On Infravacua and the Localisation of Sectors 9 Walter Kunhardt 9 Institut fu¨r Theoretische Physik der Universit¨at G¨ottingen 9 Bunsenstraße 9, 37073 G¨ottingen, Germany 1 e-mail: [email protected] n a J June 5, 1998 9 2 Abstract v 3 A certain class of superselection sectors of the free massless scalar 0 fieldin3spacedimensionsisconsidered. Itisshownthatthesesectors, 0 6 whichcannotbe localisedwith respectto the vacuum, acquirea much 0 better localisation, namely in spacelike cones, when viewed in front of 8 suitable “infravacuum” backgrounds. These background states coin- 9 cide,essentially,withaclassofstatesintroducedbyKraus,Polleyand / h Reents as models for clouds of infrared radiation. p - h 1 Introduction t a m In the analysis of superselection sectors, the localisability properties of : v charges are crucial for defining notions as charge composition and conju- i gation, statistics or a (global) gauge group. In a classical work [1], Do- X plicher, Haag and Roberts carried through such a programme for charges r a which fulfill what is now called the DHR criterion, i.e., which are compactly localised. Now this criterion is very restrictive, and Buchholz and Freden- hagen [2] established that sectors of theories in 3 space dimensions without massless particles in general only comply with the so-called BF criterion, i.e., they are localised in spacelike cones. Still, these authors could extend the analysis of [1] to charges with such a weaker localisation behaviour. The situation is more difficult for theories with massless particles. Typ- ically, these theories possess sectors whose localisation is too poor for the DHR framework to beapplicable. Motivated by what is expected to happen in QED, it has been proposed by Buchholz in [3] to improve the localisation by viewing the charges in front of some suitable background field instead of the vacuum. In QED, such background fields should correspond to clouds 1 of infrared radiation. An appropriate mathematical description of such in- frared clouds has been introduced by Kraus, Polley and Reents in [4]. Here, we want to verify this mechanism in a simpler model, namely in the theory of the free massless scalar field in 3 space dimensions [5]. More precisely, we will consider a certain class of (non-Lorentz invariant) sectors described by automorphisms of the observable algebra and analyze their localisation properties in terms of the BF criterion. In particular, we will show that the sectors under consideration do satisfy this criterion with respect to a KPR-like background but do not satisfy it with respect to the vacuum. (Calling the background fields “KPR-like” should indicate that they are very similar, yet not identical, to those of [4].) Astotheconsistencyofsuchanapproach,itshouldbekeptinmindthat, in a theory whose charges are compactly localised, the superselection struc- ture can be described without any difference with respect to the vacuum as well as with respect to so-called infravacua, the latter being generalisations of the KPR-like background states considered here. As has been shown in [6], when viewed in front of such an infravacuum, the charges remain compactly localised and have the same fusion structure and statistics as in front of the vacuum. Moreover, positivity of the energy in a sector does not depend on the background chosen, nor do the masses of massive particles possibly contained in such a theory. The above-mentioned class of sectors of the free massless field has been studied recently by Buchholz et al. [5] with the purposeof modeling charges of electromagnetic type. In the following, we will stick very closely to the notations introduced there, but we should emphasize that our point of view is slightly different from that adopted in [5]: Buchholz et al. achieved a better localisation of the sectors by restricting them to a (non-Lorentz in- variant) subnetA ⊂ Aof theobservablenet. Thesectors theneven became 0 localised