Simulating superluminal propagation of Dirac particles using trapped ions
Abstract
Simulating quantum phenomena in extreme spacetimes in the laboratory represents a powerful approach to explore fundamental physics in the interplay of quantum field theory and general relativity. Here we propose to simulate the movement of a Dirac particle propagating with a superluminal velocity caused by the emergent Alcubierre warp drive spacetime using trapped ions. We demonstrate that the platform allows observing the tilted lightcone that manifests as a superluminal velocity, which is in agreement with the prediction of general relativity. Furthermore, the Zitterbewegung effect arising from relativistic quantum mechanics persists with the superluminal propagation and is experimentally measurable. The present scheme can be extended to simulate the Dirac equation in other exotic curved spacetimes, thus provides a versatile tool to gain insights into the fundamental limit of these extreme spacetimes.
I Introduction
Quantum field theory in curved spacetime [1, 2, 3], as a semi-classical approach to study the movement of quantized matter fields in a fixed gravitational background, has predicted many striking phenomena such as the famous Hawking radiation and cosmological particle production. However, confirmation of those extremely weak quantum effects in real gravity experiments remains highly unlikely with present technology [4, 5]. Following Unruh’s pioneering work [6], much attention has turned into sound waves travelling against the classical [7] and quantum fluids [8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19], or light in nonlinear optical platforms [20, 21, 22, 23] to effectively emulate quantum field theory in curved spacetime. Remarkably, these relevant analogue experiments among many others [24, 25] has brought us a better understanding of the relationship between space-time structure and quantum theory [26, 27, 28].
Furthermore, recent developments of precise and flexible quantum control also facilitate a quantum simulation of related relativistic phenomena [29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40]. As a representative example, intriguing quantum dynamics related to a Dirac particle moving in curved spacetime has attracted increasingly investigations via the particularly promising trapped-ion platform [41, 42, 43, 44]. The exceptional controllability of quantum simulators makes it feasible to emulate a Dirac particle in "exotic" curved spacetimes [43]. As a particularly important example, the Alcubierre warp-drive, consistent with Einstein’s field equations, can result in faster-than-light propulsion of a spaceship [45]. Compared to the extreme difficulty in realizing such a time-machine model (i.e. creating warp bubbles) in the actual world [46, 47, 48], observation of analogue phenomenon with well-controllable quantum simulators in the laboratory should be more accessible. Besides, merging of fundamental concepts from different fields when implementing such novel quantum simulation, including gravitation, quantum squeezing and quantum entanglement, would provide a fruitful way to reveal unique features of quantum effects in both the exotic curved spacetimes [43] and the basic light-matter interaction models [42].
In this Letter, we propose a trapped-ion quantum simulation of a Dirac particle propagating with a superluminal velocity caused by the emergent Alcubierre warp drive spacetime [45]. The Hamiltonian from the Dirac equation in the Alcubierre (1+1)-dimensional universe is mapped onto a spin-boson interaction quantum model, which can be further realized by a combination of sideband drives and a periodic modulation of the trapping potential. Remarkably, the tunability of system parameters in the well-controlled trapped-ion platform allows us to access the crossover from flat to curved spacetimes. Using exact numerical simulations, we demonstrate that this platform is able to observe analogue superluminal travel of Dirac particles in the tilted light cones, as well as the Zitterbewegung effect of massive Dirac particles incorporating the tilt of the Dirac cone. The extension of the present scheme to simulate the Dirac equation in more general exotic curved spacetimes would make it possible to explore intriguing phenomena arising from the interplay of quantum field theory and general relativity in well-controllable quantum experiments in the laboratory.
II Simulation of Alcubierre metric using trapped ions
We start by first introducing the Alcubierre metric [45, 49], which is given by
(1) |
where is the velocity associated with a certain trajectory . In the (1+1)-dimensional case, the light cones at a point in the - plane are specified by the curves emerging from the point with , namely
(2) |
If , the the spacetime is flat. Otherwise the corresponding light cones are tipped over and the particle travelling inside the light cone can have a velocity faster than the light speed in the flat spacetime, which is theoretically consistent with the framework of general relativity [45].
