0% found this document useful (0 votes)
38 views

Emergent Time and Time Travel in Quantum Physics

The article explores the concept of time travel within the framework of quantum physics, particularly focusing on the challenges posed by the problem of time in quantum gravity. It discusses the Page-Wootters formalism as a means to understand emergent time and its implications for time travel, while also addressing various counterarguments and theories related to time. The authors aim to study a toy model involving harmonic oscillators to investigate the viability of time travel when time is considered an emergent concept.

Uploaded by

L
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
38 views

Emergent Time and Time Travel in Quantum Physics

The article explores the concept of time travel within the framework of quantum physics, particularly focusing on the challenges posed by the problem of time in quantum gravity. It discusses the Page-Wootters formalism as a means to understand emergent time and its implications for time travel, while also addressing various counterarguments and theories related to time. The authors aim to study a toy model involving harmonic oscillators to investigate the viability of time travel when time is considered an emergent concept.

Uploaded by

L
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 20

Article

Emergent Time and Time Travel in Quantum Physics


Ana Alonso-Serrano 1,2,† , Sebastian Schuster 3, *,† and Matt Visser 4,†

1 Institut für Physik, Humboldt-Universität zu Berlin, Zum Großen Windkanal 6, 12489 Berlin, Germany
2 Max-Planck-Institut für Gravitationsphysik (Albert-Einstein-Institut), Potsdam Science Park,
Am Mühlenberg 1, 14476 Potsdam, Germany; [email protected]
3 Ústav Teoretické Fyziky, Matematicko-Fyzikální Fakulta, Univerzita Karlova, V Holešovičkách 747/2,
180 00 Praha 8, Czech Republic; [email protected]
4 School of Mathematics and Statistics, Victoria University of Wellington, P.O. Box 600, Wellington 6140,
New Zealand; [email protected]
* Correspondence: [email protected]
† These authors contributed equally to this work.

Abstract: Entertaining the possibility of time travel will invariably challenge dearly held concepts of
arXiv:2312.05202v2 [gr-qc] 27 Feb 2024

fundamental physics. It becomes relatively easy to construct multiple logical contradictions using
differing starting points from various well-established fields of physics. Sometimes, the interpretation
is that only a full theory of quantum gravity will be able to settle these logical contradictions. Even
then, it remains unclear if the multitude of problems could be overcome. Yet as definitive as this
seems to the notion of time travel in physics, such a recourse to quantum gravity comes with its
own, long-standing challenge to most of these counter-arguments to time travel: These arguments
rely on time, while quantum gravity is (in)famously stuck with and dealing with the problem of time.
One attempt to answer this problem within the canonical framework resulted in the Page–Wootters
formalism, and its recent gauge-theoretic re-interpretation—as an emergent notion of time. Herein,
we will begin a programme to study toy models implementing the Hamiltonian constraint in quantum
theory, with an aim towards understanding what an emergent notion of time can tell us about the
(im)possibility of time travel.

Keywords: time travel; quantum gravity; minisuperspace; Page–Wootters formalism; emergent time;
relational dynamics

1. Introduction
As fascinating as the notion of time travel is, its long list of problematic issues—ranging
from the classical [1,2] & [3, p.212ff] to the quantum theoretical [4–7]—presents a very solid
case against it, and its absence is effectively a feature of the universe accessible to human
experience. Yet if one wants to look closely at these arguments, cracks and loopholes
start to appear—many of which shall be mentioned in the following. To an extent, it
Citation: Alonso-Serrano, A.; almost seems as if the best evidence against any notion of time travel is that of our daily
Schuster, S.; Visser, M. Emergent Time experienced notion of time and causality, our intuition. Still, the known universe is bigger
and Time Travel in Quantum Physics. than our day-to-day experience, and particularly the field of high energy physics is ripe
Preprints 2024, 10, 73. with examples experimentally, observationally, or theoretically open to inquiry—in which
https://doi.org/10.3390/universe10020073various concepts or preconceived notions of human-scale physics have to be let go. In no
field of theoretical physics is this more apparent than in the quest for quantum gravity. The
field of contenders is large, and still growing: String theory, loop quantum gravity, quantum
geometrodynamics, asymptotic safety, causal dynamical triangulation, causal set theory,
Copyright: © 2024 by the authors.
Hořava–Lifshitz gravity, . . . Each comes with its own set of new notions beyond what can
Licensee MDPI, Basel, Switzerland. be experienced directly by human senses. Yet, these proposals still have to address, in one
This article is an open access article way or another, an issue plaguing quantum gravity from its beginning: Time. This problem
distributed under the terms and of time, and its eventual resolution, then naturally will also impact what these theories have
conditions of the Creative Commons to say about (or against) time travel.
Attribution (CC BY) license (https:// The ‘problem of time’ in classical canonical gravity is settled by a careful, gauge-
creativecommons.org/licenses/by/ theoretic analysis [8] that distinguishes the different rôles played by the Hamiltonian, either
4.0/). in the constraint/gauge fixing or in the definition of observables, respectively. In quantum
2 of 20

gravity, however, the problem of time remains (at the very least) a topic of active discussion
and research [9–11]. The reason for this goes back even before a theory of quantum gravity
was the focus: While there are ways to give position and momentum a meaning in terms
of operators on (rigged) Hilbert spaces in quantum mechanics, this cannot be carried
over to time and energy. This was concerning to researchers in quantum mechanics early
on, as the time-energy uncertainty relations proved to be both reliable and instructive,
while not fitting into the standard paradigm of the Robertson–Schrödinger uncertainty
relations for (generic) Hermitian operators. Only the proof of Mandelstam and Tamm
in 1945 improved this uncertainty relation’s foundation, while begging the question of
why a different approach was needed [12]. Essentially, a series of impossibility results for
time operators canonically conjugate to a Hamiltonian was presented: Schrödinger used
an approach based on normalizability [13], similar to the argument for why plane waves
require a rigged Hilbert space. Pauli, in a footnote, used a simpler argument juxtaposing
discrete spectra and the continuous translation in energy such a time operator would
induce [14,15]. Finally (and in reply to what will be discussed shortly), Unruh and Wald
gave a general argument for semi-bounded Hamiltonians linked to unitarity [16]. Two at
first glance quite distinct approaches were developed in reply to these no-go results.
One approach was similar in spirit to the answer to early counterarguments to plane
waves, e.g., eigenstates of definite position and momentum. While with position and
momentum one relaxes the idea of being restricted to Hilbert space and instead works
in a rigged Hilbert space [17–19], here, when faced with the above-cited no-go results
concerning time operators, one relaxes the idea of Hermiticity. Instead, one introduces the
notion of a positive, operator-valued measure (POVM). It can represent not just perfect
measurements yielding a definite eigenstate, but also imprecise measurements [20]. Besides
allowing one to address the measurement problem in quantum field theory [21], phase
observables [22], or the already mentioned, imprecise measurement processes [20,23]—and
more important for the present context—this allows a notion of time measurement [24].
Since the no-go theorems relate to Hermitian operators, the more general POVMs are not
ruled out as time observables. As we will see, the context of time travel allows one to avoid
these no-go results by other means, too. Still, the bigger picture of gauge theory provides
the most pertinent background for extensions of the model to be described in this paper.
Another approach that developed independently from POVMs was to carefully dis-
tinguish between a clock and time. Paraphrasing Einstein [25], time is meaningful only
after specifying the clock measuring it. Early on, in developing canonical quantum gravity,
DeWitt [26] pointed out how different subsystem’s operators in a given Hamiltonian con-
straint could serve this purpose of an effective clock. Page and Wootters then took this idea
and made it more general and workable, by phrasing time evolution in terms of conditional
probabilities linking different subsystems [27,28].1 This Page–Wootters (PW) formalism
was soon challenged on various grounds by Unruh, Wald [16] and Kuchař [29], leading to
this approach lying dormant for some time.
Recent developments then made the latter approach re-emerge [30–35], and both
approaches converge [36,37]. This also allowed one to address and re-contextualize the
earlier criticisms of the PW formalism [37,38]. In essence, the clocks chosen within the
PW formalism become part of a gauge-theoretic picture of clocks; the ‘physical Hilbert
space’ being a ‘clock-neutral’ picture, and different clocks representing different ‘gauge
conditions’ [37]. These results form part of a larger research programme that aims to
ground the rulers and clocks at least implicitly employed in external symmetries with an
operational underpinning [39–44]. For our purposes below, the key takeaway is that the PW

