Skip to main content
Log in

Watching the Clocks: Interpreting the Page–Wootters Formalism and the Internal Quantum Reference Frame Programme

  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

We discuss some difficulties that arise in attempting to interpret the Page–Wootters and Internal Quantum Reference Frames formalisms, then use a ‘final measurement’ approach to demonstrate that there is a workable single-world realist interpretation for these formalisms. We note that it is necessary to adopt some interpretation before we can determine if the ‘reference frames’ invoked in these approaches are operationally meaningful, and we argue that without a clear operational interpretation, such reference frames might not be suitable to define an equivalence principle. We argue that the notion of superposition should take into account the way in which an instantaneous state is embedded in ongoing dynamical evolution, and this leads to a more nuanced way of thinking about the relativity of superposition in these approaches. We conclude that typically the operational content of these approaches appears only in the limit as the size of at least one reference system becomes large, and therefore these formalisms have an important role to play in showing how our macroscopic reference frames can emerge out of wholly relational facts.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Similar content being viewed by others

Data Availability

Data sharing not applicable to this article as no datasets were generated or analysed during the current study.

Notes

  1. We also note that even if one is satisfied by the Everett account of probabilities, problems remain for an Everettian account of the PW formalism. It would still be necessary to add something to the PW picture representing some kind of interaction in which clocks are actually observed, and it would still be necessary to explain how to pair measurements on systems with measurements on clocks in order to calculate correlations - the standard Everett interpretation assumes we have unitary evolution on a background spacetime, so in an Everettian version of the IQRF formalism we would have to take care to avoid illegitimately appealing to a background time metric which is no longer available. We will not discuss here how these problems might be solved in the Everett picture, but we would invite proponents of the Everett interpretation to consider the question.

  2. Of course, one possible way of responding to this problem is to move to a version of the framework which postulates neither probabilities nor relative frequencies - i.e. perhaps regard the framework as simply a description of the universal wavefunction, understood in something like an Everettian picture with no attempt to attach probabilities to branches. However, it would seem very hard to understand the physical meaning of the framework’s predictions in the absence of any probabilities or relative frequencies, and without predictions which can be linked to observations it would seem impossible for us to obtain evidence for or learn anything about the putative ‘universal wavefunction’ in the first place, so this approach would lead to serious epistemic problems similar to those which have been raised for more vanilla versions of the Everett interpretation [32]. Also, as noted above, we do need either probabilities or relative frequencies if we are to have correlations in the usual mathematical sense, so if the physical meaning of time in these frameworks is supposed to come down to correlations, then we don’t have the option of simply refraining from ascribing probabilities.

  3. We note that the use of a delta function in this kind of expression, while common in the literature, is technically problematic: trying to extract relational variables from such a constraint requires us to combine a distribution with a function in a way that is not generally well-defined. Thus it seems likely that in fact the delta functions should be replaced with something like a Gaussian, sharply peaked around the time \(t_m\). This is in any case more physically realistic, as the delta function is being used here to indicate that an event occurs when a clock reads a certain value, but since clocks generally cannot be read with infinite precision, any real implementation would necessary have some spread in the exact time of the event. However, since we will not be solving the constraints here we will continue using delta functions for simplicity.

  4. Note that in order to adopt this strategy we do not need to assume that all the clocks in the universe have some prima facie orientation which distinguishes between the two possible directions of time, because if we take it that there is some beginning of time, as indicated by our current understanding of cosmology, it follows that the readings on the clocks can go to infinity in only one of the two possible temporal directions, so the limit \(t \rightarrow \infty \) will necessarily single out the same temporal direction for all the clocks. We do have to assume that the set of all clocks is at least well enough coordinated that there are two distinguishable time orientations; but clearly in order for the PW approach to be empirically adequate it must reproduce the dimensionality of spacetime at least at an emergent level, so this assumption is reasonable.

  5. We reinforce that decoherence alone is not enough to explain the emergence of a single definite macroreality; decoherence makes density matrices diagonal, thus suppressing interference effects, but the measurement mechanism is still needed to select one of the diagonal elements so we get a unique macroscopic history.

