1 Einstein Drive, Princeton, NJ 08540 USA
Algebras, Regions, and Observers
Abstract
In ordinary quantum field theory, one can define the algebra of observables in a given region in spacetime, but in the presence of gravity, it is expected that this notion ceases to be well-defined. A substitute that appears to make sense in the presence of gravity and that also is more operationally meaningful is to consider the algebra of observables along the timelike worldline of an observer. It is known that such an algebra can be defined in quantum field theory, and the timelike tube theorem of quantum field theory suggests that such an algebra is a good substitute for what in the absence of gravity is the algebra of a region. The static patch in de Sitter space is a concrete example in which it is useful to think in these terms and to explicitly incorporate an observer in the description.
1 Introduction
In ordinary quantum mechanics, we do not usually incorporate the observer as part of the system. That is fine for many purposes, but in the presence of gravity, one has to take into account the fact that the observer gravitates. In some contexts, this is unimportant, because the observer’s gravity is negligible. For example, in an asymptotically flat spacetime, an observer more or less at rest in the asymptotic region is likely to have negligible gravitational influence on whatever is in the interior of the spacetime. On the other hand, in a closed universe, where the gravitational flux due to the observer has “nowhere to go,” one might expect that it may be essential to include the observer in the description.
We will discuss an example of this – based on the paper CLPW . But we begin with more general considerations. In ordinary quantum field theory, one can attach an algebra of observables to a rather general open set in spacetime. In the presence of gravity, there are potentially two problems with this notion. First, when spacetime fluctuates, we may generically have difficulty saying just what we mean by the spacetime region . Second, it is not clear what is the logic of discussing an algebra unless there is someone who can make the observations corresponding to elements of that algebra. In the absence of gravity, we do not usually worry about this question; we just assume, in effect, that some observer external to the system has the relevant capability. But in the presence of gravity, anticipating that sometimes it will be necessary to include the observer in the description, we should take seriously the idea that it is only well-motivated to discuss an algebra of observables if this is the algebra of observables accessible to someone.
But what algebra is accessible to an observer? We will discuss this question in the context of ordinary quantum field theory, without gravity, but hoping to learn some lessons that are useful when gravity is included. We make use of two classic but not so well known results about ordinary quantum field theory. The first result Borch , described in section 2, enables one to define an algebra of observables along the timelike worldline of an observer. The second result, described in section 3, is the “timelike tube theorem” Borch2 ; Araki ; Stroh ; SW , which among other things says that, in the absence of gravity, in a real analytic spacetime, the algebra of observables along the observer’s worldline is the same as (roughly) the algebra of observables in the region causally accessible to the observer. The first result is more elementary, and more general (since it does not require a hypothesis of real analyticity), but in a real analytic spacetime, it can potentially be viewed as a corollary of the second.
The two results indicate that instead of discussing the algebra of observables in a spacetime region, we could discuss the algebra generated by the quantum fields along the observer’s worldline. In a theory of gravity, the algebra of observables along the worldline would appear to be both better defined and operationally more meaningful than the algebra of a region.
In section 4, following CLPW , we consider the “static patch” in de Sitter space as an example in which it is necessary to explicitly include an observer in order to define a sensible algebra of observables. The algebra defined after taking the observer explicitly into account turns out to be a von Neumann algebra of Type II1. This gives an abstract explanation of why “empty de Sitter space” is a state of maximum entropy, and in what sense the density matrix of empty de Sitter space is maximally mixed.
Presumably, in a full theory of the world, an observer cannot be added from outside but must emerge as part of the theory. In that context, what it means to include an observer in the description is that one considers a “code subspace” of states in which the observer is present, and one defines operators that are well-defined on the code subspace, though they would not be well-defined on all states of the theory. For considerations of the present article, however, however, it is not necessary to have such a full theory in hand.

2 The Algebra Accessible to an Observer
In ordinary quantum field theory in a spacetime , we can arbitrarily specify any open set and define an algebra of operators in (fig. 1). In the presence of gravity, since spacetime fluctuates, it does not make sense to talk about the region unless we have an invariant way to identify it. For example, in an asymptotically flat spacetime, in the presence of a black hole, we could talk about the region outside the black hole horizon. That is invariantly defined and presumably makes sense at least perturbatively even when spacetime fluctuates. We could introduce an observer who is more or less at rest near infinity and describe the region outside the horizon as the region visible to this observer. However, as already noted, in an asymptotically flat universe we do not expect that it is essential to incorporate the observer in the description.
If we assume the existence of an observer, we can invariantly identify various regions in spacetime. For example, if the observer carries a clock, then as in fig. 2, we can discuss the region that is visible to the observer prior to a given time, or, alternatively, the region that is causally accessible to the observer in a stated time interval (meaning that the observer can both see and influence this region during the interval in question).

But what can an observer actually measure? Here we will assume a very simple model in which the observer is described by a timelike worldline, and what the observer can measure are simply the quantum fields along this worldline.111See Unruh for a classic discussion in this framework, with a somewhat different motivation. This seems like a rather minimal model of what an observer is, and one could well worry that it is too crude. A realistic observer would presumably also carry measuring equipment, and a recording device, and would have access to operators that act on all that. But it turns out that in ordinary quantum field theory without gravity, the rather crude model in which an observer is just characterized by a worldline and the observables are the quantum fields along the worldline is sufficient, for many purposes.
This model raises two immediate questions:
(1) Can well-defined operators be defined by smearing a quantum field along a timelike worldline?
(2) Given a “yes” answer to the first question, what is the algebra generated by these operators?
Let us elaborate a bit on the first question. We are accustomed in quantum field theory to talking about “local operators” , but a local operator is not really a Hilbert space operator, since acting on a Hilbert space state it takes us out of Hilbert space. In the case of the vacuum state in Minkowski space, this is clear from the fact that or equivalently
(1) |
due to a short distance singularity. Since the leading short distance singularity is universal, it is also true that for any state in any spacetime . The problem has nothing to do with the state , and purely reflects the fact the product is not well-defined, since in fact a more general product is singular for . This singularity is governed by the operator product expansion (OPE).
If we could measure , the answer would be one of its eigenvalues, but since maps us out of Hilbert space, it does not have eigenvectors or eigenvalues, and we cannot measure it. What are actually measureable are suitable smeared versions of . Which ones? Suppose we are going to smear a real scalar field over a set to get a smeared “operator” where is a complex-valued smooth function with support in . If is actually going to make sense as an operator, the smearing has to be such that the OPE singularity is integrable, when smeared in this fashion.
For example, spatial smearing will only succeed for an operator of rather low dimension. In spacetime dimensions, with space coordinates and a time coordinate , spatial smearing at, say, , produces an expression , where is a smooth function of the spatial coordinates that we can assume to have compact support. Does make sense as an operator? In computing the product , we run into the integral
(2) |
For illustrative purposes, let us consider first the case of a conformal field theory, though with minor modifications, the following remarks apply much more widely. If is a conformal field of dimension , then the leading singularity in the operator product for is proportional to , where is the dimension of the operator . The condition for this singularity to be integrable when inserted in (2) is . If is a free scalar field, then , and the condition is satisfied.222In dimension , a free scalar has and the condition is satisfied by any normal ordered polynomial . This was important in early work on constructive field theory Jaffe . But what about, say, QCD, in the real world with ? QCD is asymptotically free, so short distance singularities have the behavior just discussed up to logarithms, which are inessential except in the borderline case . In QCD, the smallest value of for any gauge-invariant operator is 3, corresponding to a quark bilinear such as . So the condition is never satisfied and QCD is an example of a theory in which no true operator can be produced by smearing of a “local operator” in space.
Smearing in Euclidean space is only slightly better. If we try to define a smeared operator (where now the integration is over all coordinates of Euclidean spacetime), we will run into the operator product singularity This is integrable if and only if , a weaker condition but one that again cannot be satisfied in QCD, for example.
How then do we get true operators by smearing of “local operators”? The secret is smearing in real time. Though smearing in space is only effective in favorable cases, smearing in real time turns a “local operator” of any dimension into a true operator. This rather old result Borch was originally proved directly on the basis of the Wightman axioms of quantum field theory for the case that the timelike curve is a timelike geodesic in Minkowski space.
