This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Probing Lorentz-Invariance-Violation Induced Nonthermal Unruh Effect in Quasi-Two-Dimensional Dipolar Condensates

Zehua Tian [email protected] CAS Key Laboratory of Microscale Magnetic Resonance and School of Physical Sciences, University of Science and Technology of China, Hefei 230026, China CAS Center for Excellence in Quantum Information and Quantum Physics, University of Science and Technology of China, Hefei 230026, China    Longhao Wu CAS Key Laboratory of Microscale Magnetic Resonance and School of Physical Sciences, University of Science and Technology of China, Hefei 230026, China CAS Center for Excellence in Quantum Information and Quantum Physics, University of Science and Technology of China, Hefei 230026, China    Liang Zhang CAS Key Laboratory of Microscale Magnetic Resonance and School of Physical Sciences, University of Science and Technology of China, Hefei 230026, China CAS Center for Excellence in Quantum Information and Quantum Physics, University of Science and Technology of China, Hefei 230026, China    Jiliang Jing Department of Physics, Key Laboratory of Low Dimensional Quantum Structures and Quantum Control of Ministry of Education, and Synergetic Innovation Center for Quantum Effects and Applications, Hunan Normal University, Changsha, Hunan 410081, P. R. China    Jiangfeng Du [email protected] CAS Key Laboratory of Microscale Magnetic Resonance and School of Physical Sciences, University of Science and Technology of China, Hefei 230026, China CAS Center for Excellence in Quantum Information and Quantum Physics, University of Science and Technology of China, Hefei 230026, China Hefei National Laboratory, University of Science and Technology of China, Hefei 230088, China
Abstract

The Unruh effect states an accelerated particle detector registers a thermal response when moving through the Minkowski vacuum, and its thermal feature is believed to be inseparable from Lorentz symmetry: Without the latter, the former disappears. Here we propose to observe analogue circular Unruh effect using an impurity atom in a quasi-two-dimensional Bose-Einstein condensate (BEC) with dominant dipole-dipole interactions between atoms or molecules in the ultracold gas. Quantum fluctuations in the condensate possess a Bogoliubov spectrum ω𝐤=c0|𝐤|f(c0|𝐤|/M)\omega_{\mathbf{k}}=c_{0}|\mathbf{k}|f(\hbar\,c_{0}|\mathbf{k}|/M_{\ast}), working as an analogue Lorentz-violating quantum field with the Lorentz-breaking scale MM_{\ast}, and the impurity acts as an effective Unruh-DeWitt detector thereof. When the detector travels close to the sound speed, observation of the Unruh effect in our quantum fluid platform becomes experimentally feasible. In particular, the deviation of the Bogoliubov spectrum from the Lorentz-invariant case is highly engineerable through the relative strength of the dipolar and contact interactions, and thus a viable laboratory tool is furnished to experimentally investigate whether the thermal characteristic of Unruh effect is robust to the breaking of Lorentz symmetry.

Introduction.— One of the surprising fundamental consequences of relativistic quantum field theory is that the concept of particle number is observer dependent. A prominent paradigm is the so-called Unruh effect Unruh (1976): In the view of an uniformly accelerating observer, the Fock vacuum state of quantum field in the Minkowski spacetime appears as a thermal state rather than a zero-particle state. The corresponding characteristic temperature is proportional to the observer’s acceleration aa, given by kBTU=a2πck_{\text{B}}T_{\text{U}}=\frac{\hbar\,a}{2\pi\,c}. To produce a measurable temperature for fundamental quantum fields, extremely huge accelerations are required (e.g., smaller than 1Kelvin1\,\mathrm{Kelvin} even for accelerations as high as 1020m/s210^{20}\,\mathrm{m/s^{2}}), and thus until now, the direct experimental confirmation of the Unruh effect still remains elusive.

Analogue gravity Unruh (1981); Barceló et al. (2011) opens up a new route to study various phenomenas predicted by relativistic quantum field theory, e.g., Hawking effect Garay et al. (2000); Nation et al. (2009); Lahav et al. (2010); Garay et al. (2000); Steinhauer (2014); Weinfurtner et al. (2011); Horstmann et al. (2010); Euvé et al. (2016); Steinhauer (2016); Roldán-Molina et al. (2017); Tian and Du (2019); Muñoz de Nova et al. (2019); Kolobov et al. (2021); Drori et al. (2019); Človečko et al. (2019); Kedem et al. (2020); Yang et al. (2020); Ribeiro et al. (2022); Tian et al. (2022), cosmological particle production Fedichev and Fischer (2004a); Hung et al. (2013); Prain et al. (2010); Fedichev and Fischer (2003); Alsing et al. (2005); Tian et al. (2017); Schützhold et al. (2007); Eckel et al. (2018); Lang and Schützhold (2019); Chä and Fischer (2017); Wittemer et al. (2019); Bhardwaj et al. (2021); Banik et al. (2022), and dynamical Casimir effect Johansson et al. (2009); Fujii et al. (2011); Wilson et al. (2011); Jaskula et al. (2012); Lähteenmäki et al. (2013); Macrì et al. (2018); Koghee and Wouters (2014); Dodonov (2020); Schneider et al. (2020), in a variety of electronic, acoustic, optical and even magnetic and superconducting settings. Recently, analogue gravity program for observing the Unruh effect has been successfully theoretically put forward Chen and Tajima (1999); Lochan et al. (2020); Schützhold et al. (2008); Retzker et al. (2008); Marino et al. (2020); Gooding et al. (2020); Sheng et al. (2021); Hegde et al. (2019); Zeng and Zubairy (2021); Cozzella et al. (2017); Barros et al. (2020); Guedes et al. (2019); Adjei et al. (2020), and through a BEC system Hu et al. experimentally realized the analogue Unruh effect relied on functional equivalence (i.e., simulating two-mode squeezed mechanics) Hu et al. (2019). Furthermore, in the quantum field theory, the contributions from the trans-Planckian modes as seen by an inertial observer are indispensable for deriving the Unruh effect with Bogoliubov transformation method. This particular feature makes the Unruh effect a potentially important arena for understanding and exploring implications of trans-Planckian physics Nicolini and Rinaldi (2011); Agulló et al. (2008); Hossain and Sardar (2015); Kajuri (2016); Hossain and Sardar (2016); Alkofer et al. (2016); Carballo-Rubio et al. (2019); Hammad et al. (2021); Louko and Upton (2018), and even the detecting means to probe some candidate theories of quantum gravity that may modify the trans-Planckian modes significantly. This modification usually accompanied by the breaking of Lorentz symmetry may even challenge the equivalence between Unruh effect and Hawking effect since the latter appears to be robust to high energy modifications of the dispersion relation Unruh (1995) while the former is not so immune and will lose its conventional thermal interpretation Nicolini and Rinaldi (2011); Agulló et al. (2008); Hossain and Sardar (2015, 2016); Carballo-Rubio et al. (2019) (see more details in the following). It is thus of great interest to experimentally investigate the consequences of trans-Planckian physics in a microscopically well-understood setup in a regime, that is inaccessible for quantum fields in real relativistic scenarios.

In this paper, we propose to study the interplay between the Unruh effect and trans-Planckian physics with an experimentally accessible platform consisting of a dipolar BEC Baranov (2008) and an immersed impurity Recati et al. (2005); Fedichev and Fischer (2003). From the perspective of analog, the density fluctuations in the condensate possessing trans-Planckian spectra leading to strong departures from Lorentz invariance Fischer (2006); Chä and Fischer (2017); Tian et al. (2018), resemble Lorentz-violating quantum field (LVQF), while the impurity, analogously dipole coupled to the density fluctuations, is modeled as an Unruh-DeWitt detector coupled to the LVQF. We will show that for the Lorentz-invariant (LI) spectrum the Unruh-DeWitt detector for an accelerated circular path indeed experiences a similar thermal response, while yields significant changes of this standard thermal feature when the spectra strongly deviate from Lorentz invariance. As far as we know, this represents the first example within analogue gravity where Unruh effect without thermality caused by the breaking of Lorentz symmetry can become experimentally manifest.

Refer to caption
Figure 1: (a) A two-level atom immersed in a quasi-2D dipolar BEC, moving along a circular trajectory with the radius of the orbit RR, an angular velocity Ω\Omega and the corresponding linear speed vv. (b) The dimensionless function ff shown in (2) as a function of ζ=c0k/M\zeta=\hbar\,c_{0}k/M_{\ast}. For the LI case, f=1f=1. R0=0R_{0}=0 denotes the contact interaction case, where ff is independent of AA. For DDI dominance, R0=π/2R_{0}=\sqrt{\pi/2}, ff dips below 1 for an interval of ζ\zeta. Note that ff becomes negative when A>Ac=3.4454A>A_{c}=3.4454, which means the spectrum of quasiparticle becomes unstable.

Lorentz-violating quantum field and analogue Unruh-DeWitt detector in dipolar BEC.— As schematically shown in Fig. 1, we establish the connection between an impurity immersed in a quasi-two-dimensional (quasi-2D) dipolar BEC with the Unruh-DeWitt detector model Unruh (1976); Birrell and Davies , inspired by the seminal atomic quantum dot idea introduced in Refs. Recati et al. (2005); Fedichev and Fischer (2003, 2004b).

We begin with the Lagrangian density of an interacting Bose gas comprising atoms or molecules of mass mm,

\displaystyle\mathcal{L} =\displaystyle= i2(ΨtΨtΨΨ)22m|Ψ|2Vext|Ψ|2\displaystyle\frac{i\hbar}{2}(\Psi^{\ast}\partial_{t}\Psi-\partial_{t}\Psi^{\ast}\Psi)-\frac{\hbar^{2}}{2m}|\nabla\Psi|^{2}-V_{\text{ext}}|\Psi|^{2} (1)
12|Ψ|2d3𝐑Vint(𝐑𝐑)|Ψ(𝐑)|2,\displaystyle-\frac{1}{2}|\Psi|^{2}\int\,d^{3}\mathbf{R}^{\prime}\,V_{\text{int}}(\mathbf{R}-\mathbf{R}^{\prime})|\Psi(\mathbf{R}^{\prime})|^{2},

where 𝐑=(𝐫,z)\mathbf{R}=(\mathbf{r},z) are spatial 3D coordinates. The interaction reads Vint(𝐑𝐑)=gcδ3(𝐑𝐑)+3gd{[13(zz)2/|𝐑𝐑|2]/|𝐑𝐑|3}/4πV_{\text{int}}(\mathbf{R}-\mathbf{R}^{\prime})=g_{c}\delta^{3}(\mathbf{R}-\mathbf{R}^{\prime})+3g_{d}\{[1-3(z-z^{\prime})^{2}/|\mathbf{R}-\mathbf{R}^{\prime}|^{2}]/|\mathbf{R}-\mathbf{R}^{\prime}|^{3}\}/4\pi, where gc=4π2ac/mBg_{c}=4\pi\hbar^{2}a_{c}/m_{B} represents the contact interaction strength, with aca_{c} being the ss-wave scattering length; the dipole-dipole interaction (DDI) strength gd=μ0μm2/3g_{d}=\mu_{0}\mu_{m}^{2}/3, with μ0\mu_{0} and μm\mu_{m} being the permeability of vacuum and the magnetic dipole moment that is polarized to the zz direction, respectively. Moreover, the gas is trapped by an external potential, Vext(𝐑)=mBω2𝐫2/2+mBωz2z2/2V_{\text{ext}}(\mathbf{R})=m_{B}\omega^{2}\mathbf{r}^{2}/2+m_{B}\omega^{2}_{z}z^{2}/2, and is strongly confined along the zz axis, with aspect ratio κ=ωz/ω1\kappa=\omega_{z}/\omega\gg 1 over the whole time evolution. As a result of that, the motion of the Bose gas along the zz axis is frozen to the ground state with a Gaussian form, ρz(z)=(πdz2)1/2exp[z2/dz2]\rho_{z}(z)=(\pi\,d^{2}_{z})^{-1/2}\exp[-z^{2}/d^{2}_{z}], with dz=/mBωzd_{z}=\sqrt{\hbar/m_{B}\omega_{z}}. Therefore, the whole system effectively reduces to a quasi-2D one which can ensure stability in the DDI-dominated regime Fischer (2006). Finally, we may assume that the Bose gas is condensed to the zero-momentum state with an area density ρ0\rho_{0}.

