<<

Universal quantum modifications to general relativistic dilation in delocalised Shishir Khandelwal1,2, Maximilian P. E. Lock3, and Mischa P. Woods1,4

1Institute for Theoretical , ETH Zurich,¨ Switzerland 2Group of Applied Physics, University of Geneva, Switzerland 3Institute for and (IQOQI), Vienna, Austria 4Department of Computer , University College , United Kingdom

The associates a proper to clarify. In the latter, time is not treated on the time with each moving object via its . same footing as other quantities in the In quantum theory however, such well-defined theory — it is a parameter, rather than being repre- trajectories are forbidden. After introducing sented by a self-adjoint . This aspect of the a general characterisation of quantum clocks, theory of troubled its founders; we demonstrate that, in the weak-field, low- Wolfgang Pauli, for example, famously noted the im- velocity limit, all “good” quantum clocks ex- possibility of a self-adjoint operator corresponding to perience as dictated by general time [5]. Specifically, a quantum observable with out- relativity when their state of is classical come t ∈ R equal to the time parameterising the sys- (i.e. Gaussian). For nonclassical states of mo- tem’s , can only be achieved in the limit- tion, on the other hand, we find that quantum ing case of infinite energy, and is therefore unphys- interference effects may give rise to a signifi- ical, as we discuss in Sec.2. It is however possible cant discrepancy between the and to construct self-adjoint operators whose outcome is the time measured by the . The univer- approximately t, with an error that can in principle sality of this discrepancy implies that it is not be made arbitrarily small, in of finite energy simply a systematic error, but rather a quan- (see e.g. [6]). tum modification to the proper time itself. We If one takes an operationalist viewpoint, one can also show how the clock’s delocalisation leads draw conclusions about time by discussing the be- to a larger uncertainty in the time it measures haviour of clocks. This underlies Einstein’s develop- — a consequence of the unavoidable entangle- ment of [7], the prototypical exam- ment between the clock time and its center- ple of operationalism [8]. In a quantum setting, the of- degrees of freedom. We demonstrate clocks must of course be quantised, leading naturally how this lost precision can be recovered by per- to a time operator. This operational approach has forming a of the clock’s state of revealed fundamental limitations to time-keeping in motion alongside its time reading. non-relativistic quantum systems [6,9–11], and novel effects in a variety of relativistic settings [12–16]. The work concerns slowly-moving generic 1 Introduction quantum clocks embedded in a weakly-curved - time. We are interested in the of time di- One of the most important programs in theoretical lation, and how quantizing the clock may result in pre- physics is the pursuit of a successful theory unify- dictions distinct from those of alone. ing quantum mechanics and general relativity. Ar- We take the operational viewpoint noted above, asso- guably, many of the difficulties arising in this pursuit ciating time with the outcomes of per- stem from a lack of understanding of the nature of arXiv:1904.02178v4 [quant-ph] 12 Aug 2020 formed on generic quantum clocks. In this setting, in time [1–3], particularly the conflict between how it contrast with the study of clocks in non-relativistic is conceived in the two theories [4]. For example, in quantum mechanics, we must consider the position general relativity a given object’s proper time is de- and momentum of the clock. Quantum theory dic- fined geometrically according to that object’s world tates that it be subject to some degree of spatial de- line, but according to quantum mechanics such well- localisation in addition to its “temporal delocalisa- defined trajectories are impossible. How do we then tion” (i.e. the indeterminacy of its time-reading), as assign a proper time to the delocalised objects de- depicted in Fig.1. scribed by quantum theory? Relativistic effects manifest in this quantum set- Confusion over the nature of time is not limited ting via a coupling between the clock’s kinematic and to the context of quantum . Indeed, even in internal (in our case, time-measuring) degrees of free- non-relativistic quantum mechanics, there is much dom. The form of this coupling has been derived in a

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 1 curately. We associate the relevant degree of free- dom with a Hilbert space Hc, and consider an ini- tial clock state ρc(0) which evolves according to the ˆ Schr¨odingerequation, with Hamiltonian Hc. We use the laboratory time to refer to the parame- ter appearing in the Schr¨odingerequation. At lab- oratory time t, a measurement is performed, which corresponds to a Positive-Operator Valued Measure (POVM) {Fˆ(s)}s∈ S with S ⊆ R. We refer to the outcome s of this measurement as the clock time. A is then defined by the  ˆ ˆ Figure 1: Depiction of a spatially-delocalised quantum clock tuple {F (s)}s∈ S , Hc, ρc(0) . In this article how-  ˆ ˆ exibiting temporal indeterminacy. In the theory of relativity, ever, it suffices to consider Tc, Hc, ρc(0) , where the time measured by a clock is defined geometrically by in- ˆ R ˆ Tc := S s F (s)ds is the first- operator of the tegrating along a well-defined path, and the clock POVM, which determines the expected value of the may be arbitrarily precise and accurate. This contrasts with measurement. In the case where the POVM is a time in quantum mechanics, where a fully quantum treat- projection-valued measure (in other words, where the ment requires clocks to be delocalised in space and experience clock are perfectly distinguishable) the defini- an indeterminacy in their time. They cannot be meaningfully ˆ assigned a single trajectory. Relativity dynamically couples tion of Tc above is simply the spectral decomposition the temporal and kinematic quantum degrees of freedom, of the self-adjoint operator corresponding to the mea- resulting in time measurements which are influenced by the surement. For simplicity, we choose the laboratory ˆ delocalised spatial trajectory. time t to be such that hTci(0) = 0. If a quantum clock satisfies the Heisenberg form of  ˆ ˆ  1 the canonical commutation relation, Tc, Hc = i 1c, number of different ways: from the relativistic disper- the mean clock time will exactly follow the labo- ˆ sion relation and mass-energy equivalence [17], from ratory time, i.e. hTci(t) = t mod T0 ∀t, for some the Klein-Gordon equation in curved spacetime [18] (possibly infinite) clock period T0 > 0. Further- (or its corresponding Lagrangian density [16]), and more, the standard deviation of such clocks is con- as a consequence of classical time dilation between stant and determined by the initial state which can the clock and the laboratory [19]. be arbitrarily small. We refer to such a clock as ide- It has been used to predict a relativistically-induced alised. If one demands only this commutation rela- decoherence effect [20], and a corresponding reduc- tion, one can construct an idealised clock on a dense tion in the visibility of quantum interference experi- subset of the Hilbert space, with finite period T0 and ˆ ments [19]. This approach treats spacetime classically, with Hc bounded below (e.g. the phase operators making no reference to a quantum theory of gravity. in [21]). If we further impose canonical commuta- It is applicable only in the low-energy, weak-gravity tion relations of the Weyl form, and consider mea- limit, so that we may consider relativistic effects as surements corresponding to projection-valued mea- ˆ perturbative corrections to non-relativistic quantum sures, we necessarily have that Hc is unbounded below mechanics. (Pauli’s theorem [5, 22]), which is clearly unphysical, We begin by introducing a way of characterising and T0 = ∞. These are both examples of covariant the accuracy of a generic quantum clock, before de- POVMs [9, 23], i.e. those satisfying the covariance ˆ ˆ scribing how to incorporate the phenomenon of time property Fˆ(s) = e−isHc Fˆ(0)eisHc , s ∈ R; where in dilation into . This is followed by the case that S is bounded, Fˆ(s) has been period- a description of the average time dilation experienced ically extended to R. Furthermore, they satisfy the by a quantum clock for two different states of mo- canonical relation of the Weyl form when tion. We then discuss how the incorporation of rel- S is unbounded, and of the Heisenberg form if ρc(0) ativity increases the uncertainty associated with the and Fˆ(0) have disjoint support (see AppendixC for measurement of time, and how this can be mitigated proof). by gaining knowledge of the clock’s location/motion, To derive results that apply universally to all quan- illustrating this with a numerical example. We finish tum clocks, regardless of whether they exhibit the with a discussion of our results and the questions that idealised behaviour discussed above, we first quan- arise from them. tify the extent to which a given clock deviates from an idealised one. To this end, we introduce a clock’s 2 Non-Relativistic Quantum Clocks error operator Eˆ(t); for any clock characterised by 1Here and throughout this article, we use units such that Before we discuss relativistic effects, it is conve- ~ = 1. The symbol 1c denotes the identity operator on Hc. nient to introduce a characterisation of the extent to which a given quantum clock measures time ac-

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 2  ˆ ˆ ˆ Tc, Hc, ρc(0) , we define E(t) via body of mass M, Φ(r) = −GM/r is the Newtonian gravitational potential, and t˜ is the proper time of h ˆ ˆ i ˆ −i Tc, Hc ρc(t) = ρc(t) + E(t) (1) a fictional observer at rest at r = ∞. We take the common approximation of the Newtonian potential ˆ ˆ −itHc itHc r Φ(r) ≈ Φ(r ) + gx g where ρc(t) = e ρc(0) e . While there are a va- around a point 0, i.e. 0 , where riety of ways one might characterise the deviations of is the gravitational at the point r = r0 −2 any clock, it will turn out that for our purposes, the (e.g. g = 9.81 m s at the surface of the ), definition of Eˆ(t) is particularly convenient. As ini- and x := r − r0. The reference frame corresponding tial motivation, observe that the error operator Eˆ(t) to the laboratory is then given by (t, x), where the allows us to quantify the discrepancy between the av- laboratory time t is the proper time of an observer at erage clock time and the laboratory time t via rest at r = r0. Then, for a classical clock with initial position x0 and velocity v0, Eq. (3) allows us to relate Z t ˆ 0 h ˆ 0 i the clock’s proper time τ to the laboratory time: hTciNR(t) = t + dt tr E(t ) (2) " # 0 v2 gx v gt 1 gt2 τ = 1 − 0 + 0 + 0 − t. (4) where the subscript NR denotes that the clock does 2c2 c2 c2 3 c not account for relativistic effects (i.e. it evolves under ˆ the of Hc alone), and we have assumed for con- If we now consider the clock to be subject to the ˆ laws of quantum mechanics, we must describe how its venience that hTciNR(0) = 0. For an idealised clock,  ˆ ˆ  ˆ we have −i Tc, Hc = 1c, and therefore E(t) = 0 ∀t. temporal degree of freedom interacts with its kine- We therefore call a clock good when Eˆ(t) is small in matic degrees of freedom. To that end, we consider norm relative to the other quantities involved. a total Hilbert space H = Hc ⊗ Hk, where the space An interesting case is when the clock is a d- Hk corresponds to the kinematic variables in one di- dimensional spin , with evenly-spaced en- mension, namely position xˆ and momentum pˆ, with 1 ergy eigenvalues (corresponding to the spin-projection [ˆx, pˆ] = i k. The coupling between the temporal and ˆ kinematic (which can be understood as a conse- states). Constructing Tc such that its eigenvectors form an appropriate mutually-unbiased basis with re- quence of mass-energy equivalence [26]) is then given by the interaction Hamiltonian [17, 18, 20] spect to the spin-projection states, and choosing ρc(0) to be one of the eigenvectors of Tˆ , one arrives at  pˆ2 gxˆ c Hˆ = Hˆ ⊗ − + , (5) the well-known Salecker-Wigner-Peres clock [12, 24] ck c 2m2c2 c2 — a quantum version of the common dial clock (see where m is the mass of the clock, and where we con- Fig.1). In this case, Eˆ(t) is not always small; one tinue the approximation in Eq. (3), neglecting terms has trEˆ(t) = −1 at regular intervals of labora- proportional to 1/c4 (see AppendixA). The two terms tory time, regardless of the dimension d or energy. in Eq. (5) represent the lowest-order contributions to This arises due to the clock states’ lack of coher- the time dilation due to motion (c.f. v2/2c2 in Eq. (3)) ence in the eigenbasis of Tˆ ; specifically [Tˆ , ρ (t)] is c c c and gravity respectively. The clock is then subject zero at regular intervals of laboratory time. This is- to the total Hamiltonian Hˆ = Hˆ + Hˆ + Hˆ , with sue can be removed by choosing ρ (0) to have some c k ck c Hˆ = mc2 + mgxˆ +p ˆ2/2m − pˆ4/8m3c2, where the fi- quantum coherence in this basis, as in the Quasi- k nal term is the usual lowest-order relativistic correc- Ideal clock [6]. In the latter case, kEˆ(t)k is expo- 1 tion to the . nentially small in both the dimensionality d and the mean energy of the clock ∀t. This is a consequence of the Quasi-Ideal clock’s approximately dispersion- 4 Time Dilation in Quantum Clocks ˆ less dynamics in the eigenbasis of Tc. In addition to the Quasi-Ideal clock, the Salecker-Wigner-Peres and We assume the initial state of the clock to be un- qubit-phase clocks, which do not posses this property, correlated across Hc and Hk, i.e. that ρck(0) = are discussed in AppendixD. ρc(0)⊗ρk(0) for some ρc(0) and ρk(0). After evolving for an amount of laboratory time t, the average time dilation experienced by the clock (i.e. the average 3 Relativistic Time Dilation clock time) is Classically, in the post-Newtonian approximation of ˆ ˆ n h ˆ io hTci(t) = hTciNR(t) + tR(t) 1 + tr E(t) , (6) general relativity, the proper time τ experienced by an observer is determined by [25] h pˆ2 gxˆ pgtˆ g2t2  i where R(t) := tr − 2m2c2 + c2 + mc2 − 3c2 ρk(0)  v2 Φ(r) is determined by the initial kinematic state. Recall dτ = dt˜ 1 − + , (3) ˆ 2c2 c2 that E(t) was defined according to the non-relativistic ˆ evolution of the clock (i.e. under Hc alone). We now 2 2 2 to first order in v /c and Φ/c , where v is the ob- consider two possibilities for ρk(0) — a classical (i.e. server’s velocity, r is the distance from a gravitating Gaussian) and a nonclassical state.

