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

Radiative processes of single and entangled detectors on
circular trajectories in (2+1)(2+1) dimensional Minkowski spacetime

Subhajit Barman subhajit.barman@physics.iitm.ac.in Centre for Strings, Gravitation and Cosmology, Department of Physics, Indian Institute of Technology Madras, Chennai 600036, India Department of Physics, Indian Institute of Technology Guwahati, Guwahati 781039, Assam, India    Bibhas Ranjan Majhi bibhas.majhi@iitg.ac.in Department of Physics, Indian Institute of Technology Guwahati, Guwahati 781039, Assam, India    L. Sriramkumar sriram@physics.iitm.ac.in Centre for Strings, Gravitation and Cosmology, Department of Physics, Indian Institute of Technology Madras, Chennai 600036, India
Abstract

We investigate the radiative processes involving two entangled Unruh-DeWitt detectors that are moving on circular trajectories in (2+1)(2+1)-dimensional Minkowski spacetime. We assume that the detectors are coupled to a massless, quantum scalar field, and calculate the transition probability rates of the detectors in the Minkowski vacuum as well as in a thermal bath. We also evaluate the transition probability rates of the detectors when they are switched on for a finite time interval with the aid of a Gaussian switching function. We begin by examining the response of a single detector before we go on to consider the case of two entangled detectors. As we shall see, working in (2+1)(2+1) spacetime dimensions makes the computations of the transition probability rates of the detectors relatively simpler. We find that the cross transition probability rates of the two entangled detectors can be comparable to the auto transition probability rates of the individual detectors. We discuss specific characteristics of the response of the entangled detectors for different values of the parameters involved and highlight the effects of the thermal bath as well as switching on the detector for a finite time interval.

I Introduction

In quantum field theory in Minkowski spacetime, the concept of a particle is covariant under Lorentz transformations. However, about half-a-century ago, it was discovered that the notion of a particle is not a generally covariant concept (for the original discussion, see Ref. Fulling (1973); for detailed discussions, see the textbooks Birrell and Davies (1984a); Fulling (1989); Mukhanov and Winitzki (2007); Parker and Toms (2009)). In general, an observer in motion along a non-inertial trajectory in flat spacetime may see the Minkowski vacuum to be populated with particles Letaw (1981); Letaw and Pfautsch (1981); Sriramkumar and Padmanabhan (2002). For instance, a uniformly accelerated observer sees the Minkowski vacuum as a thermal bath, a phenomenon that has come to be known as the Unruh effect (for the original discussion, see Refs. Unruh (1976); DeWitt (1980); for a detailed review on the phenomenon, see, for instance, Ref. Crispino et al. (2008)).

Over the last few decades, there has been a constant effort to understand the notion of a particle in a curved spacetime. The idea of detectors was originally introduced to provide an operational definition to the concept of a particle Unruh (1976); DeWitt (1980); Letaw and Pfautsch (1981). By a detector one has in mind, say, a two level system which interacts with the quantum field of interest and is excited or de-excited when it is in motion. The response of detectors that are in motion on a variety of trajectories and are coupled to the quantum field in different manner have been examined in flat and curved spacetimes (for an inexhaustive list, see Refs. Letaw and Pfautsch (1981); Bell and Leinaas (1983); Svaiter and Svaiter (1992); Higuchi et al. (1993); Sriramkumar and Padmanabhan (1996); Davies et al. (1996); Sriramkumar and Padmanabhan (2002); Korsbakken and Leinaas (2004); Gutti et al. (2011); Jaffino Stargen et al. (2017); Louko and Upton (2018); Picanço et al. (2020); Biermann et al. (2020).)

At this stage, we should clarify that, in general, the response of the detectors may not match the results obtained from more formal methods such as the Bogoliubov transformations and the effective Lagrangian, which also reflects the particle content of the field (for a discussion in this context, see Ref. Sriramkumar (2017)). Moreover, apart from depending on the trajectory, the response of the detectors depends on the nature of their interaction with the quantum field. Nevertheless, the response of the detectors has been studied extensively in a variety of situations. In particular, it has been recognized that the idea of detectors can prove to be indispensable to experimentally observe the phenomenon of the Unruh effect or its equivalents (in this context, see, for example, Refs. Schutzhold et al. (2006); Sudhir et al. (2021)). Therefore, it seems important to construct specific models of detectors which closely capture possible experimental realizations and investigate the response of these detectors under different conditions.