Extension of the Dirac equation into curved spacetimes successfully merges quantum mechanics with the general relativity. In the (1+1)-dimensional spacetime with signature , the Dirac equation reads [50, 41]
(3) |
where the Greek letter and the Latin letter denote the coordinates of curved spacetime and local rest frame respectively, is the mass of the quantized field, is the determinant of the metric tensor, and is the vielbein allowing the constant Dirac matrices to act at each spacetime point. Here, we choose the chiral representation such that , , with the Pauli matrices. In the specific case of Alcubierre metric, the Dirac equation can be transformed into the Schrödinger equation of the following form [51]

(4) |
from which the Hamiltonian that governs the above dynamical evolution can be written as
(5) |
Here we introduce the operators and
(6) |
to rescale the spatial coordinate of the simulated Dirac particle by a dimensionless factor . Note that and will be the observables of the trapped ion in our proposed simulation platform. We remark that the second term of the Hamiltonian in Eq.(5), which is linearly dependent on the momentum , is an evidence of relativistic physics [52]. Based on the standard commutation relation , the operators and can be further mapped to a bosonic field of frequency
(7) |
where is a constant with the dimension of mass. By substituting Eq.(19) into Eq. (5) and applying the Hadamard transformation (i.e. , ), the Hamiltonian can be rewritten as
(8) |
To simulate the above Hamiltonian in Eq.(20) with the tunable parameter , we consider a setup of a single ion trapped above a linear surface electrode radio-frequency trap. The radial motional mode of the trapped ion with a frequency models a harmonic oscillator and its two internal ground hyperfine levels with the transition frequency plays the role of an effective spin-1/2, denoted as and . As an example, the present scheme may employ a trapped 25Mg+ ion hyperfine qubit, with an out-of-phase radial motional mode frequency MHz [53]. The qubit states are chosen as and [53, 54], where is the total angular momentum and is its projection along the quantization axis.
The spin-harmonic oscillator coupling of the form in Eq.(20) can be realized by implementing the MølmerSørensen interaction [55] via simultaneous blue and red sideband drives at frequencies and respectively, which in real experiments can be generated by using oscillating near-field magnetic field gradients [56]. In addition, we introduce a periodic modulation of the single ion’s trapping potential of the strength at frequency to obtain the second term in Eq.(20). This can be experimentally implemented by applying an oscillating potential directly to the radio-frequency trapping electrodes, as demonstrated in Ref.[53]. Such a trapped-ion platform can be characterised by the following system Hamiltonian as [51]
where the frequencies of the blue and red sideband drives satisfy the condition . By making a displacement transformation and () and then moving to the interaction picture with respect to , the effective Hamiltonian can be approximated as [51]
(10) |
under the conditions of . Therefore, it can be seen that the Hamiltonian in Eq.(31) as implemented in the trapped-ion platform is equivalent to the Hamiltonian in Eq.(20) that governs the Dirac equation in the Alcubierre (1+1)-dimensional universe with the parameter correspondence as , , and .
We remark that the experimental tunability of the parameters in the well-controllable trap-ion platform allows access to the crossover from flat to curved spacetimes, namely the tilt of the Dirac equation can be tuned by choosing appropriate values of the parameter , see Figure 1. The angle of the simulated light cone is given by , which decreases for a larger value of , i.e. the velocity of the Dirac particle would be constrained in a smaller range. In the limit case of extremely large , and the angle of the light cone , the trajectory of the trapped ion can be seen as a counterpart of the closed timelike curves.
III Observation of analogue superluminal travel
To show the effective velocity of the the simulated Dirac particle, we solve the following Heisenberg equation of the trapped ion system [51]
(11) |
where represents the position operator of the Dirac particle, and with the evolution operator from the Hamiltonian in Eq.(5). We remark that Eq.(21) is consistent with the Alcubierre metric, since the two eigenvalues of correspond to the two opposite directions of the velocity. The mechanical degrees of freedom of the trapped ion is related to the spatial coordinate of the simulated Dirac particle by (see Eq.6), thus the velocity is in the range of .