  6. It’s worth nothing that there has been considerable debate about whether the SEP is universally true - it has been pointed out that tidal effects may spoil the Lorentz invariance unless the ‘region’ is just a point [49], and more recently it has been noted that minimal coupling may also lead to apparent violations of the equivalence principle [50]; the fact that the SEP can apparently be violated even in the classical regime casts some doubt on the idea that we should expect it to be universally true in the quantum regime.

  7. We reinforce that although it is particularly common for proponents of the dynamical perspective on relativity [53] to emphasize the role of the SEP in defining chronogeometric behaviour, this understanding of the SEP isn’t predicated on acceptance of the dynamical approach: the SEP plays a very similar operational role in the competing geometric approach [55]. If we adopt the dynamical perspective, we would say that the equivalence principle is true because the dynamics of matter fields locally exhibit the same symmetries as the metric; whereas if we adopt the geometric perspective, we would say that the equivalence principle is a consequence of the fact that the metric constrains the matter fields to be invariant with respect to its own local symmetries: either way, it is clear that the equivalence principle is important in large part because it encodes a practical connection between the metric and the dynamics of matter. The two approaches differ on the direction of explanation, but agree that the SEP is significant because of this practical connection.

References

  1. Smith, A.R.H., Ahmadi, M.: Quantum clocks observe classical and quantum time dilation. Nat. Commun. 11, 1 (2020)

    Article  ADS  Google Scholar 

  2. Hellmann, F., Mondragon, M., Perez, A., Rovelli, C.: Multiple-event probability in general-relativistic quantum mechanics. Phys. Rev. D 75, 8 (2007)

    Article  MathSciNet  Google Scholar 

  3. Kent, A.: Solution to the Lorentzian quantum reality problem. Phys. Rev. A 90(1), 012107 (2014)

    Article  ADS  MathSciNet  Google Scholar 

  4. Kent, A.: Lorentzian quantum reality: postulates and toy models. Philos. Trans. R. Soc. A 373(2047), 20140241 (2015)

    Article  ADS  Google Scholar 

  5. Giacomini, F., Castro-Ruiz, E., Brukner, Č: Quantum mechanics and the covariance of physical laws in quantum reference frames. Nat. Commun. 10, 1 (2019)

    Article  Google Scholar 

  6. Castro-Ruiz, E., Giacomini, F., Belenchia, A., Brukner, Č: Quantum clocks and the temporal localisability of events in the presence of gravitating quantum systems. Nat. Commun. 11, 1 (2020)

    Google Scholar 

  7. Vanrietvelde, A., Hoehn, P.A., Giacomini, F., Castro-Ruiz, E.: A change of perspective: switching quantum reference frames via a perspective-neutral framework. Quantum 4, 225 (2020)

    Article  Google Scholar 

  8. Höhn, P.A., Vanrietvelde, A.: How to switch between relational quantum clocks. New J. Phys. 22(12), 123048 (2020)

    Article  ADS  MathSciNet  Google Scholar 

  9. Giacomini, F., Brukner, Č: Quantum superposition of spacetimes obeys einstein’s equivalence principle. AVS Quant. Sci. 4(1), 015601 (2022)

    Article  ADS  Google Scholar 

  10. Aharonov, Y., Kaufherr, T.: Quantum frames of reference. Phys. Rev. D 30, 368–385 (1984)

    Article  ADS  MathSciNet  Google Scholar 

  11. Aharonov, Y., Susskind, L.: Charge superselection rule. Phys. Rev. 155, 1428–1431 (1967)

    Article  ADS  Google Scholar 

  12. Superselection Rules, chapter 11, pp. 149–159. Wiley, New York (2005)

  13. Dirac, P.A.M.: The Principles of Quantum Mechanics. Oxford University Press, Oxford (1930)

    MATH  Google Scholar 

  14. Kuchař, Karel V.: Canonical quantum gravity, arXiv General Relativity and Quantum Cosmology (1993)

  15. Rovelli, C.: The strange equation of quantum gravity. Class. Quant. Grav. 32(12), 124005 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  16. Hoehn, P.A., Lock, M.P.E., Ahmad, S.A., Smith, A.R.H., Galley, T.D.: Quantum relativity of subsystems (2022)