In the literature, we find that a significant amount of attention has been paid to detectors that are in uniformly accelerated motion. Evidently, this interest has been due to the fact that uniformly accelerated detectors exhibit a thermal response, which has a close analogy with Hawking radiation from black holes Birrell and Davies (1984a); Mukhanov and Winitzki (2007). But, from a practical and experimental perspective, it seems more convenient to consider detectors that are moving on circular trajectories (for early discussions, see Refs. Bell and Leinaas (1983); Davies et al. (1996); Korsbakken and Leinaas (2004); for more recent discussions in this context, see Refs. Gutti et al. (2011); Jaffino Stargen et al. (2017); Louko and Upton (2018); Biermann et al. (2020)). Moreover, often the response of the detectors has been evaluated assuming that they remain switched on for infinite time. Needless to say, if such non-trivial phenomena are to be experimentally observed, it becomes important to examine the response of detectors that are switched on for a finite time interval.

With the above motivations in mind, in this work, we examine the response of the so-called Unruh-DeWitt detectors that are coupled to a massless, quantum scalar field through a monopole interaction and are in motion on circular trajectories in Minkowski spacetime. The Unruh-DeWitt monopole detectors are the simplest of the different possible detectors in the sense that they are coupled linearly to the quantum field Unruh (1976); DeWitt (1980). We evaluate the infinite time as well as the finite time response of these detectors. We shall work with Gaussian window functions to switch the detectors on for a finite time interval (for early discussions in this context, see Ref. Svaiter and Svaiter (1992); Higuchi et al. (1993); Sriramkumar and Padmanabhan (1996); for recent discussions, see Ref. Picanço et al. (2020)). For mathematical convenience, we shall work in (2+1)(2+1)-spacetime dimensions, and calculate the transition probability rate of the detectors in the Minkowski vacuum and in a thermal bath. We should mention that our focus on the (2+1)(2+1)-dimensional case is also motivated by its extensive consideration in models of analogue gravity (in this regard, see Ref. Sánchez-Kuntz et al. (2022) and the references therein). After discussing the case of a single detector, we shall go on to calculate the transition probability rate of two detectors that are assumed to be in an entangled initial state, a situation that has drawn considerable attention in the literature over the last few years (in this regard, see, for example, Refs. Hu and Yu (2015); Menezes and Svaiter (2015); Arias et al. (2016); Flores-Hidalgo et al. (2015); Menezes and Svaiter (2016); Menezes (2016); Rizzuto et al. (2016); Rodríguez-Camargo et al. (2018); Zhou et al. (2016); Cai and Ren (2018); Liu et al. (2018); Zhou and Yu (2020); Barman and Majhi (2021); for a discussion on entangled detectors in circular motion, see, for instance, Refs. Picanço et al. (2020); Zhang and Yu (2020)). We shall focus on the excitation of the detector (i.e. it absorbs than emits quanta) due to its interaction with the quantum field and its motion. As we shall illustrate, when the detectors are in circular motion and are switched on for an infinite time interval, generically, the transition probability rate of the detectors in the Minkowski vacuum and in the thermal bath is higher when the energy gap between the two levels of the detectors is smaller and the velocity of the detector is larger. We also find that, in a thermal bath, when the detectors remain switched on for an infinite time interval, the higher the temperature of the bath, the higher is response of the detectors. Interestingly, we find that the transition probability rate of the detectors in the Minkowski vacuum are higher when they are switched on for a shorter time interval, and we should point out that similar phenomenon has also been noticed previously in the literature (see, for instance, Refs. Sriramkumar and Padmanabhan (1996); Louko and Satz (2006); Satz (2007); Brown et al. (2013)). The corresponding transition probability rate in a thermal bath exhibits a more complex behavior, with the transition probability rate being higher when the temperature is higher provided the energy gap is large, while the behavior can be reversed for lower energy gaps, depending on the temperature. We also discuss different aspects of the total transition probability rate of the detectors for specific transitions from the symmetric and anti-symmetric entangled states to the collective excited state due to the presence of the thermal bath and the Gaussian switching function.