As an illustrative example, we choose the initial state of the trapped ion as , where and represent the spatial and internal degrees of freedom of the trapped ion respectively. Without loss of generality, the spatial wavefunction satisfies a Gaussian distribution with the normalization factor. In Figure 2 (a) and (b), we show the trajectory (in the unit of ) of the simulated massless Dirac particle (i.e. ) as a function of the evolution time. According to Eq. (21), the trajectory of the evolving states from the initial spin states and (namely the two eigenstates of ) specifies the shape of the simulated light cone. It can be seen that, corresponding to massless Dirac particles, the light cones for different values of the parameter are consistent with the analytical results from classical general relativity theory [43]. In particular, the simulated light cones start to tip over as the parameter increases, which indicates that the spacetime becomes more curved. These numerical simulation results demonstrate that the quantum simulation using trapped-ions can confirm the superluminal propagation of Dirac particles in the Alcubierre warp drive spacetime.
In the cases of initial states with or , the velocity is well defined, and thus the wavepacket of the simulated Dirac particle remains localized, as shown by the blue line and green circles in Figure 2 (c). In contrast, if we start from a coherent superposition state , the wavepacket of the simulated Dirac particle has two peaks which correspond to the initial velocities respectively, resulting in an increase of the variance of the trajectory, see the orange line in Figure 2 (c).
IV Zitterbewegung effect with superluminal propagation
In addition to the analogue superluminal propagation, we also find that for massive Dirac particles with , the Zitterbewegung effect induces an oscillatory behaviour in the trajectory of the trapped ion. This effect is caused by the quantum superpositions of electron and positron solutions of the free wavepackets [52, 44] and may exist in both flat and curved spacetime [52, 42]. The equation of motion for the expectation value of the position operator in this case can be obtained as [51]
(12) |
where represents the Dirac Hamiltonian in the flat spacetime. From Eq.(12), one can see that the oscillatory behaviour is induced by the mass but independent of the parameter , and therefore the Zitterbewegung effect would persist with the superluminal propagation. In Figure 3 (a), we show that the Zitterbewegung effect can manifest for both constant and time-dependent in the trapped-ion simulator. As an example, we consider a time-dependent spacetime by choosing such that both the initial and final velocities are subluminal, which is derived from a trajectory of the simulated Dirac particle. This result is shown by the grey dashed line in Fig.3 (a), which clearly exhibits the Zitterbewegung effect.
In Figure 3 (b-c), we present the properties of the wavepacket of the simulated Dirac particle during its evolution. As shown in Figure 3 (b), we do not observe any squeezing of the density profile caused by the simulated gravity in the trapped-ion system, which represents a different phenomenon from the one in Ref.[42]. We remark that the qualitative behavior of the variance of the position is determined by the mass and is independent of the parameter . The wavepackets at ms and ms are shown in Figure 3(c), which suggests that the wavepacket of the simulated Dirac particle is stretched and the oscillatory feature becomes more prominent as the system evolves, due to the Zitterbewegung effect.
Conclusion.— To summarize, we propose a scheme to implement a trapped-ion quantum simulation of a Dirac particle moving in the Alcubierre (1+1)-dimensional universe, which is a particularly important example of extreme curved spacetime leading to superluminal propagation in consistence with general relativity. We show that the flexibility of control parameters in such a platform allows one to observe counterintuitive effects including the superluminal velocity and the Zitterbewegung effect of the simulated Dirac particle in the Alcubierre curved spacetime. Our work demonstrates the feasibility to explore interesting and counterintuitive features of exotic curved spacetime by quantum simulation of Dirac particles in the corresponding spacetime geometry. Furthermore, the analogy between particles in the setting of semiclassical quantum field theory and basic light-matter interaction quantum models would further inspire the ideas to investigate intriguing phenomena of general relativity and quantum field theory in a variety of quantum platforms that are available in laboratories.
Acknowledgements.— We thank Dr. Dongxiao Li and Prof. Jianwei Cui for very helpful discussions. This work is supported by National Natural Science Foundation of China (Grant No. 11690032), and the Open Project Program of Wuhan National Laboratory for Optoelectronics (No. 2019WNLOKF002).