  17. Krumm, M., Höhn, P.A., Müller, M.P.: Quantum reference frame transformations as symmetries and the paradox of the third particle. Quantum 5, 530 (2021)

    Article  Google Scholar 

  18. de la Hamette, A.-C., Galley, T.D., Hoehn, P.A., Loveridge, L., Mueller, M.P.: Perspective-neutral approach to quantum frame covariance for general symmetry groups (2021)

  19. Höhn, P.: Switching internal times and a new perspective on the “wave function of the universe.” Universe 5(5), 116 (2019)

    Article  ADS  Google Scholar 

  20. Baumann, V., Santo, F.D., Smith, A.R.H., Giacomini, F., Castro-Ruiz, E., Brukner, C.: Generalized probability rules from a timeless formulation of Wigner’s friend scenarios. Quantum 5, 524 (2021)

    Article  Google Scholar 

  21. Page, D.N., Wootters, W.K.: Evolution without evolution: Dynamics described by stationary observables. Phys. Rev. D 27, 2885–2892 (1983)

    Article  ADS  Google Scholar 

  22. Isham, C.: Canonical quantum gravity and the problem of time. Integrable Systems, Quantum Groups, and Quantum Field Theories, 11 (1992)

  23. KUCHAŘ, K.V.: Time and interpretations of quantum gravity. Int. J. Mod. Phys. D 20(supp01), 3–86 (2011)

  24. Moreva, E., Brida, G., Gramegna, M., Giovannetti, V., Maccone, L., Genovese, M.: Time from quantum entanglement: An experimental illustration. Phys. Rev. A 89, 5 (2014)

    Article  Google Scholar 

  25. Giovannetti, V., Lloyd, S., Maccone, L.: Quantum time. Phys. Rev. D 92, 045033 (2015)

    Article  ADS  MathSciNet  Google Scholar 

  26. Marletto, C., Vedral, V.: Evolution without evolution and without ambiguities. Phys. Rev. D 95, 043510 (2017)

    Article  ADS  Google Scholar 

  27. Höhn, P., Smith, A., Lock, M.: Trinity of relational quantum dynamics. Phys. Rev. D 104, 6 (2021)

    Article  MathSciNet  Google Scholar 

  28. Poulin, D.: Toy model for a relational formulation of quantum theory. Int. J. Theoret. Phys. 45(7), 1189–1215 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  29. Vedral, V.: Classical evolution without evolution (2022)

  30. Smith, A.R.H., Ahmadi, M.: Quantizing time: Interacting clocks and systems. Quantum 3, 160 (2019)

    Article  Google Scholar 

  31. Wallace, D.: Everett and structure. Stud. Hist. Philos. Sci. B 34(1), 87–105 (2003)

    MathSciNet  MATH  Google Scholar 

  32. Adlam, E.: The Problem of Confirmation in the Everett Interpretation. Stud. Hist. Philos. Sci. Part B 47, 21–32 (2014)

    MathSciNet  MATH  Google Scholar 

  33. Albert, D.: Probability in the everett picture. In: Saunders, S., Barrett, J., Kent, A., Wallace, D. (eds.) Many Worlds? Everett. Quantum Theory & Reality. Oxford University Press, Oxford (2010)

    Google Scholar 

  34. Kent, A.: One world versus many: the inadequacy of Everettian accounts of evolution, probability, and scientific confirmation (2009)

  35. Dolby, C.E.: The conditional probability interpretation of the Hamiltonian constraint. arXiv General Relativity and Quantum Cosmology (2004)

  36. Craig, D., Singh, P.: Consistent histories in quantum cosmology. Found. Phys. 41(3), 371–379 (2011)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  37. Callender, C., Weingard, R.: The bohmian model of quantum cosmology. PSA 218–227, 1994 (1994)

