ev 1 : Hom(R×, G) Inv(G) and the set Inv(G) of involutions of G is − → a symmetric space ([Lo69]; Appendix A.3). In many concrete situations, the centralizer of γ( 1) in G coincides with the centralizer of the whole subgroup γ(R×), so that the − g conjugacy class Cγ = γ : g G can be identified with the conjugacy class Cγ( 1) of { ∈ } − the involution γ( 1), and this manifold is a symmetric space. We are therefore led to − index sets which are ordered symmetric spaces, and these objects have been studied in ANTIUNITARY REPRESENTATIONS 7 detail in the 90s. We refer to the monograph [HO96]´ for a detailed exposition of their structure theory. Section 6 presents the second quantization process from standard subspaces V ⊆ H to pairs ( ±(V ), Ω) in a uniform way, stressing in particular the similarity between the R bosonic and the fermionic case. In the final Section 7 we briefly describe some perspectives and open problems. An- tiunitary representations occur naturally for interesting classes of groups such as the Virasoro group, conformal and affine groups related to euclidean Jordan algebras, and automorphism groups of bounded symmetric domains. For detailed results we refer to the forthcoming paper [NO17].´ In 7.6 we also explain how second quantization leads to § interesting dual pairs in the Heisenberg group Heis( ): Any standard subspace V H ⊆ H satisfying the factoriality condition V V ′ = 0 , where V ′ is the symplectic orthog- ∩ { } onal space, leads by restriction of the irreducible Fock representation of Heis( ) to a H factor representation of the subgroup Heis(V ), which forms a dual pair with Heis(V ′) in Heis( ) (both subgroups are their mutual centralizers). So far, such dual pairs have not H been exploited systematically from the perspective of unitary representations of infinite dimensional Lie groups. Some basic auxiliary lemmas and definitions have been collected in the appendix.

Notation and conventions: As customary in physics, the scalar product , on a h· ·i complex Hilbert space is linear in the second argument. H JSK denotes the closed subspace of a Hilbert space generated by the subset S. H a,b := ab + ba is the anti-commutator of two elements of an associative algebra. { } For the cyclic group of order n we write Zn = Z/nZ.

Rd 1 d 1 For x, y − , we write xy = j=1− xj yj for the scalar product and, for x = d ∈ (x0, x) R , we write [x, y] = x0y0 Pxy for the Lorentzian scalar product on the d- ∈ R1,d 1 −Rd dimensional Minkowski space − ∼= . The light cone in Minkowski space is denoted 1,d 1 V = x R − : x > 0, [x, x] > 0 . + { ∈ 0 } Here is our notation for some of the groups arising in physics: R1,d 1 ⋊ R R1,d 1 the Poincar´egroup P (d) ∼= − O1,d 1( ) of affine isometries of − , • 1,d 1 − P (d)+ = R − ⋊ SO1,d 1(R) is the subgroup of orientation preserving maps, and • 1,d 1 − P (d)↑ = R − ⋊ O1,d 1(R)↑ with O1,d 1(R)↑ = g O1,d 1(R): gV+ = V+ the • − − { ∈ − } subgroup preserving the causal structure. The corresponding conformal group is O2,d(R), acting on the conformal compacti- • 1 d 1 d fication S S − of M with the kernel 1 (see [HN12, 17.4]). × {± } § If not otherwise states, all Lie groups in this paper are finite dimensional.

2. Antiunitary representations. In this section we discuss antiunitary representa- tions of group pairs (G, G1) and criteria for a unitary representation of G1 to extend to an antiunitary representation of G. We start in 2.1 with some general remarks on § group pairs (G, G1) and how to classify twists in this context. We also take a closer look 8 K.-H. NEEB AND G. OLAFSSON´ at antiunitary representations of one-dimensional Lie groups in 2.3 and discuss antiuni- § tary representations of the affine group Aff(R), the projective group PGL2(R) and the 3-dimensional Heisenberg group in 2.4. § Definition 2.1. An antiunitary representation (U, ) of a group pair (G, G ), where H 1 G1 G is a subgroup of index 2, is a homomorphism U of G into the group AU( ) of uni- ⊆ H1 tary or antiunitary operators on a complex Hilbert space for which G = U − (U( )), H 1 G H i.e., G is represented by unitary operators and the coset G G by antiunitary operators. 1 \ 1 If G is a Lie group, then (G, G1) is called a Lie group pair. If G is a topological group with two connected components, then we obtain a canonical group pair by G1 := G0 (the identity component). In this case an antiunitary representa- tion of G is a continuous homomorphism U : G AU( ) mapping G G into antiunitary → H \ 0 operators. We start this section with a discussion of the natural class of group pairs that will show up in the context of antiunitary representations.

2.1. Involutive group pairs.

Definition 2.2. An involutive group pair is a pair (G, G1) of groups, where G1 G is 2 ⊆ a subgroup of index 2 and there exists an element g G G1 with g Z(G1). Then 1 ∈ \ ∈ τ(g1) := gg1g− defines an involutive automorphism of G1. In most examples that we encounter below G is a Lie group with two connected components and G1 is its identity component. Remark 2.3. (a) If g2 Z(G ), then other elements gh gG need not have central ∈ 1 ∈ 1 squares. From (gh)2 = ghgh = g2τ(h)h it follows that (gh)2 is central if and only if 1 τ(h)h Z(G ), which is in particular the case if τ(h)= h− . ∈ 1 (b) If G is a Lie group, then any conjugacy class C of g G G with g2 Z(G) g ∈ \ 1 ∈ carries a natural symmetric space structure (Appendix A.3). In fact, the stabilizer of g τ in G1 is G1 , so that we obtain a diffeomorphism τ τ 1 1 G /G C , hG hgh− = hτ(h)− g. 1 1 → g 1 7→ Examples 2.4. (a) Let be a complex Hilbert space and (G, G ) := (AU( ), U( )). An H 1 H H antiunitary operator J AU( ) is called a conjugation if J 2 = 1 and an anticonjugation ∈ H if J 2 = 1. Conjugations always exist and define a on in the sense − H that J = Fix(J) := ker(J 1) is a real Hilbert space whose complexification is .1 H − H Anticonjugations define on a quaternionic structure, hence do not exist if is of finite H H odd dimension. Any (anti-)conjugation J on is contained in G G and satisfies J 2 1 H \ 1 ∈ {± } ⊆ Z(U( )). H (b) If G is a group and τ Aut(G ) is an involutive automorphism, then 1 ∈ 1 G := G ⋊ 1, τ defines an involutive group pair. 1 { } 1 For the existence, fix an orthonormal basis (ej )j∈I of H and defined J to be antilinear with Jej = ej for every j ∈ I. ANTIUNITARY REPRESENTATIONS 9

Example 2.5. (A non-involutive group pair) Let σ : C Aut(C) denote the natural 4 → action of the subgroup C = 1, i T by multiplication and form the semidirect 4 {± ± } ⊆ product group G := C⋊ C . Then G := C⋊ 1 is a subgroup of index 2 but no σ 4 1 σ {± } element g G G satisfies g2 Z(G ) because g2 acts on C as idC. ∈ \ 1 ∈ 1 − Remark 2.6. (Classification of involutive group pairs) (a) Suppose we are given a group G and an involutive automorphism τ of G. We want to classify all group extensions 1 G G♯ Z 1, → → → 2 → where the corresponding involution in the group Out(G) of outer automorphisms of G is represented by τ. In view of [HN12, Thm. 18.1.13], the equivalence classes of these 2 Z extensions are parametrized by the cohomology group Hτ ( 2,Z(G)), where 1 acts on Z(G) by τ . As any cocycle f : Z Z Z(G) normalized by f(0,g)= f(g, 0) = e is |Z(G) 2 × 2 → determined by the element z := f(1, 1) Z(G) because all other values vanish, the group ∈ structure on the corresponding extension is given by an element τ G♯ G satisfying ∈ \ 2 1 τ = z and τgτ − = τ(g) for g G. ∈ b This description showsb in particular thatb b τ(z) = z, and a closer inspection of the coho- mology groups yields H2(Z ,Z(G)) = Z(G)τ /Z(G) , where Z(G) := τ(z)z : z Z(G) (2) τ 2 ∼ τ τ { ∈ } ([HN12, Ex. 18.3.5]). (b) For τ = id we have |Z(G) Z(G) Z(G) = z2 : z Z(G) and H2(Z ,Z(G)) = Z(G)/Z(G) . τ { ∈ } τ 2 ∼ τ (c) For τ = id , we have |Z(G) − Z(G) H2(Z ,Z(G)) = Z(G)τ = z Z(G): z2 = e , τ 2 ∼ { ∈ } the subgroup of central involutions. Remark . 2 Z 2.7 Although by (2) the cohomology groups Hτ ( 2,Z(G)) are elementary abelian two groups, one cannot expect any bound on the order of an element g G♯ G. ♯ ♯ ∈ n \ In the cyclic group G = Z n with G = Z n−1 , any element of G G is of order 2 . 2 ∼ 2 \ Example . R 2 Z R 2.8 (a) For G = , Remark 2.6 implies that Hτ ( 2, )= 0 for any involu- ♯ R⋊ Z { } tive automorphism τ. This implies that G ∼= τ 2. 1 (b) For G = T, the cohomology is trivial for τ = idT, but for τ(z)= z− the group H2(Z , T) = z T: z2 =1 = 1 τ 2 ∼ { ∈ } {± } is non-trivial. A concrete model for the non-trivial extension with τ 2 = 1 is given by − the subgroup b Pin (R) = exp(RI) J exp(RI) H×, 2 ∪ ⊆ where I and J are the two generators of the skew-field H of quaternions satisfying I2 = J 2 = 1 and IJ = JI ([HN12, Ex. B.3.24]). This is a 1-dimensional Lie group without − − a simply connected covering group ([HN12, Ex. 18.2.4])

Examples 2.9. Here are some concrete involutive group pairs (G, G1) that we shall be dealing with. 10 K.-H. NEEB AND G. OLAFSSON´

R R⋊R R⋊R 2 (a) G = Aff( ) ∼= × with G1 ∼= +×, the identity component. Here rx = 1 holds for the reflections rx = (2x, 1) in x R. − ∈ R P R (b) The automorphism group G = PGL2( ) of the real projective line 1( ) ∼= R , where G = PSL (R) is the identity component and reflections in GL (R) lead ∪{∞} 1 2 2 to orientation reversing involutions of S1. 1,d 1 (c) The Poincar´egroup P (d)= R − ⋊ O1,d 1(R) of d-dimensional Minkowski space 1,d 1 1,d 1 − R − contains the subgroup P (d)+ = R − ⋊ SO1,d 1(R) of orientation preserving − affine isometries. Then we obtain the involutive group pair (G, G1) with G := P (d)+ and G := P (d)↑ . In the following the involution R := diag( 1, 1, 1,..., 1) G G plays 1 + 01 − − ∈ \ 1 an important role (cf. Lemma 4.12). (d) For a (bounded) symmetric domain Cn, the group Aut( ) of biholomorphic D⊆ D automorphisms is an index 2-subgroup of the hermitian group AAut( ) of all bijections D of that are either holomorphic or antiholomorphic. There always exist antiholomorphic D involutions σ in AAut( ) (see [Ka97] for a classification covering even the infinite dimen- D sional case). For any such involution σ, we obtain by G := Aut( ) and G := G ⋊ 1, σ 1 D 0 1 { } an involutive group pair (cf. [NO17]´ and 7.3). §

2.2. Extending unitary representations. Suppose that G1 is an index two subgroup of the group G and (U, ) is a unitary representation of G . In this subsection we discuss H 1 extensions of U to antiunitary representations of G. In particular, we show that, in analogy to the classical case G = G Z , irreducible antiunitary representations fall into three 1 × 2 types that we call real, complex and quaternionic, according to their commutant. We start with the following lemma on a situation where extensions always exist be- cause the representation has been doubled in a suitable way. Lemma 2.10. (Extension Lemma) Let G G be a subgroup of index two and (U, ) be 1 ⊆ H a unitary representation of G1. Fix r G G1 and consider the automorphism τ(g) := 1 ∈ \ rgr− of G . Then the unitary representation V := U U ∗ τ on ∗ extends to an 1 ⊕ ◦ H ⊕ H antiunitary representation of G.

Proof. Let Φ: ∗, Φ(v)(w) := v, w denote the canonical antiunitary operator and H → H h i note that U ∗ Φ=Φ U for g G . We consider the antiunitary operator g ◦ ◦ g ∈ 1 1 J : ∗ ∗, J(v, λ) := (Φ− λ, ΦU 2 v). H ⊕ H →H⊕H r It satisfies 2 1 1 J (v, λ)= J(Φ− λ, ΦUr2 v) = (Ur2 v, ΦUr2 Φ− λ) = (Ur2 v,Ur∗2 λ)= Vr2 (v, λ), 2 2 2 where we have used τ(r )= r for the last equality. This proves that J = Vr2 . We now 1 show that JV J − = V for g G: g τ(g) ∈ 1 1 JVg(v, λ)= J(Ugv,Uτ∗(g)λ)=(Φ− Uτ∗(g)λ, ΦUr2 Ugv) = (Uτ(g)Φ− λ, ΦUτ 2(g)Ur2 v) 1 1 = (Uτ(g)Φ− λ, Uτ∗2(g)ΦUr2 v)= Vτ(g)(Φ− λ, ΦUr2 v)= Vτ(g)J(v, λ). 2 1 The relations J = Vr2 and JVgJ − = Vτ(g) now imply by direct calculation that the as- signment V := V J for g G defines an extension of V to an antiunitary representation gr g ∈ 1 of G (Lemma A.4). ANTIUNITARY REPRESENTATIONS 11

The following theorem implies that extensions of unitary representations of G1 to antiunitary representations (U, ) of G are always unique up to isomorphism. It also H describes the situation for irreducible representations. Note that the commutant

U ′ = A B( ): ( g G) AU = U A G { ∈ H ∀ ∈ g g } is not a complex subalgebra of B( ) because some U are antilinear. H g 1 Theorem 2.11. Let G G be a subgroup of index two, r G G and τ(g) := rgr− 1 ⊆ ∈ \ 1 for g G . ∈ 1 (a) For two antiunitary representation (U j , ) , we then have Hj j=1,2 U 1 = U 2 U 1 = U 2 . ∼ ⇐⇒ |G1 ∼ |G1 (b) For any antiunitary representation (U, ) of (G, G ), the von Neumann algebra H 1 UG′ 1 is the complexification of the real algebra UG′ . (c) An antiunitary representation (U, ) of (G, G1) is irreducible if and only if its R HC H commutant UG′ is isomorphic to , or . More specifically:

(i) If U ′ = R, then U ′ = C and U is irreducible. G ∼ G1 ∼ |G1 2 (ii) If U ′ = C, then U ′ = C and U is a direct sum of two inequivalent irre- G ∼ G1 ∼ |G1 ducible representations which do not extend to an antiunitary representation of G.

(iii) If U ′ = H, then U ′ = M (C) and U is a direct sum of two equivalent ir- G ∼ G1 ∼ 2 |G1 reducible representations which do not extend to an antiunitary representation of G. (d) For an irreducible unitary representation (U, ) of G , either H 1 (i) U extends to an antiunitary representations U of G, and then U is irreducible ′ R with U G ∼= ; or (ii) U does not extend to an antiunitary representation of G. Then V := U U ∗ τ ⊕ C ◦ extends to an irreducible antiunitary representation of G and VG′ ∼= if U ∗ τ = U and V ′ = H if U ∗ τ = U. ◦ 6∼ G ∼ ◦ ∼ 2 Proof. (a) Let Φ: 1 2 be a unitary intertwining operator for the representations j H → H j U G1 . Pick r G G1 and consider the antiunitary operators Jj := Ur AU( j ). | ∈ \ 1 1 1 ∈ H Then the unitary operator U := J1− Φ− J2Φ U( 1) commutes with UG1 . The map 1 ∈ H j1(M) := J1MJ1− defines an antilinear automorphism of the von Neumann algebra 1 (UG1 )′ satisfying 1 1 1 1 2 1 1 1 1 1 1 1 1 j1(U)=Φ− J2ΦJ1− =Φ− J2− Ur2 ΦJ1− =Φ− J2− ΦUr2 J1− =Φ− J2− ΦJ1 = U − . 1 Therefore Lemma A.1(c) implies the existence of a unitary operator V (UG1 )′ with 2 1 1 ∈ V = U − and j (V )= V − . With Ψ := Φ V , this leads to 1 ◦ 1 1 1 1 1 1 1 Ψ− J2Ψ= V − Φ− J2ΦV = V − J1UV = V − U − J1V = VJ1V = V V − J1 = J1. We conclude that the antiunitary representations U 1 and U 2 are equivalent.

2In the finite dimensional context, this was already known to E. Wigner; see [Wig59, p. 344]. 12 K.-H. NEEB AND G. OLAFSSON´

(b) Let J := Ur. Then UG′ 1 is invariant under the antilinear automorphism j(M) := 1 2 JMJ − . Since J commutes with UG′ 1 , it is involutive. As UG′ is the set of fixed points of j, it is a real form of the complex UG′ 1 . This implies the assertion. (c) The closed complex subspaces invariant under UG are precisely the closed real R subspaces of the underlying real space invariant under the group T UG. Therefore H R · U is irreducible if and only if the real representation of T U on is irreducible, which G · G H is equivalent to its commutant being isomorphic to R, C or H ([StVa02, Thm. 1]). Next we observe that the real linear commutant of T1 consists of the complex linear operators. T Therefore the real linear commutant of UG equals the complex linear commutant UG′ . · 2 Now (b) implies that U ′ = R, C, H leads to U ′ = C, C ,M (C), respectively. G ∼ G1 ∼ 2 In the first case U G1 is irreducible. In the second case = + , where | H ∼ H ⊕ H− H± are G1-invariant subspaces on which the G1-representations are irreducible and non- equivalent. As Ur permutes the G1-isotypical subspaces, Ur = . For the represen- + H± +H∓ tations U ± of G1 on , this implies that U − = (U )∗ τ. If U or U − extends to an H± ∼ ◦ antiunitary representation of G, then U has an extension to a reducible representation |G1 U of G. As U is irreducible, this contradicts (a). In the third case we have a similar decomposition with U + U . Again, the irreducibility of U, combined with (a), implies e ∼= − that U ± do not extend to G. (d) If U extends to an antiunitary representation U of G on the same space, then this ′ R representation is obviously irreducible and (c) implies that U G ∼= . If such an extension does not exist, then the Extension Lemma 2.10 provides an extension of V := U U ∗ τ ⊕ ◦ to an antiunitary representation of G by 1 Vr := J, where J(v, λ) := (Φ− λ, Ur∗2 Φv). 2 If U ∗ τ = U, then V ′ = C , and if U ∗ τ = U, then V ′ = M (C). ∼ G1 ∼ ∼ G1 ∼ 2 ◦ 6 ◦ 2 In the first case the algebra V ′ = C acts by diagonal operators T (v, λ) := G1 ∼ (a,b) C (av,bλ). Such an operator commutes with J if and only if b = a. Therefore VG′ ∼= , and thus the representation V of G is irreducible. In the second case, V ′ is a real form of V ′ = M (C). We show that the repre- G G1 ∼ 2 sentation (V, ∗) of G is irreducible. If this is not the case, there exists a proper H ⊕ H G-invariant subspace ∗. As V = U U, the G -representation on must K⊆H⊕H |G1 ∼ ⊕ 1 K be irreducible and equivalent to U. This contradicts the non-extendability of U to an H antiunitary representation of G. Therefore V is irreducible and VG′ is isomorphic to . Definition 2.12. (Three types of irreducible representations 3 ) We keep the notation of the preceding theorem. If (U, ) is an irreducible unitary representation of G1 with H 1 U = U ∗ τ, then there exists a Φ AU( ) with ΦU Φ− = U −1 for g G . By ∼ ◦ ∈ H g rgr ∈ 1 Schur’s Lemma, such an operator Φ is unique up to a scalar factor in T, so that Φ2 does not depend on the concrete choice of Φ. Therefore an antiunitary extension to G exists if 2 and only if Φ = U 2 . Then we call (U, ) of real type (with respect to τ). If this is not the r H case, but U = U ∗ τ, then (U, ) is said to be of quaternionic type (with respect to τ), and ∼ ◦ H otherwise we say that it is of complex type (with respect to τ). This terminology matches

3In a special context, this classification by three types can already be found in Wigner’s book [Wig59, §26, p. 343]. ANTIUNITARY REPRESENTATIONS 13 the type of the commutant of the corresponding irreducible antiunitary representation of G. Example 2.13. (a) If = C is one-dimensional, then AU( )= T 1,J = O (R) for any H H { } ∼ 2 conjugation J. We conclude in particular that all antiunitary operators are involutions. (b) If = C2 is two-dimensional, we can already see all types of situations for groups H generated by a single antiunitary operator, i.e., for antiunitary representations of the pair (G, G1) = (Z, 2Z). Let J AU( ) be antiunitary and J 2 U( ) be its square. If J 2 = 1, then J ∈ H ∈ H is a conjugation, so that there are proper J-invariant subspaces. If J 2 = 1, then J C2 H − is an anticonjugation defining a quaternionic structure on ∼= . In particular, the representation is irreducible with U ′ = H and U ′ = B( ) = M (C). G ∼ G1 H ∼ 2 Assume that J 4 = 1. Then J 2 is not an involution, so that it has an eigenvalue λ = 1. 6 6 ± If λ is the corresponding eigenspace, then J λ = λ, so that Spec(J 2) = λ, λ . H H H λ {λ } Choosing an orthonormal basis e1,e2 such that e1 and e2 := Je1 , we 2 ∈ H ∈ H obtain Je2 = J e1 = λe1, so that J is determined up to equivalence. The corresponding 2 2 representation on C is irreducible with U ′ = C and U ′ = C (Theorem 2.11(c)). G1 ∼ G ∼ Example 2.14. (a) For G = G Z , the concepts of real/complex/quaternionic type 1 × 2 coincides with the classical definition for G1, as the characterization in Theorem 2.11 shows. (b) For G = G ⋊ 1, τ and τ 2 = id , the extendability of an irreducible unitary 1 { } G1 representation (U, ) of G is equivalent to the existence of a conjugation J AU( ) H 1 ∈ H satisfying JU J = U for g G . g τ(g) ∈ 1 If U ∼= U ∗ τ, then a J AU( ) satisfying JUgJ = Uτ(g) for g G1 exists and 2 ◦ ∈ 2 H 2 2 ∈ J U ′ = C1, together with JJ J = J imply J 1 . Accordingly, U is of real, ∈ G1 ∈ {± } resp., quaternionic type if J 2 = 1, resp., J 2 = 1. − Example 2.15. (a) For G = Z and G = e , Theorem 2.11)(a) reproduces the fact 2 1 { } that all conjugations on are conjugate under U( ). H H (b) For G = Z = Z/4Z and G = 0, 2 , the case of antiunitary representations with 4 1 { } U2 = 1 likewise implies that all anticonjugations are conjugate under U( ). − C C2 HH (c) The irreducible unitary representation of G = SU2( ) on ∼= (by left mul- tiplication) is of quaternionic type. The complex structure on H is defined by the right multiplication with I. Then Φ(a) = aJ defines a G-equivariant anticonjugation on C2. Therefore the representation is of quaternionic type.

(d) For any compact connected Lie group G1, the irreducible unitary representations (U , ) are classified in terms of their highest weights λ with respect to a maximal λ Hλ torus T G , resp., by the orbits λ under the Weyl group . As λ is the Weyl ⊆ 1 W W −W group orbit of the dual representation, U is self-dual if and only if λ λ ([BtD85, λ − ∈ W Prop. VI.4.1]). It is of real, resp., quaternionic type if and only if an invariant symmetric, resp., skew-symmetric exists ([BtD85, Prop. II.6.4]), and this can also be read from the highest weight ([BtD85, Prop. VI.4.6]).

Further, for any automorphism σ Aut(G1), there exists an inner automorphism σ′ ∈τ such that τ := σσ′ preserves T . Then λ := λ τ is an extremal weight of U τ = U σ, ◦ |T λ ◦ ∼ λ ◦ 14 K.-H. NEEB AND G. OLAFSSON´

τ so that U ∗ τ = U if and only if λ λ. λ ◦ ∼ λ − ∈W The following lemma shows that, if only G1 and an involutive automorphism τ of G1 are given, then there always exists an extension to a group of the type G1 ⋊α Z4, where α = τ. This issue is already discussed in Wigner’s book [Wig59, 26, p. 329], where 1 § J 2 = 1 is related to spin being integral or half-integral. ± Lemma 2.16. Let (U, ) be a unitary representation of the group G1 and τ Aut(G1) H 4∈ be an involution. If U τ ∼= U ∗, then there exists a J AU( ) with J = 1 and 1 ◦ ∈ H JU J − = U for g G . g τ(g) ∈ 1 1 Proof. From U τ ∼= U ∗ we obtain a J AU( ) with JUgJ − = Uτ(g) for g G1. As 2 ◦ 2 ∈ H ∈ τ = idG1 , the unitary operator J commutes with UG1 . We therefore have a G1-invariant 2 orthogonal decomposition = + , where = ker(J + 1) and + = ⊥. Since H H ⊕ H− H− H H− both subspaces are invariant under G1 and J, we may w.l.o.g. assume that = 0 H− { } and show that there exists a conjugation commuting with G1. Conjugating with J defines an antilinear automorphism of the von Neumann algebra 2 := UG′ 1 fixing the unitary element J . Therefore Lemma A.1 implies the existence of M 2 2 1 a unitary A UG′ 1 with JAJ = A and A = J . Replacing J by J := A− J, we obtain 2 ∈ J = 1. e 1 Lemmae 2.17. Let G be an abelian group, τ(g)= g− and G := G ⋊ 1, τ . Then every 1 1 { } unitary representation of G1 extends to an antiunitary representation of G.