To understand the result, we begin again with the case of a conformal field theory. Suppose that at , we smear a “local operator” by a compactly supported function that depends only on . Thus we define . In evaluating , we now run into the operator product . If has dimension , the leading OPE singularity is , so we have to consider the integral
(3) |
There are also subleading OPE singularities; they have the same form with different exponents and can be treated just as we are about to describe. The integral is obviously well-defined for , and we want to show that no divergence appears in the limit . For this, we write333If , the following formula has to be slightly modified with a logarithmic factor on the right hand side. The derivation otherwise proceeds in the same way.
(4) |
for any integer , with a constant . Inserting this in the integral (3) and integrating by parts times, we replace the original integral with
(5) |
For large enough , this is manifestly convergent for .
In a general quantum field theory, consider the operator product expansion
(6) |
The coefficient functions are holomorphic for . This follows from positivity of energy. Normally one can assume that the singularities of the are bounded444An example of a local operator that would not satisfy this condition is , where is a scalar field in spacetime dimension . by a power law, , for some . In the context of the Wightman axioms of quantum field theory, the correlation functions are usually assumed to be tempered distributions, which implies such a bound. Alternatively, if a theory is conformally invariant or asymptotically free in the ultraviolet, and the operators considered behave as fields of definite dimension in the ultraviolet (modulo logarithms in the asymptotically free case), this again implies such a bound. Given holomorphy in the lower half plane and a power law bound,555The precise mathematical statement is that if is a function holomorphic in the lower half plane, then a necessary and sufficient condition for the boundary values of along the real axis to define a distribution is a bound for some constants . For an elementary proof (and a bound on the distribution) see Proposition 4.2 in BF . See also Theorem 1.1 in Straube for a -dimensional generalization. one can imitate the previous derivation to show the finiteness of
(7) |
For this, one writes each as the derivative with respect to , for some , of a function that remains continuous (though not smooth) for , and then one integrates by parts as in the derivation of eqn. (5).
What happens if we consider an arbitrary timelike curve in Minkowski space, not necessarily a geodesic? Parametrize by the proper time and let be the signed proper distance666The proper distance in Minkowski space between points along labeled by and by is the proper time elapsed along a geodesic between the two points (not along the path ). To get the signed proper distance , we multiply the proper distance by if and by if . in Minkowski space between points on labeled by and by . The effect of replacing a timelike geodesic by an arbitrary timelike curve is to replace in the preceding formulas by . Since is a smooth function and this does not substantially affect the preceding analysis. The singularities at are of the same general form and are harmless.
What happens if we replace Minkowski space by a general spacetime ? Intuitively, one would not expect this to matter, since everything is determined by short distance behavior. In a curved spacetime, the operator product expansion becomes more complicated, with curvature dependent terms; see HW for an axiomatic discussion. But the extra terms have similar singularities to what we have already considered, so one would expect the same result. For a proof, under reasonable assumptions about correlation functions in a curved spacetime, see K . See also further discussion in Fewster , SW .
3 The Timelike Tube Theorem
3.1 The Timelike Envelope and the Timelike Tube Theorem
In section 2, we learned that one can define operators by smearing a quantum field along the timelike worldline of an observer. Therefore we can consider the algebra generated by such operators (or more precisely by bounded functions of such operators). But what are the algebras that we make this way? In the context of quantum field theory without gravity, this question is answered by the “timelike tube theorem.” This theorem was originally formulated for timelike geodesics in Minkowski space Borch2 ; Araki . It was generalized to free field theories in curved spacetime in Stroh . For a version of the theorem suitable for non-free theories in curved spacetime, see SW ; SW2 .

If is an open set in spacetime, its “timelike envelope” consists of all points that can be reached by deforming timelike curves in through a family of timelike curves, keeping the endpoints fixed (fig. 3). (Thus is contained in the intersection of the past and future of , but in general it is smaller. See section 3.3 for an example.) The timelike tube theorem asserts that the algebra of operators777For a more precise explanation of what algebra is intended here, see section 3.2. in is the same as the algebra of operators in the possibly much larger region . We will aim to give at least a hint of why this is true, after first explaining an implication, noted in Stroh .
Suppose that we are actually interested in a timelike curve , possibly one of finite extent with endpoints or possibly an infinite or semi-infinite curve. The timelike envelope consists of all points that can be reached by deforming through a family of timelike curves, keeping fixed near its ends. If has endpoints, they are omitted from to ensure that is an open set. We can thicken (minus its endpoints, if any) to an open set , in such a way that the timelike envelope does not depend on – it is the same as the timelike envelope of . See fig. 3. The timelike tube theorem says that the algebra of operators in region does not really depend on but only on . It coincides with , the algebra of operators in . So we can define an algebra for every (possibly bounded) timelike curve . That in itself is not a surprise, since we arrived at the same conclusion in a more direct way888The timelike tube theorem depends on real analyticity of spacetime, as discussed shortly, and the reasoning in section 2 does not, so the previous analysis was more general. in section 2. But it is nice to see the consistency with what we learned from more elementary arguments.
Strictly speaking, the algebra associated to a curve in section 2 consists of the quantum fields smeared along , while the timelike tube theorem gives us an algebra of operators that can be defined in an arbitrarily small neighborhood of . Presumably, these two algebras coincide (assuming that we include all possible local operators in the construction of section 2, including derivatives of operators). A proof might start with an axiomatic characterization of local operators at a point in terms of a limit of the operators in a small ball around that point. Were the two algebras to differ, one might argue that as the notion of an observer characterized by an infinitely thin worldline is an idealization, the physically more relevant algebra would be the one provided by the timelike tube theorem.
The interpretation that we want to give is that the algebra of operators supported on a curve is a good stand-in for the algebra of the spacetime region associated to . The algebra is more operationally meaningful than , since it is more directly what an observer can measure. And it is also better defined in the presence of gravity than would appear to be. Once one takes into account fluctuations in spacetime, it is hard to see how one would directly define , but as long as an observer is included in the description, the algebra of observables along the observer’s worldline seems to be a meaningful notion, even when the observer’s worldline fluctuates. So seems like a good substitute for the algebras associated to open sets that one considers in the absence of gravity.
To try to gain some idea of why the timelike tube theorem is true, let us consider the classical limit. Suppose we have a reasonable relativistic wave equation like the Klein-Gordon equation , where is the wave operator, or possibly a nonlinear modification of this equation. In fig. 4, we are given a solution in one region of a spacetime , and we want to predict the solution in a larger region. We consider two cases. In fig. 4(a), the solution is given in a “spacelike pancake” , and we want to extend it over the domain of dependence of . In fig. 4(b), the solution is given in a “timelike tube” and we want to extend the solution over the corresponding “timelike envelope” .
There is a basic asymmetry between the two cases.999 To be precise, this asymmetry holds in any spacetime dimension . In , spacetime has one dimension of space and one dimension of time, and there is a perfect symmetry between the two cases of fig. 4. In , there are more space dimensions than time dimensions, and there is no such symmetry. The counterexample illustrated in fig. 5 fails in because in , the solution with a delta function source along a worldline does not blow up along , but instead is discontinuous across . Starting with the solution on one side of , one can drop the discontinuity and smoothly continue the solution across . In , a solution with such a delta function source blows up along and there is no way to remove the singularity without changing the solution in the original region . The Holmgren uniqueness theorem of classical partial differential equations101010For an accessible exposition of this theorem, see chapter 5 of Smoller . asserts that the extension over the larger region is unique, if it exists, both in fig. 4(a) and in fig. 4(b). But existence is more special and only holds in fig. 4(a).


A simple counterexample (fig. 5) shows that an existence result cannot possibly hold in fig. 4(b). Let be a timelike curve that passes through the region but not through . Consider a solution of the equation with a delta function source along . If we consider Maxwell’s equations rather than the Klein-Gordon equation, then can be the worldline of a point charge. The field blows up along (in dimension ; see footnote 9). Starting with the solution in the region , there is no way to extend it over as a solution, since if one tries to do this, one encounters the blowup along . The equation is not obeyed along , because of the delta function source.