1 With a bit of hindsight, already Schrödinger anticipated this in his footnote on page 245 of [13] [here, in our
translation and keeping the somewhat convoluted grammar of the original]: “An interesting application of
this is the following: if one knows of a system composed of several, coupled subsystems only the total energy,
then it is impossible to know more about the distribution of energy across the subsystems than the statistical,
time-independent data, which already follows from the knowledge of the total energy. Except for the case that
individual subsystems are in truth fully decoupled, energetically isolated from the others.”
3 of 20

formalism still works, if its defects and the criticisms of it are understood as related to issues
already familiar from electromagnetism: One has to choose whether a calculation should
be gauge-fixed or gauge-independent. Depending on the question at hand, one or the other
will be better suited to calculations or understanding. Our model below will be too simple
to fully showcase these developments, but given the criticisms the PW formalism faced, it
is important to keep the resolution in mind.
Within this special issue’s scope, many other pertinent links of time and time travel to
quantum physics could be made. A selected, non-exhaustive list would include: Deutsch
CTCs and (non/retro-)causal quantum processes [45–49], simulation of time (travel) in
analogues [50–52], semi-classical/quantum stability [4–7,53], et cetera. For the purposes of
this article, however, we would like to highlight only two further research avenues. One
is that of different notions of time. Yet another, non-exhaustive list would include at least
cosmological time, psychological time, parameter time, thermodynamic time, dynamical
versus kinematical time, . . . While this distinction is studied mostly in the context of ‘the
arrow of time’ [54], it is worth pointing out that many counterarguments to time travel
have to rely at least on some conflation of these concepts. To give just one example, if
thermodynamic arguments are brought forth, one needs to establish a more or less direct
relation between the thermodynamic notion implied by the second law2 and the notion
of time invoked in time travel. (This itself does not have to be the same notion as that of
general relativity (GR), where time travel can be identified with closed, time-like curves
(CTCs) or related concepts [56]. In theories other than GR, a different notion of time might
be relevant to describe time travel!)
The other important point to make is that despite all the counterarguments, time
travel is just one notion of many with questionable ‘physicality’ [57]. Even within the
context of GR, other notions of physicality vie for validity with the absence of time travel,
for example, various kinds of completeness (geodesic, hole free, . . . ) or the validity of
various kinds of energy conditions. These notions, however, are not all compatible with
each other [58–63]. One might find oneself in the uncomfortable situation that a dearly
held and important property (stability, completeness, fulfilled energy conditions, . . . ) will
require time travel to be permitted; and this just within GR. Other theories are unlikely
to be free from such problems, especially as many of the comforting no-go theorems are
specifically proven within the context of GR. It it therefore not surprising that in the context
of classical field theories of gravity alone, there exist many different ways to realize time
travel. This includes wormholes [3,64], warp drives [65], cosmologies [66–69], maximally
extended space-times (such as Kerr near the central ring singularity) [70], and various
more mathematical construction techniques [71–73]. The latter can also be employed
to distinguish between, for example, time travel (CTCs) and time machines (whatever
structure creates or causes CTCs) [59,60,73].
This finally brings us to the goal of this paper: We want to study what can be said
about the viability of time travel if time itself is only an emergent concept, as in the PW
formalism. Concretely, we will be studying an example of two non-interacting harmonic
oscillators similarly constrained to a fixed total energy of zero as in a Wheeler–DeWitt
(WDW) equation, i.e., it will mimic a minisuperspace model of time travel. This toy
model is by no means meant to exhibit an exhaustive description of whether and how
time travel can arise in the PW formalism. Nor should this model be taken literally as
a minisuperspace model, as no gravitational model is canonically quantized to arrive
at it; rather it demonstrates possible phenomenology. We will see that the results in
our toy model point to an extraordinarily bland version of Novikov’s self-consistency
conjecture [74–76].

2 One would also have to brush aside that standard thermodynamics has to artificially graft time onto its
formalism in the first place, leading to the very name ‘thermodynamics’ being rather unintuitive compared to
other occurrences of ‘dynamics’ in physical terminology [55].
4 of 20

1.1. Outline
Following the preceding introduction to and overview of the related research question
in section 1, we will provide an introduction into the methods employed: In section 2, we
briefly summarize the PW formalism. In section 3 we demonstrate POVM’s utility for
implementing time observables in the context of a single harmonic oscillator. The main
part of the paper is employing these concepts then, in section 4, to two harmonic oscillators
subjected to a WDW-like Hamiltonian constraint equation. We close by discussing our
results and pointing out future directions and open questions in section 5. Finally, the
appendices collect additional details and context: Appendix A reviews the mathematical
definition of a POVM, appendix B collects results concerning periodic time in the quantum
harmonic oscillator that supplement the discussion of section 3.1, while appendix C gives
additional details concerning the construction of the paper’s figures.
We will use natural units in which G = h̄ = c = 1.

2. The Page–Wootters Formalism


Let us consider a quantum system (‘the universe’) that separates into two subsystems,
a ‘clock’ system and a ‘residual’ system. The PW formalism is our first, and most important
ingredient for studying time travel without time (or rather: time travel with only an
emergent notion of time). It is a proposal to give an operational meaning to measuring
the time evolution of the residual subsystem in the larger quantum system (the ‘universe’)
with respect to the other, (mostly) decoupled ‘clock’ subsystem. Partly, this separation into
‘clock’ and ‘residual’ (in the literature often ‘rest’) systems is motivated by the problem
of time encountered in quantum gravity. In terms of the total Hilbert space H, the clock
Hilbert space H C , and H R , the residual Hilbert space, this means:

H = HC ⊗ HR . (1)

Subsequent developments and refinements sometimes further separated what we call H R


into evolving and ancillary/memory/supplementary/. . . systems, but for our purposes,
we will not follow this line of thinking—at least in this current paper. In future work,
it could be interesting to see whether such an approach could provide a notion of time
travel that—contrary to GR’s CTCs—only allows for a finite number of temporal round
trips. Each of these Hilbert spaces is equipped with a Hamiltonian acting on these separate
Hilbert spaces, such that

Ĥ |Ψ⟩ = ĤC ⊗ 1R + 1C ⊗ ĤR |Ψ⟩ = E |Ψ⟩ ,



(2)

where |Ψ⟩ is a state in the total Hilbert space H. In line with Schrödinger’s 1931 footnote,
this state |Ψ⟩ is required to be in a definite, fixed eigenstate of Ĥ with total energy E. More
generally, the PW formalism remains viable even for density matrices ρ̂, with the above
case regained by setting ρ̂ = |Ψ⟩ ⟨Ψ|.
Examples of such systems usually are constructed by symmetry-reducing the WDW
equation of quantum gravity, the quantized Hamiltonian constraint of GR:

Ĥ |Ψ⟩ = 0. (3)

In this expression, we gloss over many technicalities (such as operator ordering, domain
issues, topological concerns, configuration space considerations, . . . ) [19,26,77], as our
focus lies less on quantum gravity and more on the emergent notion of time that the PW
formalism provides. Still, we will occasionally take inspiration from quantum gravity. In
particular, one can consider our final model to be a crude example of a minisuperspace
model that has not been found through symmetry reduction. Already here, we therefore
want to point out that many questions the minisuperspace models of quantum black holes
or quantum cosmology try to answer, will not and cannot appear in our context. Most
5 of 20

notably will be an absence of gravitational singularities, whose avoidance has been a


mainstay of quantum gravity.
The key step for formulating the PW formalism now lies in describing the evolution
of the system encoded through a tuple (H R , ĤR ) in terms of the evolution of (H C , ĤC ). For
this, one picks some initial state |ψ(0∗ )⟩ C ∈ H C , which is then evolved using the clock
Hamiltonian according to

|ψ(θ∗ )⟩C := e−i ĤC θ∗ |ψ(0∗ )⟩C . (4)

This corresponds to the familiar idea of unitary time evolution of states in the Schrödinger
picture through a Hamiltonian, but—at this level of the discussion, anyway—is a defin-
ition. The Schrödinger or Heisenberg equations of motion would then be a result of this
approach [27,36].
The evolution of the subsystem ĤR is then evaluated in terms of relative probabilities
with respect to these clock states. To do this, let us first define the projector

P̂θ∗ := (|ψ(θ∗ )⟩ C C
⟨ψ(θ∗ )|) ⊗ 1R (5)

of a given state onto particular ‘clock time states’ |ψ(θ∗ )⟩ C in H C , leaving states in H R
unchanged. Then the evolution in H R by ĤR with respect to clock states is given by the
expression
tr ĤR P̂θ∗ ρ̂
E( ĤR |θ∗ ) = . (6)
tr P̂θ∗ ρ̂
This gives the conditional expectation value of the residual Hamiltonian ĤR when the clock
(state) ‘reads’ θ∗ . The technical requirements behind this interpretation are stationarity,

[ Ĥ, ρ̂] = 0, (7)

and commutativity of ĤR and ĤC , which we enforced by the block structure of Ĥ encoded
in equation (2). These assumptions are quite strong, especially once one wants to include
an interaction Hamiltonian coupling H C and H R on the right hand side of equation (2).
Given the recent clarification on this issue provided by [30,31,36,37], we want to return to
such generalizations at a later point, but stick to these more restrictive assumptions in the
present context.
Some general comments are appropriate at this point: First, fixing the initial, total
state |Ψ⟩ can, in general, greatly change the time evolution as described by equation (6).
Second, to reiterate: This is not the most sophisticated or modern implementation of the
PW formalism, see the previous paragraph. Third, without further specifying a particular
system comprised of the various Hamiltonians present in equation (2), we cannot even
begin our investigation. This holds true for the sort of time evolution people expect—one
without time travel—just as well as for the more unexpected time evolution—showcasing
a notion of time travel. The next step will be to lay the groundwork for our toy model,
specifying the state and the evolution.