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 3 4.1 A Gaussian state of motion xAAAB6HicbVDLTgJBEOzFF+IL9ehlIjHxRHbRRI9ELx4hkUcCGzI79MLI7OxmZtZICF/gxYPGePWTvPk3DrAHBSvppFLVne6uIBFcG9f9dnJr6xubW/ntws7u3v5B8fCoqeNUMWywWMSqHVCNgktsGG4EthOFNAoEtoLR7cxvPaLSPJb3ZpygH9GB5CFn1Fip/tQrltyyOwdZJV5GSpCh1it+dfsxSyOUhgmqdcdzE+NPqDKcCZwWuqnGhLIRHWDHUkkj1P5kfuiUnFmlT8JY2ZKGzNXfExMaaT2OAtsZUTPUy95M/M/rpCa89idcJqlByRaLwlQQE5PZ16TPFTIjxpZQpri9lbAhVZQZm03BhuAtv7xKmpWyd1Gu1C9L1ZssjjycwCmcgwdXUIU7qEEDGCA8wyu8OQ/Oi/PufCxac042cwx/4Hz+AOeHjQA=

We take the initial kinematic state ρk(0) to be a pure Gaussian wavepacket, denoting the mean momentum by p¯0, the standard deviation of the momentum by 2 σp, and the mean position by x¯0. This is the most AAACDHicbVDLSgMxFL3js9ZX1aWbYBFclZki6LLoxmUF+4BOLZk004ZmMkNyRyhDP8CNv+LGhSJu/QB3/o1pOwttPRA4OedcknuCRAqDrvvtrKyurW9sFraK2zu7e/ulg8OmiVPNeIPFMtbtgBouheINFCh5O9GcRoHkrWB0PfVbD1wbEas7HCe8G9GBEqFgFK3UK5V9yUP0bQSJnxjRy6oT4msxGM7Fe3u3KbfizkCWiZeTMuSo90pffj9macQVMkmN6Xhugt2MahRM8knRTw1PKBvRAe9YqmjETTebLTMhp1bpkzDW9igkM/X3REYjY8ZRYJMRxaFZ9Kbif14nxfCymwmVpMgVmz8UppJgTKbNkL7QnKEcW0KZFvavhA2ppgxtf0Vbgre48jJpViue5bfn5dpVXkcBjuEEzsCDC6jBDdShAQwe4Rle4c15cl6cd+djHl1x8pkj+APn8wdDc5u9 2 2AAAB83icbZDLSgMxFIZPvNZ6q7p0EyyCqzJTBF0W3bisYC/QGUomzbShSWZIMmIZ+hpuXCji1pdx59uYtrPQ1h8CH/85h3PyR6ngxnreN1pb39jc2i7tlHf39g8OK0fHbZNkmrIWTUSiuxExTHDFWpZbwbqpZkRGgnWi8e2s3nlk2vBEPdhJykJJhorHnBLrrKCOA8OHkvTzp2m/UvVq3lx4FfwCqlCo2a98BYOEZpIpSwUxpud7qQ1zoi2ngk3LQWZYSuiYDFnPoSKSmTCf3zzF584Z4DjR7imL5+7viZxIYyYycp2S2JFZrs3M/2q9zMbXYc5Vmlmm6GJRnAlsEzwLAA+4ZtSKiQNCNXe3YjoimlDrYiq7EPzlL69Cu17zHd9fVhs3RRwlOIUzuAAfrqABd9CEFlBI4Rle4Q1l6AW9o49F6xoqZk7gj9DnD8P4kYA= x classical choice in the sense that such states are the | | only ones with a non-negative Wigner function [27],

AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0mqoMeiF48t2FpoQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgpr6xubW8Xt0s7u3v5B+fCoreNUMWyxWMSqE1CNgktsGW4EdhKFNAoEPgTj25n/8IRK81jem0mCfkSHkoecUWOl5rBfrrhVdw6ySrycVCBHo1/+6g1ilkYoDRNU667nJsbPqDKcCZyWeqnGhLIxHWLXUkkj1H42P3RKzqwyIGGsbElD5urviYxGWk+iwHZG1Iz0sjcT//O6qQmv/YzLJDUo2WJRmApiYjL7mgy4QmbExBLKFLe3EjaiijJjsynZELzll1dJu1b1Lqq15mWlfpPHUYQTOIVz8OAK6nAHDWgBA4RneIU359F5cd6dj0VrwclnjuEPnM8fzcOM7w== g and saturate the position-momentum uncertainty re- AAAB83icbZBNS8NAEIYn9avWr6pHL4tF8FQSEfRY1IPHCrYWmlA22027dLMJuxOxhP4NLx4U8eqf8ea/cdvmoK0vLDy8M8PMvmEqhUHX/XZKK6tr6xvlzcrW9s7uXnX/oG2STDPeYolMdCekhkuheAsFSt5JNadxKPlDOLqe1h8euTYiUfc4TnkQ04ESkWAUreX7N1wiJU+93J30qjW37s5ElsEroAaFmr3ql99PWBZzhUxSY7qem2KQU42CST6p+JnhKWUjOuBdi4rG3AT57OYJObFOn0SJtk8hmbm/J3IaGzOOQ9sZUxyaxdrU/K/WzTC6DHKh0gy5YvNFUSYJJmQaAOkLzRnKsQXKtLC3EjakmjK0MVVsCN7il5ehfVb3LN+d1xpXRRxlOIJjOAUPLqABt9CEFjBI4Rle4c3JnBfn3fmYt5acYuYQ/sj5/AGH35FX x0 lation. We then have 2 2 2  2 p¯ + σ gx¯ p¯ gt 1 gt AAACDHicbVC7SgNBFL3rM8ZX1NJmMAhWYTcIWgZtLCOYB2TXMDuZTYbMPpi5K4RlP8DGX7GxUMTWD7Dzb5wkW2jigYHDOedy5x4/kUKjbX9bK6tr6xubpa3y9s7u3n7l4LCt41Qx3mKxjFXXp5pLEfEWCpS8myhOQ1/yjj++nvqdB660iKM7nCTcC+kwEoFgFI3Ur1RdyQN0TQSJm2jRz5ycuEoMR3PxPqvnJmXX7BnIMnEKUoUCzX7lyx3ELA15hExSrXuOnaCXUYWCSZ6X3VTzhLIxHfKeoRENufay2TE5OTXKgASxMi9CMlN/T2Q01HoS+iYZUhzpRW8q/uf1UgwuvUxESYo8YvNFQSoJxmTaDBkIxRnKiSGUKWH+StiIKsrQ9Fc2JTiLJy+Tdr3mGH57Xm1cFXWU4BhO4AwcuIAG3EATWsDgEZ7hFd6sJ+vFerc+5tEVq5g5gj+wPn8AQd6bvA== 1 R(t) = − 0 p + 0 + 0 − . (7) | | 2m2c2 c2 mc2 3 c