Within the Bogoliubov theory of small excitations on top of the condensate Pethick and Smith (2008); Pitaevskiĭ and Stringari (2016), density fluctuations in Heisenberg representation can be written as δρ^(t,𝐫)=ρ0[d2𝐤/(2π)2](u𝐤+v𝐤)[b^𝐤(t)ei𝐤𝐫+b^𝐤(t)ei𝐤𝐫]\delta\hat{\rho}(t,\mathbf{r})=\sqrt{\rho_{0}}\int[d^{2}\mathbf{k}/(2\pi)^{2}](u_{\mathbf{k}}+v_{\mathbf{k}})[\hat{b}_{\mathbf{k}}(t)e^{i\mathbf{k}\cdot\mathbf{r}}+\hat{b}^{\dagger}_{\mathbf{k}}(t)e^{-i\mathbf{k}\cdot\mathbf{r}}] which closely resemble the quantum field in terms of bosonic operators b^𝐤(t)=b^𝐤eiω𝐤t\hat{b}_{\mathbf{k}}(t)=\hat{b}_{\mathbf{k}}\,e^{-i\omega_{\mathbf{k}}\,t}, satisfying the usual Bose commutation rules [b^𝐤,b^𝐤]=(2π)2δ2(𝐤𝐤)[\hat{b}_{\mathbf{k}},\hat{b}_{\mathbf{k}^{\prime}}^{\dagger}]=(2\pi)^{2}\delta^{2}(\mathbf{k}-\mathbf{k}^{\prime}). Note that u𝐤=(𝐤+𝐤+2𝒜𝐤)/2(𝐤2+2𝐤𝒜𝐤)1/4u_{\mathbf{k}}=(\sqrt{\mathcal{H}_{\mathbf{k}}}+\sqrt{\mathcal{H}_{\mathbf{k}}+2\mathcal{A}_{\mathbf{k}}})/2(\mathcal{H}^{2}_{\mathbf{k}}+2\mathcal{H}_{\mathbf{k}}\mathcal{A}_{\mathbf{k}})^{1/4} and v𝐤=(𝐤𝐤+2𝒜𝐤)/2(𝐤2+2𝐤𝒜𝐤)1/4v_{\mathbf{k}}=(\sqrt{\mathcal{H}_{\mathbf{k}}}-\sqrt{\mathcal{H}_{\mathbf{k}}+2\mathcal{A}_{\mathbf{k}}})/2(\mathcal{H}^{2}_{\mathbf{k}}+2\mathcal{H}_{\mathbf{k}}\mathcal{A}_{\mathbf{k}})^{1/4} are Bogoliubov parameters, and the quasiparticle frequency ω𝐤=𝐤2+2𝐤𝒜𝐤\omega_{\mathbf{k}}=\sqrt{\mathcal{H}^{2}_{\mathbf{k}}+2\mathcal{H}_{\mathbf{k}}\mathcal{A}_{\mathbf{k}}} with 𝐤=k2/2m\mathcal{H}_{\mathbf{k}}=k^{2}/2m and 𝒜𝐤=ρ0Vint, 02D(k)\mathcal{A}_{\mathbf{k}}=\rho_{0}V^{\text{2D}}_{\text{int, 0}}(k) Tian et al. (2018). Here k=|𝐤|k=|\mathbf{k}| and the Fourier transformation of DDI Vint, 02D(k)=g0eff(13R02kdzw[kdz2])V^{\text{2D}}_{\text{int, 0}}(k)=g^{\text{eff}}_{0}(1-\frac{3R_{0}}{2}kd_{z}w[\frac{kd_{z}}{\sqrt{2}}]), with w[x]=exp[x2](1erf[x])w[x]=\exp[x^{2}](1-\operatorname{erf}[x]), an effective contact coupling g0eff=12πdz(gc+2gd)g^{\text{eff}}_{0}=\frac{1}{\sqrt{2\pi}d_{z}}(g_{c}+2g_{d}), and the dimensionless ratio R0=π/2/(1+gc/2gd)R_{0}=\sqrt{\pi/2}/(1+g_{c}/2g_{d}). The parameter R0R_{0} could be tunable via Feshbach resonance Courteille et al. (1998); Inouye et al. (1998); Chin et al. (2010); Timmermans et al. (1999) and rotating polarizing field Giovanazzi et al. (2002); Tang et al. (2018), ranging from R0=0R_{0}=0 (when gd/gc0g_{d}/g_{c}\rightarrow 0, i.e., contact dominance), to R0=π/2R_{0}=\sqrt{\pi/2} (when gd/gcg_{d}/g_{c}\rightarrow\infty, i.e., DDI dominance).

The density fluctuations described above closely resemble a LVQF with a explicit dispersion relation given by

ω𝐤=c0k13R02Aζw[A2ζ]+ζ24=c0kf(ζ),\displaystyle\omega_{\mathbf{k}}=c_{0}k\sqrt{1-\frac{3R_{0}}{2}\sqrt{A}\zeta\,w\bigg{[}\sqrt{\frac{A}{2}}\zeta\bigg{]}+\frac{\zeta^{2}}{4}}=c_{0}kf(\zeta), (2)

where ζ=c0k/M\zeta=\hbar\,c_{0}k/M_{\ast}, c0=g0effρ0/mBc_{0}=\sqrt{g^{\text{eff}}_{0}\rho_{0}/m_{B}} is the speed of sound, A=g0effρ0/ωzA=g^{\text{eff}}_{0}\rho_{0}/\hbar\omega_{z} represents the effective chemical potential as measured relative to the transverse trapping, and M=mBc02M_{\ast}=m_{B}c^{2}_{0} is the analog energy scale of Lorentz violation. This dispersion relation (2) is approximately Lorentz invariant (f(ζ)1)(f(\zeta)\simeq 1) for ζ1\zeta\ll 1. By appropriately setting the relevant parameters AA and R0R_{0}, the dispersion could be analogously superluminal (f(ζ)>1)(f(\zeta)>1) and subluminal (f(ζ)<1)(f(\zeta)<1). In Fig. 1, we plot the function f(ζ)f(\zeta) shown in (2) to see how the Lorentz invariance is violated in this dispersion. For the DDI dominance, R0=π/2R_{0}=\sqrt{\pi/2}, the analogous subluminal spectrum develops a roton minimum for sufficiently large AA, and the Lorentz invariance is strongly broken near ζ0.9\zeta\simeq 0.9 Tian et al. (2018).

In order to probe the analogue LVQF in the dipolar BEC, we use an impurity as the analogue Unruh-DeWitt detector, which consists of a two-level atom (11 and 22) and its motion is supposed to be externally imposed by a tightly confining and relatively moving trap potential, so that its only degrees of freedom are the internal ones. Furthermore, the impurity is assumed to be controlled by a driving of a monochromatic external electromagnetic field at the frequency ωL\omega_{L} close to resonance with 121\rightarrow 2 transition ωω21\omega\simeq\omega_{21}, with a Rabi frequency ω0\omega_{0}. Then the Hamiltonian of the whole system can be written as

H(t)\displaystyle H(t) =\displaystyle= d2𝐤ω𝐤b^𝐤b^𝐤+ω21|22|(ω02eiωLt|21|\displaystyle\int\,d^{2}\mathbf{k}\hbar\omega_{\bf k}\hat{b}^{\dagger}_{\bf k}\hat{b}_{\bf k}+\hbar\omega_{21}|2\rangle\langle 2|-\bigg{(}\frac{\hbar\omega_{0}}{2}e^{-i\omega_{L}t}|2\rangle\langle 1| (3)
+H.c.)+s=1,2gsρ^(𝐫A(t))|ss|,\displaystyle+\mathrm{H.c.}\bigg{)}+\sum_{s=1,2}g_{s}\hat{\rho}(\mathbf{r}_{A}(t))|s\rangle\langle\,s|,

where the last term denotes the collisional coupling between the impurity and Bose gas, and ρ^(𝐫A)=ψ^(𝐫A)ψ^(𝐫A)ρ0+δρ^(𝐫A)\hat{\rho}(\mathbf{r}_{A})=\hat{\psi}^{\dagger}(\mathbf{r}_{A})\hat{\psi}(\mathbf{r}_{A})\simeq\rho_{0}+\delta\hat{\rho}(\mathbf{r}_{A}) represents the field density operator of the Bose gas with 𝐫A(t)\mathbf{r}_{A}(t) being the time-dependent position of the impurity.

In the rotated |g,e=(|1±2)/2|g,e\rangle=(|1\rangle\pm 2\rangle)/\sqrt{2} basis, the Rabi frequency ω0\omega_{0} determines the splitting between the |g,e|g,e\rangle states, while the detuning δ=ωLω21\delta=\omega_{L}-\omega_{21} gives a coupling terms. In such case, the impurity immersed in the condensate is collisionally coupled to the Bose gas via two channels Marino et al. (2017, 2020); Tian and Du (2021): The first term resembles the interaction of a static charge to an external scalar potential and can be canceled through proper tuning of the interaction constants g1,2g_{1,2} (e.g., via Feshbach resonance Courteille et al. (1998); Inouye et al. (1998); Chin et al. (2010); Timmermans et al. (1999)), while the second term resembles a standard electric-dipole coupling. Choosing proper detuning δ\delta to exactly compensate the coupling to the average density, we can finally find the impurity-fluctuations interaction Hamiltonian

Hint=g(eiω0τσ+eiω0τσ)δρ^(𝐫A(τ),tA(τ)),\displaystyle H_{\text{int}}=g_{-}(e^{i\omega_{0}\tau}\sigma^{+}e^{-i\omega_{0}\tau}\sigma_{-})\delta\hat{\rho}({\bf r}_{A}(\tau),t_{A}(\tau)), (4)

reproducing the usual Unruh-DeWitt detector-field interaction with g=(g1g2)/2g_{-}=(g_{1}-g_{2})/2 satisfying δ/2+gρ0=0\hbar\delta/2+g_{-}\rho_{0}=0. However, here the LVQF is coupled to the detector.