The existence and uniqueness result in the case of fig. 4(a) is the basis for much of physics. It says that the solution can be predicted from initial data – physics is causal. But by contrast the uniqueness result without a guarantee of existence in fig. 4(b) is not usually useful at the classical level because generically the extension over of a solution on does not exist and it is very hard to predict when it does.
Suppose, however, that we are doing quantum field theory and for simplicity consider a free field with the action
(8) |
In this case, we can view as an operator-valued solution of the Klein-Gordon equation . If we are studying this quantum field theory on , then the field does exist throughout and therefore existence of the extension from to is not an issue.
But what does uniqueness mean? In some sense, uniqueness means “the field for is uniquely determined by for .” As explained by Borchers and Araki in the early 1960’s, the quantum meaning of this statement is really “ for is contained in the algebra generated by for ” or equivalently the operator algebras of the two regions are the same:
This is the timelike tube theorem.
Classically, we could just as well add higher order terms to the action and consider a field that satisfies a nonlinear partial differential equation. Holmgren uniqueness applies equally well to such an equation. Quantum mechanically, although a free quantum field can be viewed as an operator-valued solution of a classical field equation, that is not the case for a non-free quantum field, because of issues involving renormalization. Accordingly, proofs of the timelike tube theorem in non-free theories require additional ingredients, beyond what is needed in free theories. An approach that works for non-free theories in curved spacetime was presented in SW (see SW2 for an informal account).
In this discussion, we have skipped over two important points. First, we have not been precise about what is the algebra that is governed by the timelike tube theorem. See section 3.2. Second, we have omitted to explain a key assumption in the classical Holmgren uniqueness theorem and the quantum timelike tube theorem. These theorems hold in real analytic spacetimes. Real analyticity is not needed for existence and uniqueness in the setting of fig. 4(a), but it is needed for uniqueness in the setting of111111Actually, it suffices for the spacetime to be, in some coordinate system, real analytic in the time direction Tataru . In free field theory, this is also true for the timelike tube theorem Stroh . fig. 4(b). Is this restriction important for the application of the timelike tube theorem to semiclassical gravity? In semiclassical gravity, one is usually studying the behavior of quantum fields in a background spacetime that in practice is normally real analytic. For instance, in section 4, we consider the static patch in de Sitter space as an example in which it is important to explicitly include an observer in order to define a sensible algebra of observables, and thus in which the timelike tube theorem provides important motivation. This is a typical example in which the starting point is a real analytic spacetime. After picking a semiclassical starting point, one then expands around it. In perturbation theory, the metric certainly fluctuates away from the real analytic starting point, but one expects that in perturbation theory, the gravitational field can be treated like any other quantum field in a real analytic background. Thus it appears that for typical applications of the timelike tube theorem to semiclassical gravity, the restriction of the theorem to real analytic spacetimes is not a problem.
3.2 The Additive Algebra
For a region in spacetime, what precisely is the algebra of operators to which the timelike tube theorem applies? The smallest reasonable candidate is the algebra generated by local operators in (or more precisely, by bounded functions of smeared local operators). If is contractible, this is the natural algebra of observables in in a generic quantum field theory. However, if is not contractible, then in general, as we discuss shortly, there can be additional operators in , such as Wilson operators defined on noncontractible loops in , that cannot be constructed from local operators in .
The algebra generated by local operators has been called the additive algebra casini1 ; casini2 . One can then reserve the name for the possibly larger algebra of all operators in . The timelike tube theorem is really a theorem about . Thus a more precise statement of the theorem than we have given so far is . This is clear from the proofs, which involve studying algebras of local operators. However, as we will discuss, while the distinction between and exists in quantum field theory in general, there are reasons to believe that this distinction is absent in any theory that emerges at long distances from a full model of quantum gravity.
The motivation for the phrase “additive algebra” is as follows. Let be a small open ball in . Then as is contractible, there is no distinction between and . Now cover by open balls , with ranging over some set . The additive algebra of is the same as the algebra generated by the , :
(9) |
In what situation will we have , where is the algebra of all possible operators in , not necessarily built from local operators? For a typical example, consider a theory with a gauge field , with curvature . Let be a closed loop in , and consider the Wilson operator . Is this operator contained in ? The answer will be “yes” if is the boundary of an oriented two-manifold , for then is a bounded function of a smeared local field in , namely . But even if is not a boundary in , the answer can still be “yes,” for a more subtle reason that was explained in Harlow . Consider a theory that has a charge 1 field . Then for an open path , say with endpoints , one can consider the gauge-invariant operator In the case that is a very short path, with very near , can be expanded in terms of local operators at and so is contained in the additive algebra of any open set containing . If is a closed loop, then can be “cut” in the sense that if one omits from a number of open intervals of size , with the portion retained then being a disjoint union of closed intervals , , then appears as the most singular contribution in the operator product for . Hence, in this situation is contained in the additive algebra for any region containing .
However, if the theory has no field of unit charge, and is not a boundary in , then is contained in but not in . This phenomenon is certainly not limited to gauge theory. In a theory with any gauge group , if some representation of is “missing,” in the sense that no local operator of the theory transforms in this representation, then for suitable regions one will have . A Wilson loop in the representation can be present in but not in . A similar role can be played by a -form gauge field with . In a theory that has such a field, one can consider the operator , where is a -cycle in . If is not a boundary in and the theory does not have a string or membrane that couples to (which would enable one to “cut” the operator in small pieces), then is contained in but not in . There is also an analog of this for magnetic charges (and their analogs for strings and membranes), with ’t Hooft operators instead of Wilson operators.
We can now easily give an example that shows that it is necessary to specify that the timelike tube theorem applies to , not to . Consider gauge theory without charged fields on a spacetime with product metric, where parametrizes time and is parametrized by an angular variable . In this spacetime, there is a nontrivial Wilson operator , where is a closed loop that wraps around the . Let and be two points at the same value of but different values of , and let be the timelike geodesic that runs from to at fixed . If is a small neighborhood of , there is no nontrivial Wilson operator in , so . But if is separated in time from by more than the circumference of the , then the region includes a circle that wraps all the way around the , so . Thus in this example, we have , as guaranteed by the timelike tube theorem, but .
Let us say that a quantum field theory which may have gauge fields or -form gauge fields for is “complete” if the electrically and magnetically charged particles, strings, and membranes coupling to those gauge fields are a maximal possible set consistent with all principles of low energy physics including Dirac quantization of electric and magnetic charge. Complete theories are precisely the ones with for all , so we might also call them “additive.” It is believed that a quantum field theory that emerges at low energies from a full-fledged (ultraviolet complete) theory of quantum gravity is always complete in this sense. This was originally suggested based on experience with string theory Polchinski . Further arguments were given in BS , and the fullest known argument is based on holographic duality HO .
The claim that a theory that emerges from a full quantum gravity theory will be additive or complete has another interpretation. A theory with “missing” charges (or strings or membranes) has a -form global symmetry in the language of GKSW (where depends on what is missing). So the statement that a full-fledged theory of quantum gravity is complete can be viewed as an extension to -form global symmetry with of the statement that there are no global symmetries in a full theory of quantum gravity. For discussion of this statement, see for example BS ; EW .
In the spirit of the present article, one can motivate as follows the idea that an ultraviolet-complete theory of quantum gravity should have no missing charges. We assume that a complete gravity theory can be approximated, in an appropriate class of states, by an ordinary quantum field theory weakly coupled to gravity. Suppose that this ordinary quantum field theory is not complete, for example because it has a gauge field but no state of charge 1. Then if a loop is not a boundary in spacetime, the Wilson operator is not measurable by any observer living in the spacetime and described by the theory. After all, given that is not a boundary, cannot be measured by measuring the curvature . To measure when is not a boundary requires measuring the interference between different histories of a charge 1 particle that differ by the homology class of the worldline of the particle. But the assumption that the theory has no state of charge 1 means that such an experiment is not possible using the resources available in the theory.
What, then, do we mean in claiming that (or its hermitian part) is an “observable”? In ordinary quantum mechanics, we consider the observer to be external to the system, and we make no particular assumption about what resources the observer can use in probing the system. From that point of view, the observer might be able to introduce a massive charged particle to probe the system and measure . The observer, after all, is described by a more complete theory of the universe that might have the necessary charged particle. An ultraviolet-complete theory of quantum gravity, however, is expected to describe all that there is in the universe that it describes, with no way to add anything from outside. In the context of such a theory, the observer and any apparatus used by the observer must already be described by the theory. So in an ultraviolet complete theory of quantum gravity with a “missing” charge, we would be in the awkward situation that the theory would enable us to define an “observable” that would be well-defined in an appropriate, long distance limit, but that no one, in principle, could measure.