V Appendix
V.1 Derivation of the Hamiltonian of a trapped-ion simulated Dirac particle
In the main text, we propose to simulate a Dirac particle moving in the (1+1)-dimensional Alcubierre spacetime with a well-controlled trap-ion quantum simulator. In a semi-classical way, the Dirac particle moving in a fixed gravitational background, characterized by a general metric tensor , can be expressed as [50, 41]
(13) |
where the Greek letter and the Latin letter denote the coordinates of curved spacetime and local rest frame respectively, is the mass of the quantized field, is the determinant of the metric tensor, and is the vielbein allowing the constant Dirac matrices to act at each spacetime point. Here, we choose the chiral representation such that , , with the Pauli matrices. In the specific case of the Alcubierre spacetime, the metric tensor of which is given by
(14) |
the vielbein can be set as
(15) |
with and the column and row indices respectively. By multiplying Eq. (13) with , the Dirac equation can be expanded as
(16) |
which can be further rewritten in the form of Schrödinger equation [i.e. Eq. \textcolorred(4) in the main text] as follows
(17) |
Thus, the Hamiltonian of the Dirac particle to be simulated can be written as
(18) |
with and . We remark that the operators and , which satisfy the standard commutation relation, i.e. , can be related to the position and momentum operators of the trapped ion used to simulate the relativistic Dirac particle. Thereby, we can rewrite them using the creation and annihilation operators as
(19) |
with the mass and the trapping potential of the trapped ion. In this way, the Hamiltonian of the trapped-ion simulated Dirac particle can be expressed as
(20) |
which is equivalent to the Eq. \textcolorred(8) in the main text by considering the Hadamard transformation.
V.2 Quantum dynamics of the trapped-ion simulated Dirac particle
In this section, we derive the quantum dynamics of the simulated Dirac particle in more detail. Based on the Heisenberg equation, the time evolution of the position and spin operators is given by
(21) |
where
(22) |
with . The effective Hamiltonian and the evolution operator are
(23) |
with . In the above derivation, we have omitted the time ordering operator in due to the fact that for arbitrary and . From the above equations, we can obtain
(24) |
In order to derive , we notice that , which gives and further
(25) |
where corresponds to the Dirac Hamiltonian in the flat spacetime. Hence, we have [i.e. Eq. \textcolorred(12) in the main text]
(26) |
V.3 Trapped-ion realization of the Dirac Hamiltonian
The well-controllable trapped-ion system is a particularly promising platform for simulating the basic light-matter interaction models by exploiting its internal and motional degrees of freedom. Here we explain in more detail the trapped-ion setup to simulate the Dirac particle moving in a curved spacetime, the Hamiltonian of which (i.e. Eq. \textcolorred(8) in the main text) has been derived in the above section. We consider the radial motional mode with frequency of the trapped ion to model the bosonic field and the two internal ground hyperfine levels with transition frequency to form an effective spin-1/2, denoted as and . With blue and red sideband drives at frequencies and respectively [55, 53], which can be generated by using oscillating near-field magnetic field gradients in experiments [56], one is able to realize the spin-harmonic oscillator coupling via the Mølmer–Sørensen interaction that is described as
(27) |
In addition, it is possible to apply an oscillating potential directly to the radio-frequency trapping electrodes such that the single ion’s trapping potential is modulated periodically with an amplitude at frequency . The corresponding Hamitonian is given by
(28) |
Therefore, the total Hamiltonian of such a trapped-ion setup is [namely Eq. \textcolorred(9) in the main text]
(29) |
In order to derive the Hamiltonian of Eq. \textcolorred(8) in the main text, we first make a displace transformation to the above Hamiltonian with , , which gives
(30) |
Then we move to the interaction picture with respect to while requiring . After applying the rotating-wave approximation under the conditions of , we are left with
(31) |
the form of which is equivalent to the Hamiltonian of Eq. \textcolorred(10) in the main text by further making the Hadamard transformation.
V.4 Numerical simulation of the trapped-ion simulated Dirac particle
Without loss of generality, the initial state of the trapped ion is chosen as
(32) |
where and denote the spatial and internal degrees of freedom of the trapped ion respectively, and is the normalization factor of its wavefunction. Using the following relation
(33) |
where are the Hermite polynomials, we can rewrite the initial state in the Fock space as
(34) |
with . Hence, the average position
(35) |
and the corresponding variance
(36) |
can be obtained with respect to the time evolved state in a truncated Fock space with the maximum excitation number . We remark that this truncation is verified to guarantee converging results, which are shown in Fig.\textcolorred2 and Fig.\textcolorred3 (a-b) in the main text. In addition, the probability distribution of the wavepacket corresponding to the final state can be obtained by
(37) |
where is the reduced density matrix for the spatial degrees of freedom by tracing out the spin degree of freedom in the total final density matrix of the final state .