Note that when ζ=c0k/M1\zeta=\hbar\,c_{0}k/M_{\ast}\ll 1 the density fluctuations resemble massless scalar field with spectrum, ω𝐤c0k\omega_{\mathbf{k}}\simeq\,c_{0}k, the linearly moving impurity remains unexcited when its velocity satisfying v<c0v<c_{0} , while behaves dramatically differently when moving at a supersonic speed vc0v\gtrsim\,c_{0}. Specifically, although the “charge neutrality” of the impurity rules out Bogoliubov-Cherenkov emission Carusotto et al. (2006); Astrakharchik and Pitaevskii (2004), the anomalous Doppler effect may induce it to be excited from its ground state while emitting Bogoliubov phonon and still conserving energy. This is the analogue Ginzburg emission for superluminal moving particles Ginzburg and Frolov (1986); Ginzburg (1996), occurring in BEC. When the spectrum breaks the Lorentz symmetry satisfying ω𝐤=c0kf(ζ)\omega_{\mathbf{k}}=c_{0}kf(\zeta): If 0<f(ζ)<10<f(\zeta)<1 for an interval of ζ\zeta, the impurity would get excited when its velocity exceeds the critical vc=c0fc(ζ)v_{c}=c_{0}f_{c}(\zeta) with fc(ζ)=inff(ζ)f_{c}(\zeta)=\inf\,f(\zeta) Tian and Du (2021). This particular property provides us a potential effective tool to constrain on the possible families of modified dispersion relations with the experimental results of the Relativistic Heavy Ion Collider Husain and Louko (2016). We will in the following consider a circularly moving impurity to observer the circular Unruh effect Bell and Leinaas (1983, 1987); Unruh (1998), and in particular examine whether the circular Unruh effect is robust to the breaking of Lorentz symmetry.

Refer to caption
Figure 2: Transition rates shown in (5) for different effective chemical potential AA, in units of g02ρ0mB23\frac{g^{2}_{0}\rho_{0}m_{B}}{2\hbar^{3}}: (a), (b), (c) as a function of the detector’s energy spacing ω0\omega_{0} with fixed velocity; (d), (e), (f) as a function of the velocity vv with fixed energy spacing of the detector. Here we assume that the DDI dominance case (R=π/2R=\sqrt{\pi/2}) is vailid.

Spontaneous excitation of the circularly moving detector.— If the detector moves with a circular trajectory (c0t(τ),𝐱(τ))=(c0γτ,Rcos(Ωγτ),Rsin(Ωγτ),0)(c_{0}t(\tau),{\bf x}(\tau))=(c_{0}\gamma\tau,R\cos(\Omega\gamma\tau),R\sin(\Omega\gamma\tau),0), with constant radius RR, the angular velocity Ω\Omega, the usual relativistic factor γ=1/1R2Ω2/c02\gamma=1/\sqrt{1-R^{2}\Omega^{2}/c^{2}_{0}}, and the corresponding acceleration a=Ω2γ2Ra=\Omega^{2}\gamma^{2}R, we find the transition rate of the detector from its ground state to excited state Sup

𝒫(ω0)\displaystyle{\cal P}(\omega_{0}) =\displaystyle= g2ρ0mB230𝑑ζζ2γf(ζ)m=Jm2(M~ζ)\displaystyle\frac{g^{2}_{-}\rho_{0}m_{B}}{2\hbar^{3}}\int^{\infty}_{0}d\zeta\frac{\zeta^{2}}{\gamma\,f(\zeta)}\sum^{\infty}_{m=-\infty}J^{2}_{m}(\tilde{M}\zeta) (5)
×\displaystyle\times δ(ζf(ζ)1M~(mvE~γ)),\displaystyle\delta\bigg{(}\zeta\,f(\zeta)-\frac{1}{\tilde{M}}\bigg{(}mv-\frac{\tilde{E}}{\gamma}\bigg{)}\bigg{)},

where M~=RM/c0\tilde{M}=RM_{\ast}/\hbar\,c_{0}, E~=Rω0/c0\tilde{E}=R\omega_{0}/c_{0}, and v=RΩ/c0v=R\Omega/c_{0} are dimensionless parameters. Note that we focus on the detector’s speed in the preferred Lorentz frame satisfying 0v<10\leq\,v<1.

For the LI scenario where R0=0R_{0}=0 and ω𝐤=c0k1+ζ2/4c0k\omega_{\mathbf{k}}=c_{0}k\sqrt{1+\zeta^{2}/4}\approx\,c_{0}k for kkc=M/c0k\ll\,k_{c}=M_{\ast}/c_{0}\hbar, we find in the ultrarelativistic limit, γ1\gamma\gg 1, the equilibrium population of the upper level relative to the lower is Sup

𝒫(ω0)𝒫(ω0)=12c02ω025a2exp(245c0ω0a),\displaystyle\frac{{\cal P}(\omega_{0})}{{\cal P}(-\omega_{0})}=\frac{12c^{2}_{0}\omega^{2}_{0}}{5a^{2}}\exp\big{(}-\sqrt{\frac{24}{5}}\frac{c_{0}\omega_{0}}{a}\big{)}, (6)

when the energy splitting of the detector is not too small ω0a/c0\omega_{0}\gg\,a/c_{0} Bell and Leinaas (1983), which leads to an effective temperature

Teff=5a26kBc0.\displaystyle T_{\text{eff}}=\frac{\sqrt{5}\hbar\,a}{2\sqrt{6}k_{B}c_{0}}. (7)

This temperature is higher by a factor 5/6π\sqrt{5/6}\pi than the Unruh temperature for the linear acceleration, and higher by a factor 5/2\sqrt{5/2} than the Unruh temperature for the real massless scalar field case as a result of the Bogoliubov transformation for the quasiparticles.

If (ζf(ζ))>0(\zeta\,f(\zeta))^{\prime}>0, we can find in the limit, M~\tilde{M}\rightarrow\infty, the transition rate (5) behaves quite differently for 0<v<fc0<v<f_{c} and fc<v<1f_{c}<v<1, where fc=f(ζc)f_{c}=f(\zeta_{c}) is a global minimum at ζ=ζc>0\zeta=\zeta_{c}>0. Specifically, for the former, the transition rate 𝒫(ω0)𝒫0(ω0){\cal P}(\omega_{0})\rightarrow{\cal P}_{0}(\omega_{0}), given by

𝒫0(ω0)=Γ0m=E~vγ1γ(mvE~γ)2Jm2(mvE~γ),\displaystyle{\cal P}_{0}(\omega_{0})=\Gamma_{0}\sum^{\infty}_{m=\lceil\,\frac{\tilde{E}}{v\gamma}\rceil}\frac{1}{\gamma}\bigg{(}mv-\frac{\tilde{E}}{\gamma}\bigg{)}^{2}J^{2}_{m}\big{(}mv-\frac{\tilde{E}}{\gamma}\big{)}, (8)

where Γ0=g2ρ02MR2\Gamma_{0}=\frac{g^{2}_{-}\rho_{0}}{2\hbar\,M_{\ast}\,R^{2}}. Note this is the response for the analogue massless scalar field, and thus no low-energy Lorentz violation can been seen when v<fcv<f_{c}. While for the latter there is a correction to the former case, 𝒫(ω0)𝒫0(ω0)+Δ𝒫{\cal P}(\omega_{0})\rightarrow{\cal P}_{0}(\omega_{0})+\Delta{\cal P}, with

Δ𝒫=g2ρ0mB2π3γζζ+𝑑ζζf(ζ)v2f2(ζ),\displaystyle\Delta{\cal P}=\frac{g^{2}_{-}\rho_{0}m_{B}}{2\pi\hbar^{3}\gamma}\int^{\zeta_{+}}_{\zeta_{-}}d\zeta\frac{\zeta}{f(\zeta)\sqrt{v^{2}-f^{2}(\zeta)}}, (9)

where ζ(0,ζc)\zeta_{-}\in(0,\zeta_{c}) and ζ+(ζc,)\zeta_{+}\in(\zeta_{c},\infty) are unique solutions to f(ζ)=vf(\zeta)=v in the respective intervals. This correction means the detector sees a low-energy Lorentz violation when v>fcv>f_{c}, and alternatively the departure from the standard prediction of Unruh effect appears as a consequence of the Lorentz violation.

Experimental implementation.— Recent experimental advances have allowed for groundbreaking observations of strongly dipolar BEC, its excitation spectrum displaying roton minimum, and dynamics of impurity immersed in BEC Lu et al. (2011); Aikawa et al. (2012); Norcia et al. (2021); Chomaz et al. (2018); Petter et al. (2019); Schmidt et al. (2021); Natale et al. (2019); Hertkorn et al. (2021); Tomza et al. (2019); Weckesser et al. (2021); Skou et al. (2021); Dieterle et al. (2021). These experiments hold promise to realize our experimental scenario proposed above. Specifically, we can consider a single Rb87{}^{87}\mathrm{Rb} atom immersed in a BEC of Dy\mathrm{Dy} atom Lu et al. (2011); Schmidt et al. (2021); Norcia et al. (2021) which possesses a magnetic dipole moment of 10μB10\mu_{B} with μB\mu_{B} being the Bohr magneton. The condensate density is assumed to be ρ04.4×103μm2\rho_{0}\sim 4.4\times 10^{3}\mathrm{\mu m}^{-2}; The observer trajectory radius R10μmR\sim 10\mathrm{\mu m}; A typical trap frequency ωz=2π×103Hz\omega_{z}=2\pi\times 10^{3}\mathrm{Hz}, the corresponding harmonic oscillator width is dz0.25×106md_{z}\simeq 0.25\times 10^{-6}\mathrm{m}. Fig. 2 displays the transition rate in (5), and clearly shows the spontaneous excitation of the detector as a result of the circular acceleration in a Minkowski vacuum. This would be viewed as the circular Unruh effect: Thermal bath is predicted for an accelerated detector moving through the inertial vacuum. In addition, the transition rates clearly show the deviation from the LI field case, occurring for strongly dipolar interactions. Specifically, when the field slightly deviates from the LI case, e.g., A=0.1A=0.1 case, their corresponding transition rates share similar behaviors. When this deviation becomes stronger, the excitation rates increase and behave sharply differently compared with the LI case, especially in the high velocity and low energy spacing regimes. However, for the low velocity case and large energy spacing of the detector, the excitation rates change slightly in the presence of Lorentz violation (i.e., they are not so sensitive to the breaking of Lorentz symmetry), since in such cases the detector is harder to excite.

To further check whether the Unruh effect is robust to the breaking of Lorentz symmetry, we naively define the Unruh temperature TT to estimate the fluctuations sampled by the impurity, using the Einstein’s detailed balanced condtion,

T=ω0kB1ln1(𝒫(ω0)𝒫(ω0)).\displaystyle T=\hbar\omega_{0}k^{-1}_{B}\ln^{-1}\bigg{(}\frac{{\cal P}(-\omega_{0})}{{\cal P}(\omega_{0})}\bigg{)}. (10)

This temperature is independent of ω0\omega_{0} for uniformly linearly accelerated detectors, given by TU=a2πkBcT_{\text{U}}=\frac{\hbar\,a}{2\pi\,k_{\text{B}}\,c}, since whose response function satisfies the Kubo-Martin-Schwinger condition Kubo (1957); Martin and Schwinger (1959); Haag et al. (1967). For the circular acceleration cases, such definition of TT may depend on ω0\omega_{0}, however, keeps monotonous increase via the effective acceleration parameter for the LI case Biermann et al. (2020). We here plot this temperature registered by the impurity coupled to the analogue LVQF as shown in Fig. 3. For the LI case, the temperature increases monotonously with the increase of the detector’s speed vv as expected. If the spectrum of the field deviates from the LI case slightly, e.g., A=0.1A=0.1 case, the present temperature behaves similarly as the LI case but with a larger magnitude. Remarkably, when the spectra of the analogue LVQF strongly deviate from the LI spectrum, the temperature first increases with the increase of the detector’s speed and then oscillates if the detector’s speed exceeds a critical value which depends on the degree of the deviation. The counterintuitive oscillation phenomenon means that Lorentz violation may break the thermal characteristic of Unruh effect, and even cause the analogue anti-Unruh effect Brenna et al. (2016): Unruh temperature decreases with the increase of acceleration. Besides, the effective negative temperature Purcell and Pound (1951); Ramsey (1956) occurs during the oscillation, which implies that as a result of Lorentz violation, the corresponding detector’s state is characterized by an inverted occupation distribution, where excited state is populated more than ground state.