3.3 What Else?
In addition to the timelike tube theorem, which asserts that , one has causality, which, denoting the domain of dependence of a set as , asserts that . If the difference between and is unimportant, these statements can be combined, relating to the algebra of a set that in general is larger than either or . This was noted by Araki Araki in one of the original papers on the timelike tube theorem. However, Araki actually conjectured a further generalization. In the context of quantum fields in Minkowski space, Araki wrote, “ … If is a snake-like line, connecting two mutually timelike points and but everywhere spacelike, and is a tube around , of a small diameter, then it is rather likely that any solution of the [massless scalar] wave equation vanishing in might always vanish in the double light cone spanned by , and . If such a conjecture turns out to be true, then we immediately have the corresponding theorem for .” In our terminology, the proposal is that if is the double cone or causal diamond with vertices and , then . This follows from arguments in Araki’s paper if it is true that any solution of the massless scalar wave equation vanishing in also vanishes in . This statement does not follow from Holmgren uniqueness in any obvious way and its validity remains unclear sixty years later.
In general, what extension of the timelike tube theorem is conceivable? For any set in a spacetime , one writes for the set of points in that are spacelike separated from . Then is called the causal completion of . For any open set , since commutes with , clearly any statement that for some open set implies that commutes with . If one is hoping to make a general statement that would apply to any quantum field theory, this really forces , since local operators outside of typically do not commute with . So the most optimistic statement that one might hope for would be .
For example, if is a timelike curve in Minkowski space, then , and therefore the timelike tube theorem does indeed give . In general, , but can be strictly bigger. Indeed, in Araki’s example, is strictly bigger than . His conjecture would be a special case of , in a situation in which that does not follow directly from the known statement of the timelike tube theorem.
A statement along the lines of would be very attractive, as it would combine ordinary causality with the timelike tube theorem. Extending from a spacelike pancake to its domain of dependence (fig. 4(a)) and from a timelike tube to its timelike envelope (fig. 4(b)) can be regarded as two different cases of extending from to . However, there are some easy counterexamples to .
One type of counterexample121212This example is relatively well known but the original reference is not clear. A version of the example of fig. 6 was discussed in Stroh . arises if is not connected. In Minkowski space in any even dimension , let be a small neighborhood of the origin. Let be any open set that contains points to the future of and also points to the past of , but none that can be reached from by a null geodesic. Then , so would predict that . In the theory of a massless free scalar field , however, this statement is false. By Huygen’s principle along with commutativity at spacelike separation, the commutator function in that theory is supported on the light cone (for even), so it vanishes for , . Thus in this situation, commutes with , rather than being contained in .

In two dimensions, a small variant of this gives a counterexample with connected. In fact, the setup is closely related to that suggested by Araki. Consider the spacetime , with a product metric, where parametrizes time and parametrizes space. Let the point be slightly to the future of , and let be a spacelike curve that starts at , spirals once around to the right, and ends at . Let be an open set that is a slight thickening of , and let be the causal diamond with vertices . Finally, let be an open set small enough that any right-moving null geodesic that intersects does not intersect (fig 6). Then . In two-dimensional quantum field theory, it is possible to have a local field with the property that vanishes unless the points and are connected by a right-moving null geodesic. For example, could be a right-moving conserved current (or a chiral component of the stress tensor) in a conformally invariant theory. Let be a smearing function supported in , and consider the operator . This operator commutes with , rather than being contained in , as one would expect based on .
In the context of the present article, we are not primarily interested in applying the timelike tube theorem or a hypothetical generalization to an arbitrary open set. We are primarily interested in observables along the timelike worldline of an observer. The timelike tube theorem tells us that it does not matter whether we consider a timelike curve or an open set that is a suitable slight thickening of it, so we will express the following in terms of the algebra associated to a curve. The example just discussed can be slightly modified to give a counterexample in that context. If is farther to the future of than was assumed so far, then the curve that wraps around the (in general any number of times) en route from to can be timelike, but almost null. Then the timelike envelope is a small neighborhood of , but its causal completion , which is the same as if is suitably chosen, wraps all the way around the . The statement is false in general just as before.
In this example, is the same as the causal diamond131313Here is the intersection of the past and future of , or equivalently . with vertices , so in particular . Thus, this example shows that in the statement of the timelike tube theorem, in general cannot be replaced with141414Similarly, in a simply-connected spacetime , with metric , with , . Embedding in at and choosing as before, again is much larger than if is sufficiently far to the future of . However, as in Araki’s example, it is not clear in this case whether it is always true that . Free field theory does not provide an immediate counterexample. .
So some simple generalizations of the timelike tube theorem are false in a general quantum field theory, and others, such as Araki’s conjecture, have unclear status. However, the following somewhat fanciful remarks come to mind. Let us go back to the example in which is not connected and is, for example, the union of a small ball around a point and a small ball around a point to its future. (Similar remarks apply in the other examples.) In this case, for even, massless free field theory provides a counterexample to , but the argument was limited to massless free field theory and would not generalize in an obvious way to any other theory. Consider then in a more generic theory an observer151515In this paragraph, unlike the rest of the article, we consider an observer probing the spacetime from outside, not an observer who propagates on a worldline in the spacetime. who has the capability to manipulate the quantum fields in an arbitrary fashion, but only in . A strategy this observer can follow is to inject into spacetime a probe in the region , directed on a trajectory that will carry it to , and equipped and programmed to make some chosen measurement and record the result. Then the observer retrieves the probe at and reads the answer. Does this construction along with the timelike tube theorem prove that ? Though we assume no particular limit on the technological capabilities of the observer, we assume that whatever happens in a part of spacetime to which the observer does not have direct access is governed by the theory that is being probed. Thus in particular the probe that the observer injects into spacetime in must be something that can be built in this theory. A complex probe cannot be built in free field theory, so there is no tension between existence of this protocol and our earlier observations about free field theory. However, in a sufficiently complex theory – possibly any theory that encompasses the Standard Model, for example – this protocol does indeed hint that . Actually, the observer can choose at will (assuming that the probe can be equipped with a rocket engine and programmed to travel on a pre-chosen trajectory), and more generally the observer could inject into the system several probes with pre-chosen trajectories from to , each designed and programmed to carry out a particular quantum operation, and then the observer can collect the probes in and process their (possibly quantum) output at will. All this suggests that in a theory that is sufficiently complex, possibly in the end in general.
3.4 Causal Wedge Reconstruction
There is actually a manifestation of the timelike tube theorem that has been much discussed in the literature. This is causal wedge reconstruction, or HKLL reconstruction, in the context of the AdS/CFT correspondence BDHM ; B ; Bal ; HKLL ; HKLL2 ; KLL ; HMPS ; Mor ; H .
In the AdS/CFT correspondence, one considers a conformal field theory on a globally hyperbolic spacetime of dimension . AdS/CFT duality says that an appropriate conformal field theory on is equivalent to a gravitational theory formulated on a spacetime that has for its conformal boundary, and that is globally hyperbolic in the asymptotically AdS sense. In fact, in AdS/CFT duality, one has to consider all possible ’s that have a given conformal boundary , but for our purposes here, we can assume that a particular is important. Causal wedge reconstruction applies in a semiclassical situation in which can be viewed, in leading order, as a definite spacetime in which quantum fields are propagating.

Let be a Cauchy hypersurface in , an open set in , and the domain of dependence of in . Then is its own timelike envelope in , so the timelike tube theorem, applied directly to the region in the CFT on , says nothing of interest. However, if is the conformal boundary of and , then it makes sense to consider the timelike envelope of in (fig. 7). In its simplest form, causal wedge reconstruction then asserts that . This can be viewed as a composite of two statements: (1) the basic AdS/CFT duality expressing local operators in as limits of local operators in , and (2) the timelike tube theorem applied to quantum fields in .