in the DHR sense, which permitted them to carry through a DHR- like analysis, even though the net A does not fulfill Haag duality. Here, in 0 contrast, the subnet A will play no roˆle, and the localisation obtained will 0 be in a weaker sense. We end this Introduction by recalling the definition of the model under consideration. The observable algebra of the free massless scalar field is defined in its vacuum representation. More precisely, let K:=L2(R3,d3k) be the Hilbert space of momentum space wave functions, ω(~k):=|~k| the one-particle energy and U(t,~x) = ei(ω(~k)t−~k~x) the usual representation of the spacetime translations. The vacuum Hilbert space of our model will be the bosonic Fock space H over K; the induced unitary representation of 2 the spacetime translations will still be denoted by U(t,~x) without any risk of confusion. For any v ∈ K, W(v) ∈ B(H) will denote the corresponding Weyl operator. The normalisation is chosen such that the Weyl relations read W(u)W(v) = e−2iImhu,viW(u+v). For any real linear subspace L ⊂ K, W(L) denotes the C∗-subalgebra of B(H) generated by the operators W(f), f ∈ L. The net of observables now is given as O 7−→ A(O):=W(L(O))′′, where O 7−→ L(O) is the isotonous, local and covariant net of symplectic subspacesinK(indexedbythesetofopendoubleconesinMinkowskispace) defined as follows: If O:=({0} ×O)′′ is the causal completion of an open ball O ⊂ R3 at time t = 0, one has L(O):=ω−12D\R(O)+iω+21D\R(O), where DR(O) is the set of all real-valued smooth functions with support in O andˆdenotes the Fourier transform. For other double cones O, the space L(O) is defined by translation covariance and additivity. The symplectic form σ on L:= L(O) ⊂ K reads O S σ(f ,f ):= −Imhf ,f i, 1 2 1 2 and locality for the net L(·) just means σ(L(O ),L(O )) = 0 whenever O 1 2 1 and O are spacelike to each other. As usual, we also associate symplectic 2 subspacesofL(resp.C∗-subalgebrasofB(H))tounboundedregionsinR1+3 by additivity (resp. additivity and norm closure) and simply denote by A the quasilocal algebra A(R1+3). Thechargesunderconsiderationaregivenintermsofnetautomorphisms γ ∈ AutA which are labeled uniquely by elements of the (additive) abelian group LΓ:=ω−12D\R(R3)+iω−32D\R(R3). Any element γ ∈ L gives rise to a linear form l :L −→ C, Γ γ l (f):= −Im d3k γ(~k)f(~k) γ Z and hence to an automorphism, again denoted by γ, of A by γ(W(f)):=eilγ(f)W(f). As explained in [5], γ is indeed a well-defined automorphism of the quasi- localalgebraAsince,byHuygens’principle,itturnsouttobelocallynormal; 3 as aconsequence, itcan beextended byweak continuity fromthelocal Weyl algebras W(L(O)) to the local von Neumann algebras A(O). There will be no risk of confusion in viewing the real vector space L as Γ anabelian subgroupof AutA. Inparticular, asumγ +γ inL corresponds 1 2 Γ to the composition γ ◦γ in AutA. Moreover, γ and γ define the same 1 2 1 2 sector of A, i.e., they are unitarily equivalent in B(H), iff γ −γ ∈L ∩K. 1 2 Γ In this case, the Weyl operator W(γ −γ ) is well defined and implements 1 2 the unitary equivalence γ ∼= γ on A. 1 2 Any γ ∈ LΓ can bewritten uniquely in the formγ = ω−12σˆ+iω−23ρˆwith functions σ,ρ ∈ DR(R3). Since σˆ and ρˆ are analytic, it is obvious that γ is square integrable, i.e., γ ∈ K, iff ρˆ(0) = 0. As a consequence, the sectors considered are labeled by a single real parameter q :=ρˆ(0) = d3x ρ(~x) γ Z which is interpreted as the charge of the sector [γ]. In particular, this shows that the sectors are transportable; as a matter of fact, they even have posi- tive energy [5]. 