Proof. We consider G1 as a discrete group, so that any unitary representation (U, ) of 2 H G1 is a direct sum of cyclic representations of the form (V,L (G1,µ)), where (Vg f)(χ)= 2 χ(g)f(χ). Then Jf := f defines a conjugation on L (A,µ) with JV J = V −1 , so that we c g g obtain an extension of V to an antiunitary representation of G. b Lemma 2.18. Suppose that G = G ⋊ id, τ , where τ Aut(G) is an involution. ∼ 1 { } ∈ (i) If (U, ) is an irreducible antiunitary representations of G and x gτ satisfies H ∈ idU(x) 0, then dU(x)=0. − ≥ (ii) If (U, ) is an irreducible unitary representations of G and x gτ satisfies H 1 ∈ idU(x) 0 and dU(x) = 0, then U ∗ τ = U, i.e., U is of complex type with − ≥ 6 ◦ 6∼ respect to τ. Proof. (i) The conjugation U on satisfies U idU(x)U = idU(τx) = idU(x), so τ H τ τ − − that the positivity assumption implies dU(x) = 0. (ii) From (i) it follows that U does not extend to an antiunitary representation of G. By Theorem 2.11(d)(ii), V := U U ∗ τ extends to an irreducible antiunitary representation ⊕ ◦ of G. If V ′ = H, then U ∗ τ = U implies idV (x) 0, so that dV (x) = 0 by (i), and G ∼ ◦ ∼ − ≥ this contradicts dU(x) = 0. We conclude that V ′ = C and U ∗ τ = U. 6 G ∼ ◦ 6∼ Remark 2.19. In [OM16] the authors study a concept of a “Wigner elementary relativis- tic system” which is defined as a faithful irreducible orthogonal representation (U, ) of K the proper orthochronous Poincar´egroup G := P (4)↑ on a real Hilbert space . Writing + K (Pj )0 j 3 for the skew-adjoint generators of the unitary representation of the translations ≤ ≤ e ANTIUNITARY REPRESENTATIONS 15

e tPj groups Utej = e , the mass squared operator is defined as 3 M 2 := P 2 + P 2. − 0 j Xj=1 e e One of the main results in [OM16] is that if M 2 0, then carries a complex structure ≥ K I commuting with the image of U ([OM16, Thm. 4.3, Thm. 5.11]). This result can be obtained quite directly in our context. We consider the complex- ification (UC, C) of the representation on by extending all operators U to unitary K K g operators on C. Then the operators P := iP are selfadjoint with K j − j 3 e M 2 = P 2 P 2 0. 0 − j ≥ Xj=1 Since (U, ) is irreducible, its commutant is isomorphic to R, C or H ([StVa02, Thm. 1]). K We claim that it is isomorphic to C. If this is not the case, then UC is either irreducible (if the commutant is R) or a direct sum of two copies of the same irreducible unitary representation (U, ) of G (if the commutant is H). As M 2 0, the spectrum of the H ≥ translation group is contained in the set b b D := (x , x) R1,3 : x2 x2 . { 0 ∈ 0 ≥ } The decomposition D = D+ ˙ 0 ˙ D with D := x D : x0 > 0 is invariant under ∪{ }∪ − ± { ∈ ± +} 0 SO (R)↑, so that we obtain a corresponding decomposition U = U U U −, where 1,3 ⊕ ⊕ the spectrum of U R1,3 is supported by D . Since U is irreducible, only one summand j | j b b b b is non-zero. Further, U = U implies that the translation group acts trivially, which b 0 b is ruled out by the assumption that U is faithful. Hence we may w.l.o.g. assume that b b U = U , so that P > 0 (i.e., P 0 and ker P = 0 ) on and therefore on C. Next + 0 0 ≥ 0 { } H H we observe that the conjugation J of C with respect to commutes with P , hence b b H Hb 0 satisfies JP J = P , which leads to the contradiction P = 0 because it implies that the 0 − 0 0 e spectrum of P0 is symmetric (cf. Remark 2.24 below). This shows that the commutant U ′ is C, so that there exists an, up to sign unique, complex structure on commuting G H with UG.

2.3. One-parameter groups. We have seen in Example 2.8 that there are three types of one-dimensional Lie groups defining involutive group pairs: R R R (A) ×, resp., ( ×, +×), (B) R⋊ id , and {± } (C) Pin2(R). Before we turn to the most important case (A), we take a brief look at the other two cases. Remark 2.20. Case (B): Here any antiunitary representation (U, ) yields a conjugation H J := U(0, 1) which defines a real structure on and satisfies JUtJ = U t for t R. − H − ∈ Conversely, every unitary one-parameter group extends to an antiunitary representation of G (Lemma 2.17). 16 K.-H. NEEB AND G. OLAFSSON´

2 1 Case (C): For the group G = T 1,J = Pin (R), we have J = 1 and JzJ − = z for { } 2 − z T, so that antiunitary representations correspond to pairs (H, I), where I AU( ) ∈ ∈ H satisfies I4 = 1 and H is a selfadjoint operator satisfying IHI = H and eπiH = I2. This implies in particular that Spec(H) Z. For any such pair we put UJ := I and itH ⊆ U it := e (see [NO16,´ 4.5] for a natural occurence of such representations). e § The following simple observation is the fundamental link between modular theory and antiunitary representations.

Lemma 2.21. For every continuous antiunitary representation (U, ) of R× and the itH H infinitesimal generator H defined by Uet = e , we obtain by H ∆ := e and J := U 1 − a pair (∆,J), consisting of a positive operator ∆ and a conjugation J satisfying the modular relation 1 J∆J = ∆− . (3)

Conversely, any such pair (∆,J) defines an antiunitary representation of R× by it/2π Uet := ∆− and U 1 := J. − Proof. The only point one has to observe here is that the antiunitarity of J implies that 1 JU J = U corresponds to the relation JHJ = H, which is equivalent to J∆J = ∆− . t t − Lemma 2.21 motivates the following definition from the perspective of antiunitary representations: Definition 2.22. A pair of modular objects on a complex Hilbert space is a pair H (∆,J), where J is a conjugation, i.e., an antilinear isometric involution and ∆ > 0 is a positive selfadjoint operator satisfying the modular relation (3). Then J is called the modular conjugation and ∆ the modular operator. With this terminology, the preceding lemma immediately yields:

Corollary 2.23. For any continuous homomorphism γ : (R×, R×) (G, G ) and any + → 1 continuous antiunitary representation (U, ) of (G, G ), we obtain a pair of modular H 1 objects (∆ ,J ) from the representation U γ of R×. γ γ ◦ Remark 2.24. For a selfadjoint operator H, the existence of a conjugation J satisfying JHJ = H is equivalent to the restriction H(0, ) to the strictly positive spectral − ∞ subspace being equivalent to the restriction H( , 0) to the strictly negative spectral −∞ subspace ([Lo08]). Only such operators H arise as infinitesimal generators for antiunitary representations of R×. Example 2.25. Let (G, G ) be an involutive pair of Lie groups and r G G be such 1 ∈ \ 1 that τ := c is an involution. Then Ad(τ) is an involutive automorphism of g and if g r|G1 is non-abelian, then gτ = 0 . 6 { } (A) If r2 = 1, then any element x gτ leads to a homomorphism ∈ t γ : R× G, γ (e ) := exp(tx), γ ( 1) := r. r,x → r,x r,x − ANTIUNITARY REPRESENTATIONS 17

2 τ (B) If r = 1, then any element x g− leads to a homomorphism ∈ γ : R⋊ idR G, γ (t) := exp(tx), γ ( 1) := r. r,x {± } → r,x r,x − 4 τ 2 (C) If r = 1, then any element x g− with exp(πx)= r leads to a homomorphism ∈ γ : Pin (R)= T 1,J G, γ (eit) := exp(tx), γ (J) := r. r,x 2 { } → r,x r,x Definition 2.26. (One-parameter groups of complex type) Let (G, G1) be an involutive Lie group pair. We assume that G is a subgroup of a complex Lie group GC on which there exists an antiholomorphic involution σ such that G (GC)σ. We consider the set ⊆ := x g: 2π = min t> 0: exp(tix)= e , exp(πix) G G . Y(G,G1) { ∈ { } ∈ \ 1} We associate to each x the holomorphic homomorphism ∈ Y(G,G1) z γ : C× GC, γ (e ) := exp(zx). x → x σ Then σ(γ (w)) = γ (w) for w C× and thus γ (R×) (GC) holds automatically and x x ∈ x ⊆ r := γ ( 1) is an involution. For x , we thus obtain x x − ∈ Y(G,G1) γ Hom((R×, R×), (G, G )). x ∈ + 1 In Section 5 below we shall see that many geometric realizations of modular automor- phism groups come from elements of (G,G1), where G = P (d)+ is the Poincar´egroup 1,d 1 Y or the conformal group Conf(R − ) = O (R)/ 1 of Minkowski space (cf. [HN12, ∼ 2,d {± } 17.4]). This motivates the following discussion of examples. § Example . R R C z 2.27 (a) For (G, G1) = ( ×, +×) and GC = × and exp(z) = e , we have = 1 R = g. Y(G,G1) {± }⊆ (b) (Lorentz groups) For

a b 2 2 G = SO1,1(R) GC = SO1,1(C)= : a,b C,a b =1 , ⊆ n b a ∈ − o R C we have G ∼= × and GC ∼= ×, so that we basically have the same situation as under (a). Here a canonical generator of the Lie algebra is the boost generator 0 1 cosh z sinh z b := with ezb0 = and r = eπib0 = 1. (4) 0 1 0 sinh z cosh z b0 − We have = b . Y(G,G0) {± 0} This example embeds naturally into the higher dimensional Lorentz groups G = SO (R) GC = SO (C), where 1,d ⊆ 1,d b := E + E and r = R = diag( 1, 1, 1,..., 1). (5) 0 10 01 ∈ Y(G,G0) b0 01 − − Since the simple real Lie algebra g = so (R) (for d 2) is of real rank 1, all ad- 1,d ≥ diagonalizable elements x g are conjugate to a multiple of b . All these elements x are ∈ 0 diagonalizable matrices and im(x) is a two-dimensional Minkowski plane in which the two eigenvectors are light-like. Conversely, every triple (β,ℓ+,ℓ ) consisting of β R× and − ∈ two linearly independent light-like vectors ℓ specifies such an element x = x(ℓ+,ℓ ,β) ± − ∈ g by xℓ = βℓ and ker x = ℓ1,ℓ2 ⊥. We then have ± ± ± { } (G,G ) = Ad(G)b0 = x(ℓ+,ℓ ,β): β =1 = SO1,d(R)/(SO1,1(R) SOd 1(R)), Y 0 { − } ∼ × − 18 K.-H. NEEB AND G. OLAFSSON´

and this is a symmetric space because the centralizers of b0 and the involution rb0 share the same identity component. R R⋊R (c) For the affine group G := Aff( ) ∼= × of the real line, the coset G G0 C⋊C σ \ consists of the orientation reversing affine maps. Note that GC ∼= × and GC = G. Here = R 1 is the set of real affine vector fields X for which the vector field Y(G,G0) ∼ ×{± } iX on C generates a 2π-periodic flow (whose center lies on the real axis).

Example 2.28. We consider the real projective group G = PGL (R) GC = PGL (C) 2 ⊆ 2 acting on the real projective line S1 = R , resp., on the Riemann sphere C = ∼ ∪{∞} ∪{∞} ∼ P1(C). A subset I S1 is called an interval if it is connected, open, non-empty and not ⊂ dense. Then the interior I′ of its complement also is an interval. For every interval there is a canonical involution rI PGL2(R) fixing both endpoints and exchanging I and I′. ∈ R R R The centralizer of rI in PSL2( ) is isomorphic to PSO1,1( ) ∼= , hence connected, and there exists an element xI (G,G0) which is up to sign unique. The corresponding I R∈ Y I homomorphism γ := γxI : × G satisfies γ 1 = exp(πixI )= rI . → − For the interval I = (0, ), we have ∞ γI (z)= tz and r (z)= z. t I − I R R I cosh t z+sinh t For I = ( 1, 1), we have γ ( ×)=PO1,1( ) and γ2t(z) := sinh t·z+cosh t . This leads to − · cos t z + i sin t 1 γI (z)= · , so that γI (z)= z and r (z)= γI (z)= . 2ti i sin t z + cos t 2πi I πi z · 2.4. Some low-dimensional groups. 2.4.1. The affine group of the real line. We consider the affine group G := Aff(R) = R⋊R R⋊R × and its identity component G1 = +×. We say that a unitary representation (U, ) of G is of positive energy if U = eitP with P 0, i.e., the restriction to the H 1 (t,1) ≥ translation subgroup has non-negative spectrum. We speak of strictly positive energy if, in addition, ker P = 0 . { } Up to unitary equivalence, G1 has exactly one irreducible unitary representation with strictly positive energy and every unitary representation with strictly positive energy is a multiple of the irreducible one. The analogous statement holds for negative energy ([Lo08, Thm. 2.8]). Further, any unitary representation U of G1 decomposes uniquely as + 0 a direct sum U = U U U −, where U ± have strictly positive/negative energy and ⊕ ⊕ the translation group is contained in ker U 0. The unique irreducible representation of strictly positive energy can be realized on := L2(R+) by H itx s/2 s (U(t,es)f)(x)= e e f(e x). (6)

It obviously extends by U(0, 1)f := f to an irreducible antiunitary representation of G. − By Theorem 2.11 we thus obtain up to equivalence precisely one irreducible antiuni- tary representation of G with strictly positive energy. More generally, we have by [Lo08, Prop. 2.11] and Theorem 2.11: Proposition 2.29. Every unitary representation (U, ) of Aff(R) of strictly positive H 0 energy extends to an antiunitary representation U of Aff(R) on the same Hilbert space ANTIUNITARY REPRESENTATIONS 19 which is unique up to equivalence. The representation theory of the affine group can be used to draw some general conclusions on spectra of one-parameter groups. Proposition 2.30. Let G be a connected Lie group and (U, ) be a unitary representation H for which dU is faithful and x g. Then the following assertions hold: ∈ (a) If ad x has a non-zero real eigenvalue, then Spec(idU(x)) = R. (b) If g is semisimple and 0 = y is nilpotent, then Spec(idU(y)) R, R+, R . 6 ∈{ −} (c) If 0 = x g is such that ad x is diagonalizable and b E g is the ideal generated by 6 ∈ im(ad x) and x, then ker dU(x) = B = ξ : ( g B) U ξ = ξ holds for H { ∈ H ∀ ∈ g } the corresponding integral subgroup B E G. Proof. (a) Let 0 = y g with [x, y] = λy for some λ = 0. Then h := Rx + Ry is a 6 ∈ 6 2-dimensional non-abelian subalgebra and dU(y) = 0. Therefore the assertion follows 6 from the fact that, for all irreducible unitary representations of the corresponding 2- dimensional subgroup isomorphic to Aff(R)0, the spectrum of idU(x) coincides with R. (b) Using the Jacobson–Morozov Theorem, we find an h g with [h, x]= x, so that ∈ the Lie algebra b := Rh + Rx is isomorphic to aff(R). Then the result follows from the R classification of the irreducible representations of the group exp(b) ∼= Aff( ). (c) Let g = µ Rgµ(ad x) denote the eigenspace decomposition of g with respect to ⊕ ∈ the diagonalizable operator ad x. Then the representation theory of Aff(R)0 implies that, for µ = 0, the operators in dU(g (ad x)) vanish on ker dU(x). This shows that the Lie 6 µ subalgebra h generated by x and [x, g] acts trivially on ker dU(x). Since this subalgebra is invariant under ad(g0(ad x)) and contains the other eigenspaces of ad x, it is an ideal of g, hence coincides with b. Therefore ker dU(x) = B. H  2.4.2. The projective group of the real line. We consider the projective group G = PGL (R) and its identity component G = PSL (R). We write r(x)= x for the reflec- 2 1 2 − tion in 0 which commutes with the dilation group R× Aff(R) PGL (R) (cf. Exam- ⊆ ⊆ 2 ple 2.28). Note that r extends to an antiholomorphic automorphism r(z) := z of the − upper half plane C+, so that we obtain an identification of G with the group AAut(C+) (Example 2.9(d)).

For the generators of sl2(R), we write 0 1 0 0 1 1 0 T = , S = and E = . 0 0  1 0 2 0 1 − − They satisfy the commutation relations [E,T ]= T, [E,S]= S and [T,S]= 2E. − − In the complexification sl2(C), we have the basis 1 1 i i i 0 1 i L 1 := ∓ = E (T S),L0 := = (T + S). ± 2  i 1 ∓ 2 − −2  1 0 −2 ∓ − − These elements satisfy the relations

[L0,L 1]= L 1, [L0,L1]= L1 and [L1,L 1]= 2L0. − − − − − 20 K.-H. NEEB AND G. OLAFSSON´

Definition 2.31. The element L i sl (R) is called the conformal Hamiltonian. A uni- 0 ∈ 2 tary representation (U, ) of SL (R) is called a positive energy representation if H 2 dU(L ) 0. 0 ≥ f The following result is well known; for a proof in the spirit of the present exposition, we refer to [Lo08, Cor. 2.9]. Corollary 2.32. For every non-trivial irreducible positive energy representation U of the simply connected covering group SL2(R), the restriction to Aff(R)0 is also irreducible. If U is non-trivial, then U R is the unique irreducible representation with strictly |Aff( )0 f positive energy.

Remark 2.33. In PSL2(R), we have exp(2πiL0) = 1, so that, for every irreducible positive energy representation of PSL2(R), the spectrum of dU(L0) is contained in m+N0 for some m N0. We call m its lowest weight and write m = Cξm for the m-eigenspace ∈ Hk of L0 in . Then dU(L1)ξm = 0 and ξm+k := dU(L 1) ξm, k N0, is an orthogonal H − ∈ basis of . H Theorem 2.34. ([Lo08, Thm. 2.10]) Every unitary positive energy representation U of PSL2(R) extends to an antiunitary representation U of PGL2(R) on the same Hilbert space. This extension is unique up to isomorphism and, if U is irreducible, then J := U r (for r(x)= x) is unique up to a multiplicative factor in T. − Proof. In view of Theorem 2.11, it suffices to verify the first assertion. Here the main point is to define the antiunitary involution on the irreducible lowest weight representation U m of lowest weight m N . We specify an antiunitary involution ∈ 0 C on by Cξn = ξn for n m (cf. Remark 2.33). Then C commutes with L0 and L 1 H ≥ ± and CEC = E, CTC = T and CSC = S. − − This implies that CU mC = U m for g PSL (R). g rgr ∈ 2 Definition 2.35. (Positive energy representations) (a) A unitary representation (U, ) d 1,d 1 H of the translation group R = R − of Minkowski space is said to be a positive energy representation if idU(x) 0 for x V . − ≥ ∈ + (b) A unitary representation (U, ) of the Poincar´egroup P (d)↑ is said to be a H + positive energy representation if its restriction to the translation subgroup is of positive energy. We likewise define antiunitary positive energy representations of P (d)+. Remark . R1,1 ⋊ R 2.36 (a) For the group G := P (2)+ ∼= SO1,1( ) and the reflection r := (0, 1) inducing on G the involution τ(b,a) = ( b,a), there are similar results − − to Theorem 2.34 (cf. Theorem 3.19). Here the main point is to see that the irreducible strictly positive energy representations (U, ) of G carry a natural conjugation that we H 0 can use for the extension. In the L2-realization on the hyperbolas = (λ, µ) R2 : λ2 µ2 = m2 , m> 0, Om { ∈ − } suggested by Mackey theory, we can extend the representation simply by U(0,0, 1)f = f. − (b) For the Poincar´egroup P (d)+, the situation is more complicated. The irreducible Rd ⋊ R strictly positive energy representations of P (d)+↑ = SO1,d 1( )↑ are induced from ∼ − ANTIUNITARY REPRESENTATIONS 21

R R representations of the stabilizer group SO1,d 1( )e0 = SOd 1( ) and realized in vector- − ∼ − valued L2-spaces on the hyperboloids = (p , p) Rd : p2 p2 = m2 , m> 0. Om { 0 ∈ 0 − } Since the stabilizer group is non-trivial for d> 2, the existence of an antiunitary extension to P (d)+ depends on the existence of an antiunitary extension of the representation (ρ, V ) of SOd 1(R) to Od 1(R). We refer to [NO17]´ for a detailed analysis of these issues; see − − also [Va85, Thm. 9.10] for a discussion concerning the Poincar´egroup. 2.4.3. The Heisenberg group. In this subsection we recall the close connection between unitary representations of the 3-dimensional Heisenberg group and positive energy repre- sentations of Aff(R)0 (cf. [Lo08, Thm. 2.8]) which also extends to antiunitary extensions. We define the Heisenberg group Heis(R2) as the manifold T R2, endowed with the × group multiplication is′t (z,s,t)(z′,s′,t′) = (zz′e ,s + s′,t + t′). Note that Heis(R2) = (T R) ⋊ R for α (z,s) = (zeist,s). ∼ × α t Extending the action of R on T R to an action of R Z = R× via × × 2 ∼ is βr(z,s) = (zr ,s) and β 1(z,s) = (z, s), − − we obtain the larger group 2 2 Heis(R ) = (T R) ⋊ R× = Heis(R ) ⋊ 1, τ , with τ(z,s,t) = (z, s,t). τ ∼ × β ∼ { } − Proposition 2.37. There is a natural one-to-one correspondence between unitary repre- sentations (U, ) of Heis(R2) satisfying U = z1 and unitary strictly positive energy H (z,0,0) representations (U, ) of Aff(R) . It is established as follows: e H 0 e ibP is log P (i) If U is given and U(b,1) = e with P > 0, then we put Ws := e and

U(z,s,t) := zWsU(0,et). isA is exp A (ii) Ife U is given and Ws := U(1,s,0) = e , then we put U(s,et) := e U(1,0,t). (iii) This correspondence extends naturally to antiunitary representations of Heis(R2) e e e τ and antiunitary positive energy representations of Aff(R). t isA Proof. (i) Let Vt := U(0,et) and A := log P . Then VtP V t = e P implies that Ws = e − satisfies ist VtAV t = t1 + A and VtWsV t = e Ws. − − 2 Therefore V and W define a unitary representation of Heis(R ) via U(z,s,t) := zWsVt. isA A (ii) With Vt := U(1,0,t) and Ws = U(1,s,0) = e , the positivee operator P := e t satisfies VtP V t = e P , so that we obtain a positive energy representation of Aff(R)0 by isP− e e U(s,et) := e Vt. R R⋊R (iii) If, in addition, U is an antiunitary representation of Aff( ) ∼= × and J = U(0, 1), then JU(b,a)J = U( b,a) leads to JPJ = P and thus to JAJ = A. We − − 2 therefore obtain an antiunitary representation U of Heis(R ) = (T R) ⋊ R× by τ ∼ × α U := U U . (z,s,a) (z,s) (0,a) b b b 22 K.-H. NEEB AND G. OLAFSSON´

Remark 2.38. Write Heis(R2) as the semidirect product Heis(R2) ⋊ 1, τ , where τ { } τ(z,s,t) = (z, s,t). Then the conjugacy class C Heis(R2) is a 2-dimensional sym- − τ ⊆ τ metric space diffeomorphic to T R and the centralizer of τ in Heis(R2) is the subgroup × 1 0 R which also commutes with the whole subgroup (1, 0) R. Therefore Cτ {± }×{ }× { }× 2 can be identified with the conjugacy class of the homomorphism γ : R× Heis(R ) = → τ ∼ (T R) ⋊ R× with γ(t)=(1, 0,t). × β

3. Modular objects and standard subspaces. Besides antiunitary representations of R× (Lemma 2.21), there are other interesting ways to encode modular objects (∆,J). Below we discuss some of them. In particular, we introduce the concept of a standard subspace V which is a geometric counterpart of antiunitary representations of R× ⊆ H (Proposition 3.2). We also discuss how the embedding V can be obtained from ⊆ H the orthogonal one-parameter group ∆it on V ( 3.3), and in 3.4 we introduce half- |V § § sided modular inclusions of standard subspaces and how they are related to antiunitary representations of Aff(R), P (2) and PGL (R). Modular intersections are studied in 3.5. + 2 § 3.1. Standard subspaces. We now turn to the fundamental concept of a standard subspace V of a complex Hilbert space . The key structures on the set Stand( ) of H H standard subspaces is a natural action of the group AU( ), an order structure induced R H by inclusion, and an involution V V ′ = iV ⊥ defined by the symplectic orthogonal 7→ space. Definition 3.1. A closed real subspace V is called a standard real subspace (or ⊆ H simply a standard subspace) if V iV = 0 and V +iV is dense in . We write Stand( ) ∩ { } H H for the set of standard subspaces of . H For every standard subspace V , we obtain an antilinear unbounded operator ⊆ H S : (S) := V + iV , S(v + iw) := v iw D → H − and this operator is closed, so that ∆V := S∗S is a positive selfadjoint operator. We thus obtain the polar decomposition 1/2 S = JV ∆V , 1 1/2 1 1 1/2 1 where JV is an antilinear , and S = S− = ∆V− JV− = JV− (JV ∆V− JV− ) 1 1 leads to JV− = JV and the modular relation JV ∆V JV = ∆V− . If, conversely, (∆,J) is a pair of modular objects, then S := J∆1/2 is a densely defined antilinear involution and Fix(S) := ξ (S): Sξ = ξ { ∈ D } is a standard subspace with JV = J and ∆V = ∆. The correspondence between modular objects and standard subspaces is the core of Tomita–Takesaki Theory (see Theorem 4.2 below). Combining the preceding discussion with Lemma 2.21, we obtain:

it/2π Proposition 3.2. If (U, ) is an antiunitary representation of R× with Uet = ∆− H 1/2 for t R and J := U 1, then V := Fix(J∆ ) is a standard subspace. This defines a ∈ V − bijection V U between antiunitary representations of R× and standard subspaces. ↔ ANTIUNITARY REPRESENTATIONS 23