In this case, setting Eˆ(t) = 0 and using Eqs. (2) and (6), one finds that the average time measured by an idealised quantum clock is identical to the expecta- Figure 2: A cartoon of the (magnitude squared of) the wave- tion value of the classical proper time (Eq. (4)) for an functions used to construct delocalised initial (kinematic) observer with a momentum following the same (i.e. states of the clock. Gaussian) probability distribution. A non-idealised clock will exhibit a non-relativistic quantum correc- tion (the term in Eq. (2)) as well as a con- Note that the exponential factor in Eq. (9) is the over- tribution arising from both relativity and quantum lap of the constituent states, hψ1|ψ2ik (c.f. the grey mechanics (the final term in Eq. (6)). However, since region in Fig. (2)). both these effects are proportional to trEˆ(t), they We denote the average clock time in this case by ˆ can be made arbitrarily small, for example by us- hTcisup(t). To extract the part of the effect arising ing the Quasi-Ideal clock discussed in Sec.2 with from quantum coherence, we contrast this with the an appropriately high dimensionality and mean en- case of a classical mixture of two such states according ergy [6]. Interestingly, for the Salecker-Wigner-Peres to probabilities α and 1−α. The average clock time in   ˆ clock discussed Sec.2, tr Eˆ(t) = −1 whenever the the latter case, which we denote hTcimix(t), can easily ˆ be calculated via Eq. (6) by taking the corresponding clock is in an eigenstate of Tc, exactly cancelling the usual classical relativistic effect in Eq. (6) (see weighted sum, i.e. AppendixD). This property of the Saleker-Wigner- ˆ ˆ ˆ hTcimix(t) = αhTciψ1 (t) + (1 − α)hTciψ2 (t), (10) Peres clock states has also been identified as the cause ˆ for suboptimal performance in other areas, related to where hTciψi (t) is the expectation value for state with clocks [28]. |ψiik (t). For simplicity of expression, we consider good clocks in the sense discussed in Sec.2, so that the 4.2 A superposition of heights contribution arising from a nonzero Eˆ(t) is neglible. One then has A consideration of a nonclassical initial kinematic ˆ ˆ state can result in a significant modification of the hTcisup(t) = hTcimix(t) + Tcoh(t), (11) classical time-dilation effect found in Sec. 4.1. We where T (t), the contribution due to coherence be- consider an initial kinematic state constructed by su- coh tween the two constituent Gaussian states, is given by perposing two Gaussian wavepackets with different mean initial heights, namely " # N − 1 ∆x 2 σ2 g∆x t 1 √ √ 0 v 0  Tcoh(t) := 2 − 2 (1 − 2α) |ψi = √ α |ψ1i + 1 − α |ψ2i (8) N 2σx c c 2 k N k k (12) where σ := σ /m is their standard deviation in their for some 0 < α < 1, where |ψ1ik and |ψ2ik are Gaus- v p sian states differing only in the value of hxˆi. Specif- initial “velocity”. We recall that, since the wavepack- ets saturate the uncertainty relation, we have that ically, |ψ1ik and |ψ2ik have mean positions x¯0 and x¯0 + ∆x0 respectively, standard deviation in position σp = 1/(2 σx). σx, standard deviation in momentum σp, and for sim- In Eq. (12) we see two contributions to Tcoh(t): one 2 plicity we take both wavepackets to have the same ini- proportional to σv and one due to the average dif- tial mean momentum. This is illustrated in Fig. (2). ference in gravitational potential experienced by the The normalisation factor N is then given by wavepackets (g∆x0). In other words, we see sepa- rately how motional and gravitational time dilation 2 p − 1 ∆x0  N = 1 + 2 α(1 − α)e 2 2σx . (9) act on the wavefunction to generate quantum coher- 2 ence effects. Noting that we can write g∆x0 = vg /2,

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 4 where vg is the velocity accrued by a classsical ob- we show how the entanglement generated by the in- ject falling a distance of ∆x0 from rest, we see that teraction Hamiltonian in Eq. (5) increases the uncer- for a given ∆x0/σx and α, the relative strength of the tainty associated with measurements of the temporal motional and gravitational parts is determined by the variable. comparative magnitudes of σv and vg. We quantify the clock’s precision via the standard Furthermore, considering Eq. (9), one can see that deviation of the clock time, which we denote σT (t). Tcoh(t) decreases exponentially with increasing ∆x0, For simplicity, we ignore the contribution due to a consequence of the decreasing overlap between the gravity by setting g = 0, and consider only tempo- two wavepackets. From this we infer that Tcoh(t) ral measurements which correspond to a projection- arises due to interference between |ψ1ik and |ψ2ik. On valued measure. In order to calculate the relativistic the other hand, we evidently have Tcoh(t) = 0 when contribution to σT (t) to leading order, we now in- 4 4 ∆x0 = 0, as in that case |ψ1ik = |ψ2ik, and we return clude Hamiltonian terms up to order pˆ /(mc) (see to the classical-motion scenario described in Sec. 4.1. AppendixE). Let σT,NR(t) denote the standard devi- Consequently there exists, for a given σv and α, an in- ation of the clock time in the absence of the special termediary value of ∆x0/σx which maximises Tcoh(t) relativistic coupling. Then, assuming the clock’s kine- by finding a balance between the wavefunction over- matic and temporal degrees of freedom to be uncor- lap and the “nonclassicality” of the kinematic state. related at t = 0, one finds that σT (t) separates into Given that modern- atomic clocks are accurate and precise enough to observe classical relativistic σT (t) = σT,NR(t) + σT,I(t) + σT,NI(t), (13) time dilation in tabletop experiments [29], it is natural where the term σ I(t) is a contribution that remains to ask whether the quantum contribution to the time T, finite in the case of an idealised clock, given by dilation might also be observable. For the purposes of illustration, let us consider an atom 2 4 2 t hpˆ i + σp2 (whose mass is approximately 27 u). This is inspired σT,I(t) := 4 4 , (14) 8 σ NR(t) m c by a state-of the-art “ clock” based on T, optical transitions in a single aluminium [30]. For where σp2 is the standard deviation of the observ- clarity, we use the (Van der Waals) radius of alu- 2 able pˆ , and the term σT,NI(t) is the relativistic con- minium, specifically rAl = 184 pm [31], as a unit of tribution arising due the clock’s non-idealised na- . Then from Eq. (12), we see that for a balanced ture. Its full form is given in AppendixE. Note that superposition (i.e. α = 1/2), taking σx = 2 rAl and σT,I(t) > 0, and the effect of the relativistic cou- −16 ∆x0 = 4 rAl, we have Tcoh(t) ∼ 10 s in a labora- pling on good clocks (in the sense discussed in Sec.2) 2 tory time of t = 1 s, which is well within the measure- is therefore to increase the clock’s temporal uncer- ment capability of state-of-the-art clocks. We have tainty. In the case of an idealised clock, σT,NR(t) is of course, assumed good clocks so that the contribu- constant in laboratory time (see AppendixC) and can ˆ tion due to the error term E(t), is neglectable. While be made arbitrarily small by an appropriate choice of we have demonstrated that this is a justified approx- ρc(0). This feature can also be well approximated imation for clocks with approximately non-dispersive with the Quasi-Ideal clock. Consequently, according dynamics, such as the Quasi-Ideal clock, this approxi- to Eq. (14), the standard deviation of the clock time mation might not be so well suited to the qubit-phase increases quadratically with the laboratory time in and Seleker-Wigner-Peres clocks previously discussed. these cases. This appears very promising, though we stress that The decrease in precision is a consequence of the this example is illustrative. A careful analysis con- clock’s temporal state losing information via its entan- cerning potential experimental platforms is required glement with the kinematic degrees of freedom. We before conclusions can be drawn about the prospect now show how this effect can be reduced by recover- of observing the effect. ing some of the lost information via a measurement of the clock’s momentum. We consider course-grained 5 Clock Precision and the Coupling momentum measurements, and show how the uncer- tainty in the clock time decreases as the measurement of Temporal and Kinematic Degrees of is made more precise. We again choose a classical Freedom (i.e. Gaussian) state of motion, as in Sec. 4.1, and we consider an idealised clock for simplicity, as this In this section we show how the clock’s precision is can be approximated arbitrarily well (see the discus- modified by the relativistic coupling between its kine- sion in Sec.2). We define a set of projection operators ˆ Z matic and temporal degrees of freedom. In particular {Πn,δp}n acting on Hk, with n ∈ and δp > 0, which correspond to a partition of the range of momentum 2This is well within the coherence times achievable in both values into bins of width δp, with bin n centred on trapped-ion and neutral-atom optical atomic clocks [32]. Main- taining a coherent superposition of heights for ∼ 1 s is a sepa- rate challenge, achieved in e.g. [33].

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 5 momentum value n δp, i.e. 6 Discussion

Z (n+1/2)δp Throughout this work, we have made reference to Πˆ := dp |pihp| . (15) n,δp k t (n−1/2)δp the laboratory time , which can be interpreted as the time marked by a classical clock in the labora- The bin width δp therefore characterises the amount tory frame. One may however wish to consider a of knowledge that we gain from the measurement, fully quantum set-up, without reference to an extrin- and the degree of localisation in momentum space sic classical time, as in the Page and Wootters for- of the post-measurement clock state. As δp ap- malism [34], which can be employed in relativistic ˆ proaches zero, {Πn,δp}n approaches the set of pro- contexts [35, 36]. In our framework, one may sim- jectors onto momentum eigenstates. As δp → ∞, ply treat t as a bookkeeping coordinate, to be used to on the other hand, the projectors tend to the iden- calculate the average clock times of multiple quantum tity operator, i.e. the case where no measurement is clocks, and thus their average clock times relative to performed. Fig.3 shows examples of σT (t) condi- each other. In doing so, all our derived results can tioned on a given outcome of the measurement, at be written in terms of the average readings of quan- different laboratory times t, and for different values tum clocks without any reference to a classical clock of δp. Quantifying the coarseness of the measurement or extrinsic time. by q := δp/σp, we find that for q → 0, we recover all The effect that we describe in Sec. 4.2 is distinct of the information leaked into the kinematic state, i.e. from the manifestation of relativistic time dilation be- after the measurement, σT (t) = σT (0) = σT,NR(t). tween paths in quantum interferometry experiments. For q → ∞, on the other hand, no information is Examples of the latter include phase shifts observed recovered, as no measurement had been performed. in a neutron matter-wave interferometer [37], the prediction that redshift-induced which-path informa- 2.5 tion leads to a reduction in interferometric visibil-

qAAAB7nicbVC7SgNBFL0bXzG+opY2g0GwCrsqaGERsLGMYB6QLGF2cjcZMju7mZkVwpKPsLFQxNbvsfNvnCRbaOKBgcM59zL3nCARXBvX/XYKa+sbm1vF7dLO7t7+QfnwqKnjVDFssFjEqh1QjYJLbBhuBLYThTQKBLaC0d3Mbz2h0jyWj2aSoB/RgeQhZ9RYqTUmXRMTt1euuFV3DrJKvJxUIEe9V/7q9mOWRigNE1Trjucmxs+oMpwJnJa6qcaEshEdYMdSSSPUfjY/d0rOrNInYazsk4bM1d8bGY20nkSBnYyoGeplbyb+53VSE974GZdJalCyxUdhKoiNOMtO+lwhM2JiCWWK21sJG1JFmbENlWwJ3nLkVdK8qHqXVffhqlK7zesowgmcwjl4cA01uIc6NIDBCJ7hFd6cxHlx3p2PxWjByXeO4Q+czx9Md47c 0 ! ity [19], and the proposal to observe gravitational