2 Bad localisation of the sectors in front of the vacuum It has been shown in [5] that the automorphisms γ ∈ L do not satisfy the Γ DHRlocalisation criterion. Here,wewanttostrengthenthisresultandshow with closely related methods that they do not even satisfy the BF criterion, that is, that they are not localisable in spacelike cones. To this end, it is sufficient to prove the following Proposition 2.1 Let C ⊂ R3 be an open convex cone having 0 as its apex and denote with C:=({0}×C)′′ itscausal completion. Then, forany γ ∈ L , Γ γ|A(C) ∼= id|A(C) iff qγ = 0. The “if” part of this proposition is trivial, and before proving the “only if” part, we recall some facts about the dilation covariance of the model. The dilation group R acts unitarily on K and leaves the space L in- >0 variant. More precisely, f ∈ L(O) is mapped onto f ∈ L(λO), where λ fλ(~k):=λ23f(λ~k). Writing f = ω−12hˆ +iω+12gˆ, it is verified by a straight- forward computation that this entails for the linear form l , γ ∈ L : γ Γ d3k 1 l (f ) = ρˆ(~k/λ)hˆ(~k)− d3kσˆ(~k/λ)gˆ(~k). γ λ Z ω2 λ Z 4 In the limit λ → ∞, ~k 7−→ ρˆ(~k/λ)hˆ(~k) converges to ρˆ(0)hˆ in the space of test functions, and since 2π2 is the Fourier transform of 1 in the sense of r ω2 distributions, one obtains d3x lim l (f ) = q κ with κ :=2π2 h(~x). γ λ γ f f λ→∞ Z |~x| This allows us to prove the following Lemma 2.2 Let f ∈ L(O′), where O ⊂ R1+3 is a neighbourhood of 0. For any γ ∈ L , one then has Γ w-limγ(W(fλ)) = eiqγκf e−14kfk2 1. λ→∞ Proof: Since the dilations act geometrically, it follows by locality from the special form of thelocalisation region of f thatlim σ(f ,f′)= 0 for any λ→∞ λ f′ ∈ L. Hence, (W(f )) is a central sequence of unitaries in W(L) whose λ λ>0 set of weak limits is, by the irreducibility of the vacuum representation, a (nonempty) subsetof C1. On the other hand, unitarity of the dilations per- mits us to evaluate this limit in the vacuum state: ω0(W(fλ))−λ−→−∞→e−14kfk2. ButthismeansthatW(fλ)hase−14kfk21asitsuniqueweaklimitforλ → ∞, establishing thus the assertion for γ = 0. For arbitrary γ ∈ L , it now fol- Γ lows easily in view of the discussion in the preceding paragraph. Physically, the sequence (W(f )) is interpreted as a measurement λ λ→∞ of the asymptotic behaviour (in the spatial directions determined by the smearing function h) of the “Coulomb potential” of the “charge density” ρ. In QED, one expects that operators measuring the asymptotic electric flux distribution play a similar roˆle, cf. [3]. In the present case, the leading 1/r behaviour of the Coulomb potential is isotropic in all sectors [γ]. This fact, reflected by the factorizing of liml (f ) as seen above, is relevant in the γ λ Proof of Prop. 2.1: Let γ ∈ L with q 6= 0. Choose a nonvanishing, Γ γ nonnegativetestfunctionh ∈DR(C). Lettingf:=ω−12hˆ,thisimpliesκf 6= 0 and f ∈ L(C∩O′) for some neighbourhood O ⊂ R1+3 of 0. Since eiqγκf 6= 1 can always be achieved by a mere rescaling of h, Lemma 2.2 shows that the weak limits (as λ → ∞) of W(f ) and γ(W(f )) are different scalar λ λ multiples of the unit operator. But since W(f ) ∈ A(C) for all λ > 0, this λ implies γ|A(C) 6∼= id|A(C). 5 3 Infravacuum background states In this section we introduce a class of background states in front of which the automorphisms γ will be shown (in Section 4) to have better localisa- tion properties. Apart from two modifications necessitated by the present model, these background states are of the same type as those introduced by Kraus, Polley and Reents [4]as a modelfor infrared clouds in QEDor, more generally, in any theory containing massless particles. 3.1 Preliminaries on quasifree states First, we recall that aquasifree state on Ais a locally normalstate ω which T is, on the Weyl operators W(f)∈ A, f ∈ L of the form ωT(W(f))= e−14kTfk2. Here, T :D −→ K is a real linear, symplectic (i.e., fulfilling ImhTv,Twi = T Imhv,wi,v,w ∈ D ) operator defined on a dense, real linear subspace D T T which contains L. In the case at hand, we will have in addition TL = K, which entails that ω is a pure state. Its GNS representation π acts T T irreducibly on the vacuum Hilbert space H as π (W(f)) = W(Tf), f ∈ L. T Next, wedescribethereal linear operator T in terms of apair of complex linear operators T ,T defined on complex linear subspaces D , j =1,2. 