2.1. Time Travel in the Page–Wootters Formalism


In order to prepare this next step, let us first connect the PW formalism to some of the
notions mentioned in the introduction. In particular, we would like to be able to talk about
time travel in this setting. The simplest way to achieve this is to demand that the domain
from which our time parameter θ is taken to be S1 , i.e., periodic. By itself, this will not
guarantee a useful notion of time travel. A simple way to see this is by looking at everyday
clocks: They are all periodic. Only a calendar, also keeping track of years, would be a truly
non-periodic clock. One does not travel back in time when one wakes up at the same time
each morning, just because the alarm clock is periodic.
6 of 20

Instead, we will try to find a way to implement the simplest possible version of
time travel. This is (Novikov) self-consistency. When the clock returns to a state, so
will the remainder system return to the same state it had when the clock last was in this
state. This situation can (at least in some systems and models) also yield existence results,
though usually at the expense of uniqueness [76,78–81]. So far, however, the PW formalism
as presented here does not allow for this. The evolution of the clock as prescribed by
equation (4) is ignorant of the domain from which to draw θ∗ .
In order to alleviate this problem, we will borrow concepts from the POVM approach
to time operators already mentioned in the introduction, and to be discussed in more detail
in the next section. In order to incorporate this in the PW formalism, we note that the key
ingredient for comparing the evolution of the clock with that of the remainder system is
encoded in the definition of the projector onto clock states, equation (5). The goal, thus, will
be to not only influence the clock evolution through a convenient choice of the global state
|Ψ⟩. Rather, we also will have to make a judicious choice of the clock states themselves,
and how to identify them in a meaningful way.

3. POVMs, the Harmonic Oscillator, and Time


The second ingredient of our investigation is a particularly simple example of time
operators and POVMs: The harmonic oscillator [22,24,82–84]. As an ubiquitous model
system in all fields of classical and quantum physics, it is likely to be an instructive
ingredient even in our systems following a Hamiltonian constraint similar to the WDW
equation (3). Our discussion will follow the presentation of the two just-mentioned articles,
adapted to the notation chosen for our application. Concretely, it is possible in the POVM
framework to introduce time with a periodic domain through a phase POVM. Similar to
time operators, also an operator representing a phase variable has been difficult to construct
in a mathematically rigorous manner. Here, the problems are tied to issues of boundary
conditions, hermiticity, and the domain of operators not given in standard notation. As
with time operators, the more general setting of POVMs provides a way around this. Since
a time operator ideally would be canonically conjugate to the Hamiltonian (as suggested
by the energy-time uncertainty relation), and a phase operator ideally would be canonically
conjugate to the number operator, the connection to periodic time is easily made.
In our toy model of harmonic oscillators (further specified below in section 4), we will
identify phase and time. This is also provides the easiest way to rigorously define time
travel, making use of Novikov’s self-consistency conjecture: A system comprised of clock
and dynamical system with respect to clock time undergoes time travel, if the dynamical
system state is the same after each full clock period. This definition could, in principle, be
relaxed. For example, this could happen only a finite number of times. An emergent notion
of time furthermore allows one to make such a statement merely approximate. This will
allow us to study rigorously notions of time travel that not fully fall into this framework
of self-consistency; for example, only parts of the system could undergo time travel. We
hope that such models including the clock itself can shed light on when a system is ‘merely
periodic in time’ versus when it actually undergoes time travel. In our final section 5, we
will further expound on this. Importantly, a requirement of self-consistency will preempt
any of the traditional paradoxa constructed out of it.

3.1. Time and Phase Operators for the Harmonic Oscillator


Leaving the mathematical formalities (mostly) aside, we will now introduce the specific
POVM used in our simple model of time travel. Formal, supplementary details regarding
the definition of POVMs can be found in appendix A. For the purposes of this section, we
will drop all label indices, and only look at the simple, standard harmonic oscillator with
Hamiltonian
1 1
Ĥ = ↠â + . (8)
2 2
7 of 20

We want to construct a time operator for this. In order to do so, we look at the polar
decomposition of the ladder operator â, [85, Thm. VIII.32]:

â = Ŵ |c
a |. (9)

Here, |c
a| is easy to define as √
a| := n̂1/2 =
|c ↠â. (10)
The ‘operator’ Ŵ, however, is less simple than its occurrence in a ‘polar decomposition’
might suggest: It will not be unitary. More concretely [22]:

Ŵ Ŵ † = 1, (11)

but
Ŵ † Ŵ = 1 − |0⟩⟨0| ̸= 1. (12)
The phase or time states to be used are now constructed as eigenstates |θ ⟩ of Ŵ.

Ŵ |θ ⟩ = eiθ |θ ⟩ . (13)

While these eigenstates do exist, given the properties of Ŵ, they are now only an overcomplete
set of eigenstates.3 So different eigenstates |θ ⟩ and |θ ′ ⟩ will not be orthogonal, and express-
ing a general state |ψ⟩ in terms of these eigenstates will not be unique anymore. Each of
these states |θ ⟩ can be written as

|θ ⟩ = ∑ einθ |n⟩ . (14)


n ≥0

Now is the time to define a concrete POVM. In this instance, let us define for all Borel
sets X of [0, 2π )

1
Z
B0 ( X ) := |θ ⟩⟨θ | dθ, (15)
2π X
1
Z
= ∑ 2π X
ei(n−m)θ |n⟩⟨m| dθ. (16)
n,m≥0

The index 0 is a first indication of what is to come: Different points on the unit circle S1
can be chosen as starting points. In the above definition, one starts on the positive real
axis, i.e., we choose our angle θ from [0, 2π ), as opposed to say from [θ∗ , θ∗ + 2π ) for some
θ∗ ∈ [0, 2π ).
Using
Z 2π
1
ei(n−m)θ dθ = δnm (17)
2π o
and

∑ |n⟩⟨n| = 1 (18)
n =0

one can show that


B0 ([0, 2π )) = 1, (19)
which means that B0 is a normalized POVM. So B0 represents an observable in the sense of
POVMs. Note that in the context of the definition of a POVM, in the present example we
have that Ω = S1 .
From the POVM side of things, this gives all ingredients needed for the calculations to
follow. However, for a more comprehensive understanding of the physics of time involved

3 Such overcomplete states find ample application in, for example, the context of coherent states [23].
8 of 20

in the harmonic oscillator, it is worthwhile to point out additional considerations concerning


such periodic time variables. We have collected these observations in appendix B.

4. Time Travel in Two Harmonic Oscillators Subject to a Hamiltonian Constraint


The goals for the present article are modest. We want to study as simple a model as
possible that is subject to a constraint reminiscent of the WDW equation (3). As we also
want to capture time travel as defined above, we need for this two ingredients: A periodic
clock system (H C , ĤC ) and a system (H R , ĤR ) that evolves in the sense of the PW formalism
(at least) as periodic as the clock itself. Then it would be an instance of a quantum system
fulfilling the Novikov self-consistency conjecture. Based on the discussion of harmonic
oscillators above, the simplest such system would be to use a single harmonic oscillator for
each ĤR and ĤC , respectively. To allow for some more generality, we will not demand their
respective frequencies ωR and ωC to be the same, a priori. Then, our model will have the
following Hamiltonian in the full Hilbert space H:

ωC ωR
Ĥ |Ψ⟩ = ωC n̂ C + 1C −ωR n̂ R − 1R |Ψ⟩ = 0. (20)
| {z 2 } | {z 2 }
ĤC − ĤR

This means that the system Hamiltonian ĤR and the clock Hamiltonian ĤC are part of
a WDW-like equation. Naturally, this implies that it falls into the framework of the PW
formalism for a stationary Hamilton equation Ĥ |Ψ⟩ = E |Ψ⟩ with energy E = 0.
The previous two sections gave us the two most important ingredients in place: First,
the way to measure time evolution with respect to a clock system using the PW formalism’s
conditional probability (6). Second, the required clock states given by equation (13) with
respect to which we will measure evolution, and its representation in terms of the harmonic
oscillator eigenstates, equation (14). Now, we can make use of it by implementing the
Hamiltonian constraint (20) into this model. What remains to be done, though, is to specify
more explicitly the type of total state |Ψ⟩ ∈ H the ‘universe’ (as described by Ĥ) is in.

4.1. Important Note on the Difference to Quantum Cosmology


Having introduced the model, it is easy to recognize it as a variant of a model already
used in the context of quantum cosmology [86–89]. There is, however, an important
distinction both mathematically and interpretationally. One of the goals of quantum
cosmology is to use symmetry reduction to arrive at minisuperspace models testing aspects
of quantum gravity. In the context of (quantum) cosmology, this is in particular often the
status of Hawking-type singularity theorems in quantum gravity: Can a cosmological
singularity be avoided or not [26,77,90,91]? In many, if not most situations this leads to the
demand that the configuration space R2 of the two harmonic oscillators is limited to the
half-plane H2 , as one of the variables corresponds to the non-negative scale factor a. The
scale factor at a = 0 corresponds to a gravitational singularity. Restricting one harmonic
oscillator to only a half-plane for its configuration space variable, however, inevitably leads
to the inability to use the very convenient, powerful, and simple algebraic methods in terms
of ladder operators used so far in the present paper.
Our present goals, however, are different. We want to study time travel in a quantum-
theoretic framework, but in a manner that is as theory-agnostic as possible. Hence, we will
simply not assume that equation (20) necessarily has its origins in a Hamiltonian formu-
lation of general relativity or similar gravitational theories. This means that equation (20)
with the domain used is an explicit input. It very well could be derived from a specific
gravitational model in a specific gravitational theory through symmetry-reduction. Should
this not be possible, it still may serve as a toy model. Especially, as the process of symmetry-
reduction (to arrive at minisuperspace models) itself may be a fraught undertaking. Our
assumption will be that such an equation with the full configuration space R2 is available,
and, likewise, so are our corresponding trusty ladder operators. This also bypasses issues
9 of 20

such as the precise nature of the inner product on the quantum gravitational wavefunctions,
as already discussed in [26, §9]. Two extremal options for inner products (with many in
between) are a quantum harmonic oscillator on the one side and a wave equation on the
other. We assume the inner product structure as the one familiar from and resulting in
standard quantum mechanics. Naturally, this leads to many directions into which to extend
our present model, and we will come back to these in our concluding section 5.
In conclusion and in conscious distinction to quantum cosmology, our toy model (20)
for time travel will be treated using a ladder operator approach of quantum mechanics.
This allows us to use the methods introduced in the previous two section, and hopefully
also provides easy extensions in diverse directions.