Refer to caption
Figure 3: The temperature defined in (10) as a function of the detector’s velocity vv with fixed energy spacing of the detector, ω0=M/\omega_{0}=M_{\ast}/\hbar. Insets are the details of the corresponding oscillation parts. Here we assume that the DDI dominance case (R=π/2R=\sqrt{\pi/2}) is vailid.

Conclusions.— We present a concrete experimental proposal to test how the Lorentz violation affects the circular Unruh effect using an impurity immersed in a dipolar BEC. We find that if the spectra of quantum field deviate from the LI case strongly, the transition rates and the predicted temperature of the analogue Unruh-DeWitt detector behave quite differently compared with the LI case, and the Lorentz violation even more may induce the counter-intuitive anti-Unruh effect on certain conditions. Our preliminary estimates indicate that the proposed experimental implementation of the analogue circular Unruh effect and its interaction with the Lorentz-breaking physics is within reach of current state-of-the-art ultracold-atom experiments.

Our proposed quantum fluid platform may also allow us in the experimentally accessible regime to explore open questions concerning Unruh effect Retzker et al. (2008), and why its robustness to high energy modifications of the dispersion relation Carballo-Rubio et al. (2019) behaves differently from that of its equivalence principle dual—Hawking effect Unruh (1995). In addition, two impurities could be used as detectors to explore correlations harvest from the quantum vacuum of analogue quantum fields Pozas-Kerstjens and Martín-Martínez (2015).

Acknowledgements.
This work was supported by the National Key R&D Program of China (Grant No. 2018YFA0306600), and Anhui Initiative in Quantum Information Technologies (Grant No. AHY050000). ZT was supported by the National Natural Science Foundation of China under Grant No. 11905218, and the CAS Key Laboratory for Research in Galaxies and Cosmology, Chinese Academy of Science (No. 18010203).