1 2 Tj Lemma 3.1 Let Γ : K −→ K be an antiunitary involution. Then, the formulae T :=T 1+Γ +T 1−Γ 2 2 1 2 D :={v ∈K | 1+Γv ∈D ,1−Γv ∈ D } T 2 T2 2 T1 establish a bijection between • densely defined, Γ-invariant1 R-linear operators T :D −→ K and T • densely defined, Γ-invariant C-linear operators T : D −→ K, j Tj j = 1,2. Moreover, T is symplectic iff hT u ,T u i= hu ,u i for all u ∈ D . 1 1 2 2 1 2 j Tj 1 Here, T :D −→K being Γ-invariant means ΓD =D and [Γ,T]=0 on D . T T T T 6 Since all assertions can be checked by simple calculations, we omit the formalproofof thisLemmaandmerely pointoutthattheconverse formulae expressing T and T in terms of T read 1 2 D ={v ∈K | 1+ΓCv ⊂ D }, T = T 1+Γ −iT 1+Γ i T2 2 T 2 2 2 D ={v ∈K | 1−ΓCv ⊂ D }, T = T 1−Γ +iT 1−Γ i. T1 2 T 1 2 2 Remark: The involution Γ induces the notion of real and imaginary parts of vectors v ∈ K: Rev = 1+Γv , Imv = 1−Γv. Then, T acts on the real and 2 2i 2 T on the imaginary parts: 1 ReTv = T Rev, ImTv =T Imv, v ∈ D . 2 1 T From now on, we will fix Γ to be pointwise complex conjugation in position space. In terms of momentum space wave functions v ∈ K, this means (Γv)(~k) := v(−~k). [For the sake of completeness, we point out that Kraus et al. used pointwise conjugation in momentum space for defining their background states in [4]. In their case as well as in ours, the choice of the involution Γ is dictated by the set of sectors under consideration.] 3.2 Quasifree states with positive energy Before describingin detail the operatorsT ,T , weintroduce some notation: 1 2 for any ǫ > 0, let P : K −→ K be the projector onto the subspace P K = ǫ ǫ {v ∈ K|v(~k) =0 if |~k| <ǫ} and denote by D := P K 0 ǫ ǫ[>0 the dense subspace of functions vanishing in some neighbourhood of ~k = 0. Note that [P ,Γ] = 0 and ΓD = D . The subspace D will serve as a ǫ 0 0 0 provisional domain for T and T . 1 2 Now we follow [4] and choose i→∞ • a sequence (ǫi)i∈N in R>0 satisfying ǫi+1 < ǫi and ǫi −−−→ 0. This sequence induces a decomposition of momentum space into con- centric spherical shells. The projections onto the associated spectral subspaces of K will be denoted by P :=P − P . For notational i ǫi+1 ǫi convenience, we also put P :=P . 0 ǫ1 7 • asequence(Qi)i∈N oforthogonalprojectionsinKwithfiniterankrkQi satisfying Q Γ = ΓQ , Q P = Q . i i i i i i→∞ • a sequence (bi)i∈N in ]0,1[ satisfying bi −−−→ 0 and i bǫ2i rkQi < ∞. i If, e.g., the ǫ decrease exponentially and rkQ Pis polynomially i i bounded, this can be satisfied by b ∝ i−α, α > 0. i With these data, define C-linear operators T ,T on the subspace D by 1 2 0 n n 1 T :=1+s-lim (b −1)Q , T :=1+s-lim ( −1)Q . 1 i i 2 i n→∞Xi=1 n→∞Xi=1 bi Since, on every v ∈ D , the number of terms which contribute on the right- 0 hand side is finite, these operators are well defined and map D into itself. 0 Moreover, the relations T P = ((1−Q )+b Q )P , T P = ((1−Q )+ 1Q )P 1 i i i i i 2 i i bi i i show that the subspace P K decomposes into a subspace (1−Q )P K where i i i both T and T act trivially and an orthogonal subspace Q P K = Q K 1 2 i i i wherethey act as multiplications with the scalars b and 1, respectively. As i bi a consequence, T and T are inverses of each other. Because of lim b = 1 2 i→∞ i 0, T is bounded (kT k = 1), whereas T is not. Also, it is clear that 1 1 2 T and T are Γ-invariant and symmetric. In particular, it follows that 1 2 hT u ,T u i = hu ,T T u i = hu ,u i for any u ,u ∈ D . We are thus in 1 1 2 2 1 1 2 2 1 2 1 2 0 the situation of Lemma 3.1 and obtain an unbounded symplectic operator T :D −→ K, T = T 1+Γ +T 1−Γ. 0 2 2 1 2 In the next step, T has to be extended to a larger domain D ⊃ L. T To this end, we analyze its singular behaviour for |~k| → 0 by comparing it with powers of (a regularized version ω of ) the one-particle Hamiltonian r ω. Setting ω on (1−P )K, ω :=ω(1−P )+ǫ P = 0 r 0 1 0 (cid:26) ǫ11 on P0K 1/2 and noting that ω D ⊂ D , we obtain: r 0 0 1/2 Lemma 3.2 T ω is bounded. 