This paper is organized as follows. In Sec. II, we shall introduce and describe the response of two entangled Unruh-DeWitt detectors that are in motion along specific trajectories and are interacting with a quantum scalar field. In Sec. III, we shall discuss the response of a single Unruh-DeWitt detector that is moving on a circular trajectory in (2+1)(2+1)-dimensional Minkowski spacetime. We shall evaluate the transition probability rates of the detector in the Minkowski vacuum as well as in a thermal bath. We shall also consider the finite time transition probability rates of these detectors when they are switched on and off with the help of a Gaussian switching function. As we shall see, these calculations for a single detector prove to be helpful later when we evaluate the responses of the entangled detectors. In Sec. IV, we shall evaluate the auto and cross transition probability rates of two entangled detectors that are in motion along circular trajectories, when the field is assumed to be in the Minkowski vacuum and in a thermal bath. We shall also discuss the response of these entangled detectors when they are switched on for a finite time interval. We shall conclude in Sec. V with a summary of the results we have obtained and a discussion on the broader implications of our analysis. We shall relegate some of the additional discussions to the appendices.

A brief word on our notation is in order at this stage of our discussion. We shall work with units such that =c=1\hbar=c=1. For convenience, we shall describe the set of spacetime coordinates (t,𝒙)(t,{\bm{x}}) collectively as x~\tilde{x}.

II Radiative processes of two entangled detectors: The model

In this section, we shall briefly outline the radiative processes that arise in situations involving two entangled Unruh-DeWitt detectors. The discussion allows us to introduce the notation and also describe the quantities that we shall evaluate later. We should mention that the model we shall consider has been examined earlier in different situations (in this context, see, for instance, Refs. Arias et al. (2016); Rodríguez-Camargo et al. (2018); Picanço et al. (2020)).

The detectors we shall consider are assumed to be composed of point like atoms with two internal energy levels, which are interacting with a scalar field through a monopole interaction. For simplicity, we shall assume the field to be a massless, minimally coupled real scalar field, say, Φ\Phi. The Hamiltonian of the complete system composed of the two detectors and the scalar field is assumed to be of the form

H=HD+HF+HI,H=H_{\mathrm{D}}+H_{\mathrm{F}}+H_{\mathrm{I}}, (1)

where HDH_{\mathrm{D}} denotes the Hamiltonian of the detectors free of any interaction, HFH_{\mathrm{F}} is the Hamiltonian describing the free scalar field, and the term HIH_{\mathrm{I}} describes the interaction between the detectors and the scalar field. As initially suggested by Dicke (for the original discussion, see Ref. Dicke (1954); for a recent discussion, see Ref. Rodríguez-Camargo et al. (2018)), one may express the Hamiltonian describing the two static atoms constituting the detectors as follows:

HD=ω0[S^1z𝟙^2+𝟙^1S^2z],H_{\mathrm{D}}=\omega_{0}\,\left[\hat{S}_{1}^{z}\otimes\hat{\mathbb{1}}_{2}\,+\,\hat{\mathbb{1}}_{1}\otimes\hat{S}_{2}^{z}\,\right], (2)

where S^jz\hat{S}_{j}^{z}, with j={1,2}j=\{1,2\}, denotes the operator that determines the energy levels of the detectors. The operator S^jz\hat{S}_{j}^{z} is defined as

S^jz=12(|ejej||gjgj|),\hat{S}_{j}^{z}=\frac{1}{2}\,\left(|e_{j}\rangle\,\langle e_{j}|-|g_{j}\rangle\,\langle g_{j}|\right), (3)

where |gj|g_{j}\rangle and |ej|e_{j}\rangle represent the ground and excited states of the jj-th\mathrm{th} atom. Moreover, note that, in the Hamiltonian (2) describing the detectors, the quantity 𝟙^\hat{\mathbb{1}} represents the identity operator, and ω0\omega_{0} represents the transition energy corresponding to the collective two detector system. In particular, for identical, static detectors, the energy eigen states and eigen values for the two-atom system are given by Arias et al. (2016)