References
- Davies [1976] P. C. W. Davies, Quantum field theory in curved space–time, Nature 263, 377 (1976).
- Jacobson [2005] T. Jacobson, Introduction to quantum fields in curved spacetime and the Hawking effect, in Lectures on Quantum Gravity (Springer, 2005) pp. 39–89.
- Hollands and Wald [2015] S. Hollands and R. M. Wald, Quantum fields in curved spacetime, Physics Reports 574, 1 (2015).
- Howl et al. [2018] R. Howl, L. Hackermülller, D. E. Bruschi, and I. Fuentes, Gravity in the quantum lab, Adv. Phys.: X 3, 1383184 (2018).
- Howl et al. [2021] R. Howl, V. Vedral, D. Naik, M. Christodoulou, C. Rovelli, and A. Iyer, Non-gaussianity as a signature of a quantum theory of gravity, PRX Quantum 2, 010325 (2021).
- Unruh [1981] W. G. Unruh, Experimental black-hole evaporation?, Phys. Rev. Lett. 46, 1351 (1981).
- Weinfurtner et al. [2011] S. Weinfurtner, E. W. Tedford, M. C. J. Penrice, W. G. Unruh, and G. A. Lawrence, Measurement of stimulated hawking emission in an analogue system, Phys. Rev. Lett. 106, 021302 (2011).
- Zurek [1985] W. H. Zurek, Cosmological experiments in superfluid helium?, Nature 317, 505 (1985).
- Fedichev and Fischer [2004] P. O. Fedichev and U. R. Fischer, “cosmological” quasiparticle production in harmonically trapped superfluid gases, Phys. Rev. A 69, 033602 (2004).
- Garay et al. [2000] L. J. Garay, J. R. Anglin, J. I. Cirac, and P. Zoller, Sonic analog of gravitational black holes in Bose-Einstein condensates, Phys. Rev. Lett. 85, 4643 (2000).
- Barceló et al. [2003] C. Barceló, S. Liberati, and M. Visser, Probing semiclassical analog gravity in Bose-Einstein condensates with widely tunable interactions, Phys. Rev. A 68, 053613 (2003).
- Fedichev and Fischer [2003] P. O. Fedichev and U. R. Fischer, Gibbons-Hawking effect in the sonic de sitter space-time of an expanding Bose-Einstein-condensed gas, Phys. Rev. Lett. 91, 240407 (2003).
- Fischer and Schützhold [2004] U. R. Fischer and R. Schützhold, Quantum simulation of cosmic inflation in two-component Bose-Einstein condensates, Phys. Rev. A 70, 063615 (2004).
- Carusotto et al. [2008] I. Carusotto, S. Fagnocchi, A. Recati, R. Balbinot, and A. Fabbri, Numerical observation of Hawking radiation from acoustic black holes in atomic Bose–Einstein condensates, New Journal of Physics 10, 103001 (2008).
- Steinhauer [2014] J. Steinhauer, Observation of self-amplifying Hawking radiation in an analogue black-hole laser, Nature Physics 10, 864 (2014).
- Steinhauer [2016] J. Steinhauer, Observation of quantum Hawking radiation and its entanglement in an analogue black hole, Nature Physics 12, 959 (2016).
- Chä and Fischer [2017] S.-Y. Chä and U. R. Fischer, Probing the scale invariance of the inflationary power spectrum in expanding quasi-two-dimensional dipolar condensates, Phys. Rev. Lett. 118, 130404 (2017).
- De Nova et al. [2019] J. R. M. De Nova, K. Golubkov, V. I. Kolobov, and J. Steinhauer, Observation of thermal Hawking radiation and its temperature in an analogue black hole, Nature 569, 688 (2019).