    MATH  Google Scholar 

  38. Dowker, F., Kent, A.: On the consistent histories approach to quantum mechanics. J. Stat. Phys. 82(5–6), 1575–1646 (1996)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  39. Kent, A.: Quantum reality via late-time photodetection. Phys. Rev. A 96, 6 (2017)

    Article  Google Scholar 

  40. Rovelli, C.: Relational quantum mechanics. Int. J. Theoret. Phys. 35, 1637–1678 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  41. Dürr, D., Goldstein, S., Zanghì, N.: Quantum Spacetime without Observers: Ontological Clarity and the Conceptual Foundations of Quantum Gravity, pp. 247–261. Springer, Berlin (2013)

  42. Einstein, A.: On the electrodynamics of moving bodies. Ann. Phys. 17, 891–921 (1905)

    Article  MATH  Google Scholar 

  43. Scarani, V., Iblisdir, S., Gisin, N., Acín, A.: Quantum cloning. Reviews of Modern Physics 77, 1225–1256 (2005)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  44. Chang, H.: Operationalism. In Edward N. Zalta, editor, The Stanford Encyclopedia of Philosophy. Metaphysics Research Lab, Stanford University, Fall 2021 edition (2021)

  45. Hoehn, P.A., Krumm, M., Mueller, M.P.: Internal quantum reference frames for finite abelian groups (2021)

  46. Giacomini, F., Brukner, Č.: Einstein’s equivalence principle for superpositions of gravitational fields (2021)

  47. Okon, E., Callender, C.: Does quantum mechanics clash with the equivalence principle-and does it matter? Eur. J. Philos. Sci. 1(1), 133–145 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  48. Lehmkuhl, D.: The equivalence principle(s) (2019)

  49. Synge, J.L.: Relativity: The General Theory. Number v. 1 in North-Holland series in physics. North-Holland Publishing Company (1960)

  50. Read, J., Brown, H.R., Lehmkuhl, D.: Two miracles of general relativity,: Forthcoming in Stud. Hist. Philos. Mod. Phys. (2018)

  51. Einstein, A.: Physical relativity space-time structure from a dynamical perspective (1920)

  52. Einstein, A.: Letter to Painlevé (1921)

  53. Brown, H.R.: Fundamental Ideas and Methods of the Theory of Relativity. Presented in Their Development. Oxford University Press, Oxford (2005)

    Google Scholar 

  54. Knox, E.: Effective spacetime geometry. Stud. Hist. Philos. Sci. Part B 44(3), 346–356 (2013)

    MathSciNet  MATH  Google Scholar 

  55. Maudlin, T.: Philosophy of Physics: Space and Time. Princeton Foundations of Contemporary Philosophy. Princeton University Press, Princeton (2012)

    Book  MATH  Google Scholar 

  56. Cepollaro, C., Giacomini, F.: Quantum generalisation of Einstein’s equivalence principle can be verified with entangled clocks as quantum reference frames (2021)

  57. Albert, D.Z.: Quantum Mechanics and Experience. Harvard University Press, New York (1994)

    Book  Google Scholar 

  58. Miyadera, T., Loveridge, L., Busch, P.: Approximating relational observables by absolute quantities: a quantum accuracy-size trade-off. J. Phys. A 49(18), 185301 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  59. Loveridge, L.: A relational perspective on the wigner-araki-yanase theorem. J. Phys. 1638(1), 012009 (2020)

    Google Scholar 

  60. Hardy, L.: Implementation of the quantum equivalence principle (2019)

  61. Kretschmann, E.: Über den physikalischen sinn der relativitätspostulate. a. einsteins neue und seine ursprüngliche relativitätstheorie. Ann. Phys. 1, 575–614 (1917)

    MATH  Google Scholar 

  62. Kochen, S., Specker, E.P.: The problem of hidden variables in quantum mechanics. In: Hooker, C.A. (ed.) The Logico-Algebraic Approach to Quantum Mechanics, The University of Western Ontario Series in Philosophy of Science, pp. 293–328. Springer, Dordrecht (1975)