qAAAB6nicbVBNS8NAEJ3Ur1q/qh69LBbBU0lU0INCwYvHivYD2lA220m7dLOJuxuhhP4ELx4U8eov8ua/cdvmoNUHA4/3ZpiZFySCa+O6X05haXllda24XtrY3NreKe/uNXWcKoYNFotYtQOqUXCJDcONwHaikEaBwFYwup76rUdUmsfy3owT9CM6kDzkjBor3T1ceb1yxa26M5C/xMtJBXLUe+XPbj9maYTSMEG17nhuYvyMKsOZwEmpm2pMKBvRAXYslTRC7WezUyfkyCp9EsbKljRkpv6cyGik9TgKbGdEzVAvelPxP6+TmvDCz7hMUoOSzReFqSAmJtO/SZ8rZEaMLaFMcXsrYUOqKDM2nZINwVt8+S9pnlS906p7e1apXeZxFOEADuEYPDiHGtxAHRrAYABP8AKvjnCenTfnfd5acPKZffgF5+MbzCmNcw== =1 phase shifts in an atom interferometer [38]. We note 2.0 qAAAB7HicbVBNS8NAEJ2tX7V+VT16WSyCp5D4gR4UCl48VjBtoQ1ls920SzebuLsRSuhv8OJBEa/+IG/+G7dtDlp9MPB4b4aZeWEquDau+4VKS8srq2vl9crG5tb2TnV3r6mTTFHm00Qkqh0SzQSXzDfcCNZOFSNxKFgrHN1M/dYjU5on8t6MUxbEZCB5xCkxVvIfrj3nvFetuY47A/5LvILUoECjV/3s9hOaxUwaKojWHc9NTZATZTgVbFLpZpqlhI7IgHUslSRmOshnx07wkVX6OEqULWnwTP05kZNY63Ec2s6YmKFe9Kbif14nM9FlkHOZZoZJOl8UZQKbBE8/x32uGDVibAmhittbMR0SRaix+VRsCN7iy39J88TxTh337qxWvyriKMMBHMIxeHABdbiFBvhAgcMTvMArkugZvaH3eWsJFTP78Avo4xuuAo3q =1.5 that the latter proposal is distinct from the reinter-

qAAAB6nicbVDLSgNBEOyNrxhfUY9eBoPgKexGQQ8KAS8eI5oHJEuYnXSSIbOz68ysEJZ8ghcPinj1i7z5N06SPWhiQUNR1U13VxALro3rfju5ldW19Y38ZmFre2d3r7h/0NBRohjWWSQi1QqoRsEl1g03AluxQhoGApvB6GbqN59QaR7JBzOO0Q/pQPI+Z9RY6f7xutItltyyOwNZJl5GSpCh1i1+dXoRS0KUhgmqddtzY+OnVBnOBE4KnURjTNmIDrBtqaQhaj+dnTohJ1bpkX6kbElDZurviZSGWo/DwHaG1Az1ojcV//Paielf+imXcWJQsvmifiKIicj0b9LjCpkRY0soU9zeStiQKsqMTadgQ/AWX14mjUrZOyu7d+el6lUWRx6O4BhOwYMLqMIt1KAODAbwDK/w5gjnxXl3PuatOSebOYQ/cD5/AM2tjXQ= =2 pretation of measurements of atom-interferometry- (ns) T

qAAAB7HicbVBNS8NAEJ3Ur1q/qh69LBbBU0iqogeFghePFUxbaEPZbDft0s0m7m6EEvobvHhQxKs/yJv/xk2bg7Y+GHi8N8PMvCDhTGnH+bZKK6tr6xvlzcrW9s7uXnX/oKXiVBLqkZjHshNgRTkT1NNMc9pJJMVRwGk7GN/mfvuJSsVi8aAnCfUjPBQsZARrI3mPN3X7ol+tObYzA1ombkFqUKDZr371BjFJIyo04Viprusk2s+w1IxwOq30UkUTTMZ4SLuGChxR5WezY6foxCgDFMbSlNBopv6eyHCk1CQKTGeE9Ugtern4n9dNdXjlZ0wkqaaCzBeFKUc6RvnnaMAkJZpPDMFEMnMrIiMsMdEmn4oJwV18eZm06rZ7Zjv357XGdRFHGY7gGE7BhUtowB00wQMCDJ7hFd4sYb1Y79bHvLVkFTOH8AfW5w+viI3r =2.5 based accelerometers as a ef- AAACAnicbVDLSsNAFJ3UV62vqCtxM1iEClISFXThouDGZYW+oAllMp22QyeTMHMjlhDc+CtuXCji1q9w5984fSy09cCFwzn3cu89QSy4Bsf5tnJLyyura/n1wsbm1vaOvbvX0FGiKKvTSESqFRDNBJesDhwEa8WKkTAQrBkMb8Z+854pzSNZg1HM/JD0Je9xSsBIHfvA07wfkk5ay7B3ij1gD5CWpD7JOnbRKTsT4EXizkgRzVDt2F9eN6JJyCRQQbRuu04MfkoUcCpYVvASzWJCh6TP2oZKEjLtp5MXMnxslC7uRcqUBDxRf0+kJNR6FAamMyQw0PPeWPzPayfQu/JTLuMEmKTTRb1EYIjwOA/c5YpRECNDCFXc3IrpgChCwaRWMCG48y8vksZZ2T0vO3cXxcr1LI48OkRHqIRcdIkq6BZVUR1R9Iie0St6s56sF+vd+pi25qzZzD76A+vzB3WDlsw=

1.5 qAAAB83icbVBNS8NAEJ3Ur1q/qh69LBbBU0lU0IOHghePFWwtNKFstpt26WYTdydCCf0bXjwo4tU/481/47bNQVsfDDzem2FmXphKYdB1v53Syura+kZ5s7K1vbO7V90/aJsk04y3WCIT3Qmp4VIo3kKBkndSzWkcSv4Qjm6m/sMT10Yk6h7HKQ9iOlAiEoyilfxH4mNCfKEiHPeqNbfuzkCWiVeQGhRo9qpffj9hWcwVMkmN6XpuikFONQom+aTiZ4anlI3ogHctVTTmJshnN0/IiVX6JEq0LYVkpv6eyGlszDgObWdMcWgWvan4n9fNMLoKcqHSDLli80VRJon9cxoA6QvNGcqxJZRpYW8lbEg1ZWhjqtgQvMWXl0n7rO6d1927i1rjuoijDEdwDKfgwSU04Baa0AIGKTzDK7w5mfPivDsf89aSU8wcwh84nz+feZFk fect (see [39, 40] and subsequent criticisms [41–43]). !1 These works concern the phase accrued between dis- tinct paths in space as a consequence of relativistic

1.0 time dilation, which is a distinct phenomenon from the one that we consider in this article. Here we ask 0.00 0.02 0.04 0.06 0.08 0.10 Laboratory time (s) what time is measured by a clock delocalised across AAACA3icbVA9SwNBEN3zM8avqJ02i0GITbhTQQuLgI2FRQTzAUkIe5u5ZMne7bE7J4YjYONfsbFQxNY/Yee/cfNRaOKDgcd7M8zM82MpDLrut7OwuLS8sppZy65vbG5t53Z2q0YlmkOFK6l03WcGpIigggIl1GMNLPQl1Pz+1civ3YM2QkV3OIihFbJuJALBGVqpndtvIjxgesN8pRkqPaAoQqAFczxs5/Ju0R2DzhNvSvJkinI799XsKJ6EECGXzJiG58bYSplGwSUMs83EQMx4n3WhYWnEQjCtdPzDkB5ZpUMDpW1FSMfq74mUhcYMQt92hgx7ZtYbif95jQSDi1YqojhBiPhkUZBIioqOAqEdoYGjHFjCuBb2Vsp7TDOONrasDcGbfXmeVE+K3mnRvT3Lly6ncWTIATkkBeKRc1Ii16RMKoSTR/JMXsmb8+S8OO/Ox6R1wZnO7JE/cD5/ANFMl5Y= distinct spacetime trajectories; a phase between two Figure 3: The standard deviation of the clock time, condi- spatial components of that delocalisation affects the tioned on discovering that the clock’s momentum lies within answer to this question, but it is not itself the quantity the central (most likely) bin of variously coarse-grained mea- of interest. surements. The momentum follows a Gaussian distribution The present work has been concerned with the ef- centred around p = 0, and the (idealised) clock is initialised fect of relativity in a scenario, that is, such that its time also follows a Gaussian distribution. For we are interested in the time elapsed between two illustration, we have taken σT (0) = 1 ns, and chosen the other initial parameters to correspond to an electron with events. This contrasts with the ticking clocks dis- σx = 1 nm. The coarseness of the measurement is deter- cussed in [10, 11, 44–47]. The incorporation of rela- mined by q := δp/σp, i.e. the ratio of the bin size to the tivity into the latter is an open problem. standard deviation in momentum. The reduction in clock precision that we predict is both a quantum and relativistic effect, as the gen- eral theory of relativity allows for clocks which are always perfectly precise, and clocks in non-relativistic quantum theory are generally uncoupled from their kinematic degrees of freedom. This leads to the inter- esting statement that relativity requires us to ask how a clock is moving as well as the time it measures, or we must pay a penalty in its precision. Though our ex- plicit calculation of the decreased clock precision was carried out neglecting the effect of gravity, the same principle will hold for a nonzero gravitational field. A general initial clock state which is uncorrelated be-

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 6 tween temporal and kinematic degrees of freedom will Setting this framework into a relativistic context, only typically remain in the set of separable states if we considered the average time dilation according to either the eigenstates of the total Hamiltonian are sep- a generic quantum clock, finding that for Gaussian arable (which they are not), or if the initial state is states of motion, any quantum clock will experience particularly mixed [48]. One can therefore only escape the usual classical relativistic time dilation with some the relativistic decrease in precision by increasing the quantum error (given by the trace of the error oper- mixedness of the initial state (and thus decreasing its ator) which can be made arbitrarily small. Taking a precision anyway). The process of recovering the pre- nonclassical state of motion, namely a superposition cision via measurements of the kinematic state will of Gaussian states, we found that interference results likely be modified by gravity, however. In particular, in an average clock time which differs from the predic- the gravitational coupling between the temporal de- tion of classical relativity, even when the error opera- gree of freedom and the clock’s spatial position will tor of a clock is zero (an idealised clock). Considering mean that a perfectly accurate measurement of the the example of an aluminium ion in such a state of momentum degree of freedom will no longer totally motion, we found that this quantum relativistic time restore the clock’s precision. dilation discrepancy should be of an observable mag- The decreased clock precision is related to the effec- nitude. tive decoherence of quantum systems discussed in [20], We then discussed how relativity leads to entan- though, contrary to our work, it is the decoherence of glement between the clock’s time-measuring and mo- the kinematic degrees of freedom which is studied in tional degrees of freedom. We calculated the cor- that case. Other work has predicted a limit of the abil- responding reduction in clock precision for classical ity of multiple clocks to jointly measure time due to a states of motion, showing how it increases over time. gravity-mediated interaction between them [49], or fo- We demonstrated how the information lost from the cused on special relativistic effects [50]. Furthermore, clock’s time-measuring degrees of freedom can be re- our work shows that the qubit phase shift seen in [16] covered by performing a measurement of the clock’s is in fact a specific instance of a -dilation motion in addition to the clock time. effect. We model the effect of mass-energy equivalence It is our hope that these results will fuel interest in the same manner as these authors, and in partic- from the experimental community, so that experi- ular the resulting relativistic coupling between a sys- ments might be devised to observe combined quan- tem’s internal and centre-of-mass degrees of freedom tum relativistic effects, perhaps shedding light on the described by the interaction Hamiltonian in Eq. (5) obscure relationship between the two theories. in the limit of low velocities and a weak gravitational field. Outside of this regime, a fully relativistic theory is necessary. An analysis and rebuttal of some of the Acknowledgments criticisms of the model can be found in [51], including The authors would like to thank Alexander Smith a discussion of the validity of describing relativistic for useful discussions and Yanglin Hu for correct- effects via the Schr¨odingerequation. Since a full the- ing typos. S.K. acknowledges financial support from ory of is yet to be experimentally ETH Z¨urich during his Master’s degree (Birkigt Schol- established, all results at the interface between quan- arship). M.P.E.L. acknowledges support from the tum theory and gravity are naturally controversial. ESQ Discovery Grant Emergent time - operational- Ultimately, experiments will determine the full ism, quantum clocks and thermodynamics of the Aus- scope of validity for these results. trian Academy of (OAW),¨ as well as from the Austrian Science Fund (FWF) through the START 7 Conclusion project Y879-N27. M.P.W. acknowledges funding from the Swiss National Science Foundation (AM- In the non-relativistic limit, we characterised a generic BIZIONE Fellowship PZ00P2 179914), the NCCR quantum clock by an initial state ρc(0), a Hamilto- QSIT, and FQXi grant Finite dimensional Quantum ˆ ˆ nian Hc and a POVM {F (s)}s∈S corresponding to the Observers (No. FQXi-RFP-1623) for the programme measurement of the clock time. Not all such systems Physics of the Observer. can serve as clocks; for example, a clock where ρc(0) is an energy eigenstate will not evolve in time. An un- avoidable consequence of quantum theory is that any References finite-energy clock with perfectly distinguishable tem- [1] Isham, C. J. Canonical quantum gravity and the poral states cannot be perfectly accurate (a constraint problem of time. In Integrable systems, quan- which is distinct from the ). We tum groups, and quantum field theories, 157–287 quantified the inaccuracy of a generic clock, with ar- (Springer, 1993). URL https://doi.org/10. bitrary temporal states, by defining an error operator. 1007/978-94-011-1980-1_6. We call a clock good, when its error operator is small in norm.