Remark 3.3. The parametrization of the one-parameter group in Proposition 3.2 may appear artificial, but it turns out that it is quite natural. As V (∆1/2), for each ⊆ D v V , the orbit map U v(g) := U v has an analytic extension ∈ g v iz/2π z C: 0 Im z π , z U z := ∆− v { ∈ ≤ ≤ } → H 7→ e v 1/2 with U (iπ) = ∆ v = Jv. This fits with U 1 = J and it is compatible with the context − of Definition 2.26, where γ( 1) = exp(πix) is obtained by analytic continuation from − γ(et) = exp(tx). Remark 3.4. (a) If V = Fix(S) is a standard subspace with modular objects (∆,J), then 1/4 1/4 1/4 1/4 1/4 1/4 ∆ S∆− = ∆ J∆ = ∆ ∆− J = J (7)

1/4 1/4 J implies that V = Fix(S) = ∆− Fix(J) = ∆− . H (b) Write Stand ( ) for the set of those standard subspaces V for which V +iV = , 0 H H i.e., the antilinear involution S is bounded. Combining (7) with the fact that the unitary group U( ) acts transitively on the set of all conjugations (=antiunitary involutions), it H follows that the group GL( ) acts transitively on Stand ( ). This leads to the structure H 0 H of a Banach symmetric space on this set Stand ( ) = GL( )/ GL( J ) = GL( )/ GL( )J , 0 H ∼ H H ∼ H H where J is any conjugation on (cf. Appendix A.3 and [Kl11]). For = Cn, we obtain H H in particular Cn Cn C R Stand( ) = Stand0( ) ∼= GLn( )/ GLn( ). For elements of Stand ( ), there are no proper inclusions. As we shall see in 3.4, the 0 H § order structure on Stand( ) is non-trivial if is infinite dimensional. H H (c) To extend (b) to arbitrary standard subspaces V , we note that a dense complex subspace carries at most one Hilbert space structure (up to topological linear D ⊆ H .(isomorphism) for which the inclusion ֒ is continuous (Closed Graph Theorem D → H We consider the category whose objects are all dense subspaces carrying such G D ⊆ H Hilbert space structures and whose morphisms are the topological linear isomorphisms with respect to the intrinsic Hilbert space structures. This defines a category D1 → D2 in which all morphisms are invertible, so that we actually obtain a groupoid. As all these subspaces are isomorphic to as Hilbert spaces, this groupoid acts transitively. D H For each standard subspace V , the dense subspace V + iV carries the natural ⊆ H Hilbert structure obtained from the identification with the complex Hilbert space VC. Therefore the groupoid acts transitively on Stand( ) with stabilizer groups = G H GV ∼ GL(V ). (d) Write Conj( ) for the set of conjugations on (Examples 2.4). Then the map H H Stand( ) AU( ), V JV is surjective and AU( )-equivariant. The fiber in a fixed H → H 7→ H 1 conjugation J corresponds to the set of all positive operators ∆ satisfying J∆J = ∆− . Passing to D := i log ∆ J , it follows that it can be parametrized by the set of all skew- |H adjoint operators on the real Hilbert space J (see also Remark 3.5(b) for a different H parametrization). 24 K.-H. NEEB AND G. OLAFSSON´

The problem to describe the set of pairs (V, ), where V is a standard subspace, H ⊆ H can be addressed from two directions. One could either start with a real Hilbert space V and ask for all those complex Hilbert spaces into which V embeds as a standard subspace, or start with the pair ( ,J), respectively the real Hilbert space J , and ask H H for all standard real subspaces V with J = J. Both problems have rather explicit ⊆ H V answers that are easily explained (see [NO16]´ for details). Remark 3.5. (a) Let (V, ( , )) be a real Hilbert space. For any realization of V as a · · standard subspace of , the restriction of the scalar product of to V is a complex- H H valued hermitian form h(v, w) := v, w = (v, w)+ iω(v, w), h i where ω : V V R is continuous and skew-symmetric, hence of the form ω(v, w) = × → (v,Cw) for a skew-symmetric operator C = C⊤ on V satisfying Cv < v for any − k k k k non-zero v V ([NO16,´ Lemma A.10]). Conversely, we obtain for every such operator ∈ C on V by completion of VC with respect to h a complex Hilbert space in which V is a standard real subspace. Then C extends to a bounded skew-hermitian operator C on H satisfying b 1 iC ∆ 1 ∆= − and C = i − . 1 1 + iCb ∆+ (b) If we start with the conjugation J on b, then the standard subspaces V with b H J = J are the subspaces of the form V = (1 + iC) J , where C B( J ) is a skew- V H ∈ H symmetric operator satisfying Cv < v for 0 = v J ([NO16,´ Lemma B.2]). Writing k k k k 6 ∈ H also C for its complex linear extension to , we then have H 1 iC ∆1/2 1 ∆1/2 = − and C = i − . 1 + iC ∆1/2 + 1 Remark 3.6. If V is a standard subspace of and W V +iV is a real subspace closed H ⊆ in such that W corresponds to a standard subspace of the complex Hilbert space VC, H then W is also standard in because the closure of W + iW contains V + iV , hence all H of . H 3.2. Symplectic aspects of standard subspaces. Let V be a standard sub- ⊆ HV space and consider the corresponding antiunitary representation U : R× AU( ) with V V V it/2π it→ H U 1 = J and Uet = ∆− (Proposition 3.2). Since the operators ∆ commute with S−= J∆1/2, they leave the closed subspace V = Fix(S) invariant. Further, the relation 1/2 1/2 JSJ = ∆ J = S∗ = J∆− implies that

R JV = V ′, where V ′ := w : ( v V ) Im v, w =0 = iV ⊥ { ∈ H ∀ ∈ h i } R is the symplectic orthogonal space of V , and V ⊥ denotes the of V in the underlying real Hilbert space R ([Lo08, Prop. 3.2]). In particular, the orbit V H U × V = V, V ′ consists of at most two standard subspaces. R { } Lemma 3.7. The following assertions hold: V ′ V 1 (i) U (t) = U (t− ) for t R×, is the antiunitary representation corresponding ∈ to V ′. ANTIUNITARY REPRESENTATIONS 25

1 (ii) JV ′ = JV and ∆V ′ = ∆V− . U V V (iii) V V ′ = is the fixed point space for the antiunitary representation (U , ) ∩ H H of R×. (iv) V = V ′ is equivalent to ∆= 1. 1/2 Proof. (i) and (ii) follow immediately from V ′ = Fix(J∆− ). (iii) If v V V ′, then v = Sv = S∗v implies ∆v = v and hence Jv = v. Conversely, ∈ ∩ these two relations imply v V V ′. ∈ ∩ (iv) follows from (ii). Remark 3.8. (Direct sums of standard subspaces) (a) Suppose that V are standard subspaces for j =1, 2. Then V := V V j ⊆ Hj 1 ⊕ 2 ⊆ 1 2 is a standard subspace. We have JV = JV1 JV2 and ∆V = ∆V1 ∆V2 . H ⊕ H ⊕ V ⊕ In particular, the corresponding antiunitary representation U of R× is the direct sum U V1 U V2 . ⊕ (b) In particular, every standard subspace V can be written as such a direct sum

V = (V V ′) V , where V ′ V = 0 ∩ ⊕ 1 1 ∩ 1 { } V and (V V ′)C is the set of fixed points of the unitary representation U R× (Lemma 3.7). ∩ | + R Lemma 3.9. Let V be a standard subspace, V V be a closed subspace and V := V V ⊥ 1 ⊆ 2 ∩ 1 be its orthogonal complement in V . Then the following are equivalent: (i) V = V V is a direct sum of standard subspaces. 1 ⊕ 2 (ii) V1 V2′, i.e., iV1 V2 in . ⊆ ⊥ H it (iii) V1 is invariant under the modular automorphisms (∆V )t R. ∈ If these conditions are satisfied and V1 is also standard, then V = V1. Proof. (i) (ii) is easy to verify. ⇔ (i) (iii): Clearly, (i) implies (iii). To see the converse, consider the closed subspace ⇔ it/2π := V + iV of . Then, for each v V , the curve t ∆− v is contained in , H1 1 1 H ∈ 7→ V H1 hence the same is true for its analytic continuation to the strip 1/2 z C: 0 Im z π (Remark 3.3). Therefore ∆ v = Jv 1 and thus 1 is { ∈ ≤ ≤ } V ∈ H H invariant under the antiunitary representation U of R× corresponding to V (Proposi- V tion 3.2). Since the orthogonal decomposition = ⊥ reduces U , the standard H H1 ⊕ H1 subspace V decomposes accordingly.

If (i)-(iii) are satisfied and V1 is also standard, then (i) implies that V1 = V (cf. [Lo08, Prop. 3.10]).

3.3. Orthogonal real one-parameter groups. For any standard subspace V , the unitary operators ∆it define on the real Hilbert space V a continuous orthogonal one- parameter group (U, V ) ( 3.2). § If, conversely, (Ut)t R is a strongly continuous one-parameter group on the real Hilbert ∈ space V , then we can recover the corresponding embedding of V as a standard subspace U tD as follows. Let V0 := V be the subspace of U-fixed vectors and V1 := V0⊥. Then Ut = e with a skew-symmetric infinitesimal generator D = D⊤ satisfying V = ker D. On V we − 0 1 26 K.-H. NEEB AND G. OLAFSSON´ have the polar decomposition D = I D , where I is a complex structure and D = √ D2. | | | | − We now consider the bounded skew-symmetric operator C on V defined by C = 0 and |V0 D 1 e−| | C V1 = I − D . | 1 + e−| | Then h(v, w) := (v, w)+ i(v,Cw) leads to an embedding of V as a standard subspace V as in Remark 3.5(a). The operator D can be recovered directly from C by D =0 ⊆ H |V0 and 1 + C D = I log | 1| |V1 1 C  − | 1| (cf. [NO16,´ Rem. 4.3], where different sign conventions are used). The orthogonal one-parameter group (Ut)t R on V is trivial if and only if D = 0, ∈ which corresponds to ∆ = 1, resp., to C = 0, resp., to V = J (Lemma 3.7(iv)). H 3.4. Half-sided modular inclusions of standard subspaces. We have seen above that standard subspaces V are in one-to-one correspondence with antiunitary rep- V ⊆ H resentations U : R× AU( ) (Proposition 3.2). In this subsection we shall see how → H certain inclusions of standard subspaces can be related to antiunitary positive energy representations of Aff(R) (cf. Section 2.4.1). Here the positive energy condition for the translation group turns out to be the crucial link between the inclusion order on Stand( ) H and the affine geometry of the real line. There are two ways to approach inclusions of standard subspaces. One is to consider the interaction of a unitary one-parameter group with a standard subspace, which leads to the concept of a Borchers pairs and the other considers the modular groups of two standard subspaces and leads to the concept of a half-sided modular inclusion. These perspectives have been introduced by Borchers ([Bo92]) and Wiesbrock ([Wi93]), respec- tively, in the context of von Neumann algebras (see 4.2 for the translation to standard § subspaces and [Lo08] for the results in the context of standard subspaces).

Definition 3.10. (a) Let (Ut)t R be a continuous unitary one-parameter group on ∈ H and V be a standard subspace. We call (U, V )a (positive/negative) Borchers pair if ⊆ H U V V holds for t 0 and U = eitP with P 0. t ⊆ ≥ t ± ≥ (b) We call an inclusion K H of standard subspaces of a half-sided modular ⊆ H ± inclusion if it ∆− K K for t 0. H ⊆ ± ≥ Remark 3.11. The inclusion K H is positive half-sided modular if and only if the ⊆ inclusion H′ K′ is negative half-sided modular ([Lo08, Cor. 3.23]). ⊆ The following theorem provides a passage from Borchers pairs to antiunitary repre- sentations of Aff(R) ([BGL02, Thm. 3.2], [Lo08, Thm. 3.15]). Theorem 3.12 (Borchers’ Theorem—one particle case). If (U, V ) is a positive/negative Borchers pair, then V V 1 1 U (a)U(b)U (a)− = U(a± b) for a R×,b R, ∈ ∈ i.e., we obtain an antiunitary positive energy representation (U, ) of Aff(R) by U = H (b,a) U(b)U V (a). e e ANTIUNITARY REPRESENTATIONS 27

We are now ready to explain how inclusions of standard subspaces are related to antiunitary representations of Aff(R). The following result contains in particular a con- verse of Borchers’ Theorem. For its formulation, we recall the one-to-one correspondence between standard subspaces and antiunitary representations of R× from Proposition 3.2. Theorem 3.13. (Antiunitary positive energy representations of Aff(R) and standard subspaces) Let (U, ) be an antiunitary representation of Aff(R). For each x R, we H ∈ consider the homomorphism

γ : R× Aff(R), γ (s) := (x, 1)(0,s)( x, 1) = ((1 s)x, s) x → x − − whose range is the stabilizer group Aff(R)x and the corresponding family (Vx)x R of stan- ∈ dard subspaces determined by U Vx = U γ . Then the following assertions hold: ◦ x R (i) U(t,s)Vx = Vt+sx and U(t, s)Vx = Vt′ sx for t, x ,s> 0. − − ∈ (ii) The following are equivalent: (a) U is a positive energy representation. (b) V V for s 0. s ⊆ 0 ≥ (c) V V for s t. s ⊆ t ≥ (d) (W, V0) with Wt := U(t,1) is a positive Borchers pair. (e) V V is a +half-sided modular inclusion. 1 ⊆ 0 (iii) V = V for every x R is equivalent to U = 1 for every b R. x 0 ∈ (b,1) ∈ (iv) V := R Vt = v V0 : ( b R) U(b,1)v = v is the fixed point space for the ∞ t { ∈ ∀ ∈ } translations.T ∈ Aff(R) (v) V V ′ = = v : ( g Aff(R)) U v = v . 0 ∩ 0 H { ∈ H ∀ ∈ g } 1 Proof. (i) follows from (t,s)γx(t,s)− = γt+sx, U(0, 1)V0 = V0′ and Vx′ = U(x,1)V0′. − (ii) (a) (b): For W (s) := U we have ⇔ (s,1) it/2π it/2π t − t −t ∆V0 W (s)∆V0 = U(0,e )W (s)U(0,e ) = W (e s), so that the assertion follows from the converse of Borchers’ Theorem [Lo08, Thm. 3.17]. (b) (c) follows from V = W (t)V for t R. ⇔ t 0 ∈ By definition, (d) is equivalent to (a) and (b). (b) (e): From (b) we derive (e) by ⇔ V0 U t V1 = U(0,et)V1 = Vet = U(1,1)Vet 1 U(1,1)V0 = V1. e − ⊆ From (e) we obtain, conversely, for t 0 ≥ V0 U(1,1)V0 = V1 U t V1 = Vet = U(1,1)Vet 1, ⊇ e − and thus Vet 1 V0, which implies (b). − ⊆ (iii) If W (x) := U = 1 for every x R, then V = W (x)V = V . If, conversely, (x,1) ∈ x 0 0 W (x)V = V = V for every x R, then every W (x) commutes with ∆ and J , so 0 x 0 ∈ V0 V0 that Theorem 3.12 yields W (x)= 1 for every x R. ∈ (iv) By (i), the closed real subspace V of V0 is invariant under Aff(R)0. Hence ∞ Lemma 3.9 implies that it is a direct summand of the standard subspace V0 and therefore also invariant under J := U(0, 1). Now (iii) implies that the translation group fixes V − ∞ 28 K.-H. NEEB AND G. OLAFSSON´ pointwise. Conversely, every fixed vector v V of the translations is contained in each ∈ 0 subspace Vx = W (x)V0, hence also in V . ∞ (v) From Lemma 3.7(iv) we know that V V ′ is the space of fixed vector for the 0 ∩ 0 dilation group (U(0,a))a R× . Proposition 2.30(c) implies the translations also act trivially ∈ on this space. This proves (v). Remark 3.14. (a) If the momentum operator P from Theorem 3.12 is strictly positive, then the space of fixed points for the dilation subgroup is trivial and Theorem 3.13(iv) implies V = 0 . ∞ { } (b) If (U, V ) is a Borchers pair for which U V V for all t R, then U V = V for t ⊆ ∈ t every t R because V = U0V = UtU tV UtV . Now Theorem 3.13(iii) entails Ut = 1 ∈ − ⊆ for every t R. Therefore non-trivial representations of the translation group lead to ∈ proper inclusions. (c) For a Borchers pair (U, V ), the operators (Ut)t 0, and the modular operators it ≥ (∆ )t R act by isometries on the real Hilbert space V , so that we obtain a representation ∈ of the semigroup [0, )⋊R× by isometries on V . In this sense we may consider Borchers’ ∞ + Theorem 3.12 as a higher dimensional analog of the Lax–Phillips Theorem which provides a normal form for one-parameter semigroups of isometries on real Hilbert spaces as trans- lations acting on spaces like L2(R+, ), where is a Hilbert space (cf. Remark 3.17(b) K K and [NO15]). The connection with the Lax–Phillips Theorem can also be made more direct as fol- it lows. The subspace H := U1V is invariant under the modular automorphisms (∆− )t 0. it it ≥ More precisely, ∆− H = ∆− U1V = Ue2πt V = Ve2πt V1 = H for t 0, in the notation it ⊆ ≥ it of Theorem 3.13. This shows that t R ∆− H is dense in V and that t>0 ∆− H = V ∈ ∞ is the fixed point set for (Ut)t RSin V (Theorem 3.13(iv)). AssumingT that U has no ∈ non-zero fixed vectors (as in (a) above), we obtain V = 0 . This means that the sub- ∞ { } space H V is outgoing in the sense of Lax–Phillips for the orthogonal one-parameter ⊆it group (∆− )t R. ∈ The group Aff(R) is generated by translations and dilations, which is the structure underlying Borchers pairs. But we can also generate it by the subgroups γ0(R×) and γ (R×). For every antiunitary representation (U, ) of Aff(R), the corresponding modular 1 H objects lead to two standard subspaces V0 and V1 and we have already seen above that V V is a positive half-sided modular inclusion if U is of positive energy. The following 1 ⊆ 0 theorem provides a converse (see [Lo08, Thm. 3.21]). Theorem 3.15 (Wiesbrock Theorem—one particle case). An inclusion K H of stan- ⊆ dard subspaces is positive half-sided modular if and only if there exists an antiunitary positive energy representation (U, ) of Aff(R) with K = V and H = V . H 1 0 Proof. In view of Theorem 3.13(ii)(e), it remains to show the existence of U if the in- clusion is +half-sided modular. In view of [Lo08, Thm. 3.21], there exists a unitary positive energy representation (U, ) of the connected affine group Aff(R)0 such that H t HK t U t = U (e ) and U t = U (e ) for t R. Further, the translation unitaries γ0(e ) γ1(e ) ∈ Wt := U(t,1) satisfy W1H = K and WtH H for t 0. Therefore (W, H) is a Borchers H ⊆ ≥ pair, and thus U(b,a) := WbUa defined an extension of U to an antiunitary representation e ANTIUNITARY REPRESENTATIONS 29 of Aff(R) (Theorem 3.12). The corresponding subspaces are V0 = H by construction, and V1 = W1V0 = W1H = K. Examples 3.16. Below we provide an explicit description of a positive Borchers pair in a concrete model of the irreducible antiunitary positive energy representation of Aff(R) (cf. Theorem 3.12 and [LL14, 4]). A slight variation of (6) leads to the antiunitary R 2 §R dp representation of Aff( ) on L +, p by ibp  t (U(b,et)ψ)(p)= e ψ(e p), (U(0, 1)ψ)(p)= ψ(p). − Transforming it with the unitary operator Γ: L2(R , dp ) L2(R, dθ), Γ(ψ)(θ) = ψ(eθ), + p → transforms it into the representation ibeθ (U(b,et)ψ)(θ) := e ψ(θ + t), (U(0, 1)ψ)(θ) := ψ(θ). (8) − On the strip := z C: 0 < Im z < π we have the Hardy space Sπ { ∈ } 2 2 ( π) := ψ ( π): sup ψ(θ + iλ) dθ < , (9) H S n ∈ O S 0<λ<π ZR | | ∞o and in these terms, the standard subspace V0 corresponding to γ0(t)=(0,t) is given by V = ψ 2( ): ( z ) ψ(iπ + z)= ψ(z) . 0 { ∈ H Sπ ∀ ∈ Sπ } z On the strip , the functions B(z) := eibe satisfy Sπ b Im(ex+iy ) bex sin y B(x + iy) = e− = e− 1 because sin y 0 | | ≤ ≥ and B(iπ + z) = B(z). This shows hat, for b 0, multiplication with B defines an ≥ isometry of the Hardy space 2( ) and also of the real subspace V into itself (cf. H Sπ 0 Remark 3.14(c)). One can show that all unitary operators commuting with the repre- sentation of the one-parameter group (U(b,1))b R and mapping V0 into itself are multi- ∈ plications with bounded holomorphic functions ϕ on satisfying ϕ(iπ + z)= ϕ(z) and Sπ whose boundary values in L∞(R, C) satisfy ϕ(x) = 1 for almost every x R (cf. Re- | | ∈ mark 4.19(c)). For explicit descriptions of standard subspaces related to free fields, we refer to [FG89, p. 422ff]. Remark 3.17. (a) If K H is a proper positive half-sided modular inclusion and V is ⊆ a closed real subspace with K V H, then V is clearly standard. However, neither ⊆ ⊆ the inclusion K V nor the inclusion V H has to be half-sided modular. In fact, the ⊆ ⊆ existence of the unitary one-parameter group (Ut)t R with U1H = K implies that all the ∈ inclusions U H U H for 0 s

(b) Let V be a standard subspace. We write hsm+(V ) for the set of all standard subspaces H V for which the inclusion H V is positive half-sided modular. To ⊆ ⊆ obtain a description of this set, one can proceed as follows. First we can split off the it maximal direct summand H1 := t R ∆V H of V contained in H (Lemma 3.9). This leaves us with the situation where TH ∈= 0 . 1 { } 30 K.-H. NEEB AND G. OLAFSSON´

Decomposition of the corresponding antiunitary representation (U, ) of Aff(R) (The- H orem 3.15) into a subspaces 0 on which the translations act trivially and an orthogonal H space + on which the representation is of strict positive energy, we accordingly obtain H the direct sum V = V 0 V + of standard subspaces and H = V 0 H+. Hence our ⊕ ⊕ assumption implies V 0 = 0 and = +. Now [Lo08, Thm. 2.8] implies that U is a { } H H multiple of the unique irreducible positive energy representation of Aff(R), so that we may assume that 2 itx s/2 s = L (R+, ) and (U(t,es)f)(x)= e e f(e x),U(0, 1)f = JK f, H K − where J is a conjugation on (see 2.4.1). K K § As all antiunitary rerpresentations of Aff(R) with strictly positive energy and the same multiplicity are equivalent, we obtain all such standard subspaces H by applying elements of the group 1 K := U U( ): UV = V = U U( ): ( a R×) UU U − = U { ∈ H } { ∈ H ∀ ∈ (0,a) (0,a)} it it = U O(V ): ( a R×) U∆ = ∆ U . ∼ { ∈ ∀ ∈ V |V V } If = C, then the representation of R× on is (by Fourier transform) equivalent to the K H representation of R on L2(R, ) by K ixp (Vex ξ)(p)= e ξ(p) and (V 1ξ)(p)= ξ( p). − − Therefore any unitary operator M on commuting with VR× is of the form (Mξ)(p)= H m(p)ξ(p), where m: R T is a measurable satisfying m( p)= m(p). It would → − be interesting to see how this relates to the inner functions corresponding to endomor- phisms of one-dimensional standard pairs (see Remark 4.19(c)). Combining the preceding results with the fact that the infinite dimensional irreducible positive energy representation of Aff(R)0 extends to an antiunitary positive energy repre- sentation of PGL2(R) with lowest weight 1 (Theorem 2.34 and Corollary 2.32), we obtain ([Lo08, Cor. 4.15]): Theorem 3.18. There exists a one-to-one correspondence between (i) Positive half-sided modular inclusions K H. ⊆ (ii) Antiunitary positive energy representations of Aff(R). (iii) Positive Borchers pairs (V,U). (iv) Unitary representations of PSL2(R) which are direct sums of representations with lowest weights 0 or 1. An important aspect of the last item in the preceding theorem is that it leads to a considerable enrichment of the geometry. Starting with a positive half-sided modular inclusion K H, we obtain an antiunitary representation of PGL2(R). Accordingly, for ⊆ 1 I every interval I S , the corresponding homomorphism γ : R× PGL (R) (Exam- ⊆ → 2 ple 2.28) determines a standard subspace VI , whereas the representation of Aff(R) only leads to standard subspaces V indexed by the open half-lines I R. I ⊆ The following theorem is another result in this direction. It relates pairs of half- sided modular inclusions via the corresponding antiunitary representations of Aff(R) to representations of the two-dimensional Poincar´egroup P (2)+, resp., PGL2(R). ANTIUNITARY REPRESENTATIONS 31

Theorem 3.19. (a) Let H V be a half sided modular inclusion and H V be a 1 ⊆ − 2 ⊆ +half sided modular inclusion such that

JH1 JH2 = JV JH2 JH1 JV . (10) Then the corresponding three modular one-parameter groups combine to a faithful contin- R1,1 ⋊ R uous antiunitary representation of the proper Poincar´egroup P (2)+ ∼= SO1,1( ). (b) Let H, V be standard subspaces of such that H V V and H V H are H ∩ ⊆ ∩ ⊆ , resp., +half-sided modular inclusions satisfying JH V = V. Then the corresponding − H V H V three representation U ,U and U ∩ generate a faithful antiunitary positive energy representation of PGL2(R). Proof. (a) The version for von Neumann algebras is contained in [Wi98, Lemma 10] and we shall see in 4.2 below how the present version follows from this one. Here are § some comments on the proof. Clearly, the two half-sided modular inclusion defines two 1/2 R R⋊R antiunitary representations U of Aff( ) ∼= × that coincide on the subgroup of dilations. Now the main point is to verify that the corresponding images of the translation groups commute. That (10) is necessary can be seen as follows. If we have a unitary representation of P (2)0 as required, then 1 2 it/2π t − U(b1,b2,e ) := Ub1 Ub2 ∆V and U(0,0, 1) := JV − defines an extension to an antiunitary representation of P (2)+. Here the modular conju- gations

JH1 = U(2,0, 1) and JH2 = U(0,2, 1) − − satisfy (10) because (0, 0, 1)(0, 2, 1)(2, 0, 1)(0, 0, 1)= (2, 2, 1) holds in P (2). − − − − − (b) The von Neumann version is [Wi93b, Thm. 3] (see also [Wi93c, Lemma 2]). 3.5. Half-sided modular intersections. Definition 3.20. We consider two standard subspaces H ,H and their modular 1 2 ⊆ H objects (∆ ,J ) . We say that the pair (H ,H ) has a modular intersection if Hj Hj j=1,2 1 2 ± the following two conditions are satisfied (MI1) The intersection H H is a standard subspace and the inclusions H H H , 1 ∩ 2 1 ∩ 2 ⊆ j j =1, 2, are half-sided modular. ± it it (MI2) The strong limit S := limt ∆H1 ∆−H2 (which always exists by Remark 3.21 →±∞1 below) satisfies JH1 SJH1 = S− .