    Chapter  Google Scholar 

  63. Spekkens, R.W.: Contextuality for preparations, transformations, and unsharp measurements. Phys. Rev. A 71(5), 052108 (2005)

    Article  ADS  Google Scholar 

  64. Holland, P.R.: The Quantum Theory of Motion: An Account of the de Broglie-Bohm Causal Interpretation of Quantum Mechanics. Cambridge University Press (1995)

  65. Tumulka, R.: A relativistic grw flash process with interaction (2020)

  66. Penrose, R.: On gravity’s role in quantum state reduction. Gen. Relat. Gravit. 28, 581–600 (1996)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  67. Bacciagaluppi, G.: The Role of Decoherence in Quantum Mechanics. In: E.N. Zalta, ed., The Stanford Encyclopedia of Philosophy. Metaphysics Research Lab, Stanford University, Fall 2020 edition (2020)

  68. Anderson, E.: Geometrodynamics: Spacetime or space ? arXiv General Relativity and Quantum Cosmology (2004)

  69. Rovelli, C.: Loop quantum gravity. Living Rev. Relat. 11(1), 5 (2008)

    Article  ADS  Google Scholar 

  70. Blau, M., Theisen, S.: String theory as a theory of quantum gravity: a status report. Gen. Relat. Grav. 41(4), 743–755 (2009)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  71. Gambini, R., Porto, R.A., Pullin, J., Torterolo, S.: Conditional probabilities with Dirac observables and the problem of time in quantum gravity. Phys. Rev. D 79, 041501 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  72. Bartlett, S.D., Rudolph, T., Spekkens, R.W.: Reference frames, superselection rules, and quantum information. Rev. Mod. Phys. 79, 555–609 (2007)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  73. Smith, A.R.H.: Communicating without shared reference frames. Phys. Rev. A 99, 052315 (2019)

    Article  ADS  Google Scholar 

  74. Gour, G., Spekkens, R.W.: The resource theory of quantum reference frames: manipulations and monotones. New J. Phys. 10(3), 033023 (2008)

    Article  ADS  Google Scholar 

  75. Frauchiger, D., Renner, R.: Quantum theory cannot consistently describe the use of itself. Nat. Commun. 9, 1 (2018)

    Article  Google Scholar 

  76. Kastner, R.E.: Unitary-only quantum theory cannot consistently describe the use of itself: On the frauchiger-renner paradox. Found. Phys. 50(5), 441–456 (2020)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  77. Pikovski, I., Zych, M., Costa, F., Brukner, C.: Time dilation in quantum systems and decoherence: questions and answers (2015)

  78. Palmer, M.C., Girelli, F., Bartlett, S.D.: Changing quantum reference frames. Phys. Rev. A 89(5), 052121 (2014)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

Thanks to Flaminia Giacomini and Philipp Hoehn for their very helpful comments on a draft of this paper. This publication was made possible through the support of the ID 61466 grant from the John Templeton Foun- dation, as part of the “The Quantum Information Structure of Spacetime (QISS)” Project (qiss.fr). The opinions expressed in this publication are those of the author and do not necessarily reflect the views of the John Templeton Foundation.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Emily Adlam.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Other Relational Formalisms