References

  • Unruh (1976) W. G. Unruh, “Notes on black-hole evaporation,” Phys. Rev. D 14, 870–892 (1976).
  • Unruh (1981) W. G. Unruh, “Experimental black-hole evaporation?” Phys. Rev. Lett. 46, 1351–1353 (1981).
  • Barceló et al. (2011) Carlos Barceló, Stefano Liberati,  and Matt Visser, “Analogue gravity,” Living Reviews in Relativity 14, 3 (2011).
  • 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, 4643–4647 (2000).
  • 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).
  • Lahav et al. (2010) Oren Lahav, Amir Itah, Alex Blumkin, Carmit Gordon, Shahar Rinott, Alona Zayats,  and Jeff Steinhauer, “Realization of a sonic black hole analog in a bose-einstein condensate,” Phys. Rev. Lett. 105, 240401 (2010).
  • Steinhauer (2014) Jeff Steinhauer, “Observation of self-amplifying hawking radiation in an analogue black-hole laser,” Nature Physics 10, 864–869 (2014).
  • Weinfurtner et al. (2011) Silke Weinfurtner, Edmund W. Tedford, Matthew C. J. Penrice, William G. Unruh,  and Gregory A. Lawrence, “Measurement of stimulated hawking emission in an analogue system,” Phys. Rev. Lett. 106, 021302 (2011).
  • Horstmann et al. (2010) B. Horstmann, B. Reznik, S. Fagnocchi,  and J. I. Cirac, “Hawking radiation from an acoustic black hole on an ion ring,” Phys. Rev. Lett. 104, 250403 (2010).
  • Euvé et al. (2016) L.-P. Euvé, F. Michel, R. Parentani, T. G. Philbin,  and G. Rousseaux, “Observation of noise correlated by the hawking effect in a water tank,” Phys. Rev. Lett. 117, 121301 (2016).
  • Steinhauer (2016) Jeff Steinhauer, “Observation of quantum hawking radiation and its entanglement in an analogue black hole,” Nature Physics 12, 959–965 (2016).
  • Roldán-Molina et al. (2017) A. Roldán-Molina, Alvaro S. Nunez,  and R. A. Duine, “Magnonic black holes,” Phys. Rev. Lett. 118, 061301 (2017).
  • Tian and Du (2019) Zehua Tian and Jiangfeng Du, “Analogue hawking radiation and quantum soliton evaporation in a superconducting circuit,” The European Physical Journal C 79, 994 (2019).
  • Muñoz de Nova et al. (2019) Juan Ramón Muñoz de Nova, Katrine Golubkov, Victor I. Kolobov,  and Jeff Steinhauer, “Observation of thermal hawking radiation and its temperature in an analogue black hole,” Nature 569, 688–691 (2019).
  • Kolobov et al. (2021) Victor I. Kolobov, Katrine Golubkov, Juan Ramón Muñoz de Nova,  and Jeff Steinhauer, “Observation of stationary spontaneous hawking radiation and the time evolution of an analogue black hole,” Nature Physics 17, 362–367 (2021).
  • Drori et al. (2019) Jonathan Drori, Yuval Rosenberg, David Bermudez, Yaron Silberberg,  and Ulf Leonhardt, “Observation of stimulated hawking radiation in an optical analogue,” Phys. Rev. Lett. 122, 010404 (2019).
  • Človečko et al. (2019) M. Človečko, E. Gažo, M. Kupka,  and P. Skyba, “Magnonic analog of black- and white-hole horizons in superfluid He3b{}^{3}\mathrm{He}\text{$-$}b,” Phys. Rev. Lett. 123, 161302 (2019).
  • Kedem et al. (2020) Yaron Kedem, Emil J. Bergholtz,  and Frank Wilczek, “Black and white holes at material junctions,” Phys. Rev. Research 2, 043285 (2020).
  • Yang et al. (2020) Run-Qiu Yang, Hui Liu, Shining Zhu, Le Luo,  and Rong-Gen Cai, “Simulating quantum field theory in curved spacetime with quantum many-body systems,” Phys. Rev. Research 2, 023107 (2020).
  • Ribeiro et al. (2022) Caio C. Holanda Ribeiro, Sang-Shin Baak,  and Uwe R. Fischer, “Existence of steady-state black hole analogs in finite quasi-one-dimensional bose-einstein condensates,” Phys. Rev. D 105, 124066 (2022).
  • Tian et al. (2022) Zehua Tian, Yiheng Lin, Uwe R. Fischer,  and Jiangfeng Du, “Testing the upper bound on the speed of scrambling with an analogue of hawking radiation using trapped ions,” The European Physical Journal C 82, 212 (2022).
  • Fedichev and Fischer (2004a) Petr O. Fedichev and Uwe R. Fischer, ““cosmological” quasiparticle production in harmonically trapped superfluid gases,” Phys. Rev. A 69, 033602 (2004a).
  • Hung et al. (2013) Chen-Lung Hung, Victor Gurarie,  and Cheng Chin, “From cosmology to cold atoms: Observation of sakharov oscillations in a quenched atomic superfluid,” Science 341, 1213–1215 (2013)https://www.science.org/doi/pdf/10.1126/science.1237557 .
  • Prain et al. (2010) Angus Prain, Serena Fagnocchi,  and Stefano Liberati, “Analogue cosmological particle creation: Quantum correlations in expanding bose-einstein condensates,” Phys. Rev. D 82, 105018 (2010).
  • Fedichev and Fischer (2003) Petr O. Fedichev and Uwe 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).
  • Alsing et al. (2005) Paul M. Alsing, Jonathan P. Dowling,  and G. J. Milburn, “Ion trap simulations of quantum fields in an expanding universe,” Phys. Rev. Lett. 94, 220401 (2005).
  • Tian et al. (2017) Zehua Tian, Jiliang Jing,  and Andrzej Dragan, “Analog cosmological particle generation in a superconducting circuit,” Phys. Rev. D 95, 125003 (2017).
  • Schützhold et al. (2007) Ralf Schützhold, Michael Uhlmann, Lutz Petersen, Hector Schmitz, Axel Friedenauer,  and Tobias Schätz, “Analogue of cosmological particle creation in an ion trap,” Phys. Rev. Lett. 99, 201301 (2007).
  • Eckel et al. (2018) S. Eckel, A. Kumar, T. Jacobson, I. B. Spielman,  and G. K. Campbell, “A rapidly expanding bose-einstein condensate: An expanding universe in the lab,” Phys. Rev. X 8, 021021 (2018).
  • Lang and Schützhold (2019) Sascha Lang and Ralf Schützhold, “Analog of cosmological particle creation in electromagnetic waveguides,” Phys. Rev. D 100, 065003 (2019).
  • Chä and Fischer (2017) Seok-Yeong Chä and Uwe R. Fischer, “Probing the scale invariance of the inflationary power spectrum in expanding quasi-two-dimensional dipolar condensates,” Phys. Rev. Lett. 118, 130404 (2017).
  • Wittemer et al. (2019) Matthias Wittemer, Frederick Hakelberg, Philip Kiefer, Jan-Philipp Schröder, Christian Fey, Ralf Schützhold, Ulrich Warring,  and Tobias Schaetz, “Phonon pair creation by inflating quantum fluctuations in an ion trap,” Phys. Rev. Lett. 123, 180502 (2019).
  • Bhardwaj et al. (2021) Anshuman Bhardwaj, Dzmitry Vaido,  and Daniel E. Sheehy, “Inflationary dynamics and particle production in a toroidal bose-einstein condensate,” Phys. Rev. A 103, 023322 (2021).
  • Banik et al. (2022) S. Banik, M. Gutierrez Galan, H. Sosa-Martinez, M. Anderson, S. Eckel, I. B. Spielman,  and G. K. Campbell, “Accurate determination of hubble attenuation and amplification in expanding and contracting cold-atom universes,” Phys. Rev. Lett. 128, 090401 (2022).
  • Johansson et al. (2009) J. R. Johansson, G. Johansson, C. M. Wilson,  and Franco Nori, “Dynamical casimir effect in a superconducting coplanar waveguide,” Phys. Rev. Lett. 103, 147003 (2009).
  • Fujii et al. (2011) Toshiyuki Fujii, Shigemasa Matsuo, Noriyuki Hatakenaka, Susumu Kurihara,  and Anton Zeilinger, “Quantum circuit analog of the dynamical casimir effect,” Phys. Rev. B 84, 174521 (2011).
  • Wilson et al. (2011) C. M. Wilson, G. Johansson, A. Pourkabirian, M. Simoen, J. R. Johansson, T. Duty, F. Nori,  and P. Delsing, “Observation of the dynamical casimir effect in a superconducting circuit,” Nature 479, 376–379 (2011).
  • Jaskula et al. (2012) J.-C. Jaskula, G. B. Partridge, M. Bonneau, R. Lopes, J. Ruaudel, D. Boiron,  and C. I. Westbrook, “Acoustic analog to the dynamical casimir effect in a bose-einstein condensate,” Phys. Rev. Lett. 109, 220401 (2012).
  • Lähteenmäki et al. (2013) Pasi Lähteenmäki, G. S. Paraoanu, Juha Hassel,  and Pertti J. Hakonen, “Dynamical casimir effect in a josephson metamaterial,” Proceedings of the National Academy of Sciences 110, 4234–4238 (2013)https://www.pnas.org/content/110/11/4234.full.pdf .
  • Macrì et al. (2018) Vincenzo Macrì, Alessandro Ridolfo, Omar Di Stefano, Anton Frisk Kockum, Franco Nori,  and Salvatore Savasta, “Nonperturbative dynamical casimir effect in optomechanical systems: Vacuum casimir-rabi splittings,” Phys. Rev. X 8, 011031 (2018).
  • Koghee and Wouters (2014) Selma Koghee and Michiel Wouters, “Dynamical casimir emission from polariton condensates,” Phys. Rev. Lett. 112, 036406 (2014).
  • Dodonov (2020) Viktor Dodonov, “Fifty years of the dynamical casimir effect,” Physics 2, 67–104 (2020).
  • Schneider et al. (2020) B. H. Schneider, A. Bengtsson, I. M. Svensson, T. Aref, G. Johansson, Jonas Bylander,  and P. Delsing, “Observation of broadband entanglement in microwave radiation from a single time-varying boundary condition,” Phys. Rev. Lett. 124, 140503 (2020).
  • Chen and Tajima (1999) Pisin Chen and Toshi Tajima, “Testing unruh radiation with ultraintense lasers,” Phys. Rev. Lett. 83, 256–259 (1999).
  • Lochan et al. (2020) Kinjalk Lochan, Hendrik Ulbricht, Andrea Vinante,  and Sandeep K. Goyal, “Detecting acceleration-enhanced vacuum fluctuations with atoms inside a cavity,” Phys. Rev. Lett. 125, 241301 (2020).
  • Schützhold et al. (2008) Ralf Schützhold, Gernot Schaller,  and Dietrich Habs, “Tabletop creation of entangled multi-kev photon pairs and the unruh effect,” Phys. Rev. Lett. 100, 091301 (2008).
  • Retzker et al. (2008) A. Retzker, J. I. Cirac, M. B. Plenio,  and B. Reznik, “Methods for detecting acceleration radiation in a bose-einstein condensate,” Phys. Rev. Lett. 101, 110402 (2008).
  • Marino et al. (2020) Jamir Marino, Gabriel Menezes,  and Iacopo Carusotto, “Zero-point excitation of a circularly moving detector in an atomic condensate and phonon laser dynamical instabilities,” Phys. Rev. Research 2, 042009 (2020).
  • Gooding et al. (2020) Cisco Gooding, Steffen Biermann, Sebastian Erne, Jorma Louko, William G. Unruh, Jörg Schmiedmayer,  and Silke Weinfurtner, “Interferometric unruh detectors for bose-einstein condensates,” Phys. Rev. Lett. 125, 213603 (2020).
  • Sheng et al. (2021) Tianze Sheng, Jun Qian, Xiaolin Li, Yueping Niu,  and Shangqing Gong, “Quantum simulation of the unruh effect with a rydberg-dressed bose-einstein condensate,” Phys. Rev. A 103, 013301 (2021).
  • Hegde et al. (2019) Suraj S. Hegde, Varsha Subramanyan, Barry Bradlyn,  and Smitha Vishveshwara, “Quasinormal modes and the hawking-unruh effect in quantum hall systems: Lessons from black hole phenomena,” Phys. Rev. Lett. 123, 156802 (2019).
  • Zeng and Zubairy (2021) Xiaodong Zeng and M. Suhail Zubairy, “Graphene plasmon excitation with ground-state two-level quantum emitters,” Phys. Rev. Lett. 126, 117401 (2021).
  • Cozzella et al. (2017) Gabriel Cozzella, André G. S. Landulfo, George E. A. Matsas,  and Daniel A. T. Vanzella, “Proposal for observing the unruh effect using classical electrodynamics,” Phys. Rev. Lett. 118, 161102 (2017).
  • Barros et al. (2020) Guilherme B. Barros, João P. C. R. Rodrigues, André G. S. Landulfo,  and George E. A. Matsas, “Traces of the unruh effect in surface waves,” Phys. Rev. D 101, 065015 (2020).
  • Guedes et al. (2019) T. L. M. Guedes, M. Kizmann, D. V. Seletskiy, A. Leitenstorfer, Guido Burkard,  and A. S. Moskalenko, “Spectra of ultrabroadband squeezed pulses and the finite-time unruh-davies effect,” Phys. Rev. Lett. 122, 053604 (2019).
  • Adjei et al. (2020) Eugene Adjei, Kevin J. Resch,  and Agata M. Brańczyk, “Quantum simulation of unruh-dewitt detectors with nonlinear optics,” Phys. Rev. A 102, 033506 (2020).
  • Hu et al. (2019) Jiazhong Hu, Lei Feng, Zhendong Zhang,  and Cheng Chin, “Quantum simulation of unruh radiation,” Nature Physics 15, 785–789 (2019).
  • Nicolini and Rinaldi (2011) Piero Nicolini and Massimiliano Rinaldi, “A minimal length versus the unruh effect,” Physics Letters B 695, 303–306 (2011).
  • Agulló et al. (2008) Iván Agulló, José Navarro-Salas, Gonzalo J. Olmo,  and Leonard Parker, “Two-point functions with an invariant planck scale and thermal effects,” Phys. Rev. D 77, 124032 (2008).
  • Hossain and Sardar (2015) Golam Mortuza Hossain and Gopal Sardar, “Violation of the kubo-martin-schwinger condition along a rindler trajectory in polymer quantization,” Phys. Rev. D 92, 024018 (2015).
  • Kajuri (2016) Nirmalya Kajuri, “Polymer quantization predicts radiation in inertial frames,” Classical and Quantum Gravity 33, 055007 (2016).
  • Hossain and Sardar (2016) Golam Mortuza Hossain and Gopal Sardar, “Is there unruh effect in polymer quantization?” Classical and Quantum Gravity 33, 245016 (2016).
  • Alkofer et al. (2016) Natalia Alkofer, Giulio D’Odorico, Frank Saueressig,  and Fleur Versteegen, “Quantum gravity signatures in the unruh effect,” Phys. Rev. D 94, 104055 (2016).
  • Carballo-Rubio et al. (2019) Raúl Carballo-Rubio, Luis J. Garay, Eduardo Martín-Martínez,  and José de Ramón, “Unruh effect without thermality,” Phys. Rev. Lett. 123, 041601 (2019).
  • Hammad et al. (2021) F. Hammad, A. Landry,  and D. Dijamco, “Influence of the dispersion relation on the unruh effect according to the relativistic doppler shift method,” Phys. Rev. D 103, 085010 (2021).
  • Louko and Upton (2018) Jorma Louko and Samuel D. Upton, “Low-energy lorentz violation from high-energy modified dispersion in inertial and circular motion,” Phys. Rev. D 97, 025008 (2018).
  • Unruh (1995) W. G. Unruh, “Sonic analogue of black holes and the effects of high frequencies on black hole evaporation,” Phys. Rev. D 51, 2827–2838 (1995).
  • Baranov (2008) M.A. Baranov, “Theoretical progress in many-body physics with ultracold dipolar gases,” Physics Reports 464, 71–111 (2008).
  • Recati et al. (2005) A. Recati, P. O. Fedichev, W. Zwerger, J. von Delft,  and P. Zoller, “Atomic quantum dots coupled to a reservoir of a superfluid bose-einstein condensate,” Phys. Rev. Lett. 94, 040404 (2005).
  • Fischer (2006) Uwe R. Fischer, “Stability of quasi-two-dimensional bose-einstein condensates with dominant dipole-dipole interactions,” Phys. Rev. A 73, 031602 (2006).
  • Tian et al. (2018) Zehua Tian, Seok-Yeong Chä,  and Uwe R. Fischer, “Roton entanglement in quenched dipolar bose-einstein condensates,” Phys. Rev. A 97, 063611 (2018).
  • (72) N. D. Birrell and P. C. W. Davies, “Quantum fields in curved space,” Cambridge Monographs on Mathematical Physics (Cambridge University Press, Cambridge, England, 1982) .
  • Fedichev and Fischer (2004b) Petr O. Fedichev and Uwe R. Fischer, “Observer dependence for the phonon content of the sound field living on the effective curved space-time background of a bose-einstein condensate,” Phys. Rev. D 69, 064021 (2004b).
  • Pethick and Smith (2008) Christopher J Pethick and Henrik Smith, Bose–Einstein condensation in dilute gases (Cambridge university press, 2008).
  • Pitaevskiĭ and Stringari (2016) Lev Petrovich Pitaevskiĭ and Sandro Stringari, Bose-Einstein Condensation and Superfluidity, Vol. 164 (Oxford University Press, 2016).
  • Courteille et al. (1998) Ph. Courteille, R. S. Freeland, D. J. Heinzen, F. A. van Abeelen,  and B. J. Verhaar, “Observation of a feshbach resonance in cold atom scattering,” Phys. Rev. Lett. 81, 69–72 (1998).
  • Inouye et al. (1998) S. Inouye, M. R. Andrews, J. Stenger, H. J. Miesner, D. M. Stamper-Kurn,  and W. Ketterle, “Observation of feshbach resonances in a bose-einstein condensate,” Nature 392, 151–154 (1998).
  • Chin et al. (2010) Cheng Chin, Rudolf Grimm, Paul Julienne,  and Eite Tiesinga, “Feshbach resonances in ultracold gases,” Rev. Mod. Phys. 82, 1225–1286 (2010).
  • Timmermans et al. (1999) Eddy Timmermans, Paolo Tommasini, Mahir Hussein,  and Arthur Kerman, “Feshbach resonances in atomic bose–einstein condensates,” Physics Reports 315, 199–230 (1999).
  • Giovanazzi et al. (2002) Stefano Giovanazzi, Axel Görlitz,  and Tilman Pfau, “Tuning the dipolar interaction in quantum gases,” Phys. Rev. Lett. 89, 130401 (2002).
  • Tang et al. (2018) Yijun Tang, Wil Kao, Kuan-Yu Li,  and Benjamin L. Lev, “Tuning the dipole-dipole interaction in a quantum gas with a rotating magnetic field,” Phys. Rev. Lett. 120, 230401 (2018).
  • Marino et al. (2017) Jamir Marino, Alessio Recati,  and Iacopo Carusotto, “Casimir forces and quantum friction from ginzburg radiation in atomic bose-einstein condensates,” Phys. Rev. Lett. 118, 045301 (2017).
  • Tian and Du (2021) Zehua Tian and Jiangfeng Du, “Probing low-energy lorentz violation from high-energy modified dispersion in dipolar bose-einstein condensates,” Phys. Rev. D 103, 085014 (2021).
  • Carusotto et al. (2006) I. Carusotto, S. X. Hu, L. A. Collins,  and A. Smerzi, “Bogoliubov-Čerenkov radiation in a bose-einstein condensate flowing against an obstacle,” Phys. Rev. Lett. 97, 260403 (2006).
  • Astrakharchik and Pitaevskii (2004) G. E. Astrakharchik and L. P. Pitaevskii, “Motion of a heavy impurity through a bose-einstein condensate,” Phys. Rev. A 70, 013608 (2004).
  • Ginzburg and Frolov (1986) V. L. Ginzburg and V. P. Frolov, “Excitation and emission of a “detector” in accelerated motion in a vacuum or in uniform motion at a velocity above the velocity of light in a medium,” ZhETF Pisma Redaktsiiu 43, 265–267 (1986).
  • Ginzburg (1996) Vitalii L Ginzburg, “Radiation by uniformly moving sources (vavilov–cherenkov effect, transition radiation, and other phenomena),” Uspekhi Fizicheskikh Nauk (UFN) Journal 39, 973–982 (1996).
  • Husain and Louko (2016) Viqar Husain and Jorma Louko, “Low energy lorentz violation from modified dispersion at high energies,” Phys. Rev. Lett. 116, 061301 (2016).
  • Bell and Leinaas (1983) J.S. Bell and J.M. Leinaas, “Electrons as accelerated thermometers,” Nuclear Physics B 212, 131–150 (1983).
  • Bell and Leinaas (1987) J.S. Bell and J.M. Leinaas, “The unruh effect and quantum fluctuations of electrons in storage rings,” Nuclear Physics B 284, 488–508 (1987).
  • Unruh (1998) W.G. Unruh, “Acceleration radiation for orbiting electrons,” Physics Reports 307, 163–171 (1998).
  • (92) See Supplemental Material for detailed calculations of the detector’s transition rate.
  • Lu et al. (2011) Mingwu Lu, Nathaniel Q. Burdick, Seo Ho Youn,  and Benjamin L. Lev, “Strongly dipolar bose-einstein condensate of dysprosium,” Phys. Rev. Lett. 107, 190401 (2011).
  • Aikawa et al. (2012) K. Aikawa, A. Frisch, M. Mark, S. Baier, A. Rietzler, R. Grimm,  and F. Ferlaino, “Bose-einstein condensation of erbium,” Phys. Rev. Lett. 108, 210401 (2012).
  • Norcia et al. (2021) Matthew A. Norcia, Claudia Politi, Lauritz Klaus, Elena Poli, Maximilian Sohmen, Manfred J. Mark, Russell N. Bisset, Luis Santos,  and Francesca Ferlaino, “Two-dimensional supersolidity in a dipolar quantum gas,” Nature 596, 357–361 (2021).
  • Chomaz et al. (2018) L. Chomaz, R. M. W. van Bijnen, D. Petter, G. Faraoni, S. Baier, J. H. Becher, M. J. Mark, F. Wächtler, L. Santos,  and F. Ferlaino, “Observation of roton mode population in a dipolar quantum gas,” Nature Physics 14, 442–446 (2018).
  • Petter et al. (2019) D. Petter, G. Natale, R. M. W. van Bijnen, A. Patscheider, M. J. Mark, L. Chomaz,  and F. Ferlaino, “Probing the roton excitation spectrum of a stable dipolar bose gas,” Phys. Rev. Lett. 122, 183401 (2019).
  • Schmidt et al. (2021) J.-N. Schmidt, J. Hertkorn, M. Guo, F. Böttcher, M. Schmidt, K. S. H. Ng, S. D. Graham, T. Langen, M. Zwierlein,  and T. Pfau, “Roton excitations in an oblate dipolar quantum gas,” Phys. Rev. Lett. 126, 193002 (2021).
  • Natale et al. (2019) G. Natale, R. M. W. van Bijnen, A. Patscheider, D. Petter, M. J. Mark, L. Chomaz,  and F. Ferlaino, “Excitation spectrum of a trapped dipolar supersolid and its experimental evidence,” Phys. Rev. Lett. 123, 050402 (2019).
  • Hertkorn et al. (2021) J. Hertkorn, J.-N. Schmidt, F. Böttcher, M. Guo, M. Schmidt, K. S. H. Ng, S. D. Graham, H. P. Büchler, T. Langen, M. Zwierlein,  and T. Pfau, “Density fluctuations across the superfluid-supersolid phase transition in a dipolar quantum gas,” Phys. Rev. X 11, 011037 (2021).
  • Tomza et al. (2019) Michał Tomza, Krzysztof Jachymski, Rene Gerritsma, Antonio Negretti, Tommaso Calarco, Zbigniew Idziaszek,  and Paul S. Julienne, “Cold hybrid ion-atom systems,” Rev. Mod. Phys. 91, 035001 (2019).
  • Weckesser et al. (2021) Pascal Weckesser, Fabian Thielemann, Dariusz Wiater, Agata Wojciechowska, Leon Karpa, Krzysztof Jachymski, Michał Tomza, Thomas Walker,  and Tobias Schaetz, “Observation of feshbach resonances between a single ion and ultracold atoms,” Nature 600, 429–433 (2021).
  • Skou et al. (2021) Magnus G. Skou, Thomas G. Skov, Nils B. Jørgensen, Kristian K. Nielsen, Arturo Camacho-Guardian, Thomas Pohl, Georg M. Bruun,  and Jan J. Arlt, “Non-equilibrium quantum dynamics and formation of the bose polaron,” Nature Physics 17, 731–735 (2021).
  • Dieterle et al. (2021) T. Dieterle, M. Berngruber, C. Hölzl, R. Löw, K. Jachymski, T. Pfau,  and F. Meinert, “Transport of a single cold ion immersed in a bose-einstein condensate,” Phys. Rev. Lett. 126, 033401 (2021).
  • Kubo (1957) Ryogo Kubo, “Statistical-mechanical theory of irreversible processes. i. general theory and simple applications to magnetic and conduction problems,” Journal of the Physical Society of Japan 12, 570–586 (1957)https://doi.org/10.1143/JPSJ.12.570 .
  • Martin and Schwinger (1959) Paul C. Martin and Julian Schwinger, “Theory of many-particle systems. i,” Phys. Rev. 115, 1342–1373 (1959).
  • Haag et al. (1967) R. Haag, N. M. Hugenholtz,  and M. Winnink, “On the equilibrium states in quantum statistical mechanics,” Communications in Mathematical Physics 5, 215–236 (1967).
  • Biermann et al. (2020) Steffen Biermann, Sebastian Erne, Cisco Gooding, Jorma Louko, Jörg Schmiedmayer, William G. Unruh,  and Silke Weinfurtner, “Unruh and analogue unruh temperatures for circular motion in 3+13+1 and 2+12+1 dimensions,” Phys. Rev. D 102, 085006 (2020).
  • Brenna et al. (2016) W.G. Brenna, Robert B. Mann,  and Eduardo Martín-Martínez, “Anti-unruh phenomena,” Physics Letters B 757, 307–311 (2016).
  • Purcell and Pound (1951) E. M. Purcell and R. V. Pound, “A nuclear spin system at negative temperature,” Phys. Rev. 81, 279–280 (1951).
  • Ramsey (1956) Norman F. Ramsey, “Thermodynamics and statistical mechanics at negative absolute temperatures,” Phys. Rev. 103, 20–28 (1956).
  • Pozas-Kerstjens and Martín-Martínez (2015) Alejandro Pozas-Kerstjens and Eduardo Martín-Martínez, “Harvesting correlations from the quantum vacuum,” Phys. Rev. D 92, 064042 (2015).