Remark 3.21. (a) In Aff(R) the two multiplicative one-parameter groups γ0(r) := (0, r) and γ (r) := (1 r, r) (stabilizing the points 0 and 1, resp.) satisfy 1 − 1 1 1 γ (r)γ (r− )=(0, r)(1 r− , r− ) = (r 1, 1), 0 1 − − 1 so that limr 0 γ0(r)γ1(r− ) = ( 1, 1) exists. As a consequence, for every continuous → − unitary representation (U, ) of Aff(R) , the limit H 0 1 lim Uγ0(r)U − = U( 1,1) (11) r 0 γ1(r) − → exists. 32 K.-H. NEEB AND G. OLAFSSON´

(b) Suppose that the +-variant of (MI1) is satisfied and let (U 1, ), (U 2, ) be the H H corresponding positive energy representations of Aff(R) satisfying H H = V 1 = V 2, H = V 1 and H = V 2, 1 ∩ 2 1 1 1 0 2 0 j where (Vx )x R are the corresponding families of standard subspaces (Theorem 3.13). We j ∈ j write W (t) := U(t,1) for the representations of the translation group. Then (a) implies that 1 1 1 1 it/2π it/2π it it W ( 1) = U = lim U −t U t = lim ∆ ∆− = lim ∆ ∆− ( 1,1) t γ0(e ) γ1(e ) t H1 H1 H2 t H1 H1 H2 − − →∞ →∞ ∩ →∞ ∩ 2 it it and likewise W ( 1) = limt ∆H2 ∆H−1 H2 . This leads to − →∞ ∩ it it 1 2 S := lim ∆H ∆− = W ( 1)W (1), (12) t 1 H2 →∞ − so that the limit in (MI2) exists whenever (MI1) is satisfied. From the relation SH = W 1( 1)W 2(1)H = W 1( 1)(H H )= H (13) 2 − 2 − 1 ∩ 2 1 1 it follows that SJH2 S− = JH1 , i.e.,

SJH2 = JH1 S. (14)

As condition (MI2) means that JH1 S is an involution, and this is equivalent to SJH1 = 1 S(JH1 S)S− being an involution, (14) shows that (MI2) is equivalent to the relation 1 JH2 SJH2 = S− . (c) If the negative variant of (MI1) is satisfied, then H′ (H H )′ are +half-sided j ⊆ 1 ∩ 2 modular inclusions by Remark 3.11. Let (U 1, ), (U 2, ) be the corresponding positive H H energy representations of Aff(R) satisfying 1 2 1 2 (H H )′ = V = V , H′ = V and H′ = V 1 ∩ 2 0 0 1 1 2 1 (Theorem 3.13). With the same notation as in (b), we obtain 1 1 1 it/2π it/2π it it − W (1) = lim Uγ (e−t)Uγ (et) = lim ∆ ′ ∆ ′ = lim ∆H1 ∆− t 1 0 t H1 (H1 H2) t H1 H2 →∞ →−∞ ∩ →−∞ ∩ 2 it it and likewise W (1) = limt ∆H2 ∆H−1 H2 . This leads to →−∞ ∩ it it 1 2 S := lim ∆ ∆− = W (1)W ( 1), (15) t H1 H2 →−∞ − so that the limit in (MI2) exists. Here SH2′ = H1′ shows that (MI2) is equivalent to 1 JH2 SJH2 = S− . The following theorem extends Wiesbrock’s Theorem 3.15 from half-sided modular inclusions to general modular intersections. Theorem 3.22 (Wiesbrock’s Theorem for modular intersections—one particle version). For a pair (H1,H2) of standard subspaces, the following assertions hold:

(a) (H1,H2) has a +modular intersection if and only if there exists an antiunitary representation (U, ) of Aff(R) such that the corresponding family of standard sub- H spaces (Vx)x R from Theorem 3.15 satisfies V0 = H1 and V1 = H2. ∈ (b) (H ,H ) has a modular intersection if and only if there exists an antiunitary 1 2 − representation (U, ) of Aff(R) such that V = H′ and V = H′ . H 0 1 1 2 ANTIUNITARY REPRESENTATIONS 33

Proof. (a) Suppose first that (H1,H2) has the +modular intersection property. In the context of Remark 3.21(b) we then have 1 1 2 2 1 2 JH1 JH2 = JH1 JH1 H2 JH1 H2 JH2 = U(0, 1)U(2, 1)U(2, 1)U(0, 1) = W ( 2)W (2), ∩ ∩ − − − − − and this operator commutes with S by (MI2). From these relations Wiesbrock derives in [Wi97] that the two one-parameter groups W 1 and W 2 commute, so that W (t) := W 1(t)W 2( t) defines a unitary one-parameter group and U := W (b)U 1 defines − (b,a) γ0(a) an antiunitary representation of Aff(R) for which the corresponding standard subspaces 1 (Vx)x R satisfy V0 = H1 and V1 = W (1)V0 = S− H1 = H2.. ∈ Suppose, conversely, that (U, ) is an antiunitary representation of Aff(R) and let H (Vx)x R be the corresponding family of standard subspaces. We decompose (U, ) as a ∈ H direct sum + + 0 0 (U, ) = (U , ) (U , ) (U −, −), H H ⊕ H ⊕ H + where the representation U has strictly positive energy, U − has strictly negative energy 0 and the translation group acts trivially on . Then the subspaces Vx decompose accord- + H 0 0 0 ingly as orthogonal direct sums V = V V V −, where V = V does not depend x x ⊕ x ⊕ x x 0 on x. Theorem 3.13 now implies that V ± V ± for (x y) 0. To see that V and V x ⊆ y ± − ≥ 0 1 have a +modular intersection, we first observe that + + 0 + 0 V V = (V V ) V (V − V −)= V V V − 0 ∩ 1 0 ∩ 1 ⊕ 0 ⊕ 0 ∩ 1 1 ⊕ 0 ⊕ 0 is an orthogonal direct sum of three standard subspaces, hence a standard subspace. Its it/2π − t invariance under the modular operators ∆V0 = Uγ0(e ) for t 0 follows from the + + + 0 ≥ invariance of V under U t and the invariance of V and V under the corresponding 1 γ0(e ) 0 0 modular group. it/2π − t t t For the invariance of V0 V1 under ∆V = Uγ1(e ) = U((1 e ,e ), we likewise use ∩ 1 − that

U(1− et,et)V0− = U(1− et,1)V0− = V1− et V0− for t 0. − − − ⊆ ≥ This shows that (V0, V1) has a +modular intersection. (b) If (H1,H2) is a modular intersection, we likewise obtain with Remark 3.21(c) 1 − 1 2 that U(b,a) := W (b)Uγ1(a) and W (t) := W (t)W ( t) define an antiunitary representa- 1 − 1 tion of Aff(R) with S = W ( 1), V = V = H′ and V = W (1)V = S− H′ = H′ . − 0 1 1 1 0 1 2 Suppose, conversely, that (U, ) is an antiunitary representation of Aff(R). We use H the notation from (a). To see that V ′ and V ′ have a modular intersection, we first 0 1 − observe that + + 0 + 0 V ′ V ′ = ((V )′ (V )′) (V )′ ((V −)′ (V −)′) = (V )′ V (V −)′ 0 ∩ 1 0 ∩ 1 ⊕ 0 ⊕ 0 ∩ 1 0 ⊕ 0 ⊕ 1 it/2π to see that V0′ V1′ is standard. Its invariance under the modular operators ∆ ′ = ∩ V0 it/2π ∆− = U t for t 0 follows from the invariance of (V −) under U − t (Re- V0 γ0(e ) 1 ′ γ0(e ) ≥ + 0 mark 3.11) and the invariance of V0 and V0 under the corresponding modular group. it/2π t t t For the invariance of V0′ V1′ under ∆ ′ = Uγ1(e ) = U(1 e ,e ), we likewise use that ∩ V1 − + + + + + + U(1 et,et)(V0 )′ = U(1 et,1)(V0 )′ = (V1 et )′ (V0 )′ for t 0. − − − ⊆ ≥ 34 K.-H. NEEB AND G. OLAFSSON´

Therefore (V ′, V ′) has a modular intersection. 0 1 − The key point of modular intersections is that they no longer require any spectral condition on the corresponding representations of Aff(R). The preceding theorem shows that modular intersections are characterized as pairs of standard subspaces that can ± be obtained from arbitrary antiunitary representations of Aff(R). This is of particular relevance for representations of Lorentz groups SO1,d 1(R) which, for d> 3, never satisfy − any positive energy condition. With the same method that we used to obtain Theorem 3.19, we now obtain by transcribing [Wi97, Thm. 6] from the context of von Neumann algebras the following theorem.

Theorem 3.23. Let H1,H2,H3 be three standard subspaces such that (H1,H2) and

(H3,H1′ ) are modular intersections and (H2,H3) is a +modular intersection. Then the − Hj corresponding antiunitary representations U , j =1, 2, 3, of R× generate an anti-unitary representation of PGL2(R). Proof. For the proof we only has to observe that the group G generated by the three it R modular one-parameter groups (∆Hj )t , j = 1, 2, 3 is invariant under JH1 . For the ∈ it is subgroup G12 generated by the operators ∆H1 and ∆H2 , this follows from Theorem 3.15, and we likewise obtain the invariance of the subgroup G13 generated by the operators ∆it and ∆is . As G is generated by G G , the assertion follows. H1 H3 12 ∪ 13 4. A glimpse of modular theory. We now recall some of the key features of Tomita– Takesaki Theory. In 4.2 we discuss the translation between pairs ( , Ω) of von Neumann § M algebras with cyclic separating vectors and standard subspaces V . More specifically, we discuss this translation for half-sided modular inclusions in 4.3, and in 4.4 we take a § § closer look at the space of modular conjugations of a von Neumann algebra. 4.1. The Tomita–Takesaki Theorem. Let be a Hilbert space and B( ) be H M⊆ H a von Neumann algebra. We call a unit vector Ω ∈ H cyclic if Ω is dense in . • M H separating if the map ,M MΩ is injective. • M → H 7→ It is easy to see that Ω is separating if and only if it is cyclic for the commutant ′. M Definition 4.1. We write cs( ) for the set of cyclic and separating unit vectors for . M M Theorem 4.2 (Tomita–Takesaki Theorem). Let B( ) be a von Neumann algebra M⊆ H and Ω be a cyclic separating vector for . Write := M : M ∗ = M for ∈ H M Mh { ∈M } the real subspace of hermitian elements in . Then V := Ω is a standard subspace. M Mh The corresponding modular objects (∆,J) satisfy it it (a) J J = ′ and ∆ ∆− = for t R. M M M M ∈ (b) JΩ=Ω, ∆Ω=Ω and ∆itΩ=Ω for all t R. ∈ it it (c) For M ′, we have JMJ = M ∗ and ∆ M∆− = M for t R. ∈M∩M ∈ Proof. We only show that V is a standard subspace and refer to [BR87, Thm. 2.5.14] for the other assertions. Clearly, V is a closed real subspace for which V + iV is dense ANTIUNITARY REPRESENTATIONS 35 because it contains Ω+ i Ω= Ω. The same holds for W := Ω because Ω is Mh Mh M Mh′ also cyclic for ′. For M and M ′ ′ , we have M ∈Mh ∈Mh MΩ,M ′Ω = M ′MΩ, Ω = MM ′Ω, Ω = M ′Ω,MΩ R, h i h i h i h i ∈ which implies that ω(V, W )= 0 . Therefore V iV is a complex subspace of W ⊥ = 0 , { } ∩ { } hence trivial. Now the main point is to show that the modular objects (∆,J) associated to V satisfy (a)-(c). The key point of the Tomita–Takesaki Theorem is that it provides for each cyclic separating vector Ω cs( ) a pair (∆,J) of modular objects. The modular operators ∆ ∈ M and their spectra are the key tool in the classification of factors and in the characterization of von Neumann algebras by their natural cones by A. Connes [Co73, Co74]. Here we V emphasize that (∆,J) is encoded in an antiunitary representation U of R×. We first take a closer look at the antiunitary operators that come directly from M and its commutant. The picture will be refined in 4.4 below. § Example 4.3. Let B( ) be a von Neumann algebra, G := U( ) U( ) and M ⊆ H 1 M × M τ Aut(G ) be the flip automorphism. We consider the group G := G ⋊ 1, τ . ∈ 1 1 { } If Ω is a cyclic separating vector for and J the corresponding modular involution, M then ε U(g,h,τ ε) := gJhJJ defines an antiunitary representation of the pair (G, G1). Any other conjugation J on that we can use to extend the unitary representation H U is of the form J = Jg for some central unitary element g U( ′). |G1 e ∈ M∩M Definition 4.4. A vone Neumann algebra B( ) is said to be in symmetric form if M⊆ H there exists a conjugation J on with H J J = ′ and JZJ = Z∗ for Z ′. M M ∈M∩M According to the Tomita–Takesaki Theorem, the existence of a cyclic separating vector implies that is in symmetric form. According to [Bla06, Thm. III.4.5.6], any two M realizations of in symmetric form are unitarily equivalent. M Let S ( ) denote the set of normal states of the von Neumann algebra . By n M M the Gelfand–Naimark–Segal construction, any state ω corresponds to a cyclic normal representation (π , , Ω ) with ω(M)= Ω , π (M)Ω , which is uniquely determined ω Hω ω h ω ω ωi up to unitary equivalence of cyclic representations. By construction Ωω is cyclic, it leads to a faithful representation for which Ωω is separating if and only if the state ω is faithful, i.e., ω(M ∗M) > 0 for any non-zero M . ∈M Remark 4.5. (Existence of cyclic separating vectors) A von Neumann algebra pos- M sesses a faithful normal state if and only if it is σ-finite (also called countably decompos- able) in the sense that every family of mutually orthogonal projections in is at most M countable ([Bla06, Prop. III.4.5.3]). This is always the case if can be realized on a sep- M arable Hilbert space, but not in general. Therefore one has to generalize the concept of a state to that of a normal weight. This is an additive positively homogeneous weakly lower semicontinuous functional ω : + [0, ] on the positive cone + of that may also M → ∞ M M 36 K.-H. NEEB AND G. OLAFSSON´ take the value . A weight ω is called semifinite if the subset M : ω(M) < ∞ { ∈M+ ∞} generates as a von Neumann algebra. Every von Neumann algebra has a faithful M normal semifinite weight (cf. [Bla06, III.2.2.26]) and the GNS construction as well as Tomita–Takesaki theory extend naturally to normal semifinite weights. In particular, any such weight leads to a symmetric form realization of . M Example 4.6. (a) Let = L2(X, S,µ) for a σ-finite measure space (X, S,µ) and = H M L∞(X, S,µ), acting on by multiplication operators. Then the normal states of are H M of the form ω (f)= fhdµ, where 0 h satisfies h dµ = 1. Such a state is faithful h X ≤ X if and only if h = 0 holdsR µ-almost everywhere. ThenR Ω := √h is a corresponding 6 ∈ H cyclic separating unit vector. From S(fΩ) = fΩ, we obtain S(f)= f, which is isometric and therefore S = J and ∆ = 1. (b) Let = B ( ) be the space of Hilbert–Schmidt operators on the complex sepa- H 2 K rable Hilbert space and consider the von Neumann algebra = B( ) acting on by K op M K H left multiplications. Then ′ = B( ) acts by right multiplications. Normal states of M ∼ K are of the form ω (A) = tr(AD), where 0 D satisfies tr D = 1. Such a state is faith- M D ≤ ful if and only if ker D = 0 (which requires to be separable), and then Ω := √D { } K ∈ H is a cyclic separating unit vector. Then S(MΩ) = M ∗Ω=(ΩM)∗ implies that 2 2 1 JA = A∗ and ∆(A)=Ω AΩ− = DAD− for A B ( ). ∈ 2 K (c) The prototypical pair (∆,J) of a modular operator and a modular conjugation arises from the regular representation of a locally compact group G on the Hilbert space = L2(G, µ ) with respect to a left Haar measure µ . Here the modular operator is H G G given by the multiplication ∆f = ∆ f, G · where ∆ : G R× is the modular function of G and the modular conjugation is given G → + by 1 2 1 (Jf)(g) = ∆G(g)− f(g− ). Accordingly, we have for S = J∆1/2:

1 1 (Sf)(g) = ∆G(g)− f(g− )= f ∗(g). The corresponding von Neumann algebra is the algebra B(L2(G, µ )) generated M⊆ G by the left regular representation. If M h = f h is the left convolution with f C (G), f ∗ ∈ c then the value of the corresponding normal weight ω on is given by ω(M ) = f(e), M f so that ω corresponds to evaluation in e, which is defined on a weakly dense subalgebra of . M 4.2. Translating between standard subspaces and von Neumann algebras. We have already seen that cyclic separating vectors of a von Neumann algebra lead to M standard subspaces. In this subsection we explore some properties of this correspon- dence and describe how half-sided modular inclusions of standard subspaces translate into corresponding inclusions of von Neumann algebras. This correspondence shows that antiunitary representations of groups generated by modular one-parameter groups and conjugations from cyclic vectors of von Neumann algebras can already be studied in ANTIUNITARY REPRESENTATIONS 37 terms of standard subspaces and their inclusions, and all this can be encoded in antiuni- tary representations of pairs (G, G ), and homomorphisms R× G, resp., Aff(R) G 1 → → (Corollary 2.23 and Theorems 3.13, 3.22). Lemma 4.7. If B( ) is a von Neumann algebra and Ω a separating vector for M⊆ H ∈ H , then we associate to every von Neumann subalgebra the closed real subspace M N ⊆M V := hΩ. This assignment is injective. N N Note that the subspace V is standard if Ω is also cyclic for . N N Proof. (cf. [Lo08, Prop. 3.24]) We have to show that M h and MΩ V implies ∈ M ∈ N M . First we find a sequence A such that A Ω MΩ. For any B ′, this ∈ N n ∈ N n → ∈M leads to A BΩ = BA Ω BMΩ = MBΩ, so that A M holds pointwise on the n n → n → dense subspace := ′Ω. Since the hermitian operators A and M are bounded, is D M n D a common core for all of them. With [RS73, Thm. VIII.25] it now follows that An M 1 1 → holds in the strong resolvent sense, i.e., that (i1 + An)− (i1 + M)− in the strong 1 → operator topology. This implies that (i1 + M)− , which entails M . ∈ N ∈ N The concept of a half-sided modular inclusion was originally conceived on the level of von Neumann algebras with cyclic separating vectors, where it takes the following form ([Wi93, Wi97]). Definition 4.8. Let Ω be a cyclic separating vector for the von Neumann algebra and M be a von Neumann subalgebra for which Ω is also cyclic. The triple ( , , Ω) N ⊆M M N is called a half-sided modular inclusion 4 if ± it it ∆− ∆ for t 0. (16) M N M ⊆ N ± ≥ Note that Ω is also separating for because , so that we obtain two pairs of N N ⊆M modular objects (∆ ,J ) and (∆ ,J ). M M N N Lemma 4.9. Let B( ) be von Neumann algebras with the common cyclic N ⊆M⊆ H separating vector Ω . Then ( , , Ω) is a half-sided modular inclusion if and only ∈ H M N ± if the corresponding standard subspaces V := hΩ V := hΩ define a half-sided N N ⊆ M M ± modular inclusion. it it it Proof. Since ∆− h∆ Ω = ∆− V , relation (16) implies M N M M N it ∆− V V for t 0. (17) M N ⊆ N ± ≥ it it it If, conversely, the latter condition is satisfied, then ∆− h∆ Ω ∆− V V , so N M ⊆ N ⊆ N that Lemma 4.7 implies (16). M M The preceding lemma has a very interesting consequence because it translates directly between half-sided modular inclusions of von Neumann algebras and half-sided modular inclusions of the corresponding standard subspaces. It immediately implies that a triple ( , , Ω) consisting of two von Neumann algebras and with a common cyclic M N M N separating vector Ω defines a modular intersection in the sense of [Wi97] if and only if V and V have a modular intersection. M N 4Here we switched signs, compared to [Bo97, Wi93], to make the concept compatible with the sign convention in the context of standard subspaces [Lo08]. 38 K.-H. NEEB AND G. OLAFSSON´

Clearly, every result on half-sided modular inclusions on standard subspaces, such as Borchers’ Theorem 3.12 ([Bo00, Thms. II.5.2, VI.2.2]), Wiesbrock’s Theorem 3.15 ([Wi93, AZ05]), and Theorem 3.19 ([Wi98, Lemma 10]) yield corresponding results on half-sided modular inclusions of von Neumann algebras which preceded the corresponding results on standard subspaces. It is remarkable that this transfer also works in the other direction: every result on half-sided modular inclusions of von Neumann algebras can be used to obtain a corre- sponding result on standard subspaces. For this transfer one can use the second quan- tization procedure described in some detail in Section 6 below. It associates to every standard subspace V a von Neumann algebra (V ) B( ( )) on the bosonic ⊆ H R ⊆ F+ H Fock space ( ) for which the vacuum Ω is a cyclic separating vector and for which the F+ H modular objects are obtained by second quantization. Here we consider the antiunitary representation Γ: AU( ) AU( ( )), Γ(U)(v v ) := Uv Uv H → F+ H 1 ∨···∨ n 1 ∨···∨ n obtained by second quantization. If γ : R× AU( ) is the antiunitary representation V → H associated to V , then γ := Γ γ is the corresponding antiunitary representation on V ◦ V the Fock space +( ) (cf. Proposition 3.2). it/2π F H e If ∆− H = γ (et)H H holds for t 0, then V V ⊆ ≥ t t t (γ (e )H)= γ (e ) (H)γ (e− ) (H) R V V R V ⊆ R implies that ( (V ), (H), Ω) is a half-sided modular inclusion whenever H V is. R R ± e e ⊆ If, conversely, ( (V ), (H), Ω) is a half-sided modular inclusion, then R R ± t t t (γ (e )H)= γ (e ) (H)γ (e− ) (H) R V V R V ⊆ R implies that γ (et)H H by Theorem 6.4(i). Therefore H V is a half-sided modular V ⊆ e e ⊆ inclusion of the same type. As the subgroup of AU( ( )) generated by the correspond- F+ H ing one-parameter groups γ (R×) is contained in the subgroup Γ(AU( )) which is the V H range of the second quantization homomorphism Γ: AU( ) AU( ( )), anything H → F+ H that we can say about subgroupse generated by these groups and conditions relating to modular objects can be translated into a corresponding result on standard subspaces and the antiunitary one-parameter groups γ on . V H According to this principle, any result on half-sided modular inclusions of von Neu- mann algebras has a “one-particle version” concerning standard subspaces and vice versa (cf. 3.4, 3.5). The advantage of the one-particle version is that it has a simpler formu- §§ lation and that standard subspaces are completely encoded in the antiunitary represen- tations γV of R×, hence in an antiunitary representation of a group (G, G1) generated by the image of homomorphism (R×, R×) (G, G ). Therefore one can hope that any re- + → 1 sults on standard subspaces, half-sided modular inclusions and the corresponding groups can be expressed in terms of antiunitary representations of suitable involutive pairs of Lie groups (G, G1). This was one of the key motivations for us to write this note. Remark 4.10. (a) A typical result of this type is Wiesbrock’s Theorem on half-sided modular inclusions (cf. Theorem 3.15 and [Wi93, Wi97, AZ05]). On the level of modular inclusions of von Neumann algebras ( , , Ω), Wiesbrock provides the additional infor- M N ANTIUNITARY REPRESENTATIONS 39 mation that, if is a factor, then it is of type III (see [Wi93, Thm. 12] which uses M 1 [Lo82]). It would be interesting to see if and how this can be formulated and derived on the level of standard subspaces and antiunitary representations. The discussion of modular nuclearity in [Lo08, 6.3] may indicate a possible way how this can be done. § (b) In [GLW98, Thm. 4.11] (see also [Wi93c, Lemmas 3,4ff] and [Wi93b, Thm. 2]), similar structures related to multiple modular inclusions are studied, namely quadruples ( , , , Ω), where the are von Neumann algebras with the common separating M0 M1 M2 Mj cyclic vector Ω such that the commute pairwise and, in cyclic order, ′ Mj Mj ⊆ Mj+1 is a half sided modular inclusion. From this structure, which arises from partitions of S1 into three intervals, one derives antiunitary positive energy representations of PGL2(R) as in Theorem 3.19 ([GLW98, Thm. 1.2]). (c) In [Wi93b] it is shown that the von Neumann version of Theorem 3.19(b) char- acterizes conformal quantum fields on the circle in terms of modular data associated to three intervals. (d) In [KW01] configurations of 6 von Neumann algebras ( ij )1 i x0 . − ∈ | | To fix notation for the following, we write W = W 2 E , where R R ⊕ R d 2 E = (x , x): x = x =0 = R − R { 0 0 1 } ∼ 2 R2 is the edge of the wedge and WR is the standard right wedge in . A subset of the form W = gWR, g P (d), is called a wedge. We write for the set 1,d 1 ∈ W of wedges in R − . The following lemma contains some details on as an ordered homogeneous space. W For item (iii), we recall the generator b0 of the Lorentz boost from Example 2.27 (see also [BGL02, 2]). § Lemma 4.12. The wedge space has the following properties: W (i) The stabilizer P (d) = g P (d): gW = W of the standard right wedge has WR { ∈ R R} the form

R 2 P (d)WR = E(d 2) O1,1( )W , ∼ − × R Rd 2 where E(d 2) denotes the euclidean group on ER ∼= − . − 1 (ii) r := gR g− for W = gW and R = diag( 1, 1, 1,..., 1) yields a consistent W 01 01 − − definition of wedge reflections (rW )W . ∈W (iii) The subgroup P (d)↑ acts transitively on , and the following are equivalent for W g P (d)↑: ∈ 40 K.-H. NEEB AND G. OLAFSSON´

(a) gWR = WR.