In addition to the PW formalism, there are several other formal frameworks which enable us to obtain relative states from a global state in a similar manner. One approach uses relational Dirac observables [71], which encode how one observable evolves relative to another along the flow generated by the Hamiltonian constraint. Relational Dirac observables are often invoked in the context of classical Hamiltonian approaches, but we can obtain corresponding variables in the quantum case by first constructing variables in the kinematical Hilbert space which describe some system property conditional on some reference frame being in some orientation, and then mapping them to Dirac observables on the physical Hilbert space using a G-twirl [18]. We also have the relational Heisenberg picture [27], where we begin by trivializing the constraint so it acts only on the degrees of freedom of the reference frame (e.g. the clock) - that is, if T is the trivialization map and C the Hamiltonian constraint operator, we have \(T H T^{-1} = (H_C - \epsilon ) \otimes I_S\) for any real \(\epsilon \). This transforms physical states into product states with a fixed redundant clock factor, and we then project onto the classical gauge-fixing condition to remove the redundant reference frame degrees of freedom. These various approaches differ in their technical details but all present a roughly similar picture in which temporal notions are understood in terms of states of the universe relative to some physical reference frame. Indeed ref [27] demonstrates that the PW formalism, the relational Dirac observable formalism, and the relational Heisenberg formalism are all in fact equivalent. Ref [27] also demonstrates that in all three cases, the procedure for changing temporal reference frames amounts to transforming from a state relative to some system A to a global state which encodes all the possible relative states and then finally applying a reduction procedure to arrive at the state relative to some other system B. This demonstrates the vital importance of the global perspective-neutral picture over and above all the individual relative-state pictures, and indeed ref [19] argues that it offers a new way of interpreting the notion of the ‘quantum state of the universe’ as an encoding of all these relative states.

We should also distinguish between the IQRF approach and the related reference frame research programme in quantum information [72,73,74]. In the QI context, reference frames typically come into play in the operational setting where we have a pair of observers who wish to exchange information but who do not share any common reference frame, which is to say we have no guarantee that they will perform measurements in the same basis. The solution is to encode information into relational degrees of freedom- for example, we could encode a bit into two qubits with 0 associated with the state \(\frac{1}{\sqrt{2}} (| 00 \rangle + |11 \rangle ) \) and 1 associated with the state \(\frac{1}{\sqrt{2}} (| 01 \rangle + |10 \rangle ) \), because the answer to the question ‘are the states of the two qubits the same or different?’ will be the same regardless of what basis we measure in. The key difference between these two approaches is that the QI formalism employs observers who are external to the quantum system in question, rather than seeking to define reference frames from within the system as the IQRF programme does: hence in the QI approach the set of states which are regarded as being legitimate in the absence of an external reference frames are the incoherently group-averaged states, rather than the coherently group-averaged states as in the IQRF approach, reflecting the fact that the QI approach uses external observers who cannot themselves be entangled with parts of the system in question [17]. The QI approach therefore poses no particular interpretational challenges over and above the usual measurement problem, as all interpretations of quantum mechanics will make the same predictions for the results of the measurements performed by these external observers. Whereas the correct way to model observers as internal features of a quantum system remains highly controversial [75, 76], and therefore the internal IQRF approach lies right at the meeting point of a number of thorny conceptual issues in quantum foundations. We therefore focus in this article on the internal reference frame programme rather than the related QI approach.

Appendix B: Final-Measurement Interpretation

The final-measurement interpretation allows us to say more about the notion that time is just gauge. It is still true in this picture that specifying a state at a time or a single relative state is enough to determine the whole universal quantum state, so in that sense it remains true that time is a form of gauge, at least with respect to the universal quantum state. But in the final-measurement picture the relative states don’t really refer to anything physically real, so the fact that they are all related by gauge transformations may seem less surprising: we simply have a choice of gauge in which to express the probability rule which is used to determine the appropriate probability distribution for the final measurement. On the other hand, if we now look at the actual course of history selected by this measurement, we will see apparent indeterminism corresponding to the probabilistic results of quantum measurements, so it is not true for this actual course of history that the state at one time is sufficient to determine the state at all other times. Therefore time is not merely gauge for the real timeline as defined by the register readings - genuinely new information comes into being at different times along this timeline. This suggests that the problem of time in quantum gravity is to some extent an artefact of an approach which reifies the universal quantum state whilst sweeping the measurement problem under the rug: if one includes some actual measurement results the picture looks very different and there is less reason to be concerned about ‘timelessness.’ Note that this consequence is not unique to the final-measurement interpretation: as observed by Callender and Weingard [37], ‘contrary to the static situation described at the outset, a Bohmian approach to cosmology admits nontrivial evolution of the dynamical variables.’ So it seems that quite generally, as soon as we we add to the universal quantum state some variables describing what actually occurs, we no longer have the problem of timelessness. Indeed, from the point of view of approaches like the final-measurement interpretation and the Bohmian interpretation, the universal wavefunction is in a sense simply describing a set of counterfactual possibilities, and thus its timelessness is not at all surprising - for presumably one would not expect a set of counterfactual possibilities to undergo change or evolution in any case.