To explain this, we can proceed as follows. First, it is possible to thicken slightly to an open set such that . The timelike tube theorem says that does not depend on the precise choice of , and always equals . Since this is the case, we can take a limit in which becomes arbitrarily “thin.” The basic AdS/CFT relation between local operators on and local operators on can be interpreted as a statement
(10) |
where refers to a limit in which the bulk open set collapses down to the boundary open set . On the left hand side, is the algebra generated by bulk local operators in the bulk region , and on the right hand side, is the algebra generated by boundary local operators in the boundary region . Eqn. (10) is a way to express the fact that boundary local operators are the boundary limits of bulk local operators. Since the timelike tube theorem tells us that regardless of the choice of , the limit in eqn. (10) is trivial and we learn that
(11) |
In a globally hyperbolic spacetime that satisfies the null energy condition, is the causal wedge of , and eqn. (11) is the usual statement of causal wedge reconstruction. However, this requires some explanation,161616The following proof is not needed in the rest of the article. Facts used in the proof are explained, for example, in Wald ; GaoWald ; LightRays . On compactness of spaces of causal curves, see for instance sections 2 and 3.3 of LightRays ; on the fact that a causal curve that satisfies a promptness condition is a null geodesic without focal points, see sections 5.1 and 5.2 of that article; on the completeness of a null geodesic in an asymptotically AdS spacetime whose ends are on the conformal boundary of , see section 7.3; on the fact that a complete null geodesic along which the null energy condition is satisfied as a strict inequality must have focal points, see section 8.2. since the usual definition of the causal wedge is slightly different. The causal wedge of , which we will denote as , is usually defined to consist of all points that are contained in a causal curve between two points , say with to the future of . (Equivalently, is the intersection of the past and future of in .) Instead, contains all points that are contained in a causal curve between points such that can be deformed, through a family of causal curves with fixed endpoints, to a causal curve entirely in (and therefore in ). Thus to show that , we have to show that any causal curve that starts and ends at points can be deformed, through a family of causal curves with those fixed endpoints, to a causal curve in . To prove this, let be the space of causal curves in with initial and final endpoints . Pick a causal curve from to , and let be the space of causal curves in with initial endpoint and any final endpoint . A causal curve from to can be converted to a causal curve from to by gluing on the segment of . This operation maps continuously to ; on the other hand, . So a causal curve with endpoints can be deformed in to a curve entirely in if and only if it can be so deformed in . Hence, to prove that , it suffices to show that any causal curve with endpoints can be deformed in to a curve entirely in . This question is topological in nature: is in the connected component of that contains curves that lie entirely in ? So the answer is invariant under an infinitesimal perturbation of , and we can make such a perturbation to ensure that the null energy condition is satisfied in as a strict inequality (not as an equality). We will see later exactly where in the perturbation should be made. On , one can define the following continuous function : if has endpoints , then is the proper time elapsed from to along . An important detail is that we allow the case of a causal curve that consists of only one point. Thus in particular contains a point that corresponds to a causal curve consisting only of the point . The point is the absolute minimum of , since , and at other points in , . The purpose of including the point in the definition of is to ensure that is compact; this compactness follows from the fact that the initial endpoint of a curve is fixed and its final endpoint ranges over the compact set , along with the fact that in a globally hyperbolic spacetime, the space of causal curves with specified endpoints is compact. Compactness of ensures that in every connected component of , the nonnegative function has an absolute minimum. Any can be deformed in to the minimum of in its connected component. We will conclude the proof by showing that the point is the unique local minimum of the function , implying that the local minimum of to which can be deformed in is actually the absolute minimum . This in particular implies that can be deformed in to a causal curve entirely in , as we wished to show. To show that has no other local minimum, we note that any point in other than is a nontrivial causal curve with distinct endpoints in . In general, if the initial point of a causal curve is specified (here ) and the final point of the curve is constrained by a condition of “promptness” (here is supposed to end along , and the condition for it to be a local minimum of means that it arrives on sooner than any nearby causal curve from ; this is the promptness condition), then must be a null geodesic without focal points. However, in an asymptotically AdS spacetime, any null geodesic between distinct points on the conformal boundary is complete in both directions. As remarked earlier, we can assume that the null energy condition in is satisfied along as a strict inequality. A null geodesic that is complete in both directions and along which the null energy condition is satisfied as a strict inequality always has focal points. So the function can have no local minimum other than its absolute minimum , completing the proof.
4 An Algebra of Observables For De Sitter Space
Finally, we turn to a concrete example in which, in order to define a sensible algebra of observables, it is necessary to include an observer in the description CLPW . De Sitter space in dimensions or is the maximally symmetric solution of Einstein’s equations with a positive cosmological constant. It can be described by the metric
(12) |
where is the radius of curvature and is the metric of a round sphere of unit radius. This sphere is compact, so is an example of a closed universe. At time , the sphere has radius so it grows exponentially for or . The exponential growth for is believed to be a good approximation to what is currently beginning to happen in the real world.
In the 1970’s, Gibbons and Hawking GH studied de Sitter space as a simple example of a spacetime with a cosmological horizon – in which an observer cannot see the whole universe. They attached a temperature and entropy to the de Sitter horizon, as Bekenstein and Hawking had done not long before for the horizon of a black hole. The thermal interpretation is most obvious in Euclidean signature, where becomes simply a -sphere , with metric
(13) |
In ordinary quantum field theory in de Sitter space (and also in the presence of semiclassical gravity), there is a natural de Sitter state such that correlation functions in this state can be obtained by analytic continuation from Euclidean signature. Let be a great circle in and let be a length parameter along . is an orbit of a symmetry of (which is uniquely determined if we say that its fixed point set is a copy of ). If we normalize the generator of this to act as along , then obeys . When continued to Lorentz signature, this leads to the striking statement that correlation functions in the state have a thermal interpretation at the de Sitter temperature , where GH ; FHN . The thermal interpretation of de Sitter space has been extensively explored for nearly half a century. For a small sampling of the relevant literature, see Sewell ; Maeda ; BoussoOne ; BoussoTwo ; Banks ; BanksFischler ; BanksOne ; BanksTwo ; BFTwo ; SusskindA ; Susskind ; DF ; BD ; SB ; DST .

In Lorentz signature, de Sitter space is conveniently understood via a Penrose diagram (fig. 8). The great circle continues in Lorentz signature to a hyperbola that has two components, each of them a geodesic. Coordinates can be chosen so that these geodesics make up the left and right boundaries of the Penrose diagram. Given an observer traveling on a geodesic in , we can assume that the worldline of this observer is, say, the left boundary of the diagram. The observer then has past and future horizons which are the diagonals in the picture. The region causally accessible to the observer (the region the observer can see and also can influence) is bounded by these diagonals, along with the left boundary of the diagram. A similar region of the diagram is causally accessible to an observer on the right boundary. The operator generates a symmetry of the Penrose diagram; with a suitable choice of sign, it maps the region accessible to the left observer forward in time and the region accessible to the right observer backwards in time. Near the bifurcate horizon where the two diagonals meet, looks like the generator of a Lorentz boost.
If measures the time along the left boundary, then on the left boundary of the figure, . So it is natural for an observer whose worldline is the left boundary to interpret as a generator of time translations. With this interpretation, since is a symmetry of the region causally accessible to the left observer, this region is time-independent and thus “static.” This is why the region causally accessible to an observer in de Sitter space has been called a “static patch.” However, this view of de Sitter space as being “static” is highly observer-dependent. In a global view, the form (12) of the de Sitter metric shows that is expanding exponentially both toward the future and toward the past, so globally one would definitely not call “static.”
An ordinary quantum field theory in de Sitter space has a Hilbert space of quantum states. In such a theory, we associate to the static patch (or any region) an algebra of observables consisting of operators on that act on the quantum fields in the region in question. The algebra of any local region, and in particular the algebra of the static patch, is a possibly unfamiliar Type III von Neumann algebra. This is an algebra with an infinite amount of quantum entanglement built in, giving an abstract explanation of the fact that entanglement entropy is ultraviolet divergent in quantum field theory. Including weakly coupled gravitational fluctuations does not qualitatively change the picture. We simply include the weakly coupled gravitational field as one more field in the construction of and . What does really change the picture is that in a closed universe, such as de Sitter space, the isometries have to be treated as constraints. In the case of the static patch, the important constraint is the Hamiltonian . Imposing as a constraint means that we should replace by , its invariant subalgebra. But that does not work: the invariant subalgebra is trivial. Roughly, that is because anything that commutes with can be averaged over all the thermal fluctuations and replaced by its thermal average, a -number. A technical statement is that there are no nontrivial invariants in because generates the modular automorphism group of the state for the algebra Sewell , and therefore acts ergodically.