(b) Adg b0 = b0 and g commutes with the wedge reflection rWR = R01.

(c) Adg b0 = b0. The set of all elements satisfying these conditions is

P (d)↑ = E(d 2) SO1,1(R)↑. (18) WR ∼ − × In particular, the subgroup SO (R) is central in P (d)↑ . For d > 2, even the 1,1 ↑ WR identity component acts transitively on with stabilizer W P (d)↑ = E(d 2)+ SO1,1(R)↑. (19) +,WR ∼ − × t tb0 (iv) For γ : R× P (d) , defined by γ (e ) := e and γ ( 1) := r , we have WR → + WR WR − WR a bijection g g 1 Cγ , gWR γW := γ for g P (d)↑ , γ (t)= gγW (t)g− W → WR 7→ WR ∈ + WR R and the map Cr = rW : W , W rW W → WR { ∈ W} 7→ corresponding to evaluation in 1 is a two-fold covering map. − (v) We have a bijection Ad(P (d)↑)b , gW Ad(g)b of onto an adjoint W → 0 R 7→ 0 W orbit of P (d)↑.

(vi) The stabilizer P (d)WR is open in the centralizer of rWR in P (d). In particular

(P (d), P (d)WR ) is a symmetric pair and is a symmetric space. W ⋊ R (vii) The semigroup SWR := g P (d): gWR WR is given by WR O1,d 1( )WR . { ∈ ⊆ } − Proof. (i) The stabilizer group contains the translation group corresponding to the edge E and gW = W implies g(0) E , so that R R R ∈ R ⋊ R P (d)WR = ER O1,d 1( )WR . ∼ − Further,

R R R 2 R R O1,d 1( )WR =Od 2( ) O1,1( )W =Od 2( ) (SO1,1( )↑ 1, R1 ), − − × R − × { } where R = diag(1, 1, 1,..., 1). We thus obtain (i). 1 − (ii) follows from the fact that the wedge reflection R01 commutes with P (d)WR .

(iii) That P (d)↑ acts transitively follows from the fact that the stabilizer P (d)WR contains the reflection R satisfying R V = V . If d> 2, then the stabilizer P (d) 1 1 + − + WR intersects all four connected components of P (d), so that even P (d)+↑ acts transitively. For d = 2 we obtain two orbits because W lie in different orbits of P (2)↑ . ± R + It remains to verify the equivalence of (a), (b) and (c). From (18) we derive that (a) implies (b) and hence (c). That g = (b,a) commutes with R is equivalent to b E and 01 ∈ R a = a a with a acting on the first two coordinates and a on E . That, in addition, g 1 ⊕ 2 1 2 R commutes with b restricts a O (R)↑ to an element of SO (R)↑. Finally, we observe 0 1 ∈ 1,1 1,1 that, if g commutes with b0, then the eigenspace decomposition of ad b0 on p(d) implies that g = (b,a) with b E and a = a a with a SO (R)↑. ∈ R 1 ⊕ 2 1 ∈ 1,1 (iv) The first part follows from the equivalence of (a) and (b) in (iii). For the second part, we observe with (iii) above that the stabilizer of WR is a subgroup of index 2 in the centralizer of rWR . ANTIUNITARY REPRESENTATIONS 41

(v) follows from the equivalence of (a) and (c) in (iii). (vi) The centralizer of r in P (d) is the subgroup E(d 2) O (R) in which the WR − × 1,1 stabilizer group P (d)WR is open. This means that (P (d), P (d)WR ) is a symmetric pair. (vii) For a closed convex subset C Rd, its recession cone ⊆ lim(C) := x Rd : x + C C = x Rd : ( c C) c + R x C { ∈ ⊆ } { ∈ ∃ ∈ + ⊆ } is a closed convex cone ([Ne00, Prop. V.1.6]), and each affine map g = (b,a) Rd ⋊ R Rd ∈ GLd( ) ∼= Aff( ) satisfies lim(gC) = lim(aC)= a lim(C). (20) If g = (b,a) S , then g maps W into itself, so that b = g(0) W . Further (20) ∈ WR R ∈ R implies that W lim(gW )= aW , and hence aW W . It follows that aE E , R ⊇ R R R ⊆ R R ⊆ R so that aER = ER as a is injective and dim ER < . This in turn implies that a commutes ∞ 2 2 2 with rWR , so that a = a1 a2 as above, where a2 O(ER) and a1WR WR. As a1WR ⊕ ∈ 2 2 ⊆ is a quarter plane bounded by light rays, we get a1WR = WR, and finally aWR = WR. Definition 4.13. ([Le15, Def. 2.7]) A d-dimensional standard pair (V,U) with transla- tion symmetry relative to W consists of a standard subspace V and a strongly ∈W ⊆ H continuous unitary positive energy representation U of the translation group Rd (cf. Def- inition 2.35) such that U V V whenever x + W W . x ⊆ ⊆ Here is the corresponding concept for von Neumann algebras: Definition 4.14. ([BLS11, 4]) A (causal) Borchers triple ( ,U, Ω) relative to the § M wedge W Rd consists of ⊆ (B1) a von Neumann algebra B( ), M⊆ H (B2) a positive energy representation (U, ) of the translation group Rd such that H U U ∗ if x + W W , and xM x ⊆M ⊆ (B3) a U-invariant unit vector Ω cs( ). ∈ M Remark 4.15. Let B( ) be a von Neumann algebra, U : Rd U( ) be a con- M ⊆ H → H tinuous unitary representation and Ω U cs( ). We consider the corresponding ∈ H ∩ M standard subspace V := Ω. Then U V V is equivalent to U U ∗ by Mh x ⊆ xM x ⊆ M Lemma 4.7. Therefore ( ,U, Ω) is a Borchers triple with respect to W if and only if M (V,U) is a standard pair with respect to W . The following theorem can be obtained by translating [Bo92] from the context of Borchers triples to standard pairs by arguing as in 4.2. We give a direct proof based on § our Theorem 3.12. Theorem 4.16 (Borchers’ standard pair Theorem). Let (V,U) be a d-dimensional stan- R R dard pair with translation symmetry relative to WR and γWR : × SO1,d 1( ) be − t tb0 → the corresponding homomorphism with γWR (e ) = e and γWR ( 1) = rWR = R01 V − (Lemma 4.12(iv)). Then the antiunitary representation (U , ) of R× corresponding H to V satisfies V V Rd R U UxU −1 = Uγ (t)x for x ,t ×, t t WR ∈ ∈ so that we obtain an antiunitary representation of Rd ⋊SO (R) by (b,γ (t)) U U V . 1,1 WR 7→ b t 42 K.-H. NEEB AND G. OLAFSSON´

Conversely, every antiunitary positive energy representation (U, ) of Rd ⋊ SO (R) H 1,1 defines a standard pair (Vγ ,U Rd ). WR | d 1,1 d 2 2 d 2 2 1,1 Proof. First we write R = R R − , so that W = W R − , where W R is ⊕ R R ⊕ R ⊆ the standard right wedge. For the light-like vectors ℓ := (1, 1, 0,..., 0) we then have ± ± 2 R R itP± WR = +×ℓ+ +×ℓ . By assumption, Utℓ± = e with P 0. The strong continuity − − ± ≥ of U implies d 2 UxV V for all x W = ([0, )ℓ+ [0, )ℓ ) R − . ⊆ ∈ ∞ − ∞ − ⊕ Now Theorem 3.12 yields V V U t U U −t = U ±t for t,s R. e sℓ± e e sℓ± ∈ Further, UxV = V for x = (0, 0, x2,...,xd 1) implies that Ux commutes with ∆V . − Combing all this, the first assertion follows. For the converse, let (U, ) be an antiunitary positive energy representation of the H group Rd ⋊ SO (R) and V = V be the standard subspace corresponding to γV := 1,1 γWR U γ by (Proposition 3.2). Since γ commutes with E , the subgroup U commutes ◦ WR WR R ER with γU and leaves V invariant. That U V V for x W 2 follows from the positive x ⊆ ∈ R energy condition and Theorem 3.13(ii). Here is a variant of this concept where the translation group is replaced by the Poincar´egroup: Definition 4.17. A d-dimensional standard pair (V,U) with Poincar´esymmetry relative to W consists of a standard subspace V and a strongly continuous unitary R ∈ W ⊆ H positive energy representation U of the connected Poincar´egroup P (d)+↑ , such that

(i) U V V for all g P (d)↑ with gW W and g ⊆ ∈ + R ⊆ R (ii) U V V ′ for all g P (d)↑ with gW W . g ⊆ ∈ + R ⊆− R Lemma 4.18. Let (U, ) be an antiunitary positive energy representation of P (d) for H + d 3 and V be the standard subspace corresponding to the canonical homomorphism ≥ ⊆ H γ : R× P (d) . Then (V,U) is a standard pair with Poincar´esymmetry. WR → + Proof. If gW = W , then g P (d) commutes with γ by Lemma 4.12(iii), so that R R ∈ + WR U V = V . Further U V V for x W (and hence also for x W by continuity) g x ⊆ ∈ R ∈ R follows from the second part of Theorem 4.16. In view of Lemma 4.12(vii), this implies that U V V if gW W . g ⊆ R ⊆ R If gW W , then the element r := R = diag(1, 1, 1, 1,..., 1) P (d)↑ R ⊆ − R 12 − − ∈ + satisfies rgWR r( WR)= WR, so that the above argument leads to UgV = UrUrgV r ⊆ − ⊆ UrV . Now γW = γW∨ yields UrV = V ′ = Vγ∨ , so that UgV V ′. R R WR ⊆ Remark 4.19. (a) In [BLS11], standard pairs with Poincar´esymmetry are used to obtain Borchers triples by second quantization (cf. Section 6). Composing with a deformation process due to Rieffel, this construction yields non-free quantum fields in arbitrary large dimensions. (b) The main point of the notion of a Borchers triple is that they can be used to construct a representation of the Poincar´egroup P (d)+↑ by generating it with modular ANTIUNITARY REPRESENTATIONS 43 one-parameter groups of a finite set of von Neumann algebras with a common cyclic separating vector ([Bo96], [Wi93c, Wi98, SW00, KW01]). (c) For d = 1, we think of R = Rd as the underlying space as a light ray in Minkowski space, so that the Poincar´egroup is replaced by the affine group Aff(R). In this context unitary “endomorphisms” of irreducible one-dimensional standard pairs (Definition 4.13) are unitary operators W U( ) commuting with the one-parameter group U satisfying ∈ H W V V . If P is the momentum operator determined by U = eitP , then these are ⊆ t precisely the operators of the form W = ϕ(P ), where ϕ is a symmetric inner function on the upper half plane C C and P is the momentum operator. A symmetric inner + ⊆ function is a bounded holomorphic function on C+ satisfying 1 ϕ(p)− = ϕ(p)= ϕ( p) for almost all p R − ∈ ([LW11, Cor. 2.4]; see also Example 3.16). That these functions can be used to construct Borchers triples was shown by Tanimoto in [Ta12]. Much more could be said about the structured related to standard subspaces, half- sided modular inclusions, modular intersections etc.. For more details and an in depth study of these concepts, we refer to [Bo97] and Wiesbrock’s work [Wi93c, Wi97b, Wi98]. 4.4. Modular geometry. In this subsection we discuss some of the geometric struc- tures arising from a single von Neumann algebra B( ) which has cyclic separating M⊆ H vectors. Any such vector ξ leads to a standard subspace V = ξ and corresponding ξ Mh modular objects (∆ξ,Jξ) (Theorem 4.2). Fixing a cyclic separating vector Ω, the asso- ciated natural cone provides a means to analyze the orbits of the group generated by U( ), U( ′) and the modular conjugations on the data. M M Definition 4.20. We consider a von Neumann algebra B( ) for which the set M ⊆ H cs( ) of cyclic and separating unit vectors is non-empty. We fix an element Ω cs( ) M ∈ M and the corresponding modular objects (∆,J) (Theorem 4.2). We recall the natural pos- itive cone := Aj(A)Ω: A , where j(A) := JAJ P { ∈ M} ([BR87, Def. 2.5.25]) and write cs( ) := cs( ) M + P ∩ M for the set of cyclic separating unit vectors in . We further write P mc( ) := J : ξ cs( ) M { ξ ∈ M } for the corresponding set of modular conjugations. We further consider the set ms( )= V = ξ : ξ cs( ) Stand( ) M { ξ Mh ∈ M }⊆ H of modular standard subspaces for and note that ∆ = ∆ and J = J . M Vξ ξ Vξ ξ We write := ′ for the center of . Z M∩M M Proposition 4.21. The following assertions hold:

(i) (Polar decomposition of cs( )) The map U( ′) cs( ) cs( ), (U, ξ) Uξ M M × M + → M 7→ is a bijection. 44 K.-H. NEEB AND G. OLAFSSON´

(ii) The unitary groups U( ) and U( ′) both act transitively on mc( ) by conjuga- M M M tion. For J mc( ), the stabilizer in both groups is the discrete central subgroup ∈ M 2 U( ′) = U( ) = Inv(U( )) = z U( ): z = 1 M J M J Z { ∈ Z } 1 of central unitary involutions. The orbit map σ : U( ) mc( ), σ(U) := UJU − M → M is a covering morphism of Banach–Lie groups if we identify mc( ) with the quo- M tient U( )/ ker σ. M (iii) grp(mc( )) = Comm(U( )) mc( ) 1,J , where Comm(U( )) is the commu- M M M { } M tator subgroup of U( ). M (iv) cs( ) := ξ cs( ): J = J = Inv(U( )) cs( ) . M J { ∈ M ξ } Z M + (v) The stabilizer of J in the group U( )U( ′) is Uj(U): U U( ) Inv(U( )). M M { ∈ M } Z (vi) For ξ cs( ), we have V = V if and only if there exists a selfadjoint operator ∈ M ξ Ω Z affiliated with , i.e., commuting with U( ) U( ′), such that ξ = ZΩ. Z M M Proof. (i) For any ξ cs( ), there exists a unit vector ξ defining the same state of ∈ M ∈P ([BR87, Thm. 2.5.31]). By the GNS Theorem, there exists a U U( ′) with ξ = Uξ. M e ∈ M Since the elements of cs( ) are also separating for ′, their stabilizer in U( ′) is M + M M e trivial.

To verify injectivity, it remains to see that every U( ′)-orbit in cs( ) meets cs( ) M M M + exactly once. Let U U( ′) and ξ cs( )+ be such that Uξ . As Jξ = J for every ∈ M ∈ M 1 ∈P ξ by [BR87, Prop. 2.5.30], we obtain UJU − = JUξ = J. Then j(U) = U leads to ∈ P 1 2 U ′ and hence to U = j(U) = U − (Theorem 4.2(c)), so that U = 1. Then ∈M∩M 1 := M : UM = M are ideals of and = + is a direct sum M± { ∈ M ± } M M ∼ M ⊕M− of von Neumann algebras. Now ξ = ξ+ ξ decomposes accordingly with ξ cs( ). ⊕ − ± ∈ M± As Uξ = ξ+ ξ and = + , it follows that ξ = 0 ([BR87, − − P P ⊕P− − ∈ P− ∩ −P− { } Prop. 2.5.28]) and thus = 0 and U = 1. M− { } (ii) If Ω1, Ω2 cs( ), then [BR87, Lemma 2.5.35] implies the existence of a unitary ∈ M 1 element U U( ′) with UJ U − = J . Exchanging the roles of and ′, it also ∈ M Ω1 Ω2 M M follows that U( ) acts transitively on mc( ). M M 1 For J mc( ) we have J J = ′, so that, for U U( ), the relation UJU − = J ∈ M M M ∈ M 1 2 implies U = JUJ . As in (i), this leads to JUJ = U ∗ = U − , so that U = 1. ∈ Z Conversely, any involution in U( ) stabilizes J. Z Clearly, σ is a surjective equivariant map whose kernel is discrete in the norm topology. As the stabilizer subgroup of J in U( ) is discrete and central, the quotient U( )/ ker σ M M carries a natural Banach–Lie group structure for which σ becomes a covering homomor- phism.

(iii) We consider the group G := U( ′) and the representation of (G G) ⋊ 1, τ M × { } on given by U(g,h,τ ε)= gJhJJ ε. Then Proposition 4.21 shows that H 1 1 U(C )= gJg− JJ : g U( ′) = gJg− : g U( ′) = mc( ). (e,e,τ) { ∈ M } { ∈ M } M Now the assertion follows from Lemma A.3.

(iv) We have already seen in (i) that Jξ = J for every ξ cs( )+. If ξ = Uξ for ∈1 M some U U( ′) and ξ cs( ) as in (i), then J = UJU − equals J if and only if ∈ M ∈ M + ξ e U ′ is an involution. This proves (iv). ∈M∩M e ANTIUNITARY REPRESENTATIONS 45

(v) Since J commutes with each operator of the form JUJU = j(U)U = Uj(U), the stabilizer contains all these elements and also Inv( ), as we have already seen above. Z If, conversely, U U( ) and W U( ′) are such that UW commutes with J, then ∈ 1M ∈ 1M UW = UJUJ(JUJ)− W with (JUJ)− W U( ′) = Inv(U( )) by (ii). ∈ M J Z (vi) We shall use the theory of KMS states (cf. [BR96]). We recall that, for a continuous action α: R Aut( ) of R on a C∗-algebra , a state ω of is called an α-KMS state → A A A if, for every pair of hermitian elements A, B , the function ∈ A ψ : R C, ψ(t) := ω(Aα (B)) → t extends analyticallty to a holomorphic function on the strip := z C: 0 < Im z < 1 , S { ∈ } extends continuously to its closure and satisfies ψ(i + t)= ψ(t) for t R. ∈ First we observe that ω(A) := Ω, AΩ is a KMS state with respect to the modular ith iti automorphism group αt(A) = ∆ A∆− (Takesaki’s Theorem, [BR96, Thm. 5.3.10]). Now let ξ cs( ) with V = V , i.e., J = J and ∆ = ∆. Then, for the same reason, ∈ M ξ Ω ξ ξ the state ω (A) := ξ, Aξ is also an α-KMS state. By [BR96, Prop. 5.3.29], there exists ξ h i a unique positive selfadjoint operator T 0 affiliated with such that ≥ Z ξ, Aξ = ω (A)= ω(√T A√T )= Ω, √T A√T Ω = √T Ω, A√T Ω for A . h i ξ h i h i ∈M Therefore ξ and √T Ω define the same state. Further √T Ω is also contained in the natural cone ([BR87, Prop. 2.5.26]). P As we have seen in (i), there exists a unique U U( ′) with Uξ cs( )+. As 1 1 ∈ M ∈ M J = J = UJ U − = U − JU, it follows from (ii) that U Inv(U( )). As Uξ and Uξ ξ ∈ Z √T Ω define the same state and both are contained in , [BR87, Thm. 2.5.31] yields P Uξ = √T Ω, i.e., ξ = U√T Ω. Now the assertion follows with Z := U√T . Suppose, conversely, that ξ = ZΩ with a selfadjoint operator affiliated to . Decom- Z posing , and Ω as a direct sum corresponding to bounded spectral projections of H M Z (which are central in as well), we may w.l.o.g. assume that Z is bounded. Since M ξ is separating, ker Z = 0 , so that we may further assume that Z is invertible. As Z { } commutes with J and ∆, it commutes with S = J∆1/2, and thus Z leaves V = Fix(S) invariant. This shows that Vξ = ZV = V . Remark 4.22. (a) Proposition 4.21(iv) describes the fibers of the map cs( ) mc( ), ξ J . M → M 7→ ξ This map is U( ′)-equivariant, so that the space cs( ) is a homogeneous U( ′)-bundle M M M over the symmetric space mc( ). M (b) Proposition 4.21(vi) describes the fibers of the map cs( ) ms( ) in terms M → M of the center . If is a factor, i.e., = C1, then we see in particular that V = V Z M Z ξ implies ξ = Ω (because ξ = 1). ± k k Lemma 4.23 (Stabilizer subgroup of V = VΩ). The stabilizer of V in the group G := U( ) U( ′) consists of all elements of the form g = uj(u)z with z Inv(U( )) and M M it it ∈ Z u U( ) fixed by the modular automorphisms α (M) = ∆ M∆− . ∈ M t Proof. Since standard subspaces are completely determined by their modular objects, 46 K.-H. NEEB AND G. OLAFSSON´ the stabilizer of V in G is 1 1 G = G G = g G: gJg− = J, g∆g− = ∆ . V J ∩ ∆ { ∈ } By Proposition 4.21(v), any g G is of the form g = uj(u)z with u U( ) and z ∈ J ∈ M ∈ Inv(U( )). As z is central, it commutes with the modular unitaries ∆it (Theorem 4.2(c)), Z i.e., z G . An element of the form g = uj(u) is fixed by each α if and only if ∈ V t 1 1 αt(u)Jαt(u)J = uJuJ, resp. u− αt(u)= Juαt(u)− J. 1 Then u− α (u) J J = . To see that this implies that α (u)= u, we consider t ∈ M ∩ M Z t the commutative von Neumann subalgebra generated by the center and u. A⊆M Z As each α fixes the center pointwise, we have α ( ) = for every t R. Then the t t A A ∈ state of given by ω(A) := Ω, AΩ is a KMS state with respect to αt , so that the A h i |A restrictions αt are the unique automorphisms corresponding to this KMS states ([BR96, |A Thm. 5.3.10]). Since is abelian, the uniqueness of the automorphism group implies its A triviality. We conclude that each α fixes u if g G . t ∈ V If, conversely, αt(u) = u, then αt fixes g. This implies that g commutes with S = J∆1/2, hence preserves V = Fix(S). Example 4.24. (a) If is a factor, then Inv(U( )) = 1 , so that U( ) = 1 M Z {± } M J {± } and mc( ) = U( )/ 1 . M ∼ M {± } (b) For = B( ) acting on = B ( ) by left multiplications, we have JA = A∗ M K H 2 K (Example 4.6(b)) and by (a), we have mc(B( )) = U( )/ 1 . For any 0 < Ω=Ω∗ K ∼ K {± } ∈ cs( ) we have = A B ( ): A 0 . M P { ∈ 2 K ≥ } If Ω = √D holds for the trace class operator D > 0, then the centralizer of ∆ in U( ) is M 1 1 g U( ): g∆g− = ∆ = g U( ): gDg− = D , { ∈ M } ∼ { ∈ K } and since D is diagonalizable, this subgroup consists of those unitaries leaving all eigen- spaces of D invariant. In particular