Supplementary Material

I Unruh effect of accelerated detector

We here simply review the Unruh effect corresponding to the linear acceleration and circular motion cases. In quantum field theory, usually quantum field is probed with a linearly coupled Unruh-DeWitt detector Unruh (1976); Birrell and Davies , which is described by a localized system with internal levels |g|g\rangle and |e=σ+|g|e\rangle=\sigma^{+}|g\rangle and the energy gap ω0\omega_{0}, moving along a trajectory (t(τ),𝐱(τ))(t(\tau),{\bf x}(\tau)) with τ\tau being the detector’s proper time. The detector couples with a scalar field ϕ\phi, initially in its vacuum state, through

Hint=gχ(τ)(eiω0τσ+eiω0τσ)ϕ(𝐱(τ),t(τ)),\displaystyle H_{\text{int}}=g\chi(\tau)(e^{i\omega_{0}\tau}\sigma^{+}e^{-i\omega_{0}\tau}\sigma_{-})\phi({\bf x}(\tau),t(\tau)), (S1)

where gg is the coupling parameter and χ(τ)\chi(\tau) is a real-valued smooth switching function that specifies how the interaction is turned on and off. In the first-order perturbation theory, the probability for the detector to be excited from its ground |g|g\rangle to the excited state |e|e\rangle is proportional to the response function,

(ω0)=g22𝑑τ𝑑τχ(τ)χ(τ)eiω0(ττ)𝒲(τ,τ),\displaystyle{\cal F}(\omega_{0})=\frac{g^{2}}{\hbar^{2}}\int\int\,d\tau\,d\tau^{\prime}\chi(\tau)\chi(\tau^{\prime})e^{-i\omega_{0}(\tau-\tau^{\prime})}{\cal W}(\tau,\tau^{\prime}),

where 𝒲(τ,τ)=0|ϕ(t(τ),𝐱(τ))ϕ(t(τ),𝐱(τ))|0{\cal W}(\tau,\tau^{\prime})=\langle 0|\phi(t(\tau),{\bf x}(\tau))\phi(t(\tau^{\prime}),{\bf x}(\tau^{\prime}))|0\rangle denotes the Wightman function evaluated along the detector’s trajectory. If we consider the scenario where the detectors’s trajectories and quantum field state are stationary, in this sense that 𝒲(τ,τ){\cal W}(\tau,\tau^{\prime}) depends on its arguments only through the difference s=ττs=\tau-\tau^{\prime}. We may then calculate the transition probability per unit time (or the transition rate) by dividing (ω0){\cal F}(\omega_{0}) with respective to the total interaction time and letting this interaction time tend to infinity, finally find

𝒫(ω0)=rate(ω0)=g22𝑑seiω0s𝒲(s,0).\displaystyle{\cal P}(\omega_{0})={\cal F}_{\text{rate}}(\omega_{0})=\frac{g^{2}}{\hbar^{2}}\int_{-\infty}^{\infty}\,ds\,e^{-i\omega_{0}s}{\cal W}(s,0). (S2)

Note (S2) denotes the excitation rate from the detector’s ground state to its excited state.

The Wightman function for a massless scalar field in Minkowski spacetime is analytically 𝒲(τ,τ)=c04π21(c0t(τ)c0t(τ))2(𝐱(τ)𝐱(τ))2{\cal W}(\tau,\tau^{\prime})=-\frac{\hbar\,c_{0}}{4\pi^{2}}\frac{1}{(c_{0}t(\tau)-c_{0}t^{\prime}(\tau^{\prime}))^{2}-({\bf x}(\tau)-{\bf x}^{\prime}(\tau^{\prime}))^{2}}. For the linearly accelerated trajectory, x(τ)=c02acoshaτc0,t(τ)=c0asinhaτc0x(\tau)=\frac{c^{2}_{0}}{a}\cosh\frac{a\tau}{c_{0}},t(\tau)=\frac{c_{0}}{a}\sinh\frac{a\tau}{c_{0}}, it reduces to

𝒲l(τ,τ)\displaystyle{\cal W}_{l}(\tau,\tau^{\prime}) =\displaystyle= a216π2c03{sinh[a2c0(ττ)]}2\displaystyle-\frac{a^{2}\hbar}{16\pi^{2}c^{3}_{0}}\bigg{\{}\sinh\bigg{[}\frac{a}{2c_{0}}(\tau-\tau^{\prime})\bigg{]}\bigg{\}}^{-2} (S3)
=\displaystyle= 4π2c0(ττ)2(1+112[ac0(ττ)]2+1360[ac0(ττ)]4+)1.\displaystyle-\frac{\hbar}{4\pi^{2}c_{0}}(\tau-\tau^{\prime})^{-2}\bigg{(}1+\frac{1}{12}\bigg{[}\frac{a}{c_{0}}(\tau-\tau^{\prime})\bigg{]}^{2}+\frac{1}{360}\bigg{[}\frac{a}{c_{0}}(\tau-\tau^{\prime})\bigg{]}^{4}+\dots\bigg{)}^{-1}.