4.2. A Boring Model of Time Travel


Looking for a square-integrable solution to equation (20) introduces a new, additional
condition on this equation. Only if

ωR 2n C + 1
= , where n C , n R ∈ N0 , (21)
ωC 2n R + 1

will this be possible. This commensurability condition will have far-reaching consequences
for our simple toy model in what is to come.4 In fact, these results will (for this toy model,
at least) be so far reaching, that the precise nature of the time operators employed or their
respective states used as initial state of the clock becomes immaterial. The interested reader
can find more details on a possible choice of time operators useful for the current purposes
in appendix B.
Quite generally, the wave function of the ‘universe’, |Ψ⟩, will have the following form:

|Ψ⟩ = ∑′ An,n′ |n⟩C ⊗ |n′ ⟩R . (22)


n,n

Our calculation will proceed in three steps based on this general form.
1. The simplest possibility is a diagonal An,n′ . We will demonstrate that this leads, at
least according to the PW formalism, to the simplest time evolution possible. We will
demonstrate how this time evolution can be interpreted as a particularly boring case
of the Novikov self-consistency conjecture.
2. We derive a quite general statement about An,n′ when it is non-diagonal. Concretely,
introducing two sparse matrices ∆1 , ∆2 and a diagonal matrix D, it will take the form

A = ∆1T D∆2 . (23)

3. A general identity for the conditional probabilities (6) in our toy model is derived.
4. Using the previous two steps, we will prove that the results of the first step carry over
to non-diagonal An,n′ , too.
In the first step, let us calculate the components of equation (6). This will tell us
straightforwardly what we want to know.

tr ĤR P̂θ∗ ρ̂ = tr 1C ⊗ ĤR [(|ψ(θ∗ )⟩ C C ⟨ψ(θ∗ )|) ⊗ 1R ] [|Ψ⟩⟨Ψ|] ,



(24)
= Ψ [(|ψ(θ∗ )⟩C C ⟨ψ(θ∗ )|) ⊗ 1R ] 1C ⊗ ĤR Ψ .

(25)

4 Deviating from the commensurability condition even a little would lead to non-normalizable wave functions.
In the context of quantum cosmology such wave functions can be encountered, such as the no-boundary
proposal [77, p.282]. For the sake of our (not necessarily gravitational) toy model, we believe that this would
add another layer of interpretational problems best left for a separate study.
10 of 20

With the currently assumed diagonal form for An,n′ , we can evaluate

1
(1C ⊗ ĤR ) |Ψ⟩ = ∑ An n + |n⟩C ⊗ |n⟩R , (26)
n 2

which in turn—using R ⟨m|n⟩ R = δmn —turns equation (25) into



1
tr ĤR P̂θ∗ ρ̂ = ∑ | An | n +
2
| ⟨n|ψ(θ∗ )⟩C |2 .

(27)
n 2 C

Going back to the definition of |ψ(θ∗ )⟩ C , equation (4), we note that


1
C
⟨n|ψ(θ∗ )⟩C = e−i(n+ 2 )θ∗ C ⟨n|ψ(0∗ )⟩C . (28)

This finally leads to


|C ⟨n|ψ(θ∗ )⟩C |2 = |C ⟨n|ψ(0∗ )⟩C |2 , (29)
from which we can deduce that

tr ĤR P̂θ∗ ρ̂ = tr ĤR P̂0∗ ρ̂ , (30)

and, after a similar lead-up, likewise for the denominator of equation (6) that

tr P̂θ∗ ρ̂ = tr P̂0∗ ρ̂ . (31)

It follows immediately that


E( ĤR |θ∗ ) = E( ĤR |0∗ ). (32)
Put differently, the result (32) means that with respect to such diagonal states of the
‘universe’, no time will pass, irrespective of the chosen clock states. As was frequently
pointed out in the literature, much of the evolution is fixed by the choice of the full state
|Ψ⟩ ∈ H, so at this point this might just be an artefact of an overly specific type of states.
Still, even for these states it is interesting that changing clock states cannot change the
result. It is important to note that we make no claims concerning more general models;
this analysis heavily relies on the structure of our two-harmonic-oscillator model. We will
postpone the discussion of the implications to time travel for until after the more general
results of ‘step three’ for non-diagonal An,n′ .
Two examples of states |Ψ⟩ with diagonal An,n′ and isotropic oscillators (i.e., ωC =
ωR =: ω) are visualized in figure 1, with different choices for a dispersion parameter β. The
coordinates a, χ can in this case be assigned to either system (clock or residual), and were
chosen to allow easy comparison with the closely related examples of quantum cosmology
found in [87]. More details about the construction of the wave packets is relegated to
appendix C.
As a preliminary, second step for the discussion of non-diagonal An,n′ , it is instructive
to reexamine the commensurability condition. Plugging the ansatz (22) into the Hamilto-
nian constraint (20) we get for all values of n and n′ that
ωC ωR
ωC n + − ωR n′ − An,n′ = 0. (33)
2 2
This implies that either An,n′ = 0 or if An,n′ ̸= 0 that
ωC ωR
ωC n + − ωR n′ − = 0, (34)
2 2
which is equivalent to the original commensurability condition (21). Given that this condi-
tion yields a ratio of two odd numbers, we can rephrase n and n′ as

n = s(2p + 1) + p; n′ = s(2q + 1) + q; s, p, q ∈ N0 . (35)


11 of 20

(a) β = 1 (b) β = 0.03


Figure 1. Illustration of wave packets solving our toy model for time travel with ωC = ωR . Details
regarding the construction of these solutions to the Hamiltonian constraint equation (20) can be
found in appendix C. In comparison, quantum cosmological applications of this equation constrain
themselves to only a half-plane a ≥ 0. More general wave functions with ωC ̸= ωR will have less
symmetry than the pictured examples. Left: Dispersion-less packets, corresponding to a parameter
choice β = 1. Right: Wave packets with strong dispersion, β = 0.03.

This equation gives additional insight about the coefficients An,n′ . In components, we can
write

An,n′ = ∑ δn,s(2p+1)+ p ds δn′ ,s(2q+1)+q (36)
s =0

which in ‘matrix form’ appears as follows:

A = ∆1T D∆2 , (37)

where D really is diagonal and ∆1 , ∆2 are sparse matrices connecting s with n and n′ ,
respectively.
The third step takes a look at the numerator of the conditional probability of the PW
formalism, equation (6). Note that equation (25) is fully general, and needed no diagonality
of An,n′ in its derivation. Its last two factors can now be evaluated with our general
ansatz (22):
1
(1C ⊗ ĤR ) |Ψ⟩ = ∑ An,n′ n′ + |n⟩ C ⊗ |n′ ⟩ R . (38)
n,n ′ 2

As in the first step, we can then use orthonormality of the Fock states |n′ ⟩ C , R and cyclicity
of the trace to show that

1

tr ĤR P̂θ∗ ρ̂ = ∑ C ⟨m|ψ(θ∗ )⟩ C Am,n′
∗ ′

An,n′ n + ⟨ψ(θ∗ )|n⟩C . (39)
m,n,n′
2 C

Next, let us employ the definition of |ψ(θ∗ )⟩ C , equation (4):

C
⟨m|ψ(θ∗ )⟩C = C ⟨m| e−i ĤC θ∗ |ψ(0∗ )⟩C , (40)
−i (m+ 12 )θ∗
=e C
⟨m|ψ(0∗ )⟩C . (41)

This allows us to rearrange equation (39) to

∑ Qm,n e+i(n−m)θ∗

tr ĤR P̂θ∗ ρ̂ = (42)
m,n
12 of 20

where " # !
Qm,n := ∑′ A∗m,n′ An,n′ (n′ + 1/2) C
⟨m|ψ(0∗ )⟩C C ⟨ψ(0∗ )|n⟩C (43)
n

This effectively demonstrates how the choice of the universe’s state |Ψ⟩ and the chosen
initial clock state |ψ(0∗ )⟩ C together fully determine the time evolution of our model. Equa-
tion (43) can equivalenty be expressed in ‘matrix form’ as

Q = D1† A† D0 AD1 , (44)

where D1 is a complex diagonal ‘matrix’, and D0 a real diagonal ‘matrix’.