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 7 [2] Kuchaˇr,K. V. Time and interpretations of quan- shift of free-falling particles. Class. Quantum tum gravity. In 4th Canadian Conference on Gravity 35, 035011 (2018). URL https://doi. General Relativity and Relativistic Astrophysics, org/10.1088/1361-6382/aaa0e8. 211–314 (World Scientific, 1992). URL https: [17] Zych, M. Quantum Systems under Grav- //doi.org/10.1142/S0218271811019347. itational Time Dilation. Springer Theses [3] Anderson, E. Quantum Mechanics Versus Gen- (Springer, 2017). URL https://doi.org/10. eral Relativity (Springer International Publish- 1007/978-3-319-53192-2. ing, 2017), 1st edn. URL https://doi.org/10. [18] L¨ammerzahl,C. A Hamilton operator for quan- 1007/978-3-319-58848-3. tum optics in gravitational fields. Phys. Lett. A [4] Lock, M. P. E. & Fuentes, I. Relativistic 203, 12–17 (1995). URL https://doi.org/10. quantum clocks. In , 51–68 1016/0375-9601(95)00345-4. (Springer, 2017). URL https://doi.org/10. [19] Zych, M., Costa, F., Pikovski, I. & Brukner, C.ˇ 1007%2F978-3-319-68655-4_5. Quantum interferometric visibility as a witness of [5] Pauli, W. Die allgemeinen prinzipien der wellen- general relativistic proper time. Nat. Commun. mechanik. Handbuch der Physik 5, 1–168 (1958). 2, 505 (2011). URL https://doi.org/10.1038/ [6] Woods, M. P., Silva, R. & Oppenheim, ncomms1498. J. Autonomous Quantum Machines and [20] Pikovski, I., Zych, M., Costa, F. & Brukner, C.ˇ Finite-Sized Clocks. Annales Henri Poincar´e Universal decoherence due to gravitational time (2018). URL https://doi.org/10.1007/ dilation. Nat. Phys. 11, 668–672 (2015). URL s00023-018-0736-9. https://doi.org/10.1038/nphys3366. [7] Einstein, A. Zur elektrodynamik bewegter [21] Garrison, J. C. & Wong, J. Canonically conju- k¨orper. Annalen der physik 322, 891–921 gate pairs, uncertainty relations, and phase op- (1905). URL https://doi.org/10.1002/andp. erators. J. Math. Phys. 11, 2242–2249 (1970). 19053221004. URL https://doi.org/10.1063/1.1665388. [8] Bridgman, P. W. The logic of modern physics [22] Srinivas, M. & Vijayalakshmi, R. The time (Macmillan New York, 1927). of occurrence in quantum mechanics. Pramana [9] Buˇzek, V., Derka, R. & Massar, S. Optimal 16, 173–199 (1981). URL https://doi.org/10. quantum clocks. Phys. Rev. Lett. 82, 2207– 1007/BF02848181. 2210 (1999). URL https://doi.org/10.1103/ [23] Holevo, A. Covariant measurements and un- PhysRevLett.82.2207. certainty relations. Reports on Mathematical [10] Erker, P. The Quantum : ap- Physics 16, 385 – 400 (1979). URL https: proaching time measurement with quantum in- //doi.org/10.1016/0034-4877(79)90072-7. formation theory. Master’s thesis, ETH [24] Peres, A. Measurement of time by quantum Z¨urich (2014). URL https://doi.org/10. clocks. Am. J. Phys 48, 552 (1980). URL 3929/ethz-a-010514644. https://doi.org/10.1119/1.12061. [11] Woods, M. P., Silva, R., P¨utz, G., Stupar, S. & [25] Weinberg, S. Gravitation and cosmology: Prin- Renner, R. Quantum clocks are more accurate ciples and Applications of the General Theory of than classical ones. arXiv: 1806.00491 (2018). Relativity. ed. John Wiley and Sons, New York URL https://arxiv.org/abs/1806.00491. (1972). URL https://www.wiley.com/en-us/ [12] Salecker, H. & Wigner, E. P. Quantum limita- Gravitation+and+Cosmology%3A+Principles+ tions of the measurement of space-time distances. and+Applications+of+the+General+Theory+ Phys. Rev. 109, 571–577 (1958). URL https: of+Relativity-p-9780471925675. //doi.org/10.1007/978-3-662-09203-3_15. [26] Zych, M. & Brukner, C.ˇ Quantum formula- [13] Lindkvist, J. et al. with macro- tion of the einstein equivalence principle. Nat. scopic clocks in superconducting circuits. Phys. Phys. 1 (2018). URL https://doi.org/10. Rev. A 90, 052113 (2014). URL https://doi. 1038/s41567-018-0197-6. org/10.1103/PhysRevA.90.052113. [27] Hudson, R. L. When is the Wigner quasi- [14] Lorek, K., Louko, J. & Dragan, A. Ideal clocks probability density non-negative? Rep. Math. — a convenient fiction. Class. Quantum Gravity Phys. 6, 249–252 (1974). URL https://doi. 32, 175003 (2015). URL https://doi.org/10. org/10.1016/0034-4877(74)90007-X. 1088/0264-9381/32/17/175003. [28] Woods, M. P. & Alhambra, A.´ M. Continu- [15] Lock, M. P. E. & Fuentes, I. Quantum ous groups of transversal gates for quantum er- and classical effects in light-clock falling in ror correcting codes from finite clock reference schwarzschild geometry. Class. Quantum Gravity frames. Quantum 4, 245 (2020). URL https: 36, 175007 (2019). URL https://doi.org/10. //doi.org/10.22331/q-2020-03-23-245. 1088/1361-6382/ab32b1. [29] Chou, C.-W., Hume, D., Rosenband, T. & [16] Anastopoulos, C. & Hu, B.-L. Equivalence prin- Wineland, D. Optical clocks and relativity. Sci- ciple for quantum systems: dephasing and phase

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 8 ence 329, 1630–1633 (2010). URL https://doi. [43] Sinha, S. & Samuel, J. Atom interferometry and org/10.1126/science.1192720. the gravitational redshift. Class. Quantum Grav- [30] Brewer, S. M. et al. 27al+ quantum-logic clock ity 28, 145018 (2011). URL https://doi.org/ with a systematic uncertainty below 10−18. Phys. 10.1088%2F0264-9381%2F28%2F14%2F145018. Rev. Lett. 123, 033201 (2019). URL https:// [44] Rankovi´c,S., Liang, Y.-C. & Renner, R. Quan- doi.org/10.1103/PhysRevLett.123.033201. tum clocks and their synchronisation - the Alter- [31] Mantina, M., Chamberlin, A. C., Valero, R., nate Ticks Game. arXiv:1506.01373 (2015). URL Cramer, C. J. & Truhlar, D. G. Consistent van http://arxiv.org/abs/1506.01373. der waals radii for the whole main group. J. [45] Erker, P. et al. Autonomous quantum clocks: Phys. Chem. A 113, 5806–5812 (2009). URL does thermodynamics limit our ability to mea- https://doi.org/10.1021/jp8111556. sure time? Phys. Rev. X 7, 031022 (2017). [32] Ludlow, A. D., Boyd, M. M., Ye, J., Peik, E. & URL https://doi.org/10.1103/PhysRevX.7. Schmidt, P. O. Optical atomic clocks. Rev. Mod. 031022. Phys. 87, 637 (2015). URL https://doi.org/ [46] Woods, M. P. Autonomous Ticking Clocks from 10.1103/RevModPhys.87.637. Axiomatic Principles. arXiv: 2005.04628 (2020). [33] Kovachy, T. et al. at the URL https://arxiv.org/abs/2005.04628. half-metre scale. Nature 528, 530 (2015). URL [47] Schwarzhans, E., Lock, M. P. E., Erker, P., Friis, https://doi.org/10.1038/nature16155. N. & Huber, M. Autonomous temporal prob- [34] Page, D. N. & Wootters, W. K. Evolu- ability concentration: Clockworks and the sec- tion without evolution: Dynamics described ond law of thermodynamics. arXiv:2007.01307 by stationary . Phys. Rev. D 27, (2020). URL https://arxiv.org/abs/2007. 2885 (1983). URL https://doi.org/10.1103/ 01307. PhysRevD.27.2885. [48] Zyczkowski,˙ K., Horodecki, P., Sanpera, A. & [35] Smith, A. R. H. & Ahmadi, M. Relativistic Lewenstein, M. Volume of the set of separable quantum clocks observe classical and quantum states. Phys. Rev. A 58, 883 (1998). URL https: time dilation. arXiv:1904.12390 (2019). URL //doi.org/10.1103/PhysRevA.58.883. https://arxiv.org/abs/1904.12390. [49] Castro-Ruiz, E., Giacomini, F. & Brukner, C.ˇ [36] Hoehn, P. A., Smith, A. R. H. & Lock, M. Entanglement of quantum clocks through grav- P. E. Equivalence of approaches to rela- ity. Proc. Natl. Acad. Sci. U.S.A. 201616427 tional quantum dynamics in relativistic settings. (2017). URL https://doi.org/10.1073/pnas. arXiv:2007.00580 (2020). URL https://arxiv. 1616427114. org/abs/2007.00580. [50] Paige, A. J., Plato, A. D. K. & Kim, [37] Colella, R., Overhauser, A. W. & Werner, M. S. Classical and nonclassical time dila- S. A. Observation of gravitationally induced tion for quantum clocks. Phys. Rev. Lett. 124, quantum interference. Phys. Rev. Lett. 34, 160602 (2020). URL https://doi.org/10. 1472 (1975). URL https://doi.org/10.1103/ 1103/PhysRevLett.124.160602. PhysRevLett.34.1472. [51] Pikovski, I., Zych, M., Costa, F. & Brukner, [38] Roura, A. Gravitational redshift in quantum- C. Time Dilation in Quantum Systems clock interferometry. Phys. Rev. X 10, and Decoherence: Questions and Answers. 021014 (2020). URL https://doi.org/10. arXiv:1508.03296 (2015). URL http://arxiv. 1103/PhysRevX.10.021014. org/abs/1508.03296. [39] Peters, Achim and Chung, Keng Yeow and Chu, [52] Shankar, R. Principles of Quantum Mechanics Steven. Measurement of gravitational acceler- (Springer US, Boston, MA, 1994). URL https: ation by dropping atoms. Nature 400, 849– //doi.org/10.1007/978-1-4757-0576-8. 852 (1999). URL https://doi.org/10.1038/ [53] Hall, B. Lie Groups, Lie Algebras, and Repre- 23655. sentations: An Elementary Introduction. Gradu- [40] M¨uller,H., Peters, A. & Chu, S. A precision ate Texts in Mathematics (Springer International measurement of the gravitational redshift by the Publishing, 2015). URL https://doi.org/10. interference of matter waves. Nature 463, 926– 1007/978-3-319-13467-3. 929 (2010). URL https://doi.org/10.1038/ [54] Hoehn, P. A., Smith, A. R. H. & Lock, M. P. E. nature08776. The Trinity of Relational Quantum Dynamics. [41] Wolf, P. et al. Atom gravimeters and gravita- arXiv:1912.00033 (2019). URL https://arxiv. tional redshift. Nature 467, E1–E1 (2010). URL org/abs/1912.00033. https://doi.org/10.1038/nature09340. [55] Holevo, A. S. Probabilistic and Statistical As- [42] Does an atom interferometer test the gravita- pects of Quantum Theory, vol. 1 (North-Holland, tional redshift at the compton frequency? URL Amsterdam, 1982). URL https://doi.org/10. https://doi.org/10.1088%2F0264-9381% 1007/978-88-7642-378-9. 2F28%2F14%2F145017.