- Isoard and Pavloff [2020] M. Isoard and N. Pavloff, Departing from thermality of analogue Hawking radiation in a Bose-Einstein condensate, Phys. Rev. Lett. 124, 060401 (2020).
- Belgiorno et al. [2010] F. Belgiorno, S. L. Cacciatori, M. Clerici, V. Gorini, G. Ortenzi, L. Rizzi, E. Rubino, V. G. Sala, and D. Faccio, Hawking radiation from ultrashort laser pulse filaments, Phys. Rev. Lett. 105, 203901 (2010).
- Philbin et al. [2008] T. G. Philbin, C. Kuklewicz, S. Robertson, S. Hill, F. König, and U. Leonhardt, Fiber-optical analog of the event horizon, Science 319, 1367 (2008).
- Drori et al. [2019] J. Drori, Y. Rosenberg, D. Bermudez, Y. Silberberg, and U. Leonhardt, Observation of stimulated Hawking radiation in an optical analogue, Phys. Rev. Lett. 122, 010404 (2019).
- Lang and Schützhold [2019] S. Lang and R. Schützhold, Analog of cosmological particle creation in electromagnetic waveguides, Phys. Rev. D 100, 065003 (2019).
- Boiron et al. [2015] D. Boiron, A. Fabbri, P.-E. Larré, N. Pavloff, C. I. Westbrook, and P. Ziń, Quantum signature of analog Hawking radiation in momentum space, Phys. Rev. Lett. 115, 025301 (2015).
- Roldán-Molina et al. [2017] A. Roldán-Molina, A. S. Nunez, and R. A. Duine, Magnonic black holes, Phys. Rev. Lett. 118, 061301 (2017).
- Chuang et al. [1991] I. Chuang, R. Durrer, N. Turok, and B. Yurke, Cosmology in the laboratory: Defect dynamics in liquid crystals, Science 251, 1336 (1991).
- Barceló et al. [2011] C. Barceló, S. Liberati, and M. Visser, Analogue gravity, Living Rev. Relativity 14, 3 (2011).
- Faccio et al. [2013] D. Faccio, F. Belgiorno, S. Cacciatori, V. Gorini, S. Liberati, and U. Moschella, Analogue gravity phenomenology: analogue spacetimes and horizons, from theory to experiment, Vol. 870 (Springer, 2013).
- Alsing et al. [2005] P. M. Alsing, J. P. Dowling, and G. J. Milburn, Ion trap simulations of quantum fields in an expanding universe, Phys. Rev. Lett. 94, 220401 (2005).
- Schützhold et al. [2007] R. Schützhold, M. Uhlmann, L. Petersen, H. Schmitz, A. Friedenauer, and T. Schätz, Analogue of cosmological particle creation in an ion trap, Phys. Rev. Lett. 99, 201301 (2007).
- Menicucci et al. [2010] N. C. Menicucci, S. J. Olson, and G. J. Milburn, Simulating quantum effects of cosmological expansion using a static ion trap, New Journal of Physics 12, 095019 (2010).
- Casanova et al. [2010] J. Casanova, J. J. García-Ripoll, R. Gerritsma, C. F. Roos, and E. Solano, Klein tunneling and Dirac potentials in trapped ions, Phys. Rev. A 82, 020101 (2010).
- Gerritsma et al. [2011] R. Gerritsma, B. P. Lanyon, G. Kirchmair, F. Zähringer, C. Hempel, J. Casanova, J. J. García-Ripoll, E. Solano, R. Blatt, and C. F. Roos, Quantum simulation of the Klein paradox with trapped ions, Phys. Rev. Lett. 106, 060503 (2011).
- Wittemer et al. [2019] M. Wittemer, F. Hakelberg, P. Kiefer, J.-P. Schröder, C. Fey, R. Schützhold, U. Warring, and T. Schaetz, Phonon pair creation by inflating quantum fluctuations in an ion trap, Phys. Rev. Lett. 123, 180502 (2019).
- Tian et al. [2020] Z. Tian, Y. Lin, U. R. Fischer, and J. Du, Verifying the upper bound on the speed of scrambling with the analogue Hawking radiation of trapped ions, arXiv:2007.05949 (2020).