A closely related identity can also be found—using similar arguments—for the de-
nominator of equation (6), reading:
" # !
tr P̂θ∗ ρ̂ = ∑ ∑ Am,n′ An,n′ C ⟨m|ψ(0∗ )⟩C C ⟨ψ(0∗ )|n⟩C e
∗ +i ( n − m ) θ∗

. (45)
m,n n′

Note that this is again of the form (42), though with a simpler ‘Q’ for which D0 = 1. This
means any general statement that can be shown for Q will hold for the corresponding part
appearing in equation (45).
In our fourth and final step, we shall bring together our result (37) for An,n′ and
the result (44) for Qm,n to greatly constrain the conditional probabilities of equation (6).
Concretely, inserting the former into the latter gives:

Q = D1† (∆1T D∆2 )† D0 (∆1T D∆2 ) D1 , (46)


= D1† ∆2† D ∗ (∆1∗ D0 ∆1T ) D∆2 D1 . (47)

In this expression, let us investigate the central, bracketed term, ∆1∗ D0 ∆1T :

∆1∗ D0 ∆1T = ∑ δn′ ,s(2q+1)+q [ D0 ]n′ δn′ ,s′ (2q+1)+q , (48)
ss′
n′
=: ( D3/2 )ss′ , (49)

another diagonal matrix. This in turn means that even more terms in the middle of
equation (47) reduce to yet another diagonal matrix D2 , defined as

D2 := D ∗ (∆1∗ D0 ∆1T ) D = D ∗ D3/2 D. (50)

This, then, allows us to greatly simplify the structure of Q itself:

Q = D1† (∆2† D2 ∆2 ) D1 , (51)

whose middle part in brackets can, using the same approach as in equation (48), be shown
to be diagonal. Finally, it follows that Q itself is diagonal.
Essential for this result was the specific form of A, equation (37). Modifying the
underlying model will therefore likely change this result. For the PW formalism in our
context, though, the result is far-reaching: If we take equation (42), we see that for diagonal
Q, the term ei(n−m)θ∗ = 1, as n = m. The same observation holds for the denominator of
the PW conditional probabilities (6), equation (45). So, as in our first step when considering
diagonal An,n′ , even the general case produces trivial time evolution.
At first glance, this seems a very unremarkable result, and therefore of little use for any
interesting physics question. Ironically, this is not the case in our present situation, in which
we are discussing time travel: An evolution that is simply constant is certainly still trivially
a case of Novikov’s self-consistency conjecture. Given that nothing is happening, one
might even go so far as to say that it gives an odd entry to the intersection of the Novikov
13 of 20

self-consistency conjecture and the ‘boring physics conjecture’ [3]. It is simultaneously both
time travel and the absence of all evolution.
This result is also a marked departure from results in quantum cosmology [89,92]. The
reason for this disparity is the limited domain of the scale parameter in the models of the
form (20). Most importantly, limiting a harmonic oscillator to half of its configuration space
will prevent the application of ladder operators as we used them in the above analysis. This
again demonstrates how ‘simple’ changes to the model can quite radically change results.

5. Interpretation, Discussion and Outlook


Despite the disappointingly boring outcome for our particular model, our toy model
serves its purpose as a proof-of-concept study. It demonstrates the feasibility for future
model building of quantum systems with periodic, emergent notions of time as a tool
to study time travel. The obvious model-dependence of the PW formalism through its
dependence on |Ψ⟩, the ‘state of the universe’, means that different models might have
very different conclusions as to what sort of time evolution is possible. As even collections
of harmonic oscillators become increasingly complex once interactions start to enter, neces-
sitating many different heuristics or approaches across fields [23,93], it is unlikely that the
PW formalism can be fully general even when discussing whether time travel is possible
based on first principles of quantum theory.
In the absence of a fully general result, we still believe that some general conclusions
can nevertheless be achieved. Analytically tractable models beyond the one presented in
our study should be possible, especially given that POVMs and the PW formalism feature
prominently in the modern, gauge theoretic consolidation of emergent time in quantum
theory. The use of POVMs in future extensions was part of the motivation to highlight
them as much in our discussion as we did, despite the final result not relying on them.
The goals of such a future model building task would contain the following:
• More and different subsystems; more in the sense of more harmonic oscillators, differ-
ent in adding various combinations of other systems. Also ancillary systems could be
used. For example, in the context of time travel, one can ask the following question:
Can ancillary or memory systems allow for (more or less self-consistent) time travel
being limited to just a fixed, finite number of round trips as opposed to the infinite
traversability of GR’s CTCs?
• In a closely related step, one would introduce various interactions among the subsys-
tems. These will greatly change the outcome of our above analysis.
• Ultimately, the combination of these two model building drives can try to investig-
ate entropic arguments in the context of time travel. If self-consistent time travel in
quantum systems with an emergent notion of time is possible, are such configura-
tions potentially favoured or disfavoured on thermodynamic grounds? In particular,
entropy concepts geared towards constrained or fixed energy systems, like ‘observa-
tional entropy’ [94–96], promise access to new arguments for or against time travel
without having to rely on either mixing different concepts of time (GR’s dynamic time
vs. thermodynamic time vs. parameter time vs. . . . ) or, in fact, without relying on
any kind of explicit notion of time; the entropic arguments in such systems would
similarly be based on emergent notions, just as the notion of time itself is.
• Similarly, this exercise should be able to further clarify the difference between a
periodic clock time and time travel as such. Obviously, some kind of synchronicity
between residual system and clock system has to hold for the latter. Yet the above
mentioned research directions would allow one to turn concepts like self-consistency
into less of a binary. After all, if time is emergent, its absence might be ill-suited for
human survival but still allow a valid physical system that describes something ‘close’
to a quantum system with self-consistent, observable time travel. In a sense, this
would be a first step towards rigorously testing Hawking’s chronology protection.
• Much of the current research which led to a reimagining of the PW formalism as a
gauge fixed picture aims to ask what happens when one changes frames, i.e., clocks.
14 of 20

Once the model systems for time travel are complex enough (and, thus, more complex
than our equation (20)), a natural question is: What happens to time travel in such
quantum reference frame transformations? And, linking to the previous point, does
this differ from the situation for merely periodic clocks?
• Lastly, the previous two points hint at the possibility of making time travel more a
local, emergent notion in a large system. This is, for one, the explicit way to study
time travel classically as CTCs in a space-time manifold, see, for example, [59,81]. For
another, this is similarly done in quantum physics, ranging from the rigorously and
explicitly local [46,97] to the more imprecise and only implicitly local [45,98]. Yet, so
far, these approaches all had to rely on an ad-hoc introduction of ‘background’ time.
There are other avenues along similar lines that warrant a closer look in the framework
of quantum gravity. In the canonical approach to it, for instance, different inner products
can be considered, so at least changing the mathematical framework drastically—begging
the question how much of the physical interpretation changes or has to change in tandem.
Other approaches to quantum gravity besides the canonical ones can also test emergent
notions of time in their respective frameworks. Besides a question of tractability, this was
part of the motivation to keep our toy model somewhat agnostic and conservative with
respect to a theory of quantum gravity.
Independent of which direction one would want to take, we believe that our first
result shows that emergent notions of time in quantum physics can lead to a multitude of
questions and interpretational conundra pertaining to time travel. The very introduction
of such an emergent time concepts undermines most traditional arguments against time
travel. The concepts evoked in these arguments simply have to rely on notions that are not
valid in more ab initio approaches.
Funding: A.A.-S. is funded by the Deutsche Forschungsgemeinschaft (DFG, German Research
Foundation) — Project ID 516730869. This work was also partially supported by Spanish Project No.
MICINN PID2020-118159GB-C44. S.S. was financially supported by Czech Science Foundation grant
GACR 23-07457S. M.V. was directly supported by the Marsden Fund, via a grant administered by the
Royal Society of New Zealand.
Data Availability Statement: No new data were created or analyzed in this study. Data sharing is
not applicable to this article.
Acknowledgments: S.S. acknowledges support from the technical and administrative staff at the
Charles University. He thanks Emily Adlam, Antonis Antoniou, Julian Barbour, Martin Bojowald,
Leonardo Chataignier, Fabio Costa, Ricardo Faleiro, Klaus Fredenhagen, Philipp A. Höhn, Claus
Kiefer, Jorma Louko, Jessica Santiago, Alexander R. H. Smith, Reinhard Werner, Magdalena Zych, and
the participants of the 781st WE-Heraeus-Seminar ‘Time and Clocks’ in Bad Honnef, the Geometric
Foundations of Gravity 2023 conference in Tartu, the 13th RQI-North meeting in Chania, and the
Odborné soustředění ÚTF in Světlá pod Blaníkem for many valuable discussions and their interest.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no rôle in the design
of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or
in the decision to publish the results.

Abbreviations
The following abbreviations are used in this paper:

CTC Closed, time-like curves


GR General relativity
POVM Positive, operator-valued measure
PVM Projection-valued measure
PW Page–Wootters
WDW Wheeler–DeWitt
15 of 20

Appendix A. Formal Definition of POVMs


Let us present in this appendix the formal definition of a POVM. An operator  on a
Hilbert space H is called positive, if

∀ |ψ⟩ ∈ H : ⟨ψ| Â|ψ⟩ ≥ 0. (A1)

Positivity of an operator is often denoted as ‘ Â ≥ 0.’ An operator B is called a POVM if


and only if it is an additive map from a Borel σ-algebra A of subsets of a set Ω to operators
on H such that
(i) for all sets X ∈ A, B( X ) ≥ 0 (‘positivity’),
(ii) for disjoint X, Y ∈ A, B( X ∪ Y ) = B( X ) + B(Y ) (‘additivity’),
(iii) for all sequences of sets Xn ∈ A such that ∩∞ n=1 Xn = ∅, all operators B ( Xn ) have 0̂
(the zero operator) as greatest lower bound.
For more details, see [20, p.71, Def.4.5]. A POVM that fulfils B(Ω) = 1 is called normalized.
Normalized POVMs are a more general notion of observable compared to that of
observables being unbounded, self-adjoint operators on H. The (set-theoretic) ‘universe’
Ω corresponds to possible measurement outcomes. In this sense, additivity and being
normalized links measurements as described by POVMs to the theory of probablity as set
down in the Kolmogorov axioms. They allow for more and wider notions of imprecision
than those encoded in density matrices [20,23]. In ‘standard quantum mechanics,’ The
measurement process of an observable  according to the Born rule projects onto a subspace
formed by eigenvectors of  with set eigenvalue Ai . In the terminology of POVMs this
corresponds to ‘projection-valued measures’ (PVMs), and PVMs ⊊ POVMs.