A B ( ): A = A∗, [A, D]=0 V V ′ { ∈ 2 K }⊆ ∩ shows that V V ′ is much larger than RΩ. For every elements A = A∗ with ∩ 6∈ P ∪ −P ker A = 0 , we have JA = A but JA = J (Proposition 4.21(iv)). { } 6 2 (c) For = L∞(X, S,µ), µ finite, acting on = L (X,µ) by multiplication opera- M H tors, we find for Jf = f that U( ) is the set of involutions in U( ). As the squaring M J M map U U 2 is a morphism of Banach–Lie groups, the Banach symmetric space mc( ) 7→ M is diffeomorphic to the unitary group and we have a short exact sequence 1 Inv(U( )) U( ) mc( ) 1. → M → M → M → 5. Nets of standard subspaces and von Neumann algebras. In this section we briefly discuss some elementary properties of nets of standard subspaces (Vℓ)ℓ L and ∈ their connection with antiunitary representations (U, ). The connection with nets of H von Neumann algebras and QFT is made in 5.2. Nets of standard subspaces are con- § siderably simpler than nets of von Neumann algebras and naturally determined by an V antiunitary representation, of the group generated by all subgroups U ℓ (R×) AU( ) ⊆ H (Proposition 3.2), but this group need not be finite dimensional. ANTIUNITARY REPRESENTATIONS 47

5.1. Nets of standard subspaces. Let := (Vℓ)ℓ L be a family of standard subspaces V ∈ of the Hilbert space . We assume that the map ℓ V is injective, so that the index H 7→ ℓ set L is only a notational convenience and we could equally well work directly with the subset V : ℓ L Stand( ) (cf. [BGL02, SW03]). { ℓ ∈ }⊆ H Definition 5.1. A net automorphism is an α AU( ) permuting the standard sub- ∈ H spaces V . We write Aut( ) AU( ) for the subgroup of net automorphisms. A net ℓ V ⊆ H automorphism is called an internal symmetry if it preserves each Vℓ separately. The cor- responding subgroup of Aut( ) is denoted Inn( ). V V We write (∆ℓ,Jℓ) for the modular objects corresponding to Vℓ and consider the mod- ular symmetry group := grp( J : ℓ L ) AU( ) J { ℓ ∈ } ⊆ H generated by the modular conjugations. A natural assumption enriching the underlying geometry is the condition of geometric modular action (CGMA) Aut( ) J ⊆ V (see [BS93, BDFS00] for the von Neumann context). The condition (MS) ∆it for all t R,ℓ L Vℓ ∈ J ∈ ∈ is called the modular stability condition ([BDFS00], [Bo00, IV.5.6/7]). §§ From now on we assume that (CGMA) is satisfied. We obtain an action σ : Aut( ) L L, (g,ℓ) σ (ℓ) on the index set L by V × → 7→ g

gVℓ = Vσg (ℓ). (21) This further implies 1 1 gJℓg− = Jσg (ℓ) and g∆ℓg− = ∆σg (ℓ). (22) In particular, (CGMA) implies that := J : ℓ L S { ℓ ∈ } is a conjugation invariant set of generators of . This fact opens the door to the con- J struction of geometric structures from the group and its generating set in specific J S situations. Lemma 5.2. The subgroup of of those elements acting trivially on L is its center J := g : ( ℓ L) gV = V = Inn( ) = ker σ. Z { ∈ J ∀ ∈ ℓ ℓ} V ∩ J Proof. For g , the relation σ = id implies that g commutes with every element ∈ J g L J by (22). As is generated by , the assertion follows. ℓ ∈ S J S By Lemma 5.2, the action of on the index set describes this group as a central J extension of the group / that acts faithfully on the set L which is supposed to carry J Z geometric information (cf. [BDFS00] and Remark 5.5 below). An immediate consequence of (CGMA) is that the net is invariant under the passage to the symplectic orthogonal space V V ′ = J V (cf. 3.2). In particular, we have a ℓ 7→ ℓ ℓ ℓ § duality map ℓ ℓ′ := σ (ℓ) on L. We also have a natural order structure on L by 7→ Jℓ ≤ ℓ ℓ if V V . 1 ≤ 2 ℓ1 ⊆ ℓ2 48 K.-H. NEEB AND G. OLAFSSON´

Remark 5.3. The key properties of the triple (L, ,′ ), given by the partial order and ≤ ≤ the duality map ℓ ℓ′ are that 7→ (A1) ℓ ℓ implies ℓ′ ℓ′ , and 1 ≤ 2 2 ≤ 1 (A2) ℓ ℓ′ if and only if ℓ ℓ′ . 1 ≤ 2 2 ≤ 1 From (A2) we immediately derive ℓ ℓ′′ and by combining this with (A1), we obtain ≤ ℓ′ = ℓ′′′ for every ℓ L, i.e., the duality map restricts to an involution on its range. ∈ 1,d 1 Examples 5.4. (a) For a subset S of the Minkowski space R − , we define the causal complement by 1,d 1 S′ := x R − : ( y S)[x y, x y] < 0 . { ∈ ∀ ∈ − − } Then S′ = s S s ′, which immediately leads to (A1/2). Here S T ′ means that S ∈ { } ⊆ and T are space-likeT separated, and S′′ is the causal completion of S. For g P (d) and 1,d 1 ∈ S R − , we have (gS)′ = gS′. ⊆ R1,d 1 For the standard right wedge WR − , we have WR′ = WR and WR′′ = WR, ⊆ 1,d −1 and for the positive light cone V we have V ′ = and V ′′ = R − . For x y V , the + + ∅ + − ∈ + causal completion

x, y ′′ = (x V ) (y + V )= { } − + ∩ + Ox,y is the closure of the double cone (cf. Remark 5.13(b)). Ox,y (b) If (X, ) is a partially ordered space, then we define ≤ x ′ := y X : x y,y x and S′ := s ′. { } { ∈ 6≤ 6≤ } { } s\S ∈ Then the set L of subsets of X, endowed with the inclusion order, satisfies (A1/2). (c) For a complex Hilbert space , the set of real subspace V , endowed with the H R ⊆ H inclusion order and the symplectic orthogonal space V ′ = iV ⊥ satisfies (A1/2). (d) For a complex Hilbert space , the set of von Neumann subalgebras B( ), H M⊆ H endowed with the inclusion order and the commutant map ′ satisfies (A1/2). M 7→ M Remark 5.5. (a) In QFT, one expects that the structure (L, ,′ ) encodes physical ≤ information and one would like to recover information on the geometry of spacetime from this structure. In this context, causal complements, resp., the notion of being space- like separated, appears more fundamental than the causal order if we want to recover a spacetime M from the triple (L, ,′ ), where L consists of certain subsets of M but does ≤ not contain one-point sets (cf. [Ke96]). If the modular stability condition is satisfied, i.e., if contains also the modular J unitaries, this group is supposed to encode the dynamics of the quantum theory, the isometry group of the corresponding spacetime and a (projective) unitary representation of this group ([Su05, 6.4]). This connects naturally with the approach of Connes and § Rovelli who “construct” the dynamics of a quantum statistical system by a modular one-parameter group ∆it ([CR94]). Here an interesting result concerning the detection of known group from this viewpoint is the characterization of the Poincar´egroup in terms of structure preserving maps on 1,d 1 the set of wedges in R − ([BDFS00], [Bo00, IV.5], Lemma 4.12). W § ANTIUNITARY REPRESENTATIONS 49

(b) In [Ke98, 4] the relation of the causal structure on spacetime and how it can be § determined by data encoded in nets of C∗-algebras is discussed very much in the spirit of this section. We refer to [Ra17] for an approach to quantum field theory based on modular theory of operator algebras that does not assume an a priori given underlying spacetime. Instead, one would like to generate the spacetime geometry from operator theoretic data. (c) In [SW03] this program is carried out to a large extent by specifying a set of axioms formulated in terms of the modular conjugations Jℓ, such that the index set L corresponds to the set of wedges in three-dimensional Minkowski space R1,2, J corresponds to the W W the orthogonal reflection r P (3) in the edge of W and = P (3) (cf. Lemma 4.12). W ∈ + J ∼ + In this case, the subset SO1,2(R)↑WR identifies naturally with the anti-deSitter 2 R R ⊆W space AdS ∼= SO1,2( )↑/ SO1,1( )↑, which can be realized as an adjoint orbit in the Lie R R algebra so1,2( ) ∼= sl2( ) (cf. Lemma 4.12(v)). The following construction is of fundamental importance in our approach. It is inspired by the modular localization approach to QFT developed in [BGL02, Thm. 2.5]: Proposition 5.6. (Nets of standard subspaces from antiunitary representations; the BGL construction) Let (U, ) be an antiunitary representation of the Lie group pair H Vγ (G, G ) and associate to γ : (R×, R×) (G, G ) the standard subspace V with U = 1 + → 1 γ U γ (Proposition 3.2). Then, for every non-empty subset Γ Hom((R×, R×), (G, G )) ◦ ⊆ + 1 invariant under conjugation with elements of G and under inversion, we thus obtain a net (Vγ )γ Γ of standard subspaces which satisfies (CGMA) for the group which is the image ∈ J under U of the subgroup of G generated by the conjugation invariant set of involutions γ( 1): γ Γ . { − ∈ } ev Γ −1 Inv(G G ) −−−−−−−−−→ \ 1 γ Vγ U 7→  V JV  Stand( ) 7→ Conj( ). y H −−−−−−−−−→ y H g 1 Proof. This follows from the observation that, for γ (t) := gγ(t)g− , we have UgVγ = Vγg for g G and U V = V ′g for g G , so that G acts naturally by automorphisms on ∈ 1 g γ γ 6∈ 1 the net (Vγ )γ Γ. ∈ Remark 5.7. (a) Evaluation in 1 leads to a fibration − R R ev 1 : Hom(( ×, +×), (G, G1)) Inv(G G1). − → \ An involution r G G is contained in the image if and only if there exists an x g ∈ \ 1 ∈ fixed by Ad(r), i.e., if gAd(r) = ker(Ad(r) 1) = 0 . This is always the case if g is − 6 { } non-abelian, i.e., if idg is not an automorphism. Then the fiber over r can be identified − with the Lie subalgebra gAd(r). (b) In many situations one considers minimal sets

g Γ= C C ∨ , where C := γ : g G . γ ∪ γ γ { ∈ } Then ev 1(Cγ )= Cr is the conjugacy class of the involution r := γ( 1) G G1, hence − − ∈ \ in particular a symmetric space (cf. Appendix A.3). An important example in QFT is

γ = γWR for G = P (d)+ (Lemma 4.12). 50 K.-H. NEEB AND G. OLAFSSON´

(c) In the context of Proposition 5.6, the relation Vγ′ = Vγ∨ shows that duality is naturally built into the construction. However, in general it may not be so easy to de- termine when V V . In [BGL02, Thm. 3.4] it is shown that, for G = P (d) and γ1 ⊆ γ2 + homomorphisms (γW )W corresponding to wedges, the relation W1 W2 is equivalent ∈W ⊆ to V V if and only if U is a positive energy representation. γW1 ⊆ γW2 The preceding discussion suggests a closer look at conjugacy classes of involutions τ G G . We write C G for the conjugacy class of G. ∈ \ 1 τ ⊆ Lemma 5.8. Let G be a connected Lie group with Lie algebra g and τ Aut(G ) be an 1 ∈ 1 involutive automorphism. Then the conjugacy class C G := G ⋊ 1, τ generates the τ ⊆ 1 { } subgroup grp(Cτ ) = B 1, τ , where B is the integral subgroup whose Lie algebra is the τ τ { τ } τ τ τ ideal b := g− + [g− , g− ]. In particular, Cτ generates G if and only if g = [g− , g− ].

Proof. Let H = grp(Cτ ) G be the subgroup generated by Cτ . As τ H, we have ⊆ ∈ 1 H = B e, τ for B := H G1. Then B is generated by the elements of the form gτ(g)− , { } ∩ τ g G1, hence in particular arcwise connected. For x g− , we therefore obtain exp(2x)= ∈ ∈ τ exp(x)τ(exp( x)) B, so that the Lie algebra b of B contains g− and hence also τ −τ τ∈ b = g− + [g− , g− ], which is an ideal of g1. Let G1 denote the universal covering group of G1. Then B := exp e b is a normal h G1 i integrale subgroup of G1, hence closed. As τ acts trivially on thee quotient group G1/B, all 1 elements of the form gτ(g)− are contained in B. Therefore B := exp b contains the e h G i e e arcwise connected subgroup H G , and thus H G = B. This implies the lemma. ∩ 1 e ∩ 1 Example 5.9. G = Aff(R) and τ = (x, 1) with C = R 1 , and this conjugacy − τ × {− } class generates G. Lemma 5.10. Let τ = r P (d) be a wedge reflection for some W . Then W ∈ + ∈W (i) The conjugacy class Cτ of τ generates P (d)+ if and only if d> 2. (ii) The conjugacy class Cτ of τ in the conformal group SO2,d(R) generates the whole group for any d> 0. Proof. (i) Since all wedges W are conjugate to the standard right wedge W , it ∈ W R suffices to consider τ = r = R = diag( 1, 1, 1,..., 1). If d = 2, then R = 1, so WR 01 − − 01 − that grp(C )= R2 ⋊ 1 is a proper subgroup of P (2) . τ {± } + The case d = 2 already implies that grp(Cτ ) contains all translations in the directions of all Lorentzian 2-planes, hence all translations. Therefore it suffices to show that the conjugacy class of R01 in SO1,d(R) generates the whole group. In view of Lemma 5.8, this follows from the simplicity of the real Lie algebra g = so1,d 1(R). 1,d 1 − (ii) We consider SO2,d(R) as a group acting on R − by rational maps (cf. [HN12, 17.4]). We have already seen above that the group grp(C ) generated by the conjugacy § τ class Cτ in SO2,d(R) contains the Poincar´egroup P (d)+, which is a parabolic subgroup of SO2,d(R) and it intersects both connected components. By the same argument, it contains the opposite parabolic subgroup, and both subgroups generate SO2,d(R) because it has only two connected components (cf. [Be96]). 1,d 1 If d is odd, then SO (R) = O (R)/ 1 is the full conformal group of R − , but 2,d ∼ 2,d {± } if d is even, then the kernel 1 of the action of O (R) is contained in the identity {± } 2,d ANTIUNITARY REPRESENTATIONS 51 component, so that Conf(R1,d) = O (R)/ 1 has four connected components ([HN12, ∼ 2,d {± } 17.4]). Therefore the conjugacy class of a wedge reflection does not generate the whole § conformal group. Remark 5.11. In [BGL02, Thm. 4.7], Brunetti, Guido and Longo describe a one-to-one correspondence between antiunitary positive energy representations of P (d)+ and certain nets of closed real subspaces V indexed by certain open subsets Rd, for which the O O⊆ subspaces (VW )W corresponding to wedges are standard and the modular covariance ∈W condition it/2π − ∆W V = VγW (t) O O holds for the homomorphisms γ : R× P (d) and the modular unitaries of V . W → + W The uniqueness of the local net, once the unitary representation is given, is discussed in [BGL02, Rem. 4.8] (see also [BGL93]). For the converse, i.e., the uniqueness of the unitary representation, once the local net is given, we refer to [BGL93]. In [Mu01], Mund shows that, for any representation (U, ) of P (d)↑ that is a finite direct sum of irreducible H + representations of strictly positive mass, there is only one covariant net of standard subspaces; which therefore coincides with the one obtained in Proposition 5.6 from any antiunitary extension of U to P (d)+. Example 5.12. (Nets arising from a single von Neumann algebra) (a) Let B( ) be a von Neumann algebra for which cs( ) = and consider the M ⊆ H M 6 ∅ corresponding set := ms( )= V : ξ cs( ) V M { ξ ∈ M } of standard subspaces (Definition 4.20). Fix fix a cyclic separating vector Ω and the corresponding modular objects (∆,J) and consider the group

G := U( ) U( ′) 1,J AU( ). M M { }⊆ H It is easy to see that this group permutes the standard subspaces in . From Proposi- V tion 4.21(ii) we derive that 1 mc( )= gJg− : g G M { ∈ } 1 is the conjugacy class of J in G. We also note that the G-orbit g g− : g G = { M ∈ } , ′ of in the set of von Neumann subalgebras of B( ) consists only of two {M M } M H elements. (b) Consider the group ♯ t it/2π G := U( ) U( ′)γ(R×) for γ( 1) := J and γ(e ) := ∆− . M M − ♯ That G is a group follows from the fact that γ(R×) normalizes U( ) and U( ′), + M M whereas conjugation by J = γ( 1) exchanges both. This group is strictly larger than − it it U( ) U( ′) 1,J if the modular automorphisms α (M) := ∆ M∆− of are not M M { } t M inner. If ξ cs( ) is different from Ω, then Connes’ Radon Nikodym Theorem5 implies ∈ M the existence of a strongly continuous path of unitaries (ut)t R in U( ) such that the ∈ M 5 See [Bla06, Thm. III.4.7.5], [BR96, Thm. 5.3.34], and in particular [Fl98] for a quite direct proof. 52 K.-H. NEEB AND G. OLAFSSON´

ξ it it corresponding modular automorphism group αt (M) = ∆ξ M∆ξ− satisfies ξ α (M)= u α (M)u∗ for M ,t R. t t t t ∈M ∈ it it ♯ it This implies that ∆− u ∆ U( ′), so that G also contains the operators ∆ . Hence ξ t ∈ M ξ the net of standard subspaces of specified by the conjugacy class of the antiunitary ♯ H ♯ representation γ Hom(R×, G ) coincides with the orbit G V = GV Stand( ). ∈ ⊆ H 5.2. Nets of von Neumann algebras. The context that actually motivates the consid- eration of families of standard subspaces are families ( ℓ)ℓ L of von Neumann algebras M ∈ on some Hilbert space . In the theory of algebras of local observables, one considers ℓ as H indicating the “laboratory” in which observables corresponding to can be measured, Mℓ and then L is the set of laboratories (cf. [Ha96, Ar99, Bo97]). We write B( ) for the von Neumann algebra generated by all the algebras . M⊆ H Mℓ We shall discuss several properties of these families and relate them to antiunitary rep- resentations and some results in Algebraic Quantum Field Theory (AQFT). Our first assumption is the Reeh–Schlieder property: (RS) There exists a unit vector Ω that is cyclic and separating for each . Mℓ By the Tomita–Takesaki Theorem, (RS) leads to a family of standard subspaces given V := Ω ℓ Mℓ,h and the map ℓ V is injective if and only if the map ℓ is injective (Lemma 4.7). 7→ ℓ 7→ Mℓ This leads us to the setting of the preceding subsection, so that everything said there applies in particular here. As each Jℓ fixes Ω, it is fixed by the whole group . For g 1 g J g and := g g− , we therefore have gV = Ω, so that Lemma 4.7 implies ∈ J Mℓ Mℓ ℓ Mℓ,h 1 that gV = Ve for some ℓ L is equivalent to g g− = e. Hence the condition ℓ ℓ ∈ Mℓ Mℓ of geometric modular action (CGMA) from the preceding section is equivalent to the e following ([BDFS00]): (CGMA) Conjugation with elements of the group permutes the von Neumann algebras J ( ℓ)ℓ L. M ∈ The relation Jℓ ℓJℓ = ′ then implies that the net ( ℓ)ℓ L is invariant under the M Mℓ M ∈ passage to the commutant. Remark 5.13. (a) In quantum field theory, where L often is the set of wedges in W Minkowski space R1,3, [BDFS00, Thm. 5.2.6] asserts that (CGMA) basically is equivalent to the duality condition (W ′) = (W )′ for every W . Then one obtains an M M ∈ W antiunitary representation of the Poincar´egroup P (4)+ fixing Ω and acting covariantly on the net. Further, UrW = JW (cf. Lemma 4.12) and the spectrum of the translation subgroup is either contained in V or in V , i.e., we either have positive or negative + − + energy representations. (b) For x, y R1,3 and x y V , the open causal interval ∈ − ∈ + := (x V ) (y + V ) Ox,y − + ∩ + is called a double cone. There are various Reeh–Schlieder Theorems, that provide suffi- cient conditions for the vacuum vector to be cyclic and separating for an algebras ( ) M O ANTIUNITARY REPRESENTATIONS 53 of local observables attached to an open subset of Minkowski space ([Bo92, RS61, Bo68]). The most classical results concern the cyclicity of the vacuum for double cone algebras ( ). Since every wedge contains double cones, the vacuum is also cyclic and sepa- M Ox,y rating for wedge algebras (W ), W . This leads to modular objects (∆ ,J ), so M ∈W W W that the condition (RS) in 5.2 holds for the index set L = . § W (c) For nets of von Neumann algebras ( ) of local oberservables associated to M O regions in some spacetime M, it is important to specify those regions behaving well O with respect to our assumptions. In [Sa97] they are called test regions. This requires in particular that the vacuum vector Ω should be cyclic for ( ) (the Reeh–Schlieder M O property) and that a suitable duality holds ( )′ = ( ′), where ′ is the (interior M O M O O of the) causal complement of . Prototypical examples of test domains are wedges W in O Minkowski space (or its conformal completion) [BGL02, Thm. 2.5], but in many situations larger classes also have these properties, such as double cones or spacelike cones, i.e., translates of convex cones R , where is a double cone not containing 0. In this +D D context the CGMA is a natural additional requirement for test regions that ties the corresponding modular structure to spacetime geometry. (d) For a Haag–Kastler net ( ) (as in the introduction), the (CGMA) for the net A O π ( ( ))′′ of von Neumann algebras specified by a state ω of can be seen as a require- ω A O A ment that selects states which are particularly natural (cf. [BDFS00, p. 485]).

Under certain assumptions on the corresponding net of local observables, the Bisog- nano–Wichmann Theorem ([BW76, So10]) asserts that the antiunitary representation (U, ) of P (4)+ obtained from the PCT Theorem, where Θ = U 1 is the antiunitary H − PCT operator, has the property that the boost generator b0 from (5) in Example 2.27 satisfies

it/2π tb0 ∆− = U(e ) and JW = U(rW )=ΘU(diag(1, 1, 1, 1)). WR R R − − The first relation is called the modular covariance relation (cf. [Mu01, p. 911]). In [GL95, Props. 2.8,2.9] Guido and Longo show that modular covariance implies covariance of the corresponding modular conjugations which in turn implies the PCT Theorem. In the context of standard subspaces, the Bisognano–Wichmann Theorem and the PCT Theorem were derived in by Mund ([Mu01, Thm. 5]).