Ee\displaystyle E_{e} =\displaystyle= ω0,|e=|e1|e2,\displaystyle\omega_{0},~{}~{}~{}~{}|e\rangle=|e_{1}\rangle\,|e_{2}\rangle,\quad (4a)
Es\displaystyle E_{s} =\displaystyle= 0,|s=12(|e1|g2+|g1|e2),\displaystyle 0,~{}~{}~{}~{}~{}~{}|s\rangle=\frac{1}{\sqrt{2}}\,\left(|e_{1}\rangle\,|g_{2}\rangle+|g_{1}\rangle\,|e_{2}\rangle\right), (4b)
Ea\displaystyle E_{a} =\displaystyle= 0,|a=12(|e1|g2|g1|e2),\displaystyle 0,~{}~{}~{}~{}~{}~{}|a\rangle=\frac{1}{\sqrt{2}}\,\left(|e_{1}\rangle\,|g_{2}\rangle-|g_{1}\rangle\,|e_{2}\rangle\right),\quad (4c)
Eg\displaystyle E_{g} =\displaystyle= ω0,|g=|g1|g2,\displaystyle-\omega_{0},~{}~{}|g\rangle=|g_{1}\rangle|g_{2}\rangle, (4d)

where |g|g\rangle and |e|e\rangle correspond to the ground and the excited states of the collective system, while |s|s\rangle and |a|a\rangle denote the symmetric and anti-symmetric maximally entangled Bell states. A pictorial representation of the different states and the associated energy levels of the two entangled detectors is illustrated in Fig. 1.

Refer to caption
Figure 1: An illustration of the energy levels corresponding to the eigen states of the two entangled detectors with two levels each. (We should mention that this figure has been taken from Ref. Picanço et al. (2020).) The contributions from the monopole moment for each transition have also been indicated in the figure.

In (2+1)(2+1)-dimensional Minkowski spacetime, the Hamiltonian of the massless, free scalar field is given by

HF=12d2𝒙[Φ˙2(x~)+|Φ(x~)|2],H_{\mathrm{F}}=\frac{1}{2}\int{\rm d}^{2}{\bm{x}}\,\left[\dot{\Phi}^{2}(\tilde{x})+\left|\bm{\nabla}\Phi(\tilde{x})\right|^{2}\right], (5)

where the overdot denotes differentiation with respect to the time coordinate, and \bm{\nabla} denotes the spatial gradient. The interaction Hamiltonian describing the monopole detectors and the scalar field is assumed to be

HI=j=12μjmj(τj)κj(τj)Φ[x~j(τj)],H_{\mathrm{I}}=\sum_{j=1}^{2}\mu_{j}\,m_{j}(\tau_{j})\,\kappa_{j}(\tau_{j})\,\Phi[\tilde{x}_{j}(\tau_{j})], (6)

where μj\mu_{j} denotes the strengths of the individual coupling between the detectors and the scalar field, while mj(τj)m_{j}(\tau_{j}) and κj(τj)\kappa_{j}(\tau_{j}) denote the monopole operators of the detectors and the switching functions, respectively. For identical atomic detectors, the coupling strengths between the detectors and the scalar field can be assumed to be the same, i.e. we can set μ1=μ2=μ\mu_{1}=\mu_{2}=\mu. In such a case, the time evolution operator can be expressed as

U^\displaystyle\hat{U} =\displaystyle= 𝒯exp{iμ[m^1(τ1)κ1(τ1)Φ^[x~1(τ1)]dτ1\displaystyle\mathcal{T}\;\exp\,\biggl{\{}-i\,\mu\int_{-\infty}^{\infty}\,\biggl{[}\hat{m}_{1}(\tau_{1})\,\kappa_{1}(\tau_{1})\,\hat{\Phi}[\tilde{x}_{1}(\tau_{1})]\,{\rm d}\tau_{1} (7)
+m^2(τ2)κ2(τ2)Φ^[x~2(τ2)]dτ2]},\displaystyle+\,\hat{m}_{2}(\tau_{2})\,\kappa_{2}(\tau_{2})\,\hat{\Phi}[\tilde{x}_{2}(\tau_{2})]\,{\rm d}\tau_{2}\biggr{]}\biggr{\}},

where 𝒯\mathcal{T} implies time ordering. Let |ω|\omega\rangle be the collective initial state of the two detector system, and |ω¯|\bar{\omega}\rangle be the collective final state. Also, let the initial state of the scalar field be the Minkowski vacuum |0M|0_{{}_{\mathrm{M}}}\rangle, and let |Θ|\Theta\rangle be the final state of the scalar field. Under these conditions, the transition amplitude from the initial state |ω,0M|\omega,0_{{}_{\mathrm{M}}}\rangle to the final state |ω¯,Θ|\bar{\omega},\Theta\rangle at the first order (when expanded in the strength of the coupling constant μ\mu) in perturbation theory is given by