Appendix B. Time Operators for a Harmonic Oscillator


In this appendix, we want to draw attention to some additional background regarding
POVMs and phase or time operators in the context of harmonic oscillators.
With the POVM B0 of equation (15) in place, we can now define a large number of
new, self-adjoint operators [24]. To do this, we associate to each real-valued, bounded, and
measurable function f on [0, 2π ) the operator B̂0 [ f ] defined as
Z 2π
1
B̂0 [ f ] := f (θ ) |θ ⟩⟨θ | dθ, (A2)
2π 0
Z 2π
1
= ∑ 2π 0
ei(n−m)θ f (θ ) |n⟩⟨m| dθ. (A3)
n,m≥0

This procedure fixes the set X in the original POVM B0 , and then uses it to construct
various Toeplitz operators. Different choices for f can be found in the literature (see, for
example, [83]), but we will focus on one of the simplest options:

f (θ ) = θ. (A4)

This will give us a time operator T̂0 , which we calculate with the above expressions to be

1
T̂0 = B̂0 [θ ] = ∑ i ( n − m)
|n⟩⟨m| + π1. (A5)
n ̸ = m ≥0

However, this operator is not unique. First of all, we could have chosen a different
starting point for our angles, changing the index 0 at that point. There is also a different,
but more physically motivated point of arriving at additional time operators. The time
operator T̂0 can be covariantly shifted to a new operator T̂θ∗ :

T̂θ∗ := ei Ĥθ∗ T̂0 e−i Ĥθ∗ . (A6)


16 of 20

This covariant shift property is a general feature of such time operators from POVMs, see [24],
that is increasingly prominent and clear in their modern implementations [37,99]. Import-
antly, while all such T̂θ∗ fulfil canonical commutation relations with Ĥ (when the resulting
operator is well-defined), different time operators will not commute with each other:

θ∗ ̸= θ∗′ mod 2π =⇒ [ T̂θ∗ T̂θ∗′ ] ̸= 0. (A7)

These results deserve some commentary. The reason why these time operators are
still self-adjoint, despite the illustrious list of no-go theorems given in the introduction,
is two-fold [24]: One, they are hardly unique. Any change in starting point (phase) will
yield a different time operator. A covariant shift similarly yields new time operators. Two,
the domain of time is now periodic. The no-go theorems usually aim for R as domain for
time θ, or that the time would be unique in some sense. In the present context, it is more a
question of when one started the periodic clock, and what period the clock has. It is also
worth pointing out that our time operators subtly fail to satisfy the expected time-energy
uncertainty: It will not hold for all states in H, but only for a densely defined set [83,84].
This latter restriction is not an issue for our purposes, at least in the present article, and
the other reasons for evading the no-go theorems are what allows our application in the
context of time travel.

Appendix C. Detailed Construction of Depicted Wave Packets


For our figure 1, we adapted the plots of [87] to our present context with a larger
configuration space. For ease of comparison, we kept the notation for the wave functions
the same. We do, however, emphasize that in our context one should not interpret a as
a scale factor. Our investigation is not quantum cosmology, just inspired by a particular
model of it due to its simplicity.
A general solution of equation (20) can be expressed in configuration space variables
a, χ instead of the Fock space picture adapted in the remainder of the article. For illustration
purposes, this is preferable. We will restrict attention to symmetric wavefunctions, in
which a and χ can be assigned equally well to the clock system ĤC or the residual system
ĤR . While our enlarged configuration space (compared to that of quantum cosmology)
might change the overall normalization, it will not change the qualitative behaviour of the
wavefunctions. We therefore also copy the normalization of reference [87].
A wavefunction with two Gaussians centered around χ = ±χ0 , each of width β, at
time/position a = 0, is given by

H2m ( a) H2m (χ) 1 2 1 2
Ψ( a, χ) = ∑ 2m 22m (2m)!H (0)
c p exp −
2
a −
2
χ , (A8)
m =0 2m

where Hn ( x ) is the n-th Hermite polynomial and the expansion coefficients are
s ! s !n !
1 2β χ20 1 − β2 χ0
cn = √ exp − Hn . (A9)
1 + β2 2(1 + β2 ) 1 + β2
p
2n n! 1 − β4

These coefficients cn simplify further for the ‘coherent case’, i.e., when β = 1, to

χn

1 2
cn = exp − χ0 √ 0 . (A10)
4 2n n!

Note that the index m in equation (A8) is meant to make more explicit that the sum runs
only over even n.

References
1. Roman, T.A. Inflating Lorentzian wormholes. Physical Review D 1993, 47, 1370–1379, [arXiv:gr-qc/9211012 [gr-qc]]. https:
//doi.org/10.1103/PhysRevD.47.1370.
17 of 20

2. Krasnikov, S. No time machines in classical general relativity. Classical and Quantum Gravity 2002, 19, 4109–4129, [arXiv:gr-
qc/0111054 [gr-qc]]. Erratum in Classical and Quantum Gravity 2014 31(7), 079503. https://doi.org/10.1088/0264-9381/31/7/079
503, https://doi.org/10.1088/0264-9381/19/15/316.
3. Visser, M. Lorentzian Wormholes: From Einstein to Hawking; AIP Series in Computational and Applied Mathematical Physics,
American Institute of Physics, 1996.
4. Hawking, S.W. Chronology protection conjecture. Physical Review D 1992, 46, 603–611. https://doi.org/10.1103/PhysRevD.46.603.
5. Klinkhammer, G. Vacuum polarization of scalar and spinor fields near closed null geodesics. Physical Review D 1992, 46, 3388–3394.
https://doi.org/10.1103/PhysRevD.46.3388.
6. Visser, M. From wormhole to time machine: Remarks on Hawking’s chronology protection conjecture. Physical Review D 1993,
47, 554–565, [arXiv:hep-th/9202090 [hep-th]]. https://doi.org/10.1103/PhysRevD.47.554.
7. Emparan, R.; Tomašević, M. Quantum backreaction on chronology horizons. Journal of High Energy Physics 2022, 02, 182,
[arXiv:2109.03611 [hep-th]]. https://doi.org/10.1007/JHEP02(2022)182.
8. Pons, J.M.; Sundermeyer, K.A.; Salisbury, D.C. Observables in classical canonical gravity: folklore demystified. Journal of Physics:
Conference Series 2010, 222, 012018, [arXiv:1001.2726 [gr-qc]]. https://doi.org/10.1088/1742-6596/222/1/012018.
9. Anderson, E. The Problem of Time; Vol. 190, Fundamental Theories of Physics, Springer, 2017. https://doi.org/10.1007/978-3-319-58
848-3.
10. Gioia, F.D.; Maniccia, G.; Montani, G.; Niedda, J. Non-Unitarity Problem in Quantum Gravity Corrections to Quantum Field
Theory with Born-Oppenheimer Approximation. Physical Review D 2021, 103, 103511, [arXiv:1912.09945 [gr-qc]]. https:
//doi.org/10.1103/PhysRevD.103.103511.
11. Kiefer, C.; Peter, P. Time in Quantum Cosmology. Universe 2022, 8, 36, [arXiv:2112.05788 [gr-qc]]. https://doi.org/10.3390/
universe8010036.
12. Mandelstam, L.I.; Tamm, I.Y. The uncertainty relation between energy and time in nonrelativistic quantum mechanics. Journal
of Physics 1945, IX, 249–254. English translation of Л. И. Мандельштам, И. Е. Тамм “Соотношение неопределённости
энергия-время внерелятивистской квантовой механике”, Изв. Дкад. Наук СССР (сер. физ.) 9, 122–128 (1945).
13. Schrödinger, E. Spezielle Relativitätstheorie und Quantenmechanik. Sitzungsberichte der Preußischen Akademie der Wissenschaften.
Physikalisch-mathematische Klasse 1931, pp. 238–247.
14. Pauli, W. Die allgemeinen Prinzipien der Wellenmechanik; Springer, 1990. Neu herausgegeben und mit historischen Anmerkungen
versehen von Norbert Straumann, https://doi.org/10.1007/978-3-642-62187-9.
15. Pauli, W. General Principles of Quantum Mechanics; Springer, 1980. English translation of [14], https://doi.org/10.1007/978-3-642-
61840-6.
16. Unruh, W.G.; Wald, R.M. Time and the interpretation of canonical quantum gravity. Physical Review D 1989, 40, 2598–2614.
https://doi.org/10.1103/PhysRevD.40.2598.
17. Böhm, A. Quantum Mechanics – Foundations and Applications, third ed.; Springer, 1993.
18. de la Madrid, R.; Böhm, A.; Gadella, M. Rigged Hilbert Space Treatment of Continuous Spectrum. Fortschritte der Physik 2002,
50, 185–216, [quant-ph/0109154]. https://doi.org/10.1002/1521-3978(200203)50:2<185::AID-PROP185>3.0.CO;2-S.
19. Thiemann, T. Modern Canonical Quantum General Relativity; Cambridge Monographs on Mathematical Physics, Cambridge
University Press, 2008.
20. Busch, P.; Lahti, P.; Pellonpää, J.P.; Ylinen, K. Quantum Measurement; Theoretical and Mathematical Physics, Springer, 2016.
21. Fewster, C.J.; Verch, R. Quantum Fields and Local Measurements. Communications in Mathematical Physics 2020, 378, 851–889,
[arXiv:1810.06512 [math-ph]]. https://doi.org/10.1007/s00220-020-03800-6.
22. Busch, P.; Grabowski, M.; Lahti, P.J. Who Is Afraid of POV Measures? Unified Approach to Quantum Phase Observables. Annals
of Physics 1995, 237, 1–11. https://doi.org/10.1006/aphy.1995.1001.
23. Auletta, G.; Fortunato, M.; Parisi, G. Quantum Mechanics; Cambridge University Press, 2009.
24. Busch, P.; Grabowski, M.; Lahti, P.J. Time observables in quantum theory. Physics Letters A 1994, 191, 357–361. https:
//doi.org/10.1016/0375-9601(94)90785-4.
25. Einstein, A. Zur Elektrodynamik bewegter Körper. Annalen der Physik 1905, 322, 891–921. https://doi.org/10.1002/andp.190532
21004.
26. DeWitt, B.S. Quantum Theory of Gravity. 1. The Canonical Theory. Physical Review 1967, 160, 1113–1148. https://doi.org/10.110
3/PhysRev.160.1113.
27. Page, D.N.; Wootters, W.K. Evolution without evolution: Dynamics described by stationary observables. Physical Review D 1983,
27, 2885–2892. https://doi.org/10.1103/PhysRevD.27.2885.
28. Wootters, W.K. “Time” replaced by quantum correlations. International Journal of Theoretical Physics 1984, 23, 701–711. https:
//doi.org/10.1007/BF02214098.
29. Kuchař, K.V. Time and Interpretations of Quantum Gravity. International Journal of Modern Physics D 2011, 20, 3–86. Contribution
to: 4th Canadian Conference on General Relativity and Relativistic Astrophysics, https://doi.org/10.1142/S0218271811019347.
30. Giovannetti, V.; Lloyd, S.; Maccone, L. Quantum Time. Physical Review D 2015, 92, 045033, [arXiv:1504.04215 [quant-ph]].
https://doi.org/10.1103/PhysRevD.92.045033.
31. Marletto, C.; Vedral, V. Evolution without evolution, and without ambiguities. Physical Review D 2017, 95, 043510,
[arXiv:1610.04773 [quant-ph]]. https://doi.org/10.1103/PhysRevD.95.043510.
18 of 20