Given that the final measurement on the register will necessarily take the state outside the physical Hilbert space, one might perhaps have concerns that this interpretation would break the symmetries which have been used to set up the physical Hilbert space. And indeed, it is true that if we define a post-measurement state in the usual way this state will no longer belong to the physical Hilbert space; but by definition the final-measurement occurs after all meaningful dynamical evolution is complete, so we are not required to actually perform a projection into a post-measurement state since nothing happens after the final measurement. The constraint quantisation ensures that the probability distribution defined over possible register readings by the quantum state is invariant under diffeomorphisms and global time translations, and therefore the process of selecting one set of register readings according to that probability distribution is also necessarily invariant under diffeomorphisms and global time translations. Moreover, the constraints also ensure that the ‘course of history’ thus selected is defined only up to diffeomorphisms and global time translations, which ensures that we are not ‘double counting’ - i.e. two courses of history related by a global time translation are not treated as distinct possibilities by the probability distribution, they are assigned the some probability and the sum over histories required to normalise the probability distribution counts this probability only once. So the final-measurement is compatible with the symmetries used to define the quantised theory, and therefore the final-measurement approach does not violate the founding idea of the IQRF approach that all physical quantities are ultimately relational: the course of history that we select in the final measurement is definite and observer-independent, but it does not presuppose an absolute background spacetime.

Appendix C: Penrose

Ref [9] suggests an application for the quantum equivalence principle in the context of Penrose’se argument for gravitationally induced collapse [66]. That is, ref [9] contends that Penrose’s argument can be blocked by observing that he is using the wrong Equivalence Principle - he is working with the classical Equivalence Principle whereas he should in fact be using the Quantum Equivalence Principle. Now if this were correct, it would clearly be a counterexample to our worry that the quantum equivalence principle fails to have operationally meaningful content that goes beyond the classical SEP. But let us consider in more detail the way in which Penrose’s argument uses the Equivalence principle. Penrose’s fundamental concern is that in cases where we have a superposition of two different spacetimes there is no unique mapping from the points of one spacetime to the points of another, since the principle of general covariance prevents us from assigning primitive identities to spacetime points which would identify them across branches of the wavefunction. Penrose then worries that the global time-translation operator does not seem to be well-defined unless we can find a natural way of mapping the points of one superposed spacetime to points of the other, and therefore asks how we might go about constructing such a map. This is where Penrose invokes the Principle of Equivalence - he notes that ‘In accordance with the principle of equivalence, it is the notion of free fall which is locally defined, so the most natural local identification between a local region of one space-time and a corresponding local region of the other would bef that in which the free falls (i.e. spacetime geodesics) agree.’

Penrose does seem to be invoking a version of the SEP, but note that he does not appear to be insisting that the principle of equivalence will be violated if there does not exist a way of identifying the two spacetimes such that the geodesics agree. Rather, he is making use of the operational construal of the principle of equivalence, which he understands as the statement that the metric can be accessed by means of local observables in the form of geodesic motions: if this is accepted, presumably any sensible mapping between the two spacetimes should map the geodesic motions of one onto the geodesic motions of the other. Note that Penrose in fact already seems to be employing ‘quantum reference frames’ in some sense - that is, he seems to take it that the correct way to use the SEP in this situation is to apply it separately within each branch of the wavefunction to draw conclusions about our local access to the metric inside each branch. This application of the equivalence principle is entirely separate to his questions about whether there exists a way of mapping the locally accessible features of the metric of one onto the locally accessible features of the metric of the other. He does not say anywhere that the existence of such a mapping is required by the equivalence principle, or that he thinks there must be one unified Lorentzian frame of reference encapsulating both branches of the wavefunction - indeed, such a thing would be meaningless from the operational point of view where frames of reference are associated with actual or possible measurements, since measurements surveying two macroscopically distinct metrics must necessarily take place in one branch or another, so it would be nonsensical to postulate a frame of reference associated with measurements surveying both metrics at once. This indicates that Penrose’s argument can’t be blocked by simply altering the equivalence principle to apply separately within different branches of the wavefunction, because that is exactly how Penrose is already using it.