- Yang et al. [2020] R.-Q. Yang, H. Liu, S. Zhu, L. Luo, and R.-G. Cai, Simulating quantum field theory in curved spacetime with quantum many-body systems, Phys. Rev. Research 2, 023107 (2020).
- Nation et al. [2009] P. D. Nation, M. P. Blencowe, A. J. Rimberg, and E. Buks, Analogue Hawking radiation in a dc-SQUID array transmission line, Phys. Rev. Lett. 103, 087004 (2009).
- Sabín [2016] C. Sabín, Quantum simulation of traversable wormhole spacetimes in a dc-SQUID array, Phys. Rev. D 94, 081501 (2016).
- Tian et al. [2017] Z. Tian, J. Jing, and A. Dragan, Analog cosmological particle generation in a superconducting circuit, Phys. Rev. D 95, 125003 (2017).
- Tian and Du [2019] Z. Tian and J. Du, Analogue Hawking radiation and quantum soliton evaporation in a superconducting circuit, The European Physical Journal C 79, 1 (2019).
- Collas and Klein [2019] P. Collas and D. Klein, The Dirac equation in curved spacetime (Springer International Publishing, 2019).
- Pedernales et al. [2018] J. S. Pedernales, M. Beau, S. M. Pittman, I. L. Egusquiza, L. Lamata, E. Solano, and A. del Campo, Dirac equation in ()-dimensional curved spacetime and the multiphoton quantum Rabi model, Phys. Rev. Lett. 120, 160403 (2018).
- García and Sabín [2019] J. F. García and C. Sabín, Dirac equation in exotic spacetimes, Phys. Rev. D 99, 025008 (2019).
- Lamata et al. [2007] L. Lamata, J. León, T. Schätz, and E. Solano, Dirac equation and quantum relativistic effects in a single trapped ion, Phys. Rev. Lett. 98, 253005 (2007).
- Alcubierre [1994] M. Alcubierre, The warp drive: hyper-fast travel within general relativity, Classical and Quantum Gravity 11, L73 (1994).
- Finazzi et al. [2009] S. Finazzi, S. Liberati, and C. Barceló, Semiclassical instability of dynamical warp drives, Phys. Rev. D 79, 124017 (2009).
- White [2003] H. White, A discussion of space-time metric engineering, Gen. Relativ. Gravit. 35, 2025 (2003).
- White [2011] H. White, Warp field mechanics 101, J. Br. Interplanet. Soc. 66, 242 (2011).
- González-Díaz [2000] P. F. González-Díaz, Warp drive space-time, Phys. Rev. D 62, 044005 (2000).
- Mann et al. [1991] R. B. Mann, S. M. Morsink, A. E. Sikkema, and T. G. Steele, Semiclassical gravity in 1+1 dimensions, Phys. Rev. D 43, 3948 (1991).
- [51] Further details of analysis and calculation areare available as supplementary material, which includes Refs. [50, 41, 53, 55, 56] .
- Thaller [2013] B. Thaller, The Dirac equation (Springer Science & Business Media, 2013).
- Burd et al. [2021] S. C. Burd, R. Srinivas, H. M. Knaack, W. Ge, A. C. Wilson, D. J. Wineland, D. Leibfried, J. J. Bollinger, D. T. C. Allcock, and D. H. Slichter, Quantum amplification of boson-mediated interactions, Nature Physics 10.1038/s41567-021-01237-9 (2021).
- Burd et al. [2019] S. C. Burd, R. Srinivas, J. J. Bollinger, A. C. Wilson, D. J. Wineland, D. Leibfried, D. H. Slichter, and D. T. C. Allcock, Quantum amplification of mechanical oscillator motion, Science 364, 1163 (2019).
- Sørensen and Mølmer [1999] A. Sørensen and K. Mølmer, Quantum computation with ions in thermal motion, Phys. Rev. Lett. 82, 1971 (1999).
- Srinivas et al. [2019] R. Srinivas, S. C. Burd, R. T. Sutherland, A. C. Wilson, D. J. Wineland, D. Leibfried, D. T. C. Allcock, and D. H. Slichter, Trapped-ion spin-motion coupling with microwaves and a near-motional oscillating magnetic field gradient, Phys. Rev. Lett. 122, 163201 (2019).