32. Diaz, N.L.; Rossignoli, R. History state formalism for Dirac’s theory. Physical Review D 2019, 99, 045008, [arXiv:1806.01472
[quant-ph]]. https://doi.org/10.1103/PhysRevD.99.045008.
33. Diaz, N.L.; Matera, J.M.; Rossignoli, R. History state formalism for scalar particles. Physical Review D 2019, 100, 125020,
[arXiv:1910.04004 [quant-ph]]. https://doi.org/10.1103/PhysRevD.100.125020.
34. Diaz, N.L.; Matera, J.M.; Rossignoli, R. Spacetime quantum actions. Physical Review D 2021, 103, 065011, [arxiv:quant-
ph/2010.09136]. https://doi.org/10.1103/PhysRevD.103.065011.
35. Gemsheim, S.; Rost, J.M. Emergence of Time from Quantum Interaction with the Environment. Physical Review Letters 2023,
131, 140202, [arXiv:2309.05159 [quant-ph]]. https://doi.org/10.1103/PhysRevLett.131.140202.
36. Smith, A.R.H.; Ahmadi, M. Quantizing time: Interacting clocks and systems. Quantum 2019, 3, 160, [arXiv:1712.00081 [quant-ph]].
https://doi.org/10.22331/q-2019-07-08-160.
37. Höhn, P.A.; Smith, A.R.H.; Lock, M.P.E. The Trinity of Relational Quantum Dynamics. Physical Review D 2021, 104, 066001,
[arXiv:1912.00033 [quant-ph]]. https://doi.org/10.1103/PhysRevD.104.066001.
38. Höhn, P.A.; Smith, A.R.H.; Lock, M.P.E. Equivalence of Approaches to Relational Quantum Dynamics in Relativistic Settings.
Frontiers in Physics 2021, 9, 181, [arXiv:2007.00580 [gr-qc]]. https://doi.org/10.3389/fphy.2021.587083.
39. Krumm, M.; Höhn, P.A.; Müller, M.P. Quantum reference frame transformations as symmetries and the paradox of the third
particle. Quantum 2021, 5, 530, [arXiv:2011.01951 [quant-ph]]. https://doi.org/10.22331/q-2021-08-27-530.
40. Goeller, C.; Höhn, P.A.; Kirklin, J. Diffeomorphism-invariant observables and dynamical frames in gravity: reconciling bulk
locality with general covariance 2022. [arXiv:2206.01193 [hep-th]].
41. Adlam, E. Watching the Clocks: Interpreting the Page–Wootters Formalism and the Internal Quantum Reference Frame
Programme. Foundations of Physics 2022, 52, 99, [arXiv:2203.06755 [physics.hist-ph]]. https://doi.org/10.1007/s10701-022-00620-7.
42. Höhn, P.A.; Russo, A.; Smith, A.R.H. Matter relative to quantum hypersurfaces 2023. [arXiv:2308.12912 [quant-ph]].
43. Hausmann, L.; Schmidhuber, A.; Castro-Ruiz, E. Measurement events relative to temporal quantum reference frames 2023.
[arXiv:2308.10967 [quant-ph]].
44. Höhn, P.A.; Kotecha, I.; Mele, F.M. Quantum Frame Relativity of Subsystems, Correlations and Thermodynamics 2023.
[arXiv:2308.09131 [quant-ph]].
45. Deutsch, D. Quantum mechanics near closed timelike lines. Physical Review A 1991, 44, 3197–3217. https://doi.org/10.1103/
PhysRevD.44.3197.
46. Bishop, L.G.; Costa, F.; Ralph, T.C. Time-travelling billiard-ball clocks: a quantum model. Physical Review A 2021, 103, 042223,
[arXiv:2007.12677 [quant-ph]]. https://doi.org/10.1103/PhysRevA.103.042223.
47. Baumann, V.; Krumm, M.; Guérin, P.A.; Časlav Brukner. Noncausal Page–Wootters circuits. Physical Review Research 2022,
4, 013180, [arXiv:2105.02304 [quant-ph]]. https://doi.org/10.1103/PhysRevResearch.4.013180.
48. Vilasini, V. Approaches to causality and multi-agent paradoxes in non-classical theories. PhD thesis, University of York, 2021,
[arXiv:2102.02393 [quant-ph]].
49. Adlam, E. Two roads to retrocausality. Synthese 2022, 200, 422, [arXiv:2201.12934 [physics.hist-ph]]. https://doi.org/10.1007/s1
1229-022-03919-0.
50. Jannes, G. Condensed matter lessons about the origin of time. Foundations of Physics 2015, 45, 279–294, [arXiv:0904.3627 [gr-qc]].
https://doi.org/10.1007/s10701-014-9864-3.
51. Barceló, C.; Sánchez, J.E.; García-Moreno, G.; Jannes, G. Chronology protection implementation in analogue gravity. European
Physical Journal C 2022, 82, 299, [2201.11072]. https://doi.org/10.1140/epjc/s10052-022-10275-3.
52. Sabín, C. Analogue Non-Causal Null Curves and Chronology Protection in a dc-SQUID Array. Universe 2022, 8, 452, [arXiv:quant-
ph/2207.14164]. https://doi.org/10.3390/universe8090452.
53. Visser, M. The reliability horizon for semi-classical quantum gravity: Metric fluctuations are often more important than
back-reaction. Physics Letters B 1997, 415, 8–14, [arXiv:gr-qc/9702041 [gr-qc]]. https://doi.org/10.1016/S0370-2693(97)01226-4.
54. Zeh, H.D. The Physical Basis of the Direction of Time, fifth ed.; The Frontiers Collection, Springer, 2007. https://doi.org/10.1007/97
8-3-540-68001-7.
55. Truesdell, C. Tragicomedy of Classical Thermodynamics; Springer, 1971.
56. Minguzzi, E.; Sánchez, M. The causal hierarchy of spacetimes. In Recent developments in pseudo-Riemannian geometry; Baum, H.;
Alekseevsky, D., Eds.; EMS Publishing House, 2006; p. 299–358, [arXiv:gr-qc/0609119 [gr-qc]].
57. Schuster, S. Frenemies with Physicality: Manufacturing Manifold Metrics, [arXiv:2305.08725 [gr-qc]]. Essay written for the
Gravity Research Foundation 2023 Awards for Essays on Gravitation.
58. Barceló, C.; Visser, M. Twilight for the energy conditions? International Journal of Modern Physics D 2002, 11, 1553–1560,
[arXiv:gr-qc/0205066 [gr-qc]]. https://doi.org/10.1142/S0218271802002888.
59. Earman, J.; Smeenk, C.; Wüthrich, C. Do the laws of physics forbid the operation of time machines? Synthese 2009, 169, 91–124.
https://doi.org/10.1007/s11229-008-9338-2.
60. Manchak, J.B. On the Existence of “Time Machines” in General Relativity. Philosophy of Science 2009, 76, 1020–1026. https:
//doi.org/10.1086/605806.
61. Manchak, J.B. Some “No Hole” Spacetime Properties are Unstable. Foundations of Physics 2018, 48, 1539–1545. https://doi.org/10
.1007/s10701-018-0211-y.
19 of 20