Together with (S2), one can find that the accelerated detector becomes thermalized with the transition probabilities satisfying 𝒫excitation/𝒫de-excitation=𝒫(ω0)/𝒫(ω0)=eω0/kBTU{\cal P}_{\text{excitation}}/{\cal P}_{\text{de-excitation}}={\cal P}(\omega_{0})/{\cal P}(-\omega_{0})=e^{-\hbar\omega_{0}/k_{\text{B}}T_{\text{U}}}, where TU=a2πc0kBT_{\text{U}}=\frac{\hbar\,a}{2\pi\,c_{0}k_{B}} is the Unruh temperature. Note that observation of Unruh effect in the practice remains a challenged problem because of the extreme requirement of high linearly acceleration for typical detector’s transition.

Going beyond the linear acceleration scenario, the circular motion with a constant radial acceleration could also produce an approximately thermal spectrum Bell and Leinaas (1983, 1987); Unruh (1998), presenting the circular Unruh effect. Specifically, if the detector moves along a circular trajectory (c0t(τ),𝐱(τ))=(c0γτ,Rcos(Ωγτ),Rsin(Ωγτ),0)(c_{0}t(\tau),{\bf x}(\tau))=(c_{0}\gamma\tau,R\cos(\Omega\gamma\tau),R\sin(\Omega\gamma\tau),0), with constant radius RR, the usual relativistic factor γ=(1v2/c02)1/2=(1β2)1/2\gamma=(1-v^{2}/c^{2}_{0})^{-1/2}=(1-\beta^{2})^{-1/2}, the angular velocity Ω=v/R\Omega=v/R and the corresponding acceleration a=v2γ2/R=Ω2γ2Ra=v^{2}\gamma^{2}/R=\Omega^{2}\gamma^{2}R, one can find the corresponding Wightman function

𝒲c(τ,τ)\displaystyle{\cal W}_{c}(\tau,\tau^{\prime}) =\displaystyle= 4π2c0{γ2(ττ)2(2c0aβ2γ2)2(sin[a2c0ττβγ])2}1\displaystyle-\frac{\hbar}{4\pi^{2}c_{0}}\bigg{\{}\gamma^{2}(\tau-\tau^{\prime})^{2}-\bigg{(}2\frac{c_{0}}{a}\beta^{2}\gamma^{2}\bigg{)}^{2}\bigg{(}\sin\bigg{[}\frac{a}{2c_{0}}\frac{\tau-\tau^{\prime}}{\beta\gamma}\bigg{]}\bigg{)}^{2}\bigg{\}}^{-1} (S4)
=\displaystyle= 4π2c0(ττ)2(1+112[ac0(ττ)]21360β2γ2[ac0(ττ)]4+)1.\displaystyle-\frac{\hbar}{4\pi^{2}c_{0}}(\tau-\tau^{\prime})^{-2}\bigg{(}1+\frac{1}{12}\bigg{[}\frac{a}{c_{0}}(\tau-\tau^{\prime})\bigg{]}^{2}-\frac{1}{360\beta^{2}\gamma^{2}}\bigg{[}\frac{a}{c_{0}}(\tau-\tau^{\prime})\bigg{]}^{4}+\dots\bigg{)}^{-1}.

Comparing the Wightman function (S4) with (S3), there is distinct difference: an extra parameter, β\beta or γ\gamma, appears in the circular motion case. This is because for circular motion the radius of the circle can be varied independently of the acceleration. Unlike the linear accelerated case, the Fourier transformation of (S4) is not analytical. However, it, in the ultrarelativistic limit, γ1\gamma\gg 1, can be simplified. Then the corresponding equilibrium transition probabilities yields Bell and Leinaas (1983)

𝒫excitation𝒫de-excitationa43ω0c0e23ω0c0a\displaystyle\frac{{\cal P}_{\text{excitation}}}{{\cal P}_{\text{de-excitation}}}\approx\frac{a}{4\sqrt{3}\omega_{0}c_{0}}e^{-2\sqrt{3}\frac{\omega_{0}c_{0}}{a}} (S5)

for the a/c0ω01a/c_{0}\omega_{0}\ll 1 case. It leads to an effective temperature Teff=a23kBc0T_{\text{eff}}=\frac{\hbar\,a}{2\sqrt{3}k_{B}c_{0}}, which is higher by a factor π/3\pi/\sqrt{3} than the Unruh temperature for the linear acceleration.

Note that compared with the linear acceleration case, circular motion allows the accelerating system to remain within a finite-size laboratory for an arbitrarily long interaction time. Furthermore, unlike what happens in uniform linear acceleration, the proper and coordinate time for the circular motion are related by a time-independent gamma faction, which will be crucial when estimating the experimental feasibility for detecting the analogue circular Unruh effect.

II Derivation of the transition rate

The density fluctuations of dipolar BEC is of the form,

δρ^(t,𝐫)=ρ0d2𝐤(2π)2(u𝐤+v𝐤)[b^𝐤(t)ei𝐤𝐫+b^𝐤(t)ei𝐤𝐫].\displaystyle\delta\hat{\rho}(t,\mathbf{r})=\sqrt{\rho_{0}}\int\frac{d^{2}\mathbf{k}}{(2\pi)^{2}}(u_{\mathbf{k}}+v_{\mathbf{k}})\big{[}\hat{b}_{\mathbf{k}}(t)e^{i\mathbf{k}\cdot\mathbf{r}}+\hat{b}^{\dagger}_{\mathbf{k}}(t)e^{-i\mathbf{k}\cdot\mathbf{r}}\big{]}. (S6)

We can use it to calculate the analogue correlation function of the density fluctuations

0|δρ^(t,𝐫)δρ^(t,𝐫)|0\displaystyle\langle 0|\delta\hat{\rho}(t,\mathbf{r})\delta\hat{\rho}(t^{\prime},\mathbf{r}^{\prime})|0\rangle =\displaystyle= ρ0(2π)4d2𝐤d2𝐤(uk+vk)(uk+vk)0|[b^𝐤(t)ei𝐤𝐫+h.c.][b^𝐤(t)ei𝐤𝐫+h.c.]|0\displaystyle\frac{\rho_{0}}{(2\pi)^{4}}\int\int\,d^{2}\mathbf{k}\,d^{2}\mathbf{k}^{\prime}(u_{k}+v_{k})(u_{k^{\prime}}+v_{k^{\prime}})\langle 0|\big{[}\hat{b}_{\mathbf{k}}(t)e^{i\mathbf{k}\cdot\mathbf{r}}+\mathrm{h.c.}\big{]}\big{[}\hat{b}_{\mathbf{k}^{\prime}}(t^{\prime})e^{i\mathbf{k}^{\prime}\cdot\mathbf{r}^{\prime}}+\mathrm{h.c.}\big{]}|0\rangle (S7)
=\displaystyle= ρ0(2π)2d2𝐤d2𝐤(uk+vk)(uk+vk)eiω𝐤t+iω𝐤tei𝐤𝐫i𝐤𝐫δ2(𝐤𝐤)\displaystyle\frac{\rho_{0}}{(2\pi)^{2}}\int\int\,d^{2}\mathbf{k}\,d^{2}\mathbf{k}^{\prime}(u_{k}+v_{k})(u_{k^{\prime}}+v_{k^{\prime}})e^{-i\omega_{\mathbf{k}}\,t+i\omega_{\mathbf{k}^{\prime}}t^{\prime}}e^{i\mathbf{k}\cdot\mathbf{r}-i\mathbf{k}^{\prime}\cdot\mathbf{r}^{\prime}}\delta^{2}(\mathbf{k}-\mathbf{k}^{\prime})
=\displaystyle= ρ0(2π)2d2𝐤(uk+vk)2eiω𝐤(tt)ei𝐤(𝐫𝐫)\displaystyle\frac{\rho_{0}}{(2\pi)^{2}}\int\,d^{2}\mathbf{k}(u_{k}+v_{k})^{2}e^{-i\omega_{\mathbf{k}}(t-t^{\prime})}e^{i\mathbf{k}\cdot(\mathbf{r}-\mathbf{r}^{\prime})}
=\displaystyle= ρ0(2π)20𝑑k02π𝑑θk(uk+vk)2eiω𝐤(tt)ei|𝐤||𝐫𝐫|cosθ\displaystyle\frac{\rho_{0}}{(2\pi)^{2}}\int^{\infty}_{0}dk\int^{2\pi}_{0}d\theta\,k(u_{k}+v_{k})^{2}e^{-i\omega_{\mathbf{k}}(t-t^{\prime})}e^{i|\mathbf{k}||\mathbf{r}-\mathbf{r}^{\prime}|\cos\theta}
=\displaystyle= ρ0(2π)0𝑑kk(uk+vk)2eiω𝐤(tt)J0(|𝐤||𝐫𝐫|),\displaystyle\frac{\rho_{0}}{(2\pi)}\int^{\infty}_{0}dk\,k(u_{k}+v_{k})^{2}e^{-i\omega_{\mathbf{k}}(t-t^{\prime})}J_{0}\big{(}|\mathbf{k}||\mathbf{r}-\mathbf{r}^{\prime}|\big{)},

where J0(x)J_{0}(x) is the Bessel function of the first kind. For the uniform circular motion case, the trajectory is (t(τ),𝐫(τ))=(γτ,Rcos(γΩτ),Rsin(γΩτ))(t(\tau),\mathbf{r}(\tau))=(\gamma\tau,R\cos(\gamma\Omega\tau),R\sin(\gamma\Omega\tau)), and it is easy to find tt=γ(ττ)=γst-t^{\prime}=\gamma(\tau-\tau^{\prime})=\gamma\,s and |𝐫𝐫|=2RsinγΩ2s|\mathbf{r}-\mathbf{r}^{\prime}|=2R\sin\frac{\gamma\Omega}{2}s. Inserting this trajectory into the above correlation function of the density fluctuations, we can calculate the transition rate

𝒫(ω0)\displaystyle{\cal P}(\omega_{0}) =\displaystyle= g2ρ02π2𝑑s0𝑑kk(uk+vk)2J0(2RksinγΩ2s)ei(ω0+γω𝐤)s\displaystyle\frac{g^{2}_{-}\rho_{0}}{2\pi\hbar^{2}}\int\,ds\int^{\infty}_{0}dk\,k(u_{k}+v_{k})^{2}J_{0}\bigg{(}2Rk\sin\frac{\gamma\Omega}{2}s\bigg{)}e^{-i(\omega_{0}+\gamma\omega_{\mathbf{k}})s} (S8)
=\displaystyle= g2ρ02π2𝑑s0𝑑kk(uk+vk)2m=Jm2(Rk)eimγΩsei(ω0+γω𝐤)s\displaystyle\frac{g^{2}_{-}\rho_{0}}{2\pi\hbar^{2}}\int\,ds\int^{\infty}_{0}dk\,k(u_{k}+v_{k})^{2}\sum^{\infty}_{m=-\infty}J^{2}_{m}(Rk)e^{im\gamma\Omega\,s}e^{-i(\omega_{0}+\gamma\omega_{\mathbf{k}})s}
=\displaystyle= g2ρ02γ0𝑑kk(uk+vk)2m=Jm2(Rk)δ(ω0γ+ω𝐤mΩ)\displaystyle\frac{g^{2}_{-}\rho_{0}}{\hbar^{2}\gamma}\int^{\infty}_{0}dk\,k(u_{k}+v_{k})^{2}\sum^{\infty}_{m=-\infty}J^{2}_{m}(Rk)\delta\big{(}\frac{\omega_{0}}{\gamma}+\omega_{\mathbf{k}}-m\Omega\big{)}
=\displaystyle= g2ρ0mB230𝑑ζζ2γf(ζ)m=Jm2(M~ζ)δ(ζf(ζ)1M~(mvE~γ)),\displaystyle\frac{g^{2}_{-}\rho_{0}m_{B}}{2\hbar^{3}}\int^{\infty}_{0}d\zeta\frac{\zeta^{2}}{\gamma\,f(\zeta)}\sum^{\infty}_{m=-\infty}J^{2}_{m}(\tilde{M}\zeta)\delta\bigg{(}\zeta\,f(\zeta)-\frac{1}{\tilde{M}}\bigg{(}mv-\frac{\tilde{E}}{\gamma}\bigg{)}\bigg{)},

where ζ=c0k/M\zeta=\hbar\,c_{0}k/M_{\ast}, M~=RM/c0\tilde{M}=RM_{\ast}/\hbar\,c_{0}, E~=Rω0/c0\tilde{E}=R\omega_{0}/c_{0}, and v=RΩ/c0v=R\Omega/c_{0} are dimensionless parameters. Note that to derive the second equality the identity, J0(2asinx)=m𝐙Jm2(a)e2imxJ_{0}(2a\sin\,x)=\sum_{m\in\mathbf{Z}}J^{2}_{m}(a)e^{2imx}, has been used.