𝒜|ω,0M|ω¯,Θ\displaystyle\mathcal{A}_{|\omega,0_{{}_{\mathrm{M}}}\rangle\to|\bar{\omega},\Theta\rangle}\!\! =\displaystyle= Θ,ω¯|U^|ω,0M\displaystyle\!\!\langle\Theta,\bar{\omega}|\hat{U}|\omega,0_{{}_{\mathrm{M}}}\rangle (8)
\displaystyle\simeq iμΘ,ω¯|[κ1m^1Φ^(x~1)dτ1\displaystyle-i\,\mu\,\langle\Theta,\bar{\omega}|\int_{-\infty}^{\infty}\,\biggl{[}\kappa_{1}\,\hat{m}_{1}\,\hat{\Phi}(\tilde{x}_{1})\,{\rm d}\tau_{1}
+κ2m^2Φ^(x~2)dτ2]|ω,0M.\displaystyle+\,\kappa_{2}\,\hat{m}_{2}\,\hat{\Phi}(\tilde{x}_{2})\,{\rm d}\tau_{2}\biggr{]}|\omega,0_{{}_{\mathrm{M}}}\rangle.

Note that we shall be interested in examining the final state of the detectors. The total transition probability of the detectors can be arrived at from the above transition amplitude by summing over all the final states {|Θ}\left\{|\Theta\rangle\right\} of the field. The total transition probability of the two detectors can be expressed as

Γ|ω|ω¯()\displaystyle\Gamma_{|\omega\rangle\to|\bar{\omega}\rangle}(\mathcal{E}) =\displaystyle= {|Θ}𝒜|ω,0M|ω¯,Θ𝒜|ω,0M|ω¯,Θ\displaystyle\sum_{\left\{|\Theta\rangle\right\}}\mathcal{A}_{|\omega,0_{{}_{\mathrm{M}}}\rangle\to|\bar{\omega},\Theta\rangle}\,\mathcal{A}_{|\omega,0_{{}_{\mathrm{M}}}\rangle\to|\bar{\omega},\Theta\rangle}^{\ast} (9)
\displaystyle\simeq μ2j,l=12mjω¯ωmlω¯ωFjl(),\displaystyle\mu^{2}\sum_{j,l=1}^{2}m_{j}^{\bar{\omega}\omega\ast}\,m_{l}^{\bar{\omega}\omega}\,F_{jl}(\mathcal{E}),

where =Eω¯Eω\mathcal{E}=E_{\bar{\omega}}-E_{\omega}, with EωE_{\omega} and Eω¯E_{\bar{\omega}} denoting the energy eigen values associated with the states |ω|\omega\rangle and |ω¯|\bar{\omega}\rangle, and mjω¯ω=ω¯|m^j(0)|ωm_{j}^{\bar{\omega}\omega}=\langle\bar{\omega}|\hat{m}_{j}(0)|\omega\rangle. As we shall discuss below, the quantities Fjl()F_{jl}(\mathcal{E})—which we shall refer to as the auto or the cross transition probabilities—depend on the trajectory of the detectors.

Meanwhile, let us understand the values that the quantity mjω¯ωm_{j}^{\bar{\omega}\omega} can take. We shall assume that the operator describing the monopole moment of the detectors is given by

m^j(0)=|ejgj|+|gjej|.\hat{m}_{j}(0)=|e_{j}\rangle\langle g_{j}|+|g_{j}\rangle\langle e_{j}|. (10)