Note that it is not our intention to claim that Penrose’s argument is right. There are other possible objections to it - for example, ref [9] also notes (correctly, in our view) that there is no obvious reason why it should be possible to define a universal time-translation operator which applies across all the different branches of a spacetime superposition, since we can simply allow that time flows at different rates within the different branches. But whether or not there exists such an operator is nothing to do with the principle of equivalence: the idea that the stability of a state depends on the existence of a well-defined time evolution operator is a further assumption of Penrose’s argument which is a consequence of the standard quantum-mechanical definition of a stationary state, so it is actually an assumption that comes from quantum mechanics rather than general relativity, and therefore it is not related to his use of the principle of equivalence to identify the local observables providing access to the metric.

Appeindix D: Relational Superposition

On top of the claims about relational superposition made in the IQRF research programme, there are several other ways in which superposition might be said to be relational. First, of course superposition is relational in a trivial sense of being relative to a basis: for example, a qubit in the state \(|0\rangle \) is not in a superposition relative to the computational basis \(\{ | 0 \rangle , |1 \rangle \}\), but it is in a superposition relative to the Hadamard basis \(\{ | + \rangle , | - \rangle \}\), since \(|0 \rangle = \frac{1}{\sqrt{2}} ( | + \rangle + |- \rangle )\). In that sense, the assertion that a quantum system is ‘in a superposition’ is meaningless without the specification of a basis. But the claims about the relativity of superposition in the IQRF research programme are stronger than that: for there we find that given a certain fixed basis, whether or not a system is in a superposition relative to that basis is itself relative to a choice of reference frame. For example, in ref [5] we are given an example where relative to particle C, the particles A and B are in the entangled state \(\int dx |x \rangle _A |x + X \rangle _B\) so neither of them has a sharp state in the position basis relative to C, but when we switch to A’s reference frame we find the state of B is now sharp in the position basis.

Second, relativistic effects can cause superpositions to look different from different reference frames in settings where frames are moving quickly relative to one another. For example, in ref [77] it is noted that a Hamiltonian is defined with respect to a given slicing of space-time in equal-time surfaces, so observers in different references frames will assign different interaction Hamiltonians to a given experiment and as a result will come to different conclusions about the amount of superposition and/or entanglement present in a given physical situation. While these effects are very interesting and suggestive, they don’t seem to require any major conceptual revolutions, as the ‘reference frames’ in question are the ordinary kinds of reference frames commonly invoked within special relativity, and we are already very familiar with the idea that unusual frame-relative effects may occur when such frames are moving quickly relative to one another. On the other hand, the relativity of superposition invoked by the IQRF research programme occurs even in the absence of relative motion, so it can’t be understood in the same way as relativistic frame-dependence.

Third, we can also see frame-dependence of superposition and entanglement in the QI approach to reference frames based on the existence of an external reference frames; in this setting frame-dependence arises because the state of a system undergoes decoherence when we discard one reference frame during the switch to another reference frame [72, 78]. ‘Reference frame’ is here being understood in a similar way as in the IQRF reference frame, i.e. reference frames are associated with individual physical objects, and therefore much of what we are about to say regarding the relativity of superposition in the IQRF framework also apply here. However, it’s important to note that the frame-dependent effects in the QI case can be understood as a consequence of a system (the initial reference frame) being discarded, whereas in the IQRF framework this explanation is not available since we are dealing with the whole universe at once and therefore nothing can be discarded.

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Adlam, E. Watching the Clocks: Interpreting the Page–Wootters Formalism and the Internal Quantum Reference Frame Programme. Found Phys 52, 99 (2022). https://doi.org/10.1007/s10701-022-00620-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10701-022-00620-7

Keywords

Navigation