II.1 Thermal circular Unruh effect

If R0=0R_{0}=0, it means only contact interaction happens between atoms or molecules in Bose gas, we can find the excitation spectrum has the form, ω𝐤=c0k1+ζ2/4\omega_{\mathbf{k}}=c_{0}k\sqrt{1+\zeta^{2}/4}, which is “relativistic” Retzker et al. (2008). ω𝐤c0k\omega_{\mathbf{k}}\approx\,c_{0}k, for kkc=M/c0k\ll\,k_{c}=M_{\ast}/c_{0}\hbar. Then, (S8) reduces to,

𝒫(ω0)\displaystyle{\cal P}(\omega_{0}) =\displaystyle= g2ρ02π2𝑑s0𝑑kk(uk+vk)2J0(2RksinγΩ2s)ei(ω0+γω𝐤)s\displaystyle\frac{g^{2}_{-}\rho_{0}}{2\pi\hbar^{2}}\int\,ds\int^{\infty}_{0}dk\,k(u_{k}+v_{k})^{2}J_{0}\bigg{(}2Rk\sin\frac{\gamma\Omega}{2}s\bigg{)}e^{-i(\omega_{0}+\gamma\omega_{\mathbf{k}})s} (S9)
=\displaystyle= g2ρ02π2𝑑s0𝑑kk2k2/2mBc0kJ0(2RksinγΩ2s)eiγc0kseiω0s\displaystyle\frac{g^{2}_{-}\rho_{0}}{2\pi\hbar^{2}}\int\,ds\int^{\infty}_{0}dk\,k\frac{\hbar^{2}k^{2}/2m_{B}}{\hbar\,c_{0}k}J_{0}\bigg{(}2Rk\sin\frac{\gamma\Omega}{2}s\bigg{)}e^{-i\gamma\,c_{0}ks}e^{-i\omega_{0}s}
=\displaystyle= g2ρ04πmBc0𝑑s0𝑑kk2J0(2RksinγΩ2s)eiγc0kseiω0s\displaystyle\frac{g^{2}_{-}\rho_{0}}{4\pi\hbar\,m_{B}c_{0}}\int\,ds\int^{\infty}_{0}dk\,k^{2}J_{0}\bigg{(}2Rk\sin\frac{\gamma\Omega}{2}s\bigg{)}e^{-i\gamma\,c_{0}ks}e^{-i\omega_{0}s}
=\displaystyle= g2ρ04πmBc0𝑑s2γ2c02s24R2sin2γΩ2s(γ2c02s2+4R2sin2γΩ2s)5/2eiω0s.\displaystyle\frac{g^{2}_{-}\rho_{0}}{4\pi\hbar\,m_{B}c_{0}}\int\,ds\frac{-2\gamma^{2}c^{2}_{0}s^{2}-4R^{2}\sin^{2}\frac{\gamma\Omega}{2}s}{\big{(}-\gamma^{2}c^{2}_{0}s^{2}+4R^{2}\sin^{2}\frac{\gamma\Omega}{2}s\big{)}^{5/2}}e^{-i\omega_{0}s}.

In the ultrarelativistic limit, γ1\gamma\gg 1, we can do the same process as in Bell and Leinaas (1983) to further calculate the above transition rate, which is given by

𝒫(ω0)\displaystyle{\cal P}(\omega_{0}) =\displaystyle= g2ρ04πmBc0𝑑s24i(1+3γ2)24c03s3+5a2c0s5eiω0s\displaystyle\frac{g^{2}_{-}\rho_{0}}{4\pi\hbar\,m_{B}c_{0}}\int\,ds\frac{24i(-1+3\gamma^{2})}{24c^{3}_{0}s^{3}+5a^{2}c_{0}s^{5}}e^{-i\omega_{0}s} (S10)
=\displaystyle= 5g2ρ0962πmBc06(3γ21)a2[1exp(245c0ω0a)+12c02ω025a2].\displaystyle\frac{5g^{2}_{-}\rho_{0}}{96\sqrt{2\pi}\hbar\,m_{B}c^{6}_{0}}(3\gamma^{2}-1)a^{2}\bigg{[}1-\exp\big{(}-\sqrt{\frac{24}{5}}\frac{c_{0}\omega_{0}}{a}\big{)}+\frac{12c^{2}_{0}\omega^{2}_{0}}{5a^{2}}\bigg{]}.

Furthermore, assuming the energy splitting of the detector to be not too small ω0a/c0\omega_{0}\gg\,a/c_{0}, we can find the equilibrium population of the upper level relative to the lower is

𝒫(ω0)𝒫(ω0)=12c02ω025a2exp(245c0ω0a),\displaystyle\frac{{\cal P}(\omega_{0})}{{\cal P}(-\omega_{0})}=\frac{12c^{2}_{0}\omega^{2}_{0}}{5a^{2}}\exp\big{(}-\sqrt{\frac{24}{5}}\frac{c_{0}\omega_{0}}{a}\big{)}, (S11)

leading to an effective temperature

Teff=5a26kBc0.\displaystyle T_{\text{eff}}=\frac{\sqrt{5}\hbar\,a}{2\sqrt{6}k_{B}c_{0}}. (S12)

II.2 Correction to the LI case

As shown above, the transition rate of the detector from its ground state to excited state is

𝒫(ω0)=g2ρ0mB230𝑑ζζ2γf(ζ)m=Jm2(M~ζ)δ(ζf(ζ)1M~(mvE~γ)),\displaystyle{\cal P}(\omega_{0})=\frac{g^{2}_{-}\rho_{0}m_{B}}{2\hbar^{3}}\int^{\infty}_{0}d\zeta\frac{\zeta^{2}}{\gamma\,f(\zeta)}\sum^{\infty}_{m=-\infty}J^{2}_{m}(\tilde{M}\zeta)\delta\bigg{(}\zeta\,f(\zeta)-\frac{1}{\tilde{M}}\bigg{(}mv-\frac{\tilde{E}}{\gamma}\bigg{)}\bigg{)}, (S13)

where M~=RM/c0\tilde{M}=RM_{\ast}/\hbar\,c_{0}, E~=Rω0/c0\tilde{E}=R\omega_{0}/c_{0}, and v=RΩ/c0v=R\Omega/c_{0} are dimensionless parameters.

We consider the stable dipolar BEC and thus f(ζ)f(\zeta) is smooth and strictly positive, and f(ζ)1f(\zeta)\rightarrow 1 as ζ0\zeta\rightarrow 0. We also consider the scenario where the only stationary point of ff is a global minimum at g=gc>0g=g_{c}>0, written as fc=f(gc)f_{c}=f(g_{c}) with 0<fc<10<f_{c}<1. In such case, f(ζ)<0f^{\prime}(\zeta)<0 for 0<ζ<ζc0<\zeta<\zeta_{c} and f(ζ)>0f^{\prime}(\zeta)>0 for ζ>ζc\zeta>\zeta_{c}. In addition to that, we also assume that ζf(ζ)\zeta\,f(\zeta) is a monotonely increasing function of ζ\zeta, which means (ζf(ζ))=ζf(ζ)+f(ζ)>0(\zeta\,f(\zeta))^{\prime}=\zeta\,f^{\prime}(\zeta)+f(\zeta)>0 for ζ>0\zeta>0. Then we can perform the integral in Eq. (S13), and find

𝒫(ω0)=g2ρ0mB23γm=E~vγζm2/f(ζm)(ζf(ζ))|ζ=ζmJm2(M~ζm),\displaystyle{\cal P}(\omega_{0})=\frac{g^{2}_{-}\rho_{0}m_{B}}{2\hbar^{3}\gamma}\sum^{\infty}_{m=\lceil\,\frac{\tilde{E}}{v\gamma}\rceil}\frac{\zeta^{2}_{m}/f(\zeta_{m})}{(\zeta\,f(\zeta))^{\prime}|_{\zeta=\zeta_{m}}}J^{2}_{m}(\tilde{M}\zeta_{m}), (S14)

where ζm\zeta_{m} is the unique solution to ζf(ζ)=1M~(mvE~γ)\zeta\,f(\zeta)=\frac{1}{\tilde{M}}\big{(}mv-\frac{\tilde{E}}{\gamma}\big{)}. In the limit M~\tilde{M}\rightarrow\infty, we find that the limit is qualitatively different for 0<v<fc0<v<f_{c} and fc<v<1f_{c}<v<1, as found in Refs. Louko and Upton (2018):

𝒫0(ω0)=g2ρ02MR2γm=E~vγ(mvE~γ)2Jm2(mvE~γ),\displaystyle{\cal P}_{0}(\omega_{0})=\frac{g^{2}_{-}\rho_{0}}{2\hbar\,M_{\ast}\,R^{2}\gamma}\sum^{\infty}_{m=\lceil\,\frac{\tilde{E}}{v\gamma}\rceil}\bigg{(}mv-\frac{\tilde{E}}{\gamma}\bigg{)}^{2}J^{2}_{m}\big{(}mv-\frac{\tilde{E}}{\gamma}\big{)}, (S15)

for the 0<v<fc0<v<f_{c} case. Note that in this case the detector sees no low-energy Lorentz violation: the corresponding response is the same as that for the usual massless scalar field. For fc<v<1f_{c}<v<1, 𝒫(ω0)𝒫0(ω0)+Δ𝒫{\cal P}(\omega_{0})\rightarrow{\cal P}_{0}(\omega_{0})+\Delta{\cal P} as M~\tilde{M}\rightarrow\infty, where

Δ𝒫=g2ρ0mB2π3γζζ+𝑑ζζf(ζ)v2f2(ζ),\displaystyle\Delta{\cal P}=\frac{g^{2}_{-}\rho_{0}m_{B}}{2\pi\hbar^{3}\gamma}\int^{\zeta_{+}}_{\zeta_{-}}d\zeta\frac{\zeta}{f(\zeta)\sqrt{v^{2}-f^{2}(\zeta)}}, (S16)

and ζ(0,gc)\zeta_{-}\in(0,g_{c}) and ζ+(gc,)\zeta_{+}\in(g_{c},\infty) are unique solutions to f(ζ)=vf(\zeta)=v in the respective intervals.

Note that when v>fcv>f_{c} and M~\tilde{M} is large, the Lorentz-breaking contribution to the sum in Eq. (S14) comes from values of mm that are comparable to M~\tilde{M}. In conclusion, Lorentz violation of quantum fields would affect the transition rate of the Unruh-DeWitt detector: The detector in circular motion in the preferred inertial frame sees a large low-energy Lorentz violation when its orbital speed exceeds the critical value fcf_{c}.