This expression for the monopole operators can be utilized to determine the contributions due to specific transitions between the collective initial and final states of the two detectors. For instance, it can be shown that the transition from the collective ground state |g|g\rangle to the collective excited state |e|e\rangle (or the other way around) of the entangled detectors is not possible since mjge=mjeg=0m_{j}^{ge}=m_{j}^{eg}=0. Also, one finds that m1se=m2se=1/2m_{1}^{se}=m_{2}^{se}=1/\sqrt{2} and m1ae=m2ae=1/2m_{1}^{ae}=-m_{2}^{ae}=-1/\sqrt{2}, which denote the amplitudes for transitions between the symmetric and anti-symmetric Bell states (viz. |s|s\rangle and |a|a\rangle) and the excited state of the detectors, respectively. Moreover, one can show that the amplitudes for the transitions from the collective ground state to the symmetric and anti-symmetric Bell states are given by m1gs=m2gs=1/2m_{1}^{gs}=m_{2}^{gs}=1/\sqrt{2} and m1ga=m2ga=1/2m_{1}^{ga}=-m_{2}^{ga}=1/\sqrt{2}. The energy levels associated with the different states of the two entangled detectors and the various possible transitions are illustrated diagrammatically in Fig. 1 (taken from Ref. Picanço et al. (2020)).

Let us now shift our attention to the transition probabilities Fjl()F_{jl}(\mathcal{E}) in Eq. (9). The explicit form of the transition probabilities Fjl()F_{jl}(\mathcal{E}) are found to be

Fjl()\displaystyle F_{jl}(\mathcal{E}) =\displaystyle= dτldτjei(τjτl)\displaystyle\int_{-\infty}^{\infty}{\rm d}\tau_{l}^{\prime}\int_{-\infty}^{\infty}{\rm d}\tau_{j}\,\mathrm{e}^{-i\,\mathcal{E}\,(\tau_{j}-\tau_{l}^{\prime})} (11)
×Gjl+[x~j(τj),x~l(τl)]κj(τj)κl(τl),\displaystyle\times\,G_{jl}^{+}[\tilde{x}_{j}(\tau_{j}),\tilde{x}_{l}(\tau_{l}^{\prime})]\,\kappa_{j}(\tau_{j})\,\kappa_{l}(\tau_{l}^{\prime}),

where the quantity Gjl+[x~j(τj),x~l(τl)]G_{jl}^{+}[\tilde{x}_{j}(\tau_{j}),\tilde{x}_{l}(\tau_{l}^{\prime})] denotes the positive frequency Wightman function evaluated along the trajectories of the detectors. The positive frequency Wightman function is defined as

Gjl+[x~j(τj),x~l(τl)]=0M|Φ^[x~j(τj)]Φ^[x~l(τl)]|0M.G_{jl}^{+}[\tilde{x}_{j}(\tau_{j}),\tilde{x}_{l}(\tau_{l}^{\prime})]=\langle 0_{{}_{\mathrm{M}}}|\hat{\Phi}[\tilde{x}_{j}(\tau_{j})]\,\hat{\Phi}[\tilde{x}_{l}(\tau_{l}^{\prime})]|0_{{}_{\mathrm{M}}}\rangle. (12)

In the following sections, we shall evaluate the transition probabilities of detectors that are in motion on circular trajectories in (2+1)(2+1)-dimensional Minkowski spacetime. We shall evaluate the responses of the detectors in the Minkowski vacuum as well as in a thermal bath. As we shall see, it proves to be convenient to work in terms of the polar coordinates to arrive at the Wightman function along the trajectories of the detectors when they are in circular motion.

III Response of a detector in circular motion

In this section, we shall derive the response of a single Unruh-DeWitt detector that is interacting with a scalar field. The results we obtain in this situation will prove to be helpful for understanding the results in the case of the two entangled detectors. Consider a massless and minimally coupled scalar field Φ\Phi that is described by the action

S[Φ]=d3xg12gμνμΦνΦ.S[\Phi]=-\int{\rm d}^{3}x\,\sqrt{-g}\,\frac{1}{2}\,g^{\mu\nu}\,{\partial}_{\mu}\Phi\,{\partial}_{\nu}\Phi. (13)

On varying the action, we can obtain the equation of motion of the scalar field to be

Φ=1gμ(ggμνν)Φ=0.\Box\Phi=\frac{1}{\sqrt{-g}}\,{\partial}_{\mu}\left(\sqrt{-g}\,g^{\mu\nu}\,{\partial}_{\nu}\right)\Phi=0. (14)