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 9 [56] Giovannetti, V., Lloyd, S. & Maccone, L. Ad- 5, 222 (2011). URL https://doi.org/10.1038/ vances in quantum . Nature photonics nphoton.2011.35.

Appendices A Relativistic coupling of the clock’s temporal and kinematic degrees of freedom

For completeness, we give a derivation of the coupling between the clock’s kinematic and time-measuring degrees of freedom, based on the one in [17]. We use the post-Newtonian metric to approximate a weak gravitational field. In coordinates x˜µ = t,˜ ~r corresponding to an observer at rest at r = ∞, this metric is given by ! !  2Φ(r) 2Φ(r)2  Φ(r)3  2Φ(r) Φ(r)3 g˜ = − 1 + + + O and g˜ = δ 1 − + O , (16) 00 c2 c4 c2 ij ij c2 c2 where r = |~r|, i and j label spatial coordinates, Φ(r) = −GM/r is the gravitational potential of the Earth and M is its mass. Let us now consider the laboratory frame xµ = (t, ~r), where t is the proper time of an observer at rest at ~r = ~r0. Transforming the metric in Eq. (16) to the laboratory frame, we have, !  Φ 2 Φ(r)3 g = 1 − 0 g˜ + O and g =g ˜ , (17) 00 c2 00 c2 ij ij where Φ0 := Φ(r0). In the laboratory frame the clock has energy Elab and spatial momentum ~p, with p = |~p|. Now consider the rest frame of the clock, in which it has energy Erest and zero spatial momentum. The norm of the clock’s four-momentum is a quantity, and therefore the same in both frames. Equating this norm in both frames and rearranging, one finds

g¯00E2 − p pjc2 E2 = rest j = −g E2 + p pjc2 , (18) lab g00 00 rest j

00 j ij where g¯ = −1 is the 00-component of the metric in the clock’s rest frame, pjp = g pipj (using the Einstein 00 summation convention), and we have used the fact that g00 = 1/g . Now, in order to mark the passage of time the clock must have some evolving internal structure, which is associated with some internal energy. Denoting the clock’s rest mass by m and its internal energy by Ec, then by mass-energy equivalence,

2 Erest = mc + Ec. (19)

Let us now restrict to one spatial dimension, namely the radial one, define the coordinate x := r − r0 and take the common approximation of the Newtonian potential around the point x = 0, i.e. Φ(x) ≈ Φ0 + gx, where g is the gravitational acceleration at x = 0. Then, inserting Eqs. (16), (17) and (19) into Eq. (18), one finds that

p2 p4  p2 gx E = mc2 + mgx + − + E 1 − + , (20) lab 2m 8m3c2 c 2m2c2 c2

2 2 4 2 4 2 2 4 where we have neglected terms of order (Ec/mc ) , (p/mc) , (gx) /c and gxp /m c and higher. In the following, we refer to such terms with the notation O(1/c4). Quantising Eq. (20), one finds that the clock evolves subject to the Hamiltonian ˆ ˆ ˆ ˆ H = Hc + Hk + Hck, (21) ˆ ˆ where upon quantisation Elab → H (which acts upon the space Hc ⊗ Hk) and Ec → Hc (which acts upon the space Hc), pˆ2 pˆ4 Hˆ = mc2 + mgxˆ + − (22) k 2m 8m3c2 represents the rest-mass-energy, gravitational potential energy and kinetic energy (with a lowest-order relativistic correction) of the clock, and  pˆ2 gxˆ Hˆ = Hˆ ⊗ − + , (23) ck c 2m2c2 c2 is a relativistic coupling of the temporal and kinematic degrees of freedom. We stress that Hˆ generates evolution with respect to the proper time of a classical observer at x = 0, i.e. the laboratory time t.

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 10 B Time dilation in generic quantum clocks

We now calculate the average time dilation experienced by a generic quantum clock evolving under the Hamil- tonian given in Eq. (21). We begin by employing perturbation theory to calculate the evolution of the clock state in the low-velocity, weak-gravity limit described in AppendixA (i.e. neglecting O(1/c4) terms). As in the main text, we choose units such that ~ = 1. ˆ ˆ ˆ ˆ ˆ ˆ Defining a free Hamiltonian by H0 := Hc +Hk, so the total Hamiltonian can be written H = H0 +Hck, where ˆ Hck is given in Eq. (23), one can find the evolution of the clock’s state via the . Denoting the evolution of the state in the interaction picture by ρI (t), and following standard perturbation theory (see, e.g. [52]), we have that it is related to the Schr¨odingerpicture by

I † † ρ (t) = U0 (t)ρ(t)U0(t), ρ(t) = U(t)ρ(0)U(t) , (24)

−itHˆ0 −itHˆ I where U0(t) = e and U(t) = e . Then ρ (t) satisfies, ∂ h i ρI (t) = Hˆ I (t), ρI (t) , (25) ∂t ck

ˆ ˆ ˆ I iH0t ˆ −iH0t where Hck(t) := e Hcke is the interaction Hamiltonian in the interaction picture. A solution to this dif- ferential equation can be obtained using the Dyson series. Neglecting terms of order O(1/c4), as in AppendixA, it yields Z t I 0 ˆ I 0 ρ (t) = ρ(0) − i dt [Hck(t ), ρ(0)]. (26) 0 Thus going back into the Schr¨odingerpicture,  t  ˆ ˆ ˆ Z ˆ −iH0t iH0t −iH0t 0 ˆ I 0 iH0t ρ(t) = e ρ(0)e − i e dt [Hck(t ), ρ(0)] e . (27) 0

Tracing over the kinematic space Hk, we find the reduced state corresponding to the temporal degree of freedom ρc(t) := trk[ρ(t)], which we write as

(1) ρc(t) = ρc,NR(t) + ρc (t), (28)

−iHˆct iHˆct where ρc,NR(t) := e ρc(0) e is the non-relativistic evolution of the reduced state, and the lowest-order relativistic correction to it is given by

t Z ˆ ˆ (1) 0  −iH0t ˆ I 0 iH0t ρc (t) = −i dt trk e [Hck(t ), ρ(0)]e , (29) 0 Now, we consider a clock whose temporal and kinematic degrees of freedom are initially uncorrelated, writing

ρ(0) = ρc(0) ⊗ ρk(0), (30)

ˆ ˆ ˆ ˆ pˆ2 Φ(ˆx) for some initial reduced states ρc(0) and ρk(0). Then, writing Hck = Hc ⊗ Vk, i.e. Vk := − 2m2c2 + c2 , we calculate the integrand of Eq. (29) to find

ˆ ˆ h i h ˆ ˆ i  −iH0t ˆ I 0 iH0t ˆ iHkt ˆ −iHkt trk e [Hck(t ), ρ(0)]e = Hc, ρc,NR(t) tr e Vke ρk(0) . (31)

Using the relation h i 1 1 eXˆ Y e−Xˆ = Yˆ + X,ˆ Yˆ + [X,ˆ [X,ˆ Yˆ ]] + [X,ˆ [X,ˆ [X,ˆ Yˆ ]]] + ··· (32) 2! 3! for some Xˆ and Yˆ (see Prop. 3.35 in [53]), one finds, by neglecting terms of order 1/c4 and higher, that

 2  ˆ ˆ 1 pˆ 2gt eiHktVˆ e−iHkt = − + gxˆ + pˆ − g2t2 . (33) k c2 2m2 m

Performing the integration in Eq. (29), we find

(1) h ˆ i ρc (t) = −i Hc, ρc,NR(t) tR(t), (34)

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 11 h pˆ2 gxˆ pgtˆ g2t2  i where R(t) := tr − 2m2c2 + c2 + mc2 − 3c2 ρk(0) . Having obtained the evolution of the clock’s temporal state, one can calculate the average clock time, giving ˆ ˆ n  ˆ o hTci(t) = hTciNR(t) + tR(t) 1 + tr E(t) (35)

ˆ  ˆ  where hTciNR := tr Tc ρc,NR(t) is the clock time in the absence of relativistic effects, and where the error ˆ ˆ ˆ operator E(t) := −i[Tc, Hc]ρc,NR(t) − ρc,NR(t) was introduced in Sec.2 of the main text.