To get a reasonable algebra of observables, we include an observer in the analysis. Of course, as noted in the introduction, in principle an observer should really be described by the theory, not injected from outside. What it means to include an observer is that we consider a “code subspace” of states in which an observer is present in the static patch, and then we consider operators that can be defined in the low energy effective field theory in this code subspace, though they are not well-defined on the whole Hilbert space.
Should we be surprised that we need to include the observer in the analysis to get a sensible answer? As was also noted in the introduction, a gravitating system in a closed universe is the situation in which we are most likely to need to explicitly incorporate the observer in the analysis. That is exactly the situation here because de Sitter space is a simple model of a closed universe, that is, a universe with compact spatial sections.
Once we include an observer in the analysis, there is a rationale to study a particular static patch, namely the region that is causally accessible to that given observer. Moreover, the timelike tube theorem tells us that the algebra of observables in the causally accessible region can be interpreted as the algebra generated by the quantum fields along the observer’s worldline. Thus the algebra of the static patch becomes, in a sense, operationally meaningful once the observer is included.
It turns out that to get a sensible answer, it suffices to consider a minimal model in which the observer is characterized just by a clock with Hamiltonian
(14) |
It is physically reasonable to assume that the observer’s energy is bounded below. The precise lower bound will not be important and we will just take it171717More realistically, the observer would have a mass and minimum energy . If the observer is minimally coupled to gravity, then the observer worldline will be a geodesic. In the presence of an observer, the relevant static patch is the region causally accessible to the observer. to be 0. With that choice, the effect of including the observer is to modify the Hilbert space by
(15) |
where is a multiplication operator on the positive half-line . The algebra is likewise extended from to
(16) |
The last factor is the (Type I) algebra of all bounded operators on .
Finally the constraint becomes the total Hamiltonian of the quantum fields plus the observer:
(17) |
The “correct” algebra of observables taking account of the presence of the observer is therefore
(18) |
that is, the -invariant part of . To be more exact, this is the algebra of observables accessible to the observer in the limit . In higher order in , it will be necessary to incorporate direct couplings between the observer and the quantum fields; the constraint operator will not be a simple sum. Of course, the notion of the algebra of a spacetime region is presumably only well-defined in the limit ; the timelike tube theorem suggests that in going beyond that limit, we should reinterpret as the algebra of observables along the observer’s worldline. As explained presently, the important conclusions that we will draw about are expected to be robust against perturbative corrections in .
The answer (18) makes sense, unlike the previous one. The reason is that once an observer is present, we can “gravitationally dress” any operator to the observer’s world-line. It is easiest to explain this if we momentarily ignore the lower bound . In fact, in the field of a black hole, there is a problem that is mathematically quite similar to what we are discussing here, but without the lower bound on CLPW . If we do ignore the lower bound on , then the Hilbert space including the observer is , and the algebra is . Here is generated by (bounded functions of) and . If is the projection operator onto states with , then , . Similarly the -invariant subalgebras , are related by . So is easily constructed once we understand .
To construct , we reason as follows. For any , the operator
commutes with the constraint . One more operator that commutes with the constraint is itself (or equivalently , which equals modulo the constraint). It follows from a classic result of Takesaki Takesaki that by coincidence was proved in the early days of black hole thermodynamics that (1) there are no more operators in that commute with the constraint, and (2) the algebra that is generated by the operators and is actually a von Neumann algebra of Type II∞, with a trivial center (that is, its center consists only of -numbers). In the case of the static patch in de Sitter space, we want to impose the constraint , so the appropriate algebra is not but . This is an algebra of Type II1, again with trivial center.
A basic introduction to the relevant facts about von Neumann algebras can be found in CLPW , among other places. For much more depth, see Sorce . An important fact for our purposes is that a Type II algebra (unlike one of Type III, which we would have in the absence of gravity) has a trace, that is a complex-valued linear function obeying , for all . Moreover, the trace in a Type II algebra is positive, in the sense that for all .
A von Neumann algebra with trivial center is called a factor. Thus the above analysis implies that in the limit , the algebra of observables in the static patch is a factor of Type II1. A factor is the von Neumann algebra analog of a simple Lie group. A simple Lie group is rigid, in the sense that it has has no infinitesimal deformations. That is not true for a non-simple Lie group; for example, the symmetry group of is non-simple, and can be deformed, by an arbitrarily small perturbation of the commutation relations in its Lie algebra, to the symmetry group of . A similar statement holds for von Neumann algebras: an algebra with a non-trivial center can potentially be deformed to an algebra of a different type, while making the center smaller,181818For example, the Type III algebra described by Leutheusser and Liu in the limit in the field of a black hole LL ; LLtwo has a nontrivial center and is modified by perturbative corrections to a factor of Type II∞ CLPW ; Witt . but a factor is rigid and has no infinitesimal deformations. Though we have only analyzed the algebra of observables in the static patch in the limit , because the answer that we obtained is a factor, perturbative corrections in are not expected to modify the algebra up to isomorphism (they will modify the commutation relations among operators along the observer’s worldline, but not the isomorphism class of the algebra that those operators generate). Nonperturbatively, matters are unclear and indeed it is quite unclear whether quantum de Sitter space makes sense nonperturbatively. If quantum de Sitter space does make sense nonperturbatively, then one expects to describe it by a finite-dimensional Hilbert space BanksOne ; BanksTwo , and the algebra will have to be of Type I.
A Type II algebra does not have pure states, but because it does have a trace, familiar ideas like density matrices and entropies make sense for a state of such an algebra Segal ; LW . To a global state of de Sitter space plus the observer, reduced to the static patch, one can associate a density matrix , characterized by
(19) |
Therefore, we can define a von Neumann entropy
(20) |
There is no such definition in the absence of gravity, because without gravity, the observables in the static patch constitute the Type III algebra . The fact that gravity turns the Type III algebra into a Type II algebra gives an abstract explanation for why entropy is better defined in the presence of gravity than in ordinary quantum field theory. However, from a physical point of view, Type II entropy is a renormalized entropy from which an infinite constant has been subtracted.
As already remarked, if we put no constraint on the observer’s energy, we get an algebra of Type II∞; if we assume the observer’s energy is bounded below, we get an algebra of Type II1. In a Type II1 algebra, the trace is defined for all elements of the algebra, while in an algebra of Type II∞, the trace is finite only for a dense subset of elements of the algebra. In particular, for Type II1, the identity element has a finite trace and we can normalize the trace so that the trace of the identity is 1, while in Type II∞, the trace of the identity element is .
For our purposes, the important difference is that a Type II1 algebra has a state of maximum possible entropy, which is the “maximally mixed” state with density matrix . This is consistent with , assuming that the trace has been appropriately normalized. Evaluating for , we see that the state with has entropy 0. It is not difficult to prove that all other states have negative entropy (for example, see CLPW ; LW ). This tells us the meaning of the subtraction that is involved in defining the entropy of a state of a Type II1 algebra so that for : the constant that is subtracted is the maximum possible entropy. In general, entropy of a state of a Type II1 algebra is the entropy difference of that state relative to the entropy of the maximum entropy state. The maximum entropy state with is the Type II1 analog of a maximally mixed state in ordinary quantum mechanics, in which the density matrix is a multiple of the identity. By contrast, there is no upper bound on the entropy of a state of a Type II∞ algebra.
It is felicitous that a Type II1 algebra has a state of maximum entropy, because in fact de Sitter space is believed to have a state of maximum entropy – namely “empty de Sitter space,” with all the entropy in the cosmological horizon Maeda ; BoussoOne ; BoussoTwo . So to get a reasonable model of de Sitter space, it is important to assume that the observer’s energy is bounded below, which is a more reasonable assumption anyway, thereby ensuring that the algebra of observables is of Type II1, not Type II∞. One can explicitly construct a state that has maximum entropy (and thus density matrix ) when reduced to the static patch. In fact, one can pick . Thus empty de Sitter space, tensored with a state in which the observer’s energy has a thermal distribution at the de Sitter temperature, has maximum entropy from the point of view of the Type II1 algebra. In this sense the analysis based on the Type II1 algebra agrees with the claim that empty de Sitter space has maximum entropy.