62. Manchak, J.B. General Relativity as a Collection of Collections of Models. In Hajnal Andréka and István Németi on the Unity
of Science: From Computing to Relativity Theory Through Algebraic Logic; Madarász, J.; Székely, G., Eds.; Springer, 2021; Vol. 19,
Outstanding Contributions to Logic, chapter 19, pp. 409–425. https://doi.org/10.1007/978-3-030-64187-0_17.
63. Santiago, J.; Schuster, S.; Visser, M. Generic warp drives violate the null energy condition. Physical Review D 2022, 105, 064038,
[arXiv:2105.03079 [gr-qc]]. https://doi.org/10.1103/PhysRevD.105.064038.
64. González-Díaz, P.F.; Alonso-Serrano, A. Observing other universe through ringholes and Klein-bottle holes. Physical Review D
2011, 84, 023008, [arXiv:1102.3784 [astro-ph.CO]]. https://doi.org/10.1103/PhysRevD.84.023008.
65. Everett, A.E. Warp drive and causality. Physical Review D 1996, 53, 7365. https://doi.org/10.1103/PhysRevD.53.7365.
66. Gödel, K. An Example of a New Type of Cosmological Solutions of Einstein’s Field Equations of Gravitation. Review of Modern
Physics 1949, 21, 447–450. https://doi.org/10.1103/RevModPhys.21.447.
67. Hawking, S.W.; Ellis, G.F.R. The large scale structure of space-time; Cambridge Monographs on Mathematical Physics, Cambridge
University Press, 1974. https://doi.org/10.1017/CBO9780511524646.
68. Griffiths, J.B.; Podolský, J. Exact Space-Times in Einstein’s General Relativity, first paperback ed.; Cambridge Monographs on
Mathematical Physics, Cambridge University Press, 2012.
69. Gray, F.; Santiago, J.; Schuster, S.; Visser, M. “Twisted” black holes are unphysical. Modern Physics Letters A 2017, 32, 1771001,
[1610.06135]. https://doi.org/10.1142/S0217732317710018.
70. Wald, R.M. General Relativity; The University of Chicago Press, 1984.
71. Penrose, R. Techniques of Differential Topology in General Relativity; Society for Industrial & Applied Mathematics, 1972.
72. Krasnikov, S. Hyperfast Interstellar Travel in General Relativity. Physical Review D 1998, 57, 4760–4766, [arXiv:gr-qc/9511068
[gr-qc]]. https://doi.org/10.1103/PhysRevD.57.4760.
73. Manchak, J.B. No no-go: A remark on time machines. Studies in History and Philosophy of Modern Physics 2011, 42, 74–76.
https://doi.org/10.1016/j.shpsb.2011.01.001.
74. Novikov, I.D. Time machine and selfconsistent evolution in problems with selfinteraction. Physical Review D 1992, 45, 1989–1994.
https://doi.org/10.1103/PhysRevD.45.1989.
75. Sklar, L. How Free are Initial Conditions? In Proceedings of the Proceedings of the Biennial Meeting of the Philosophy of
Science Association Vol. 1990, Volume Two: Symposia and Invited Papers. University of Chicago Press, 1990, pp. 551–564.
https://doi.org/10.1086/psaprocbienmeetp.1990.2.193097.
76. Friedman, J.; Morris, M.S.; Novikov, I.D.; Echeverria, F.; Klinkhammer, G.; Thorne, K.S.; Yurtsever, U. Cauchy problem in
spacetimes with closed timelike curves. Physical Review D 1990, 42, 1915–1930. https://doi.org/10.1103/PhysRevD.42.1915.
77. Kiefer, C. Quantum Gravity, 3rd ed.; Vol. 136, International Series of Monographs on Physics, Oxford University Press, 2012.
78. Echeverria, F.; Klinkhammer, G.; Thorne, K.S. Billiard balls in wormhole space-times with closed timelike curves: Classical theory.
Physical Review D 1991, 44, 1077–1099. https://doi.org/10.1103/PhysRevD.44.1077.
79. Fewster, C.J.; Higuchi, A.; Wells, C.G. Classical and Quantum Initial Value Problems for Models of Chronology Violation. Physical
Review D 1996, 54, 3806, [arXiv:gr-qc/9603045 [gr-qc]]. https://doi.org/10.1103/PhysRevD.54.3806.
80. Bachelot, A. Global properties of the wave equation on non-globally hyperbolic manifolds. Journal de Mathématique Pures et
Appliquées 2002, 81, 35–65. https://doi.org/10.1016/S0021-7824(01)01229-6.
81. Dolanský, J.; Krtouš, P. Billiard in the space with a time machine. Physical Review D 2010, 82, 124056, [arXiv:1011.2881 [gr-qc]].
https://doi.org/10.1103/PhysRevD.82.124056.
82. Susskind, L.; Glogower, J. Quantum mechanical phase and time operator. Physics Physique Fizika 1964, 1, 49–61. https:
//doi.org/10.1103/PhysicsPhysiqueFizika.1.49.
83. Garrison, J.C.; Wong, J. Canonically Conjugate Pairs, Uncertainty Relations, and Phase Operators. Journal of Mathematical Physics
1970, 11, 2242–2249. https://doi.org/10.1063/1.1665388.
84. Galindo, A. Phase and Number. Letters in Mathematical Physics 1984, 8, 495–500. https://doi.org/10.1007/BF00400979.
85. Reed, M.; Simon, B. Methods of Modern Mathematical Physics: I. Functional Analysis, revised and enlarged ed.; Academic Press, 1980.
86. Kiefer, C. Non-minimally coupled scalar fields and the initial value problem in quantum gravity. Physics Letters B 1989,
225, 227–232. https://doi.org/10.1016/0370-2693(89)90810-1.
87. Kiefer, C. Wave packets in quantum cosmology and the cosmological constant. Nuclear Physics B 1990, 341, 273–293. https:
//doi.org/10.1016/0550-3213(90)90271-E.
88. Page, D.N. Minisuperspaces with conformally and minimally coupled scalar fields. Journal of Mathematical Physics 1991,
32, 3427–3438. https://doi.org/10.1063/1.529457.
89. Brunetti, R.; Fredenhagen, K.; Hoge, M. Time in Quantum Physics: From an External Parameter to an Intrinsic Observable.
Foundation of Physics 2010, 40, 1368–1378, [arXiv:0909.1899 [math-ph]]. https://doi.org/10.1007/s10701-009-9400-z.
90. Hawking, S.W.; Ellis, G.F.R. Singularities in homogeneous world models. Physics Letters 1965, 17, 246–247. Typo in first author of
published version., https://doi.org/10.1016/0031-9163(65)90510-X.
91. O’Neill, B. Semi-Riemannian Geometry with Applications to Relativity; Pure and Applied Mathematics, Academic Press, 1983.
92. Conradi, H.D.; Zeh, H.D. Quantum cosmology as an initial value problem. Physics Letters A 1991, 154, 321–326. https:
//doi.org/10.1016/0375-9601(91)90026-5.
93. Zee, A. Quantum Field Theory in a Nutshell, second ed.; Princeton University Press, 2010.
20 of 20

94. Šafránek, D.; Deutsch, J.M.; Aguirre, A. Quantum coarse-grained entropy and thermodynamics. Physical Review A 2019,
99, 010101, [arXiv:1707.09722 [quant-ph]]. https://doi.org/10.1103/PhysRevA.99.010101.
95. Šafránek, D.; Deutsch, J.M.; Aguirre, A. Quantum coarse-grained entropy and thermalization in closed systems. Physical Review
A 2019, 99, 012103, [arXiv:1803.00665 [quant-ph]]. https://doi.org/10.1103/PhysRevA.99.012103.
96. Šafránek, D.; Aguirre, A.; Schindler, J.; Deutsch, J.M. A Brief Introduction to Observational Entropy. Foundations of Physics 2021,
51, 101, [arXiv:2008.04409 [quant-ph]]. https://doi.org/10.1007/s10701-021-00498-x.
97. Alonso-Serrano, A.; Tjoa, E.; Garay, L.J.; Martín-Martínez, E. The time traveler’s guide to the quantization of zero modes. Journal
of High Energy Physics 2021, 12, 170, [arXiv:2108.07274 [quant-ph]]. https://doi.org/10.1007/JHEP12(2021)170.
98. Lloyd, S.; Maccone, L.; Garcia-Patron, R.; Giovannetti, V.; Shikano, Y. Quantum mechanics of time travel through post-selected
teleportation. Physical Review D 2011, 84, 025007, [arXiv:1007.2615 [quant-ph]]. https://doi.org/10.1103/PhysRevD.84.025007.
99. Smith, A.R.S. Communicating without shared reference frames. Physical Review A 2019, 99, 052315, [arXiv:1812.08053 [quant-ph]].
https://doi.org/10.1103/PhysRevA.99.052315.

You might also like