C Idealised clocks, covariant POVMs and higher moments (in the absence of relativistic effects)

In the main text, we defined an idealised clock as one whose first-moment operator satisfies the canonical  ˆ ˆ  1 commutation relations Tc, Hc = i c (in the Heisenberg form), and used this to motivate the definition of the error operator Eˆ(t) in Eq. (1). We now find covariant POVMs {Fˆ(s)}s∈S which are absolutely continuous, that either form an idealised clock, or whose deviation from this condition is covered by a particular form of the error operator Eˆ(t). In doing so, we show that our choice of the definition of the error operator is well motivated. We will look at both cases in which S is bounded or unbounded. We start with the case that S is bounded, and denote the boundaries by S = [s0, s1]. We expand the domain of Fˆ to R by periodic extension: Fˆ(s + n(s1 − s0)) := Fˆ(s), s ∈ S, n ∈ Z. To start with, we will only need the covariance property of the POVMs and the assumption that S is bounded. ˆ ˆ ˆ † Writing the evolution of Tc in the , and using the covariant property F (s) = Uc(s)F (0)Uc (s); −iHˆcs Uc(s) := e , we find for all t ∈ R:

Z Z s1 ˆ † ˆ ˆ ˆ Tc(t) = Uc (t)TcUc(t) = ds s F (s − t) = ds f(s)F (s − t) S s 0 (36) Z s1−t Z s1 = ds f(t + s)Fˆ(s) = ds f(t + s)Fˆ(s), s0−t s0 where f(s) is the sawtooth function: f(s + n(s1 − s0)) := s, s ∈ S, n ∈ Z. By taking limits s0 → −∞ and/or R ˆ 1 s1 → +∞ in Eq. (36) and noting that POVMs form a resolution of the identity, s∈S F (s) = c; we find ˆ 1 ˆ Tc(t) = t c + Tc, (37) for the three unbounded cases:

1) S = [s0, ∞) and t ≥ 0. 2) S = (−∞, s1] and t ≤ 0. 3) S = (−∞, ∞) and t ∈ R. d ˆ †  ˆ ˆ  On the other hand, the Heisenberg-picture equation of motion is dt Tc(t) = −i Uc (t) Tc, Hc Uc(t). Thus, equating ˆ 1 ˆ this with the time derivative of Tc(t) = t c + Tc, we arrive at h i ˆ ˆ 1 Tc, Hc = i c (38) for cases 1), 2) and 3) above. Eq. (38) represents the canonical commutation relations of a specialised Heisenberg form called the Weyl form. We now investigate the case in which S is bounded. From Eq. (36) we find

Z s1−s0−t Z s1 ˆ ˆ ˆ Tc(t) = ds (t + s)F (s) + ds (t + s − (s1 − s0)) F (s) s0 s1−s2−t (39) Z s1 1 ˆ ˆ = t c + Tc + (s1 − s0) dsF (s). s1−s0−t d ˆ 1 ˆ 1 † ˆ Differentiating under the integral sign, we find dt Tc(t) = c+(s0−s1)F (s1−s0−t) = c+(s0−s1)Uc (t)F (0)Uc(t). Hence taking into account Heisenberg-picture equation of motion as before, we obtain h i ˆ ˆ 1 ˆ Tc, Hc = i c + (s0 − s1)F (0). (40)

 ˆ ˆ   The family of quantum states for which the commutator Tc, Hc is well defined, is the set SComm := ρ ∈ ˆ ˆ ˆ ˆ D(Hc) Tcρ ∈ Dom(Hc)& Hcρ ∈ Dom(Tc) , where D(Hc) is the set of density operators on Hc and Dom is

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 12 the domain of the self-adjoint operator in question. Thus Eq. (40) satisfies the canonical commutation relations ˆ on the subset of states C := {ρ ∈ Scomm | F (0)ρ = 0} ⊆ Scomm. If C is a dense subset of Scomm, one says that ˆ ˆ the pair Tc and Hc satisfy the Heisenberg form of the canonical commutation relation. The special case of S bounded and POVMs which are projective values measures, was studied in [21], where it was shown that C is indeed a dense subset of Scomm. Another example of a Heisenberg form of the canonical commutation relations is a “particle in a box” which is also reviewed in [21]. These (and other) aspects of covariant POVMs are discussed in detail in [54]. Moreover, our definition of an error operator Eˆ(t) in Eq. (1) readily accommodates the situation in Eq. (40). In particular, we have Eˆ(t) = −i(s0 − s1)Fˆ(0)ρ(t). We now move on to examine higher moments of covariant POVMs. Assuming all moments up to and including the nth moment are well defined, we have

ˆ(n) h ˆ(n) i h † ˆ(n) i hTc i(t) := tr Tc ρc(t) = tr Uc (t)Tc Uc(t)ρc(0) Z  n ˆ † = tr ds f(s) Uc(s − t)F (0)Uc (s − t)ρc(0) S (41) n Z  X n = tr ds0 f(s0 + t)n U (s0)Fˆ(0)U †(s0)ρ (0) = tn−khTˆ(k)i(0), c c c k c S k=0 ˆ(0) where hTc i(0) = 1 for all initial states. From this Eq., it follows, for example, that the standard deviation of ˆ Tc is time independent and only depends in the initial state ρc. For an idealised clock, a projection-valued measure is an example of a covariant POVM. To see this, ob- ˆ ˆ(1) serve that one can write in this case F (s) = |sihs|c, with Tc |si = s |si. From there, one can use the fact † ˆ(1) (1) 0 0 that an idealised clock satisfies Uc (s)Tc Uc(s) = Tc + s to show that Uc(s ) |si = |s + s i, and therefore ˆ † ˆ Uc(s)F (0)Uc (s) = F (s), i.e. the covariance property. If, in addition, the possible clock times span the entire real line (as in the case considered by Pauli [5]), then the higher moments satisfy Eq. (41), and again the variance is therefore time-independent, regardless of the initial clock state.

D Examples of relativistic time dilation in quantum clocks D.1 An idealised quantum clock ˆ ˆ Recall that an idealised clock is one satisfying the canonical commutation relation, [Tc, Hc] = i, and consequently ˆ ˆ ˆ E(t) = 0. Let Tc =x ˆc/c + a1c and Hc = cpˆc, where xˆc and pˆc are position and momentum operators and ˆ a ∈ R is chosen such that hTci(0) = 0. We first consider the case of “classical” motion, i.e. where the initial kinematic state of the clock is a pure Gaussian wavepacket with mean momentum p¯0, standard deviation of the momentum σp and mean position x¯0, i.e. ρk(0) = |ψihψ|k with

2 Z p−p¯0  1 − −ix¯ (p−p¯ ) |ψi = dp ψ(p) |pi ; ψ(p) := e 2σp e 0 0 . (42) k k 2 1/4 (2πσp)

2 2 2 p¯0+σp gx¯0 p¯0gt 1 gt  We then have that R(t) = − 2m2c2 + c2 + mc2 − 3 c and therefore " # p¯2 + σ2 gx¯ p¯ gt 1 gt2 hTˆ i(t) = 1 − 0 p + 0 + 0 − t. (43) c 2m2c2 c2 mc2 3 c

This expression is the average proper time of a classical clock (see Eq. (4) in the main text) with position x¯0 and a random momentum following a Gaussian distribution with mean p¯0 and standard deviation σp. To summarise, an idealised clock in a classical (i.e. Gaussian) state of motion experiences the usual classical relativistic time dilation. Let us now consider a nonclassical initial kinematic state, specifically a superposition of two Gaussian states, sometimes referred to as a cat state. We will contrast this with the case of comparable classical probabilistic mixture of the two states. Let

Z p−p¯ 2 1 − −ix¯ (p−p¯) |ψ i := dp ψ (p) |pi ; ψ (p) := e 2σp e j ; j ∈ {1, 2}. (44) j k j k j 2 1/4 (2πσp) √ √ We now take the initial kinematic state to be in the superposition |ψi = √1 α |ψ i + eiθ 1 − α |ψ i  k N 1 k 2 k ˆ for some α ∈ (0, 1) and θ ∈ R, and find the average clock time, which we denote hTcisup, using Eq. (35). We

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 13 then repeat this procedure for the initial kinematic state in the probabilistic mixture, ρk(0) = α |ψ1ihψ1| + (1 − ˆ α) |ψ2ihψ2|, and denote the average clock time in that case by hTcimix(t). From the linearity of the trace, we see immediately that ˆ ˆ ˆ hTcimix(t) = αhTciψ1 (t) + (1 − α)hTciψ2 (t), (45) ˆ where hTciψj (t) is the average clock time of a clock with initial kinematic state |ψjik. One then finds that ˆ ˆ hTcisup(t) = hTcimix(t) + Tcoh(t), (46) where " 2 2 2 # N − 1 ∆x0 σv g∆x0 σv t Tcoh(t) := 2 − 2 (1 − 2α) − 2 2 ∆x0(p − mgt) tan θ (47) N 2σx c c c 2 is a contribution to the clock time arising due to interference between the two Gaussian states constituting the coherent superposition (see main text), and where σv = σp/m and σx = 1/2σp are respectively the standard deviation of the velocity and position of the Gaussian wavepackets used to construct the initial state, and ∆x0 =x ¯2 − x¯1. We recall that ~ = 1, and note that the normalisation factor N is given by

2 p − 1 ∆x0  N = 1 + 2 α(1 − α)e 2 2σx cos θ. (48)

Setting θ = 0 in Eq. (47), one arrives at Eq. (12) in the main text.

D.2 Salecker-Wigner-Peres clock Consider a d-dimensional Hilbert space and a clock Hamiltonian with equally-spaced energy levels, i.e. ˆ Pd−1 Hc = j=0 jω |ejihej|, for some ω > 0. This corresponds, for example, to the case where the clock is a spin-j system, and then d = 2j + 1. One can obtain a so-called time basis {|θmi} by taking the discrete Fourier transform of the energy eigenstates,

d−1 1 X −i2πjm/d |θmi = √ e |eji , m = 0, 1, . . . , d − 1 (49) d j=0

ˆ Pd−1 T0 2π which are used to construct a clock-time operator Tc = m=0 m d |θmihθm|, with T0 := ω . When the clock ˆ is in the state |θmi, we have hTci = mT0/d. Choosing an initial state ρc(0) = |θ0ihθ0|, then the clock state Z ˆ will at regular time intervals tm := mT0/d, m ∈ “focus” into one of the time eigenstates, i.e. hTciNR (tm) = mT0/d (m mod d). At intermediate times t 6= tm, on the other hand, the clock will be in a superposition of its time eigenstates (see e.g. [6], Appendix B). This setup is known as the Salecker-Wigner-Peres clock (or alternatively, the Larmor clock) [12, 24]. ˆ ˆ A straightforward calculation gives tr[[Tc, Hc] |θmihθm|] = 0 for all m = 0, 1, . . . , d − 1, and therefore from the definition of the error operator Eˆ(t), we have tr[Eˆ(tm)] = −1 for all m ∈ Z and for all clock dimensions d ∈ N+. Consequently, when the Salecker-Wigner-Peres clock is in a time eigenstate, Eq. (35) tells us that ˆ ˆ hTci(tm) = hTciNR(tm) for any state of the clock’s kinematic degrees of freedom. To reiterate, this clock is a special case where the average relativistic time-dilation is exactly cancelled by the quantum error of the clock at regular time intervals regardless of how large d is and regardless of the motional state. In this regard, we note that the Salecker-Wigner-Peres clock was used in [50], where it was concluded that “quantum clocks do not witness classical time dilation”.