We can now compare with some further claims in the previous literature. First of all, since the maximum entropy state has , it has a “flat entanglement spectrum” (all eigenvalues of the density matrix are equal) and accordingly the Rényi entropies all vanish:
(21) |
This matches with a result that has been found using Euclidean path integrals to analyze the Rényi entropies of the static patch DST . Given the assertion that de Sitter space has a state of maximum entropy, this result is what one should expect. In ordinary quantum mechanics, the maximum entropy state of a system is “maximally mixed,” with a “flat entanglement spectrum” (the density matrix is a multiple of the identity and all its eigenvalues are equal) and its Rényi entropies are independent of .
Now, suppose that the observer makes a measurement with two outcomes that correspond to the projection operators and . The probabilities of the two outcomes are and . All values are possible. If the outcome corresponding to is observed, then after this measurement, the density matrix is
Since the two eigenvalues of are 0 and , one has so the entropy after the observation is
The entropy reduction from knowing the outcome is therefore , and this is related to the probability of the given outcome by
However, the probability of a (low entropy) energy fluctuation of the static patch is
according to the thermal interpretation of de Sitter space. Since also , we must have for consistency of the two descriptions
In other words, “‘thermal” suppression of a fluctuation can be understood as purely entropic suppression. This is surprising, but it has been argued before on other grounds, notably by considering the case that the “fluctuation” is a small black hole at the center of the static patch Susskind .
Which part of this is unexpected? The formula for the probability of an outcome is an inevitable consequence of having a maximum entropy state in which all states are equally probable. In other words, if all states are equally likely, then the probability of a given outcome is just proportional to the number of microstates that are compatible with that outcome. Here we use language appropriate for an ordinary quantum system with a finite-dimensional Hilbert space. For a system described by a Type II1 algebra, the number of microstates compatible with any given outcome is infinite, and one has to express the argument in terms of traces, as we did earlier. In short, the surprise is not that , which one should expect for a maximum entropy state, but that after coupling to gravity and including the observer, the thermal state can be promoted to a maximal entropy state .
Let us see explicitly how this happens at the level of correlation functions.191919An error in a previous claim about this matter was pointed out by G. Penington. First we recall some basic facts about time-dependent correlation functions in a thermal ensemble. The time dependence of an operator is defined in the usual way by . A typical time-dependent two-point function is
(22) |
Here is the trace in the Hilbert space of a thermal system with Hamiltonian , and is the partition function. It follows immediately from the definition that is holomorphic in a strip and moreover that the boundary value at is the thermal correlator with the opposite ordering of the operators. In other words, let
(23) |
This function is related to by
(24) |
The precise meaning of this statement is that there is a function holomorphic in the strip whose boundary value on the upper boundary is , while its boundary value on the lower boundary is .
Let us express this in terms of Fourier transforms of the two functions. Suppose that
(25) | ||||
(26) |
Then eqn. (24) becomes
(27) |
These facts about an ordinary quantum system also hold in de Sitter space for time-dependent correlation functions in the state , though more sophisticated proofs are required. To make precisely the same argument for thermal correlators in the static patch of de Sitter space that one can make for an ordinary thermal system, one should have a Hilbert space that describes the static patch, and should be the trace in this Hilbert space. Such a “one-sided” Hilbert space does not exist for quantum fields in de Sitter space, because the algebra of the static patch is of Type III, not Type I. Instead one only has a global Hilbert space describing all of de Sitter space. To deal with this situation requires more careful arguments using Tomita-Takesaki theory Sewell .
What happens after coupling to gravity? As described earlier, we introduce an observer with energy , and canonical momentum . Let be the function that is 1 for and vanishes for , so that is the projection operator onto states of . To get a model of de Sitter space with weakly coupled gravity, we replace operators , of the original thermal system by gravitationally dressed versions
(28) |
If we do not include the projection operators , the algebra would be of Type II∞, with no maximum entropy state. Including the projection operators gives a Type II1 algebra with a maximum entropy state. As claimed earlier, this state is
(29) |
where is the natural de Sitter invariant state in the absence of gravity, and
(30) |
is a state in which the observer’s energy has a thermal distribution. Time dependence is introduced in the usual way, by, for example, .
In the limit that gravity is weakly coupled, the de Sitter analogs of , are supposed to be
(31) | ||||
(32) |
We have used the facts that , and . We observe now that
(33) |
so eqn. (31) simplifies to
(34) | ||||
(35) |
We expect , since either of these functions is supposed to be , where is the trace of the Type II1 algebra (as opposed to the trace in the Hilbert space of an underlying thermal system, which has been denoted – and which anyway does not really exist in the case of de Sitter space).
Now let us write a Fourier-transformed version of these last two equations. In the formula for , we see that a contribution that varies with as comes from intermediate states with , but in the formula for , such a contribution comes from states with . The upshot of this is that Fourier-transformed formulas can be written for and that are just analogous to those of eqn. (25), but with an extra factor involving the expectation value in the state of :
(36) | ||||
(37) |
Using the definition of , we can find explicit formulas for and :
(38) | ||||
(39) |
Note that
(40) |
We have then
(41) | ||||
(42) |
with
(43) |
So using eqns. (27) and (40), we have
(44) |
implying that . This confirms that the coupling to gravity has converted the thermal expectation value of an operator into a trace, and moreover that this trace is the expectation value in the maximum entropy state .
In sum, we have identified a concrete example of including an observer in order to get a sensible answer in a cosmological model with a closed universe. And, at least in the example of de Sitter space, we have understood that gravity makes the notion of entropy better defined than it is in ordinary quantum field theory. It is possible CLPW to probe more deeply and show that entropy defined via the Type II1 algebra agrees, up to an additive constant independent of the state, with the generalized gravitational entropy, as usually computed via quantum extremal surfaces. It is also possible to give an analogous treatment of a black hole CLPW ; Witt ; CPW , involving in this case an algebra of Type II∞ and therefore no upper bound on the entropy.
Acknowledgements I thank V. Hubeny, A. Jaffe, J. Kohn, H. Maxfield, R. Mazzeo, G. Penington, and A. Strohmaier for comments and advice. Research supported in part by NSF Grant PHY-2207584.
References
- [1] V. Chandrasekharan, R. Longo, G. Penington, and E. Witten, “An Algebra of Observables for De Sitter Space,” arXiv:2206.10790.
- [2] H. J. Borchers, “Field Operators as Functions In Spacelike Directions,” Il Nuovo Cimento 33 (1964) 1.
- [3] H. J. Borchers, “Uber die Vollständigkeit lorentzinvarianter Felder in einer zeitartigen Röhre,” Il Nuovo Cimento 19 (1961) 787.
- [4] H. Araki, “A Generalization Of Borchers’ Theorem,” Helv. Phys. Acta 36 (1963) 132-9.
- [5] A. Strohmaier, “On the Local Structure of the Klein-Gordon Field on Curved Spacetimes,” Lett. Math. Phys. 54 (2000) 249-61.
- [6] A. Strohmaier and E. Witten, “Analytic States in Quantum Field Theory on Curved Spacetimes,” arXiv:2302.02709.
- [7] A. Strohmaier and E. WItten, “The Timelike Tube Theorem in Curved Spacetime,” arXiv:2303.16380.
- [8] W. G. Unruh, “Notes on Black Hole Evaporation,” Phys. Rev. D14 (1976) 870.
- [9] A. Jaffe, “Wick Polynomials at a Fixed Time,” J. Math. Phys. 7 (1966) 1250.
- [10] H. Bostelman and C. J. Fewster, “Quantum Inequalities From Operator Product Expansions,” Commun. Math. Phys. 292 (2009) 761-95, arXiv:0812.4760.
- [11] E. J. Straube, “Harmonic and Analytic Functions Admitting a Distribution Boundary Value,” Annali della Scuola Normale Superiore di Pisa 4 Serie, 11 (1984) 559-91.