Examples 5.14. (Conformal invariance) Beyond the Bisognano–Wichmann Theorem, the following geometric implementations of modular automorphism groups are known: 1,d 1 (a) In [Bu78], Buchholz shows that, for a free scalar massless field on R − (which automatically enjoys conformal symmetry), for d > 2 the dilation group γV+ (a)(x) = ax, a R×, corresponds to the modular objects of the light cone algebra (V ). ∈ M + As we shall see in (b) below, the light cone is conformally equivalent to the right wedge WR. Therefore γV+ is conjugate in the conformal group to the homomorphism 1,d 1 γ : R× Conf(R − ), corresponding to the right wedge W , which occurs in the WR → R Bisognano–Wichmann Theorem. (b) In [HL82], Hislop and Longo obtain similar results for double cones in the context of massless scalar fields by conjugating them conformally to light cones and then apply 54 K.-H. NEEB AND G. OLAFSSON´

[Bu78]. More concretely, the relativistic ray inversion 1 ρ: x = (t, x) (t, x), [x, x]= t2 x2 7→ [x, x] − (which is an involution), exchanges the translated right wedge r r WR + e1 = (x0, x): x1 > + x0 2 n 2 | |o with the double cone e0 e1 e0 + e1 e0−e1 , −e0−e1 = − V+ + V+ . O r r  r −  ∩  − r  It also exchanges the double cone = (re V ) V and the light cone e0 +V (see Ore0,0 0 − + ∩ + r + [Gu11, p. 111]). With these explicit transformations, one also obtains the corresponding one-parameter groups of automorphisms and the corresponding conformal involutions. For the light cone V+, we know from (a) that the corresponding automorphism group is given by the dilations γV+ (t)x = tx. So it follows in particular, that it is confor- mally conjugate to the Lorentz boosts γ : R× P (4) corresponding to a wedge W W → + (Lemma 4.12). As a consequence of this discussion, the modular automorphism groups corresponding to the local observable algebras associated to double cones, light cones and wedges are R1,d 1 R conjugate under the conformal group Conf( − ) ∼= O2,d( )/ 1 . In particular, they {± } 1,d 1 correspond to a single conjugacy class of homomorphism γ : R× Conf(R − ) which → is most simply represented by γV+ . Example 5.15. (cf. Example 2.28) In the one-dimensional Minkowski space R, the order intervals are represented by the open interval ( 1, 1) transformed by the Cayley map 1+x − 1 c(x) := to (0, )= V and the involution σ(x) := x− maps ( 1, 1) to its (confor- 1 x ∞ + − mal) complement.− These are the geometric transformations corresponding to the modular operators on the double cone algebra ( 1, 1) for d = 1. M O − Example 5.16. Interesting examples of nets of von Neumann algebras with (CGMA) arise from [BDFS00, Thm. 4.3.9], where the index set is the set of wedges in R1,3. Under W suitable continuity assumptions, one obtains a continuous antiunitary representation of P (4)+ with U = J and = U . rW W J P (4)+ Here a key point is that P (4)+ is generated by the conjugacy class of the wedge reflection rWR (Lemma 5.10). Remark 5.17. A key observation in the work of Borchers and Wiesbrock is that von Neumann algebras of local observables corresponding to two wedges having a light ray in common define modular intersections ([Wi98, Prop. 7]). That one can deal with them as pairs without any direct reference to the intersection (cf. Theorem 3.23) is crucial because the modular group of the intersection need not be implemented geometrically [Bo96]. This is of particular interest for QFT on de Sitter space dSd whose isometry group O1,d(R) has no positive energy representations for d> 2. ANTIUNITARY REPRESENTATIONS 55

6. Second quantization and modular localization. In this section we explain how Second Quantization, i.e., the passage from a (one-particle) Hilbert space to the corre- H sponding Fock spaces ( ) (bosonic and fermionic) provides for each standard subspace F± H V pairs ( ±(V ), Ω), where ±(V ) is a von Neumann algebra on ( ) and the ⊆ H R R F± H vacuum vector Ω ( ) is cyclic. ∈F± H Let be a complex Hilbert space and let H ∞ bn ( ) := ⊗ F H n=0H Md be the full Fock space over . We write ( ) for the subspace of symmetric tensors, the H F+ H bosonic Fock space, and ( ) for the subspace of skew-symmetric tensors, the fermionic F− H Fock space. Both spaces carry a natural representation Γ : AU( ) AU( ( )) ± H → F± H of the antiunitary group AU( ) given by H Γ+(U)(v1 vn) := Uv1 Uvn, Γ (U)(v1 vn) := Uv1 Uvn. ∨···∨ ∨···∨ − ∧···∧ ∧···∧ Moreover, the bosonic Fock space carries a unitary representation of the Heisenberg group Heis( ) ( 6.1) and its subgroups can be used to derive a net of von Neumann algebras on H § ( ). A similar construction can be carried out for the fermionic Fock space in terms F+ H of the natural representation of the C∗-algebra CAR( ), a C∗-algebra defined by the H canonical anticommutation operators. Both constructions are functorial and associate to every antiunitary representation (U, ) of (G, G1) on a covariant family ( γ )γ Γ of H H M ∈ von Neumann algebras on ( ), where Γ is as in Proposition 5.6. F± H 6.1. Bosonic Fock space. We start with the construction of the von Neumann algebras on the bosonic Fock space. For v ,...,v , we define 1 n ∈ H 1 v1 vn := v1 vn := vσ(1) vσ(n) ··· ∨···∨ √n! ⊗···⊗ σXSn ∈ n n and v := v∨ , so that v v , w w = v , w v , w . (23) h 1 ∨···∨ n 1 ∨···∨ ni h σ(1) 1i···h σ(n) mi σXSn ∈ 1 For every v , the series Exp(v) := ∞ vn defines an element in ( ) and the ∈ H n=0 n! F+ H scalar product of two such elements is givenP by

∞ n! n v,w Exp(v), Exp(w) = v, w = eh i. h i (n!)2 h i nX=0 These elements span a dense subspace of ( ), and therefore we have for each x F+ H ∈ H a unitary operator on ( ) determined by the relation F+ H 2 x,v kxk U Exp(v)= e−h i− 2 Exp(v + x) for x, v . (24) x ∈ H A direct calculation then shows that i Im x,y U U = e− h iU for x, y . (25) x y x+y ∈ H 56 K.-H. NEEB AND G. OLAFSSON´

To obtain a unitary representation, we have to replace the additive group of by the H Heisenberg group i Im v,v′ Heis( ) := T with (z, v)(z′, v′) := (zz′e− h i, v + v′). H × H For this group, we obtain with (25) a unitary representation U : Heis( ) U( ( )) by U := zU . H → F+ H (z,v) v In this physics literature, all this is expressed in terms of the so-called Weyl operators W (v) := U , v iv/√2 ∈ H satisfying the Weyl relations i Im v,w /2 W (v)W (w)= e− h i W (v + w), v,w . (26) ∈ H Definition 6.1. To each real subspace V , we assign the von Neumann algebra + ⊆ H (V ) := (V ) := W (V )′′ B( ( )) on the bosonic Fock space of . R R ⊆ F+ H H Lemma 6.2. We have (i) ( )= B( ( )), resp., the representation of Heis( ) on ( ) is irreducible. R H F+ H H F+ H (ii) (V ) (W )′ if and only if V W ′ (locality). R ⊆ R ⊆ (iii) (V )= (V ). R R (iv) Ω = Exp(0) ( ) is cyclic for (V ) if and only if V + iV is dense in . ∈F+ H R H (v) Ω ( ) is separating for (V ) if and only if V iV = 0 . ∈F+ H R ∩ { } (vi) Ω cs( (V )) if and only if V is standard. ∈ R Proof. (i) is well-known ([BR96, Prop. 5.2.4(3)]). (ii) follows directly from the Weyl relations (26). (iii) follows from the fact that B( ( )), v W is strongly continuous and H → F+ H 7→ v (V ) is closed in the weak operator topology. R (iv) Assume that := V + iV = . Then (V )Ω ( ), so that Ω cannot be K 6 H R ⊆ F+ K cyclic. Suppose, conversely, that = and that f ( (V )Ω)⊥. Then the holomorphic K H ∈ R function f(v) := f, Exp(v) on vanishes on V , hence also on V + iV , and since this h i H subspace is dense in , we obtain f = 0 because Exp( ) is total in ( ). We conclude b H H F+ H that Ω is cyclic. (v) In view of (iii), we may assume that V is closed. Let 0 = w V iV . To see that 6 ∈ ∩ Ω is not separating for (V ), it suffices to show that, for the one-dimensional Hilbert R space := Cw, the vector Ω is not separating for (Cw)= B( (Cw)) (which follows H0 R F+ from the irreducibility of the representation of Heis(Cw) on (Cw)). This is obviously F+ the case because dim (Cw) > 1. F+ Suppose that = 0 . As = V ′′ (iV ′′) = (V ′ + iV ′)′, it follows that V ′ + iV ′ K { } K ∩ is dense in . By (ii), Ω is cyclic for (V ′) which commutes with (V ). Therefore Ω is H R R separating for (V ). R (vi) follows from (iv) and (v).

Remark 6.3. (a) (V ) is commutative if and only if V V ′. For a standard subspace R ⊆ V the relation V ′ = JV shows that this is equivalent to V = V ′, respectively to ∆ = 1 ANTIUNITARY REPRESENTATIONS 57

(Lemma 3.7). (b) The imaginary part ω(ξ, η) := Im ξ, η turns into a symplectic manifold ( ,ω). h i H H From this perspective, we may consider the algebras (V ) as “quantizations” of the R J algebra of measurable functions on the Lagrangian subspace E := . If V = V ′, 2 H then ( ) = L (E∗,γ), where γ is a Gaussian probability measure on the algebraic F+ H ∼ E∗ of E, endowed with the smallest σ-algebra for which all evaluation maps are measurable. Then the commutative von Neumann algebra (V ) is isomorphic to R L∞(E∗,γ). In general, if V V ′, then (V ) is non-commutative and the degree of non- 6⊆ R commutativity depends on the non-degeneracy of ω on V . It is “maximal” if V V ′ = 0 , ∩ { } which implies that (V ) is a factor by the following theorem. R Theorem 6.4. ([Ar63, Thm. 1]) For closed real subspaces V, W, V of , the following j H assertions hold: (i) (V ) (W ) if and only if V W (isotony). R ⊆ R ⊆ (ii) R j J Vj = j J (Vj ), where j J Vj denotes the closed subspace generated by ∈ ∈ R ∈ the WVj and  j WJ (Vj ) denotes theW von Neumann algebra generated by the (Vj ). ∈ R R (iii) R j J Vj W= j J (Vj ). ∈ ∈ R (iv) (VT)′ = (V ′)T(duality). R R (v) (V ) (V ′)= (V V ′). In particular, the algebra (V ) is a factor if and only R ∩ R R ∩ R if V V ′ = 0 . ∩ { } 6.2. Fermionic Fock space. On the fermionic Fock space, the construction of the von Neumann algebras is slightly different but similar in spirit. For v ,...,v , we define 1 n ∈ H 1 v1 vn := sgn(σ)vσ(1) vσ(n), ∧···∧ √n! ⊗···⊗ σXSn ∈ so that v v , w w = sgn(σ) v , w v , w (27) h 1 ∧···∧ n 1 ∧···∧ ni h σ(1) 1i···h σ(n) mi σXSn ∈ In 0 ( ) = C we pick a unit vector Ω, called the vacuum. F− H ∼ Definition 6.5. The CAR-algebra CAR( ) of is a C∗-algebra, together with a contin- H H uous antilinear map a: CAR( ) satisfying the canonical anticommutation relations H → H

a(f),a(g)∗ = f,g 1 and a(f),a(g) =0 for f,g (28) { } h i { } ∈ H and which has the universal property that, for any C∗-algebra and any antilinear A map a′ : satisfying the above anticommutation relations, there exists a unique H → A homomorphism ϕ: CAR( ) with ϕ a = a′. This determines the pair (CAR( ),a) H → A ◦ H up to isomorphism ([BR96, Thm. 5.2.8]). We write a∗(f) := a(f)∗ and observe that this defines a complex a∗ : CAR( ). H → H Remark 6.6. The C∗-algebra CAR( ) has an irreducible representation (π0, ( )) on H F− H the fermionic Fock space ( ) ([BR96, Prop. 5.2.2(3)]). The image c(f) := π0(a(f)) F− H 58 K.-H. NEEB AND G. OLAFSSON´ acts by c(f)Ω = 0 and n j 1 c(f)(f1 fn)= ( 1) − f,fj f1 fj 1 fj+1 fn. ∧···∧ − h i ∧···∧ − ∧ ∧···∧ Xj=1 Accordingly, we have

c∗(f)Ω = f and c∗(f)(f f )= f f f . 1 ∧···∧ n ∧ 1 ∧···∧ n Consider the hermitian operators

b(f) := c(f)+ c∗(f) CAR( ) (29) ∈ H and note that

b(f),b(g) = c(f),c∗(g) + c∗(f),c(g) = f,g 1 + g,f 1 =2β(f,g)1, (30) { } { } { } h i h i where β(f,g)=Re f,g for f,g h i ∈ H is the real scalar product on . H Definition 6.7. Let = be a 2-graded Hilbert space. Accordingly, B( ) H H0 ⊕ H1 H inherits a grading and therefore a Lie superbracket which on homogeneous elements is given by A B [A, B] := AB ( 1)| || |BA, τ − − where A denotes the degree of a homogeneous element A. For a subset E B( ), we | | ⊆ H accordingly define the super-commutant by E♯ := A B( ): ( M E)[A, M] =0 . { ∈ H ∀ ∈ τ } For each homogeneous M B( ), the operator D (A) := [M, A] is a superderiva- ∈ H M τ tion of the Z -graded associative algebra B( ) in the sense that 2 H M A D (AB)= D (A)B + ( 1)| || |AD (B). (31) M M − M It follows in particular that, if E is spanned by homogeneous elements, then E♯ is a von v Neumann algebra adapted to the 2-grading of B( ). Let Zv = ( 1)| |v ( v 0, 1 ) Hv − | |∈{ } denote the parity operator on and Zv = ( i)| |v (also known as the Klein twist H − Z = 1+iZ ) which satisfies Z2 = Z. For A and M odd we then have 1+i1 e 1 1 e [Z± AeZ∓ ,M]= iZ A, M = iZ[A, M] . ± { } − τ This leads to e e ♯ 1 1 E = ZE′Z− = Z− E′Z (32) for any graded subspace E B( ). e e e e ⊆ H As in [Fo83], we associated to every real linear subspace V a von Neumann ⊆ H subalgebra (V ) := −(V ) := b(V )′′ B( ( )). R R ⊆ F− H We list some properties of this assignment (cf. [Fo83, Prop. 2.5] for (iv) and (v)): Lemma 6.8. We have (i) ( )= B( ( )), resp., the representation of CAR( ) on ( ) is irreducible. R H F− H H F− H ANTIUNITARY REPRESENTATIONS 59

(ii) (V )= (V ). R R (iii) (V ) and (W ) super-commute if and only if V W (twisted duality). R R ⊥β (iv) The vacuum Ω is cyclic for (V ) if and only if V + iV is dense in . R H (v) The vacuum Ω is separating for (V ) if and only if V iV = 0 . R ∩ { } (vi) Ω cs( (V )) if and only if V is standard. ∈ R Proof. (i) is well-known ([BR96, Prop. 5.2.2(3)]). (ii) follows from the fact that b: B( ( )) is continuous. H → F− H (iii) follows immediately from (30). (iv) We explain how this can be derived from [BJL02, Prop. 3.4], where a different setting is used: Consider a conjugation Γ on a complex Hilbert space and a corre- K sponding basis projection P , i.e., ΓP Γ= 1 P . For v Γ we then have the orthogonal − ∈ K decomposition v = P v +(1 P )v, where both summands are exchanged by Γ, hence have − the same length. Therefore the map Φ: Γ P , Φ(v)= √2P v K → K is an isometry between the real Hilbert space Γ and the complex Hilbert space := P . K H K The antilinear map

a: CAR( ), a(f) := c∗(P Γf)+ c(Pf) K → H then satisfies a(Γf)= a(f)∗ for f ∈ K Γ and a is the unique antilinear extension of the map a Γ = b P : CAR( ). |K ◦ K → H For any Γ-invariant subspace , we therefore have V ⊆ K Γ Γ a( )= b(P )C = b(Φ( ))C (33) V V V and thus, for the real subspace V := Φ( Γ)= P ( Γ) , V V ⊆ H Γ Γ a( )′′ = a( )′′ = b(Φ( ))′′ = (V )′′. (34) V V V R Γ As V + iV = P ( C ) = P ( ), [BJL02, Prop. 3.4] implies that P ( ) is dense in P ( ) if V V V H and only if Ω is (V )-cyclic, and (iv) follows. R (v) In view of (ii), we may assume that V is closed. Let 0 = w W := V iV . To see 6 ∈ ∩ that Ω is not separating for (V ), it suffices to show that, for the one-dimensional Hilbert R space 0 := Cw, the vector Ω is not separating for (Cw) = B( (Cw)). This follows H R F− 2 from the irreducibility of the representation of CAR(Cw) = M2(C) on (Cw) = C ∼ F− ∼ which has no separating vector (see (i)). Suppose, conversely, that W = 0 . As W = (V ⊥ + iV ⊥)⊥, the subspace V ⊥ + iV ⊥ { } is dense in . By (iii), Ω is cyclic for (V ⊥) which anticommutes with (V ). Therefore H R R Ω is separating for (V ). R The following theorem is the fermionic version of the duality result in Theorem 6.4(iii) ([BJL02, Thm. 7.1], [Fo83, Thm. 2.4(v)]). Theorem 6.9 (Fermionic Duality Theorem).

β ♯ 1 (V ⊥ )= (V ) = A B( ( )): ( v V )[A, b(v)]τ =0 = Z− (V )′Z R R { ∈ F− H ∀ ∈ } R for every real linear subspace V . e e ⊆ H 60 K.-H. NEEB AND G. OLAFSSON´

To match our notation with Foit’s in [Fo83], we note that Foit’s operator 1 πi/4 1 1 V := (1 iZ) satisfies V = e− Z− , so that Z− AZ = V AV ∗ for every opera- √2 − tor A on ( ). F− H e e e 6.3. From antiunitary representations to local nets. For a closed real subspace V of the Hilbert space , we write ±(V ) B( ( )) for the associated von Neumann H R ⊆ F± H algebras on the bosonic and fermionic Fock space. Proposition 6.10. For a closed real subspace V , the vacuum Ω is cyclic and ⊆ H separating for the von Neumann algebras ±(V ) B( ( )) if and only if V is a R ⊆ F± H standard subspace of . The corresponding modular objects (∆±,J ±) on ( ) are H V V F± H obtained by second quantization from the modular objects (∆V ,JV ) associated to V , in the sense that + ∆± =Γ (∆V ), J =Γ+(JV ) and J − = ZΓ (iJV ). (35) V ± V V − Proof. The first assertion follows from Lemmas 6.2 and 6.8. Fore the identification of the modular objects, we refer to [FG89, Thm. 1.4] (see also [EO73]) in the bosonic case and to [Fo83, Prop. 2.8] for the fermionic case (see also [BJL02, Cor. 5.4], [Lle09, Thm. 4.13]). Remark 6.11. (a) The twists arising in Theorem 6.9 and Proposition 6.10 arise from the fact that the fermionic situation has to take the 2-grading on ( ) into account. In F− H♯ particular Theorem 6.9 takes its most natural form (V ⊥β )= (V ) if the commutant R R is defined in terms of the super bracket. (b) If is a Z -graded von Neumann algebra on the Z -graded Hilbert space = M 2 2 H 0 1 and Ω 0 is a cyclic separating vector, then the theory of Lie superalgebras H ⊕ H ∈ H ♯ suggests to consider the antilinear involution (x +x ) := x∗ ix∗ instead of the operator 0 1 0 − 1 adjoint. Then the corresponding unitary Lie superalgebra is ♯ u( )= x : x = x = x = x + x : x∗ = x , x∗ = ix . M { ∈M − } { 0 1 ∈M 0 − 0 1 − 1} Accordingly, modular theory can be based on the unbounded antilinear operator defined by S(MΩ) := M ♯Ω= ZS(MΩ) for M . The polar decomposition S = J∆1/2 results ∈M in the pair (J, ∆) of modular objects, where ∆ is unchanged, but J = ZJ. This leads to e e e e the relation e 1 ♯ e e J J = Z ′Z− = , M M M which is a super version of J Je = e ′. e e M M We also obtain with (35)

2 J − = Z Γ (iJV )= ZΓ (iJV )=Γ ( iJV ). V − − − − To obtain a situation wheref e the modular objects on −( ) are simply given by second F H quantization, one may consider the von Neumann algebras −(V ) := −(ζV ) for ζ := R R eπi/4 instead. The standard subspace V := ζV satisfies ∆ e = ∆ and J e = iJ , so that V e V V V the modular conjugation corresponding to −(V ) is e R J − =Γ ( iJeζV )=Γ (JV ). ζV − − − Remark 6.12. Let (U, ) beg an antiunitary representation of (G, G ) on and H 1 H γ : R× G be a homomorphism with γ( 1) G , so that it specifies a standard → − 6∈ 1 ANTIUNITARY REPRESENTATIONS 61 subspace V (Proposition 5.6). Consider the antiunitary representation γ ⊆ H Γ : AU( ) AU( ( )) ± H → F± H of the antiunitary group of on the corresponding Fock spaces. Then Γ U is an an- H ± ◦ tiunitary representation of (G, G1) on ( ), so that we also obtain standard subspaces + F± H Vγ± ( ). The pair ( (Vγ ), Ω) then satisfies ⊆F± H R V + = +(V ) Ω, γ R γ h and in the fermionic case the pair ( −(V ), Ω) leads to the correct modular operator R γ ∆− , but to the modular conjugation ZΓ (iJVγ ). Vγ − e 7. Perspectives. For a detailed exposition of the results mentioned below, we refer to the forthcoming paper [NO17].´

7.1. The Virasoro group. In Diff(S1) we consider the involution on S1 = T C, given ∼ ⊆ by r(z) = z. We consider the group G := Diff(S1) = Diff(S1) ⋊ 1, r . One can show ∼ 0 { } that all projective unitary positive energy representations of Diff(S1)0 extend naturally to projective antiunitary representations of G. To obtain antiunitary representations, one has to replace G by a central extension Vir ⋊ 1, r , where Vir is the simply connected { } Virasoro group. Another closely related “infinite dimensional” group that occurs in the context of mod- ular localization is the free product PSL (R) R PSL (R) of two copies of PSL (R) 2 ∗Aff( )0 2 2 over the connected affine group ([GLW98]). 7.2. Euclidean Jordan algebras. Minkowski spaces are particular examples of simple euclidean Jordan algebras, namely those of rank 2 (cf. [FK94]). Many of the geometric structures of Minkowski spaces and their conformal completions are also available for general simple euclidean Jordan algebras, where the role of the lightcone V+ is played by the open cone of invertible squares. There also exists a natural causal compactification V which carries a causal structure. The corresponding conformal group G := Conf(V ) has an index 2 subgroup G preserving the causal structure on V ; other group elements b 1 reverse it. In V , the set c := gV : g Conf(V ) specializes for Minkowski spaces W { + ∈ } b to the set of conformal wedge domains, which include in particular the light cone and b double cones (cf. Example 5.14). Moreover, the homomorphism

γ : R× GL(V ), γ(t)v = tv V+ → R is naturally specified because γV+ ( +×) is central in the identity component of the stabi- lizer GV+ . Therefore any antiunitary positive energy representation of G yields a net of standard subspaces indexed by c. In [NO17]´ we obtain a classification of these repre- W sentations along the lines of 2.2. The subsemigroups S := g G: gV V G § V+ { ∈ + ⊆ +} ⊆ also leads to a natural generalization of Borchers pairs in this context. 7.3. Hermitian groups. The conformal group Conf(V ) of a euclidean Jordan algebra can be identified with the group AAut(TV+ ) of holomorphic and antiholomorphic auto- morphisms of the corresponding tube domain TV+ = V+ + iV . This suggests that some of the crucial structure relevant for antiunitary representations can still be obtained for 62 K.-H. NEEB AND G. OLAFSSON´ the groups G := AAut( ) of all holomorphic and antiholomorphic automorphisms of a D bounded symmetric domain . The irreducible antiunitary positive energy representa- D tions can also be parametrized in a natural way by writing G = G ⋊ id, σ , where σ 1 { } is an antiholomorphic involution of ([NO17]).´ Here there are many homomorphisms D γ : (R×, R×) (G, G ) with γ( 1) = σ, but one cannot expect γ(R×) to be central + → 1 − + in Gσ, which can be achieved for tube type domains (coming from euclidean Jordan algebras). 7.4. Analytic extension. We have seen that, for antiunitary representations of Aff(R), the positive energy condition appears quite naturally from the order structure on the set of standard subspaces. If (U, ) is an antiunitary representation of G containing copies H of Aff(R) coming from half-sided modular inclusions, it follows that the closed convex cone C := x g: idU(x) 0 U { ∈ − ≥ } is non-trivial. This further leads to an analytic extension of the representation to the domain G exp(iCU ) (see [Ne00] for details on this process). On the other hand, antiunitary representations of R× correspond to modular objects (∆,J) and the orbit maps of elements v V extend to the strip z C: 0 Im z π ∈ { ∈ ≤ ≤ } (Remark 3.3). Composing families of homomorphisms γ : R× G with an antiunitary → representation, we therefore expect analytic continuation of U to natural complex do- mains containing G in their boundary. It would be very interesting to combine these two types of analytic continuations in a uniform manner, in the same spirit as the KMS condition is a generalization of the ground state condition (corresponding to positive energy) ([BR96]). One may further expect that this leads to “euclidean realizations” of antiunitary representations of G by unitary representations of a Cartan dual group in the sense of the theory of reflection positivity developed in in [NO14,´ NO16];´ see also [Sch06] for relations with modular theory. Maybe it can even be combined with the analytic extensions to the crown of a Riemannian symmetric space ([KS05]). 7.5. Geometric standard subspaces. In QFT, the algebras (V ) are supposed to R correspond to regions in some spacetime M. Therefore one looks for standard subspaces V ( ) that are naturally associated to a domain in some spacetime, such as Hardy spaces O (Example 3.16) or the standard subspaces K( ) constructed in [FG89] for free fields. O From the perspective of antiunitary group representations, a natural class of representa- tions of a pair (G, G ) are those realized in spaces C−∞(M) of distributions on 1 HD ⊆ a manifold M on which G acts. Here is the Hilbert space completion of the space 1 HD Cc∞(M) of test functions with respect to the scalar product given by a positive definite distribution D on M M via × ξ, η D = ξ(x)η(y) dD(x, y) h i ZM M × (cf. [NO14]).´ We associate to each open subset M a subset ( ) generated by O ⊆ HD O the space C∞( ) of test functions supported in . In this context it is an interesting c O O problem to find natural antiunitary extensions of the representation of G1 to G such that ANTIUNITARY REPRESENTATIONS 63 some of the corresponding standard subspaces (Proposition 3.2) have natural geometric descriptions. In this context the detailed analysis of KMS conditions for unitary repre- sentations of R in [NO16]´ should be a crucial tool because one typically expects standard subspaces to be described in terms of analytic continuations of distributions on some domain M to a complex manifold containing , which links this problem to 7.4 (cf. O⊆ O § [NO17b]).´ As one also wants the modular unitaries to act geometrically on the manifold M, the case G1 = R acting by translations on R considered in [NO16]´ is of key importance.