D.3 The Quasi-Ideal clock Consider the same setting as for the Salecker-Wigner-Peres clock described above, namely the clock Hilbert ˆ ˆ space, Hamiltonian Hc and clock-time operator Tc, but now with the initial temporal state ρc(0) = |Ψ(m0)ihΨ(m0)|, with d−1 X |Ψ(m0)i = gm0 (m) |θmi , (50) m=0 g (m) m σ¯ ∈ (0, d) where m0 √ is a complex Gaussian distribution centred on 0 and with standard deviation c . The choice σ¯c = d corresponds to the case in which the standard deviation in both the time basis {|θmi} and energy basis time basis {|eji} is approximately the same. For other choices, one obtains a clock whose behaviour tends

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 14 to that of the idealised clock when σ¯c → ∞ and d/σ¯c → ∞ as d → ∞. This is then the Quasi-Ideal clock, introduced in [6]. We now show that the contribution of the quantum error to the relativistic time dilation, i.e. tr[Eˆ(t)] (see Eq. (35)), tends exponentially to zero with increasing clock dimensionality. In other words, we show that the relativistic time dilation experienced by the idealised clock can be approximated arbitrarily well by a Quasi-Ideal clock of high enough dimension. Physically speaking, this is because the Quasi-Ideal clock can move at constant “velocity” in the time basis with minimal dispersion. From the definition of the error operator Eˆ(t), we have ˆ  ˆ ˆ  tr[E(t)] = −itr [Tc, Hc]ρc,NR(t) − 1 (51)

Choosing the initial clock state |Ψ(0)ihΨ(0)|, the non-relativistic evolution of the state ρc,NR(t) is given by,

−iHˆct iHˆct ρc,NR(t) = e |Ψ(0)ihΨ(0)| e , (52) Now, Theorem 8.1 in [6] states that

−iHˆct e |Ψ(0)i = |Ψ(t/T0)i + |εi , (53) where the specific form of |εi is irrelevant for the present purpose, only that as d → ∞ and d/σ¯c → ∞,

π d2  − π 2  p 4 σ¯2 − σ¯  hε|εi = O poly(d) e c + e 4 c , (54) and furthermore Theorem 11.1 in [6] states that ˆ ˆ [Tc, Hc] |Ψ(m)i = i |Ψ(m)i + |εcommi ∀ m ∈ R (55) where, again, the specific form of |εcommi is irrelevant, only that as d → ∞,

π d2  − π 2  p 4 σ¯2 − σ¯  hεcomm|εcommi = O poly(d) e c + e 4 c . (56)

Using these two theorems, one finds that   ˆ ˆ ˆ   tr[E(t)] = −itr |εcommi hΨ(t/T0)| + [Tc, Hc] |εihΨ(t/T0)| + |Ψ(t/T0)ihε| + |εihε| (57)

h i h  i ˆ ˆ ≤ tr |εcommi hΨ(t/T0)| + tr [Tc, Hc] |εihΨ(t/T0)| + |Ψ(t/T0)ihε| + |εihε| (58) We can now bound the terms on the r.h.s. of Eq. (58). To start with p tr[|εcommihΨ(t/T0)|] = hεcomm|Ψ(t/T0)i ≤ hεcomm|εcommi

π d2 (59)  − π 2  4 σ¯2 − σ¯  = O poly(d) e c + e 4 c

2 ˆ ˆ ˆ† ˆ † ˆ ˆ Using the Cauchy-Schwarz inequality, tr[AB] ≤ tr[A A]tr[B B] for bounded linear operators A, B; one finds that for two kets |ai, |bi h i 2 h i ˆ ˆ ˆ ˆ 2 tr [Tc, Hc] |aihb| ≤ −tr [Tc, Hc] ha|ai hb|bi (60)  h ˆ ˆ 2i h ˆ2 ˆ 2i  ≤ 2 tr (TcHc) + tr Tc Hc ha|ai hb|bi (61)   ˆ ˆ 2 ˆ2 ˆ 2 ≤ 2 (TcHc) + Tc Hc ha|ai hb|bi (62) 2 2 ˆ ˆ 2 2 2 ≤ 4 Tc Hc ha|ai hb|bi ≤ 4(T0d) (d ω) ha|ai hb|bi (63) ≤ 16π2d6 ha|ai hb|bi . (64) Therefore, using Eqs. (59), (64), and (54); one finds from Eq. (58)

2 π d 2  − 4 2 − π σ¯  ˆ σ¯c 4 c  tr[E(t)] = O poly(d) e + e (65) R η for all t ∈ . We therefore find that for any initial clock width of the form σ¯c = dc for some fixed η > 0, tr[Eˆ(t)] is upper-bounded by a quantity which decreases exponentially with dimension. Furthermore, note that ˆ for 0 < η < 1/2, the standard deviation of Tc for the initial Quasi-Ideal clock state approaches zero as d → ∞. Thus in the limit of large dimension, the relativistic time dilation (Eq. (35)) is that of the idealised clock, up to errors that decay exponentially in the dimension.

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 15 D.4 The qubit-phase clock Another clock model sometimes found in the literature (e.g. [16, 19, 49]) uses the phase difference between the two states of a qubit system to estimate elapsed time. This can be considered a simplified model of the Ramsey method sometimes used in atomic clocks (see e.g. [32] Sec. IVA for a discussion of clock interrogation schemes), where it is assumed that the is perfectly tuned to the reference frequency. The qubit is put into the initial √1 (|0i + |1i) Hˆ = − ω |0ih0| + ω |1ih1| state 2 , whose non-relativistic evolution according to the Hamiltonian c 2 2 results in the state 1 |θi := √ |0i + e−iθ |1i (66) 2 where θ = ωt. The covariant (see Sec.2 and AppendixC) POVM corresponding to the estimation of this ˆ 1 phase is given by {F (θ)}θ∈[0,2π) with F (θ) = π |θihθ|. This measurement corresponds to a maximum-likelihood estimation of θ, and therefore the laboratory time (see e.g. [55], Sec. 4.4). One can then directly calculate the error operator: ˆ E(t) = −(1c + σx)ρc(t). (67) Since tr[Eˆ(t)] = 1 + cos(ωt), we see that this is only a good clock in the sense defined in Sec.2 around θ = π, which corresponds to the time at which the clock state |θi is orthogonal to the initial state. This limitation can be overcome by performing the phase estimation on many copies of the qubit system (see e.g. [56]). Now, carrying this analysis into a relativistic context, Eq. (6) gives the average time dilation experienced by the qubit-phase clock as ˆ ˆ hTci(t) = hTciNR(t) + tR(t) [2 + cos(ωt)] , (68) which is of course equal to the average time dilation experienced by the idealised clock when θ = π.

E The effect of the coupling on a clock’s precision q ˆ2 ˆ 2 We now show how a clock’s precision, as quantified by the standard deviation σT := hTc i − hTci , changes over time as a result of the relativistic coupling between temporal and kinematic degrees of freedom. For simplicity, we now consider flat (Minkowski) spacetime. Here we work to a precision one order higher than before, i.e. neglecting only O (ˆp/mc)6 terms. This is because, as we shall see, relativistic effects on the precision of the idealised clock vanish below O(1/c4). Repeating the procedure outlined in AppendixA, one finds that the relativistic coupling between temporal and kinematic degrees of freedom is now !  pˆ 6 Hˆ = Hˆ ⊗ Wˆ + O (69) ck c k mc

ˆ 2 2 2 4 4 4 with Wk := −pˆ /2m c + 3ˆp /8m c . ˆ ˆ ˆ ˆ Now, since in this case [Hck, Hk] = [Hck, Hc] = 0, we can simply calculate the relevant observables without going into the interaction picture. Using Eq. (32), we have

" ∞ # X 1 hTˆni(t) = tr [i(1 + Wˆ )t]q [Hˆ , Tˆn] (70) c q! k k c c q q=0

th where [· , ·]q denotes the q -order nested commutator, i.e. [A,ˆ Bˆ]0 := Bˆ and [A,ˆ Bˆ]q := [A,ˆ [A,ˆ Bˆ]q−1] for q > 0. ˆ q Applying the Binomial theorem to (1k + Wk) and truncating at the appropriate order, one finds, after some algebra ˆn ˆn h n ˆ ˆn ˆ 2 ˆ ˆ ˆn ˆ 2o i hTc i(t) = hTc iNR(t) + tr it[Hc, Tc ] ⊗ Wk − t [Hc, [Hc, Tc ]] ⊗ Wk ρNR(t) (71)

−i(Hˆc+Hˆk)t i(Hˆc+Hˆk)t where ρNR(t) := e ρ(0)e is the non-relativistic evolution of the clock’s total state and ˆn ˆn ˆ 2 6 hTc iNR(t) := tr[Tc ρc,NR(t)], and where it is understood that Wk contains an O (ˆp/mc) term to be ne- glected. Assuming an uncorrelated initial state as in AppendixB, i.e. ρ(0) = ρc(0) ⊗ ρk(0), and further defining ˆ n ˆ n hWk i0 := tr[Wk ρk(0)], we have ˆn ˆn ˆ  ˆ ˆn  2 ˆ 2  ˆ ˆ ˆn  hTc i(t) = hTc iNR(t) + ithWki0 tr [Hc, Tc ]ρc,NR(t) − t hWk i0 tr [Hc, [Hc, Tc ]]ρc,NR(t) . (72)

Then by a straightforward, if lengthy, calculation, one finally finds that σT (t), separates into

σT (t) = σT,NR(t) + σT,I(t) + σT,NI(t), (73)

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 16 where σT,NR(t) is the clock’s standard deviation in the absence of relativistic effects, σT,I(t) is a contribution which remains in the case of an idealised clock, given by

2 4 2 t hpˆ i + σp2 σT,I(t) := 4 4 (74) 8σT,NR(t) m c

2 where σp2 is the standard deviation of pˆ , and σT,NI(t) is a contribution arising due to the non-idealised nature of the clock, given by ˆ hWki0 t n ˆ ˆ† ˆ ˆ ˆ o σT,NI(t) := tr[(E(t) + E (t))Tc] − 2hTciNR(t)tr[E(t)] 2σT,NR(t) 2  ˆ  hWki0 t 2 n ˆ ˆ† ˆ ˆ ˆ o − 3 tr[(E(t) + E (t))Tc] − 2hTciNR(t)tr[E(t)] 8σT,NR(t)  2 ˆ (75) hWki0 t n o − 2tr[Eˆ(t)] + tr[Eˆ(t)]2 2σT,NR(t) ˆ 2 2 n hWk i0 t ˆ ˆ ˆ ˆ ˆ ˆ ˆ ˆ ˆ† ˆ ˆ − 2tr[E(t)] + itr[(HceˆTc − TceˆHc)ρc,NR(t) + HcTcE(t) − E (t)TcHc] 2σT,NR(t) ˆ ˆ ˆ ˆ† o + 2ihTciNR(t)tr[Hc(E(t) − E (t))]

ˆ ˆ ˆ 1 4 where E(t) was defined above, and eˆ := i[Hc, Tc] − c. Note that σT,I(t) = O (ˆp/mc) , which necessitated that we work to one order higher of precision than when we calculated the clock time (AppendixB).

Accepted in Quantum 2020-08-07, click title to verify. Published under CC-BY 4.0. 17