2 r 8 Proof: Making use of kω P k = ǫ for i ∈ N, one obtains for v ∈ D r i i 0 1 2 1 2 1 1 (T −1)ω2v = (1 −1)Q ω2v = (1 −1)2 ω2v,Q ω2v (cid:13)(cid:13) 2 r (cid:13)(cid:13) (cid:13)(cid:13)Xi bi i r (cid:13)(cid:13) Xi bi D r i r E (cid:13) (cid:13) (cid:13) 1 (cid:13) 1 ≤ (1 −1)2rkQ ω2v,P ω2v ≤ (1 −1)2rkQ ǫ kvk2. Xi bi iD r i r E Xi bi i i From the conditions imposed on the b , it follows that (1 −1)2rkQ ǫ is i i bi i i finite. Thus (T −1)ω1/2 is bounded, hence also T ω1/P2. 2 r 2 r We now can extend T by continuity to all of K =: D and T by the 1 T1 2 formula 1 −1 1 T v:=T ω2 ω 2v, v ∈ ω2K 2 2 r r r 1 1 to the dense subspace ω2K =:D . (Strictly speaking, the symbol T ω2 on r T2 2 r the right-hand side stands for the continuous extension to K of the operator consideredinthepreviousLemma.) NotethatT andT stillareΓ-invariant. 1 2 We collect the relevant properties in the following Lemma: Lemma 3.3 1 1. D :={v ∈ K|1+Γv ∈ ω2K} is a real linear dense subspace of K. T 2 r 2. T = T 1+Γ +T 1−Γ is well defined on D . 2 2 1 2 T 3. T :D −→ K is a symplectic operator. T 4. L ⊂ D and TL is dense in K. T Proof: Part 1 is obvious, since D ⊂ D ; part 2 has been shown in the 0 T previous paragraph. For 3, we have to show that hT u ,T u i = hu ,u i 1 1 2 2 1 2 remains true for all u ∈ D and u ∈ D . First, assume u ∈ D . Since 1 T1 2 T2 1 0 1/2 (n) D isdenseinKandinvariantunderω , thereexistsasequenceu ∈ D , 0 r 2 0 n ∈ N such that ωr−21u2 = limωr−21u(2n), implying u2 = limu(2n). Using the 1/2 boundedness of T ω , we can compute 2 r hT u ,T u i = hT u ,T ω12 ω−21u i = hT u ,T ω12 lim ω−21u(n)i 1 1 2 2 1 1 2 r r 2 1 1 2 r r 2 n→∞ = lim hT u ,T ω12 ω−21u(n)i= lim hu ,u(n)i = hu ,u i. 1 1 2 r r 2 1 2 1 2 n→∞ n→∞ 9 SinceT is bounded,therestriction onu cannowbedroppedbycontinuity, 1 1 thus yielding the assertion. Finally, L ⊂ D is obvious, and the remaining T part of 4 is equivalent, in terms of T and T , to 1 2 1+2ΓTL = T21+2ΓL = T2ω−12DR is dense in 1+2ΓK 1−2iΓTL = T11−2iΓL = T1ω21DcR is dense in 1−2iΓK. By C-linearity, this in turn is equivalenct to T2ω−21DC = T2ωr12ωr−21ω−21DC 1 and T1ω2DC both being dense in K. But this is imcplied by the fact thcat, 1/2 on the one hand, both operators T ω and T are boundedand have dense c 2 r 1 images (since they are invertible on the dense, invariant subspace D ) and 0 that, on the other hand, the subspaces ωr−21ω−21DC and ω12DC are dense in K (by the spectral calculus of ω). c c With the above preparations, we can define a state ω : A −→ C and T analyze its main properties. Proposition 3.4 The quasifree state ω , defined on W(L) by T ωT(W(f))= e−41kTfk2, f ∈ L extends to a unique locally normal state ω over the quasilocal algebra A. T This state is pure and has positive energy. Proof: The difficult part of this proof is to obtain local normality of ω on the net O 7−→ W(L(O)) of Weyl algebras. To this end, recall that T T is (on L) the strong limit of symplectic operators T such that T − 1 n n have finite rank. As a consequence, the associated quasifree states ω are Tn vector states in the vacuum representation and converge weakly to ω on T W(L). Now since the Fredenhagen-Hertel compactness condition C [7, 8] ♯ is known to be fulfilled in the present model, we can conclude that ω is T locally normal if the sequence (ωTn)n∈N is bounded with respect to some exponential energy norm k·k , β > 0 defined by kωk2 :=ω(e2βH). But this β β follows from ǫi rkQ < ∞, as F. Hars has shown in [9], adapting ideas i b2 i i from[4]. (AltPhoughourinvolutionΓdiffersfromthatof[4,9],thearguments leading to this conclusion are still valid.) Hence, ω is locally normal on T W(L) and thus extends uniquely to a locally normal state on A. Since it is a weak limit of states in the vacuum representation with positive energy, the arguments of Buchholz and Doplicher [10] can be applied to show that ω has positive energy, too. Finally, the relation TL = K, established in T Lemma3.3, implies that ω is pure,as has beennoted atthevery beginning T of this section. 10