- [12] S. Hollands and R. M. Wald, “Axiomatic Quantum Field Theory in Curved Spacetime,” Commun. Math. Phys. 293 (2010) 85-125.
- [13] M. Keyl, “Quantum Fields on Timelike Curves,” arXiv:math-ph/0012024.
- [14] C. J. Fewster, “Lectures on Quantum Energy Inequalities,” arXiv:1208.5399.
- [15] J. Smoller, Shock Waves and Reaction-Diffusion Equations, second edition (Springer-Verlag, 2012).
- [16] D. Tataru, “Unique Continuation For Operators With Partially Analytic Coefficients,” J. Math. Pures Appl. 78 (1999) 505-21.
- [17] H. Casini, M. Huerta, J. M. Magan, and D. Pontello, “Entropic Order Parameters for the Phases of QFT,” JHEP 04 (2021) 277, arXiv:2008.11748.
- [18] H. Casini and J. M. Magan, “On Completeness and Generalized Symmetries in Quantum Field Theory,” arXiv:2110.11358.
- [19] D. Harlow, “Wormholes, Emergent Gauge Fields, and the Weak Gravity Conjecture,” JHEP 01 (2016) 122, arXiv:1510.07911
- [20] J. Polchinski, “Monopoles, Duality, and String Theory,” Int. J. Mod. Phys. A 19S1 (2004) 145-56, hep-th/0304042.
- [21] T. Banks and N. Seiberg, “Symmetries and Strings in Field Theory and Gravity,” Phys. Rev. D83 (2011) 084019, arXiv:1011.5120.
- [22] D. Harlow and H. Ooguri, “Symmetries in Quantum Field Theory and Quantum Gravity,” Commun. Math. Phys. 383 (2021) 3, 1669-1804, arXiv:1810.05338.
- [23] D. Gaiotto, A. Kapustin, N. Seiberg, and B. Willett, “Generalized Global Symmetries,” JHEP 02 (2015) 172, arXiv:1412.5148.
- [24] E. Witten, “Symmetry and Emergence,” Nat. Phys. 14 (2018) 116-9, arXiv:1710.01791.
- [25] T. Banks, M. R. Douglas, G. T. Horowitz, and E. Martinec, “AdS Dynamics From Conformal Field Theory,” hep-th/9808016.
- [26] I. Bena, “On the Construction of Local Fields in the Bulk of AdS5 and Other Spaces,” Phys. Rev. D62 (2000) 066007, hep-th/9905186.
- [27] V. Balasubramanian, P. Kraus and A.E. Lawrence, “Bulk Versus Boundary Dynamics in Anti-de Sitter Space-Time,’ Phys. Rev. D59 (1999) 046003, hep-th/9805171.
- [28] A. Hamilton, D. Kabat, G. Lifschytz, and D. A. Lowe, “Holographic Representation of Local Bulk Operators,” Phys.Rev. D74 (2006) 066009, hep-th/0606141.
- [29] A. Hamilton, D. Kabat, G. Lifschytz, and D. A. Lowe, “Local Bulk Operators in AdS/CFT: A Holographic Description of the Black Hole Interior,” Phys. Rev. D75 (2007) 106001, hep-th/0612053.
- [30] D. Kabat, G. Lifschytz, and D. A. Lowe, “Constructing Local Bulk Observables in Interacting AdS/CFT,” Phys. Rev. D83 (2011) 106009, arXiv:1102.2910.
- [31] I. Heemskerk, D. Marolf, J. Polchinski, and J. Sully, “Bulk and Transhorizon Measurements in AdS/CFT,” JHEP 10 (2012) 165, arXiv:1201.3664.
- [32] I. A. Morrison, “Boundary-to-bulk Maps for AdS Causal Wedges and the Reeh-Schlieder Property in Holography,” JHEP 05 (2014) 053, arXiv:1403.3426.
- [33] V. E. Hubeny, “Covariant Residual Entropy,” JHEP 09 (2014) 156, arXiv:1406.4611.
- [34] R. M. Wald, General Relativity (University of Chicago, 1984).
- [35] S. Gao and R. M. Wald, “Theorems on Gravitational Time Delay and Related Issues,” Class. Quant. Grav. 17 (2000) 4999-5008, gr-qc/0007021.
- [36] E. Witten, “Light Rays, Singularities, and All That,” Rev. Mod. Phys. 92 (2020) 045004, arXiv:1901.03928.
- [37] G. W. Gibbons and S. W. Hawking, “Cosmological Event Horizons, Thermodynamics, and Particle Creation,” Phys. Rev. D15 (1977) 2738-51.
- [38] R. Figari, R. Hoegh-Krohn, and C. R. Nappi, “Interacting Relativistic Boson Fields in the De Sitter Universe With Two Space-Time Dimensions,” Commun. Math. Phys. 44 (1975) 265-278.
- [39] G. L. Sewell, “Quantum Fields On Manifolds: PCT and Gravitationally Induced Thermal States,” Ann. Phys. 141 (1982) 201-24.
- [40] K. Maeda, T. Koike, M. Narita, A. Ishibashi, “Upper Bound for Entropy in Asymptotically de Sitter Space-time,” Phys. Rev. D 57(6) (1998) 3503.
- [41] R. Bousso, “Positive Vacuum Energy and the Bound,” JHEP 11 (2000) 038, hep-th/0012052.
- [42] R. Bousso, “Bekenstein Bounds in de Sitter and Flat Space,” JHEP 04 (2001) 035, hep-th/0010252.
- [43] T. Banks, “Cosmological Breaking of Supersymmetry? Or Little Lambda Goes Back to the Future, II,” Int. J. Mod. Phys. A16 (2001) 910-921, hep-th/0007146.
- [44] T. Banks and W. Fischler, “-Theory Observables For Cosmological Spacetimes,” arXiv:hep-th/0102077.
- [45] T. Banks, “More Thoughts on the Quantum Theory of Stable de Sitter Space,” arXiv:hep-th/0503066.
- [46] T. Banks, B. Fiol, and A. Morisse, “Towards a Quantum Theory of de Sitter Space,” arXiv:hep-th/0609062.
- [47] T. Banks and W. Fischler, “The Holographic Model Of Cosmology,” arXiv:1806.01749.
- [48] L. Susskind, “De Sitter Holography: Fluctuations, Anomalous Symmetry, and Wormholes,” Universe 7 (2021) 464, arXiv:2106.03964.
- [49] L. Susskind, “Black Holes Hint Towards De Sitter-Matrix Theory,” arXiv:2109.01322.
- [50] X. Dong, E. Silverstein, and G. Torroba, “De Sitter Holography and Entanglement Entropy,” arXiv:1804.08623.
- [51] P. Draper and S. Farkas, “De Sitter Black Holes as Constrained States in the Euclidean Path Integral,” Phys,. Rev. D105 (2022) 126022.
- [52] T. Banks and P. Draper, “Comments on the Entanglement Spectrum of de Sitter Space,” JHEP 01 (2023) 135, arXiv:2209.08991.
- [53] H. Lin and L. Susskind, “Infinite Temperature’s Not So Hot,” arXiv:2206.01083.
- [54] M. Takesaki, “Duality for Crossed Products and the Structure of von Neumann Algebras of Type III,” Acta Mathematica 131 (1973) 249-310.
- [55] J. Sorce, “Notes on the Type Classification of von Neumann Algebras,” arXiv:2302.01958.
- [56] S. Leutheusser and H. Liu, “Causal Connectability Between Quantum Systems and the Black Hole Interior in Holographic Duality,” arXiv:2110.05497.
- [57] S. Leutheusser and H. Liu, “Emergent Times in Holographic Duality,” arXiv:2112.12156.
- [58] E. Witten, “Gravity and the Crossed Product,” JHEP 10 (2022) 008, arXiv:2112.12828.
- [59] I. E. Segal, “A Note on the Concept of Entropy,” J. Math. Mech. 9 (1960) 623-9.
- [60] R. Longo and E. Witten, “A Note On Continuous Entropy,” arXiv:2202.03357.
- [61] V. Chandraskharan, G. Penington, and E. Witten, “Large Algebras and Generalized Entropy,” arXiv:2206.10780.