Conversely, one may also consider Hilbert spaces of holomorphic functions on a H complex manifold M on which G acts in such a way that G1 acts by holomorphic maps and G G by antiholomorphic ones. Then any γ Hom((R×, R×), (G, G )) leads to a \ 1 ∈ + 1 standard subspace of . Many natural examples of this type arise from 7.2 and 7.3. In H §§ particular, the representation of AU( ) on ( ) is of this type if we identify ( ) with H F+ H F+ H the Hilbert space of holomorphic functions on with the reproducing kernel K(ξ, η)= ξ,η H eh i (cf. [Ne00]). 7.6. Dual pairs in the Heisenberg group. Let be a complex Hilbert space and H V be a real linear subspace. We consider the corresponding subgroup Heis(V ) := ⊆ H T V Heis( ) ( 6.1). The centralizer of this subgroup in Heis( ) coincides with × ⊆ H § H Heis(V ′). If V is closed, we thus obtain a dual pair (Heis(V ), Heis(V ′)) of subgroups in Heis( ) in the sense that both subgroups are their mutual centralizers. H Remark 7.1. For a closed linear subspace V , we have Herm( )= Heis(V ) Heis(V ) ⊆ H H ′ if and only if V + V ′ is dense in , which is equivalent to (V + V ′)′ = V V ′ = 0 H ∩ { } (cf. Lemma 3.7). If this is the case, then (V ) B( ( )) is a factor by Theorem 6.4. R ⊆ F+ H Accordingly, the restriction of the irreducible Fock representation (U, ( )) of Heis( ) F+ H H to Heis(V ) is a factor representation and the same holds for Heis(V ′). We thus obtain many interesting types of factor representations of Heisenberg groups of the type Heis(V ) simply by restricting an irreducible representation of Heis( ). In [vD71] this approach is H used to realize quasi-free representations of Heis(V ) in a natural way.

Remark 7.2. (a) Suppose that G is a group which is the product G = G1G2 of two subgroups G and G such that G = Z (G ) and G = Z (G ). Then G G = Z(G) 1 2 1 G 2 2 G 1 1 ∩ 2 and every unitary representation (U, ) of G restricts to factor representations of the H subgroups Gj . (b) A typical example arises from a von Neumann algebra B( ) in symmetric M⊆ H form, i.e., there exists a conjugation J with J J = ′ (Definition 4.4). Then G := M M U( ) U( ′) is a product of two subgroups G := U( ) and G := U( ′) satisfying M M 1 M 2 M this condition. The representation of G on is multiplicity free because G′ = ′ H M∩M is the center of , hence abelian. It is irreducible if and only if is a factor, and then M M the representations of the subgroups G1 and G2 are factor representations. Note that the representation of G extends to an antiunitary representation of G ⋊ 1, j , where { } j(g)= JgJ. Remark 7.3. Similar structures also arise for infinite dimensional Lie groups such as Diff(S1), (doubly extended) loop groups and oscillator groups because modular objects 64 K.-H. NEEB AND G. OLAFSSON´ provide information on restrictions of irreducible representations to factorial representa- tions of subgroups (cf. [Wa98] for loop groups). So one should also try to develop the theory of modular localization for antiunitary representations of infinite dimensional Lie groups. 7.7. A representation theoretic perspective on modular localization. The anal- ysis of ordered families of von Neumann algebras with a common cyclic separating vector carried out by Borchers in [Bo97] should also have a natural counterpart in the context of standard subspaces, in the spirit of the translation mechanism described in Subsec- tion 4.2. It would be interesting to see if the corresponding results can be formulated en- tirely in group theoretic terms, concerning multiplicative one-parameter groups of some pair (G, G ) (cf. Proposition 5.6). As we have seen in 3.4 and 3.5, this works perfectly 1 §§ well for half-sided modular inclusions and modular intersections, The same could be said about Wiesbrock’s program, concerning the generation of Haag–Kastler nets from finite configurations of von Neumann algebras with common cyclic separating vectors ([Wi93c, Wi97b, Wi98, KW01]).

A. Appendices. In this short appendix we collect some general lemmas used in the main text. A.1. A lemma on von Neumann algebras. Lemma A.1. Let be a von Neumann algebra, α: a real-linear weakly M ⊆ H M→M continuous automorphism and U U( ) be a unitary element. Then the following as- ∈ M sertions hold: (a) If α is complex linear and α(U)= U, then there exists a V U( ) with α(V )= V ∈ M and V 2 = U. 1 (b) If α is complex linear and α(U)= U − with ker(U + 1)= 0 , then there exists a 1 2 { } V U( ) with α(V )= V − and V = U. ∈ M 1 1 (c) If α is antilinear and α(U)= U − , then there exists a V U( ) with α(V )= V − ∈ M and V 2 = U. (d) If α is antilinear and α(U) = U with ker(U + 1) = 0 , then there exists a V { } ∈ U( ) with α(V )= V and V 2 = U. M Proof. (a) Let P be the spectral measure of U on the circle T C with U = z dP (z). ⊆ T As P is uniquely determined by U and α is complex linear, we obtain α(P (E)) =R P (E) for every measurable subset E T. For every measurable function S : T T with S(z)2 = z ⊆ → for z T, we obtain by V := S(z) dP (z) a square root of U fixed by α. ∈ T (b) Now our assumptions impliesR that P ( 1 ) = 0, so that we find a unique spectral {− } π iθ 1 measure P on the open interval ( π, π) with U = π e dP (θ). From α(U) = U − we − − derive α(P (E)) = P ( E) for every measurable subsetR E ( π, π). Therefore V := π e − e⊆ − eiθ/2 dP (θ) is a square root of U satisfying α(V )= V 1. π e e − R− (c) We consider the von Neumann algebra generated by U. Then is abelian N ⊆M N and α-invariant. Therefore β(N) := α(N ∗) defines a complex linear automorphism of . N It satisfies β(U)= U, so that the existence of V follows from (a). (d) We argue as in (c), but now (b) applies. ANTIUNITARY REPRESENTATIONS 65

Example A.2. For = C, α = idC and U = 1 there exists no unitary V U( ) 2 M 1 − ∈ M with V = U and V = α(V ) = V − . This shows that the extra assumption in (b) is really needed. A.2. Two lemmas on groups. Lemma A.3. Let G be a group and G♯ := (G G) ⋊ 1, τ , where τ(g,h) = (h,g) is the × { 1} flip automorphism. We consider the subset D := (g,g− ): g G . Then the following { ∈ } assertions hold:

(a) grp(D) = (G′ e )D, where G′ G is the commutator subgroup. ×{ } ⊆ (b) The conjugacy class C of the involution (e,e,τ) G♯ coincides with D τ (e,e,τ) ∈ ×{ } and grp(C )= (G′ e )D 1, τ . (e,τ) ×{ } ×{ } Proof. (a) For g,h G, the relation  ∈ 1 1 1 1 1 1 1 (g,g− )(h,h− ) = (gh,g− h− ) = (ghg− h− ,e)(hg, (hg)− ) implies that G′ e grp(D). Conversely, it is easy to see that (G′ e )D is a subgroup ×{ }⊆ ×{ } of G G. × 1 1 1 (b) The first assertion follows from (g, h, 1)(e,e,τ)(g, h, 1)− = (gh− ,hg− , τ). The second assertion follows from (a).

Lemma A.4. Let G1 G be a subgroup of index two, r G G1, and ϕ: G1 H be ⊆ 2 2 ∈ \ 1 →1 a group homomorphism. If h H satisfies h = ϕ(r ) and hϕ(g)h− = ϕ(rgr− ) for ∈ g G , then ϕ(gr) := ϕ(g)h and ϕ(g) := ϕ(g) for g G defines an extension of ϕ to a ∈ 1 ∈ 1 homomorphism ϕ: G H. b → b Proof. First we observe that ϕ(gu) = ϕ(g)ϕ(u) obviously holds for g G and u G. b ∈ 1 ∈ For g G1 we further have ∈ b b b 1 1 1 ϕ(rg)= ϕ(rgr− r)= ϕ(rgr− )h = hϕ(g)h− h = hϕ(g)= ϕ(r)ϕ(g). Finally, we note that, for g G , b b ∈ 1 b b 1 2 1 2 1 2 ϕ(rgr)= ϕ(rgr− r )= ϕ(rgr− )ϕ(r )= hϕ(g)h− ϕ(r )= hϕ(g)h = ϕ(r)ϕ(gr).

Thisb implies thatb ϕ is a group homomorphism. b b A.3. Symmetricb spaces. Definition A.5. (a) Let M be a set and µ: M M M, (x, y) x y =: s (y) × → 7→ · x be a map with the following properties: (S1) x x = x for all x M, i.e., s (x)= x. · ∈ x (S2) x (x y)= y for all x, y M, i.e., s2 = id . · · ∈ x M (S3) s (y z)= s (y) s (z) for all x, y M, i.e., s Aut(M,µ). x · x · x ∈ x ∈ Then we call (M,µ) a reflection space. (b) If M be a smooth manifold and µ: M M M is a smooth map turning (M,µ) × → into a reflection space, then it is called a symmetric space (in the sense of Loos) if, in addition, each x is an isolated fixed point of sx ([Lo69]). 66 K.-H. NEEB AND G. OLAFSSON´

Example A.6. (a) Any group G is a reflection space with respect to the product 1 g h := s (h) := gh− g. • g The subset Inv(G) of involutions in G is a reflection subspace on which the product takes 1 the form sg(h) := ghg = ghg− . More generally, the subset G := g G: g2 Z(G) 2 { ∈ ∈ } is a reflection subspace of G because g2,h2 Z(G) implies ∈ 1 2 1 2 1 2 2 2 (gh− g) = gh− g h− g = g h− g Z(G). ∈ This calculation even shows that the square map G Z(G),g g2 is a homomorphism 2 → 7→ of reflection spaces. (b) Suppose that G is a Lie group and τ Aut(G) an involution. For any open ∈ subgroup H Gτ = Fix(τ), we obtain on the coset space M = G/H the structure of a ⊆ symmetric space by 1 xH yH := xτ(x)− τ(y)H. • Acknowledgments. K.-H. Neeb acknowledges supported by DFG-grant NE 413/9-1. The research of G. Olafsson´ was supported by NSF grant DMS-1101337. Both authors thank Arthur Jaffe for supporting a visit at Harvard University, where some of this work was done.

References

[Ar63] Araki, H., A lattice of von Neumann algebras associated with the quantum theory of a free Bose field, J. Math. Phys. 4 (1963), 1343–1362 [Ar64] —, von Neumann algebras of local observables for free scalar field, J. Mathematical Phys. 5 (1964), 1–13 [Ar99] —, “Mathematical Theory of Quantum Fields,” Int. Series of Monographs on Physics, Oxford Univ. Press, Oxford, 1999 [AW63] Araki, H., and E. J. Woods, Representations of the canonical commutation relations describing a nonrelativistic infinite free Bose gas, J. Math. Phys. 4 (1963), 637–662 [AW68] —, A classification of factors, Publ. RIMS, Kyoto Univ. Ser. A 3 (1968), 51–130 [AZ05] Araki, H., and L. Zsid´o, Extension of the structure theorem of Borchers and its application to half-sided modular inclusions, Rev. Math. Phys. 17:5 (2005), 491–543 [Ba64] Bargmann, V., Note on Wigner’s Theorem on Symmetry Operations, Journal of Mathematical Physics 5:7 (1964), 862–868 [BJL02] Baumg¨artel, H., Jurke, M., and F. Lledo, Twisted duality of the CAR-algebra, J. Math. Physics 43:8 (2002), 4158–4179 [BCL10] Bertozzini, P., R. Conti, and W. Lewkeeratiyutkul, Modular theory, Non- commutative Geometry and Quantum Gravity, in “Symmetry, Integrability and Ge- ometry: Methods and Applications,” SIGMA 6 (2010), 067; 47pp [Be96] Bertram, W., Un th´eor`eme de Liouville pour les alg`ebres de Jordan, Bull. Soc. Math. France 124:2 (1996), 299–327 [BW76] Bisognano, J. J., and E. H. Wichmann, On the duality condition for quantum fields, J. Math. Phys. 17 (1976), 303–321 ANTIUNITARY REPRESENTATIONS 67

[Bla06] Blackadar, B., “Operator Algebras,” Encyclopaedia of Mathematical Sciences Vol. 122, Springer-Verlag, Berlin, 2006. [Bo68] Borchers, H.-J., On the converse of the Reeh–Schlieder theorem, Comm. Math. Phys. 10 (1968), 269–273 [Bo92] —, The CPT-Theorem in two-dimensional theories of local observables, Comm. Math. Phys. 143 (1992), 315–332 [Bo95] —, On the use of modular groups in quantum field theory, Ann. Ins. H. Poincar´e 63:4 (1995), 331–382 [Bo96] —, Half-sided modular inclusion and the construction of the Poincar´egroup, Com- mun. Math. Phys. 179 (1996), 703–723 [Bo97] —, On the lattice of subalgebras associated with the principle of half-sided modular inclusion, Lett. Math. Phys. 40:4 (1997), 371–390 [Bo98] —, Half-sided translations and the type of von Neumann algebras, Lett. Math. Phys. 44 (1998), 283–290 [Bo00] —, On revolutionizing quantum field theory with Tomita’s modular theory, J. Math. Phys. 41 (2000), 3604–3673; extended version with complete proofs available under ftp://ftp.theorie.physik.uni-goettingen.de/pub/papers/lqp/99/04/esi773.ps.gz [BR87] Bratteli, O., and D. W. Robinson, “Operator Algebras and Quantum Statistical Mechanics I,” 2nd ed., Texts and Monographs in Physics, Springer-Verlag, 1987 [BR96] —, “Operator Algebras and Quantum Statistical Mechanics II,” 2nd ed., Texts and Monographs in Physics, Springer-Verlag, 1996 [BtD85] Br¨ocker, T., and T. tom Dieck, “Representations of Compact Lie Groups,” Springer, Graduate Texts in Mathematics 98, 1985 [BGL93] Brunetti, R., Guido, D., and R. Longo, Modular structure and duality in conformal quantum field theory, Comm. Math. Phys. 156 (1993), 210–219 [BGL02] —, Modular localization and Wigner particles, Rev. Math. Phys. 14 (2002), 759–785 [Bu78] Buchholz, D., On the structure of local quantum fields with non-trivial interaction, in “Proceedings of the Int. Conf. on Operator Algebras, Ideals and their Application in Theoretical Physics,” Baumg¨artel, Lassner, Pietsch, Uhlmann (eds.), 146–153; Teubner Verlagsgesellschaft, Leipzig [BDFS00] Buchholz, D., O. Dreyer, M. Florig, and S. J. Summers, Geometric modular action and spacetime symmetry groups, Rev. Math. Phys. 12:4 (2000), 475–560 [BLS11] Buchholz, D., Lechner, G., and S. J. Summers, Warped convolutions, Rieffel defor- mations and the construction of quantum field theories, Comm. Math. Phys. 304:1 (2011), 95–123 [BS93] Buchholz, D., and S. J. Summers, An algebraic characterization of vacuum states in Minkowski space, Comm. Math. Phys. 155:3 (1993), 449–458 [Co73] Connes, A., Une classification des facteurs de type III, Annales scientifiques de lE.N.S.´ 4i´eme s´erie 6:2 (1973), 133–252 [Co74] —, Caract´erisation des espaces vectoriels ordonn´es sous-jacents aux alg`ebres de von Neumann, Annales de l’institut Fourier 24:4 (1974), 121–155 [CR94] Connes, A., and C. Rovelli, Von Neumann Algebra Automorphisms and Time- Thermodynamics Relation in General Covariant Quantum Theories, Class. Quant. Grav. 11 (1994), 2899–2918 [vD71] van Daele, Quasi-equivalence of quasi-free states on the Weyl algebra, Commun. Math. Phys. 21 (1971), 171–191 [EO73] Eckmann, J.-P., and K. Osterwalder, An application of Tomita’s theory of modular 68 K.-H. NEEB AND G. OLAFSSON´

Hilbert algebras: duality for free Bose Fields, J. Funct. Analysis 13:1 (1973), 1–12 [FK94] Faraut, J., and A. Koranyi, “Analysis on symmetric cones,” Oxford Math. Mono- graphs, Oxford University Press, 1994 [FG89] Figliolini, F., and D. Guido, The Tomita operator for the free scalar field, Ann. Inst. H. Poincar´e, Phys. Th´eor. 51 (1989), 419–435 [Fl98] Florig, M., On Borchers’ Theorem, Lett. Math. Phys. 46 (1998), 289–293 [Fo83] Foit, J., Abstract twisted duality for quantum free Fermi fields, Publ. Res. Inst. Math. Sci. 19:2 (1983), 729–741 [Gu11] Guido, D., Modular theory for the von Neumann algebra of local quantum physics, Contemp. Math. 534 (2011), 97-120 [GL95] Guido, D., and R. Longo, An algebraic spin and statistics theorem, Commun. Math. Physics 172 (1995), 517–533 [GLW98] Guido, D., Longo, R., and H.-W. Wiesbrock, Extensions of conformal nets and superselection structures, Commun. Math. Physics 192 (1998), 217–244 [Ha96] Haag, R., “Local Quantum Physics. Fields, Particles, Algebras,” Second edition, Texts and Monographs in Physics, Springer-Verlag, Berlin, 1996 [HN12] Hilgert, J., and K.-H. Neeb, “Structure and Geometry of Lie Groups”, Springer, 2012 [HO96]´ Hilgert, J., and G. Olafsson.´ Causal Symmetric Spaces, Geometry and Harmonic Analysis. Perspectives in Mathematics 18, Academic Press, 1996 [HL82] Hislop, P., and R. Longo, Modular structure of the local observables associated with the free massless scalar field theory, Comm. Math. Phys. 84:1 (1982), 71–85 [JM17] J¨akel, C., and J. Mund, The Haag–Kastler axioms for the P (ϕ)2 model on the de Sitter space, arXiv:math-ph:1701.08231 [KW01] K¨ahler, R., and H.-W. Wiesbrock Modular theory and the reconstruction of four- dimensional quantum field theories, J. Math. Phys. 42:1 (2001), 74–86 [Ka97] Kaup, W., On real Cartan factors, manuscripta math. 92 (1997), 191–222 [Ke96] Keyl, M., Causal spaces, causal complements and their relations to quantum field theory, Rev. Math. Phys. 8:2 (1996), 229–270 [Ke98] —, How to describe the space-time structure with nets of C∗-algebras, Internat. J. of Theoretical Physics 37:1 (1998), 375–385 [Kl11] Klotz, M., Banach Symmetric Spaces, Preprint, arXiv:math.DG:0911.2089 [KS05] Kr¨otz, B., and R. J. Stanton, Holomorphic extension of representations: (II) geom- etry and harmonic analysis, Geom. Funct. Anal. 15 (2005), 190–245 [Le15] Lechner, G., Algebraic constructive quantum field theory: Integrable models and de- formation techniques, Preprint, arXiv:math.ph:1503.03822 [LL14] Lechner, G., and R. Longo, Localization in Nets of Standard Spaces, Comm. Math. Phys., to appear [Lle09] Lled´o, F., Modular theory by example, Preprint, arXiv:math.OA:0901.1004v1 [Lo82] Longo, R., Algebraic and modular structure of von Neumann algebras of physics, in “Operator algebras and applications, Part 2 (Kingston, Ont., 1980),” pp. 551–566, Proc. Sympos. Pure Math., 38, Amer. Math. Soc., Providence, R.I., 1982 [Lo08] —, Real Hilbert subspaces, modular theory, SL(2, R) and CFT in “Von Neumann Algebras in Sibiu”, 33-91, Theta Ser. Adv. Math. 10, Theta, Bucharest, [LW11] Longo, R., and E. Witten, An algebraic construction of boundary quantum field theory, Comm. Math. Phys. 303:1 (2011), 213–232 [Lo69] Loos, O., “Symmetric spaces I: General theory,” W. A. Benjamin, Inc., New York, ANTIUNITARY REPRESENTATIONS 69

Amsterdam, 1969 [Mo17] Morinelli, V., The Bisognano–Wichmann property on nets of standard subspaces, some sufficient conditions, arXiv:math-ph:1703.06831 [Mu01] Mund, J., The Bisognano–Wichmann theorem for massive theories, Ann. Henri Poincar´e 2:5 (2001), 907–926 [MSY06] Mund, J., Schroer, B., and J. Yngvason, String-localized quantum fields and modular localization, Comm. Math. Phys. 268:3 (2006), 621–672 The Bisognano–Wichmann theorem for massive theories, Ann. Henri Poincar´e 2:5 (2001), 907–926 [Ne00] Neeb, K.-H., “Holomorphy and Convexity in Lie Theory,” Expositions in Mathe- matics 28, de Gruyter Verlag, Berlin, 2000 [NO14]´ Neeb, K.-H., G. Olafsson,´ Reflection positivity and conformal symmetry, J. Funct. Anal. 266 (2014), 2174–2224 [NO15] —, Reflection positive one-parameter groups and dilations, Complex Analysis and Operator Theory 9:3 (2015), 653–721 [NO16]´ —, KMS conditions, standard subspaces and reflection positivity on the circle group, arXiv:math-ph:1611.00080; submitted [NO17]´ —, Antiunitary representations of hermitian groups, in preparation [NO17b]´ —, Reflection positivity on spheres and hyperboloids, in preparation [OM16] Oppio, M., and V. Moretti, Quantum theory in real Hilbert space: How the com- plex Hilbert space structure emerges from Poincar´esymmetry, Preprint arXiv:math- ph:1611.09029v1 28 Nov 2016 [Ra17] Raasakka, M., Spacetime-free approach to quantum theory and effective spacetime structure, arXiv:1605.03942v2 [gr-qc] 24 Jan 2017 [RS73] Reed, S., and B. Simon, “Methods of Modern Mathematical Physics I: Functional Analysis,” Academic Press, New York, 1973 [RS61] Reeh, H., and S. Schlieder, Bemerkungen zur Unit¨ar¨aquivalenz von Lorentzinvari- anten Feldern, Nuovo Cimento 22 (1961), 1051–1068 [Sa97] Salehi, H., Problems of dynamics in generally covariant quantum field theory, Inter- nat. J. Theoret. Phys. 36:1 (1997), 143–155 [Sch97] Schroer, B., Wigner representation theory of the Poincar´egroup, localization, statis- tics and the S-matrix, Nuclear Phys. B 499-3 (1997), 519–546 [Sch06] —, Positivity and Integrability (Mathematical Physics at the FU-Berlin), Preprint, arXiv:hep-th/0603118 [Sch09] —, Localization and the interface between quantum mechanics, quantum field theory and quantum gravity II. The search of the interface between QFT and QG, Stud. Hist. Philos. Sci. B Stud. Hist. Philos. Modern Phys. 41:4 (2010), 293–308 [SW00] Schroer, B., and H.-W. Wiesbrock, Modular theory and geometry, Rev. Math. Phys. 12:1 (2000), 139–158 [So10] Solveen, C., The Bisognano-Wichmann Theorem and nets on R4, manuscript from http://math.mit.edu/ eep/CFTworkshop [StVa02] Stalder, Y., and A. Valette, Le lemme de Schur pour les repr´esentations orthogonales, Expo. Math. 20 (2002), 279–285 [Su05] Summers, S., Tomita–Takesaki modular theory, arXiv:0511.034v1 [math-ph] 9 Nov 2005 [SW03] Summers, S., and R. K. White, On deriving space-time from quantum observables and states, Commun. Math. Phys. 237 (2003), 203–220 70 K.-H. NEEB AND G. OLAFSSON´

[Ta12] Tanimoto, Y., Construction of wedge local nets of observables through Longo–Witten endomorphisms, Comm. Math. Phys. 314:2 (2012), 443–469 [Tr97] Trebels, S., Uber¨ die geometrische Wirkung modularer Automorphismen, PhD The- sis, Univ. G¨ottingen, 1997 [Va85] Varadarajan, V. S., “Geometry of Quantum Theory,” Springer Verlag, 1985 [Wa98] Wassermann, A., Operator algebras and conformal field theory. III. Fusion of posi- tive energy representations of LSU(N) using bounded operators, Invent. Math. 133 (1998), 467–538 [Wi92] Wiesbrock, H.-W., A comment on a recent work of Borchers, Lett. Math. Phys. 25 (1992), 157–159 [Wi93] —, Half-sided modular inclusions of von Neumann algebras, Commun. Math. Phys. 157 (1993), 83–92 [Wi93b] —, Conformal quantum field theory and half-sided modular inclusions of von Neu- mann algebras, Comm. Math. Phys. 158:3 (1993), 537–543 [Wi93c] —, Symmetries and half-sided modular inclusions of von Neumann algebras, Lett. Math. Phys. 28:2 (1993), 107–114 [Wi97] —, Half-sided modular inclusions of von Neumann algebras, Erratum, Commun. Math. Phys. 184 (1997), 683–685 [Wi97b] —, Symmetries and modular intersections of von Neumann algebras, Lett. Math. Phys. 39:2 (1997), 203–212 [Wi98] —, Modular intersections of von Neumann algebras in quantum field theory, Comm. Math. Phys. 193:2 (1998), 269–285 [Wig59] Wigner, E., “Group Theory and its Applications to the Quantum Mechanics of Atomic Spectra,” Academic Presse, 1959