In order to examine the behavior of a rotating detector in Minkowski spacetime, it proves to be convenient to work in the polar coordinates so that, in (2+1)(2+1)-dimensions, the spacetime coordinates are given by x~=(t,ρ,ϕ)\tilde{x}=(t,\rho,\phi). In these coordinates, the normal modes of the massless scalar field can be obtained to be

uqm(x~)=14πeiqtJm(qρ)eimϕ,u_{qm}(\tilde{x})=\frac{1}{\sqrt{4\,\pi}}\,\mathrm{e}^{-i\,q\,t}\,J_{m}(q\,\rho)\,\mathrm{e}^{i\,m\,\phi}, (15)

where 0q<0\leq q<\infty, mm is an integer, and Jn(z)J_{n}(z) denotes the Bessel function of order nn. On quantization, the scalar field can be decomposed in terms of the above normal modes uqm(x~)u_{qm}(\tilde{x}) as follows:

Φ^(x~)=0dqm=[a^qmuqm(x~)+a^qmuqm(x~)],\hat{\Phi}(\tilde{x})=\int_{0}^{\infty}{\rm d}q\sum_{m=-\infty}^{\infty}\left[\hat{a}_{qm}\,u_{qm}(\tilde{x})+\hat{a}_{qm}^{\dagger}\,u_{qm}^{\ast}(\tilde{x})\right], (16)

where a^qm\hat{a}_{qm} and a^qm\hat{a}_{qm}^{\dagger} are the creation and the annihilation operators which satisfy the following standard commutation relations:

[a^qm,a^qm]\displaystyle\left[\hat{a}_{qm},\hat{a}_{q^{\prime}m^{\prime}}\right] =\displaystyle= [a^qm,a^qm]=0,\displaystyle\left[\hat{a}_{qm}^{\dagger},\hat{a}_{q^{\prime}m^{\prime}}^{{\dagger}}\right]=0, (17a)
[a^qm,a^qm]\displaystyle\left[\hat{a}_{qm},\hat{a}_{q^{\prime}m^{\prime}}^{{\dagger}}\right] =\displaystyle= δ(1)(qq)δmm.\displaystyle\delta^{(1)}(q-q^{\prime})\,\delta_{mm^{\prime}}. (17b)

In the case of a single detector, the transition probability (11) simplifies to be

F()\displaystyle F(\mathcal{E}) =\displaystyle= dτdτei(ττ)\displaystyle\int_{-\infty}^{\infty}{\rm d}\tau^{\prime}\int_{-\infty}^{\infty}{\rm d}\tau\,\mathrm{e}^{-i\,\mathcal{E}\,(\tau-\tau^{\prime})} (18)
×G+[x~(τ),x~(τ)]κ(τ)κ(τ).\displaystyle\times\,G^{+}[\tilde{x}(\tau),\tilde{x}(\tau^{\prime})]\,\kappa(\tau)\,\kappa(\tau^{\prime}).

In our discussion, we shall be interested in examining the response of detectors that are moving on circular trajectories. Also, we shall evaluate the response of the detectors in the Minkowski vacuum and in a thermal bath. We shall utilize the mode functions (15) in the polar coordinates to arrive at the Wightman function G+(x~,x~)G^{+}(\tilde{x},\tilde{x}^{\prime}) in both these situations. One finds that, in these situations, the Wightman function along the trajectory of a detector in circular motion is invariant under the time translation in the proper time in the frame of the detector, i.e. G+[x~(τ),x~(τ)]=G+(τ,τ)=G+(u)G^{+}[\tilde{x}(\tau),\tilde{x}(\tau^{\prime})]=G^{+}(\tau,\tau^{\prime})=G^{+}(u), where u=(ττ)u=(\tau-\tau^{\prime}). It is well known that such a time translation invariance allows one to define the transition probability rate of the detector. Often, in these contexts, the Wightman function is first evaluated by summing over all the normal modes of the quantum field, before evaluating the transition probability rate of the detector. As we shall see, to arrive at the transition probability rate of the rotating detector, rather than explicitly evaluate the Wightman function, it proves to be convenient to first carry out the integral over the quantity uu, and then sum over the modes. In the following subsections, we shall evaluate the transition probability rate of a rotating detector that has been switched on for infinite as well as a finite time interval.

III.1 Detector switched on for infinite duration

Let us first consider the situation wherein the detector remains switched on for all times. In such a situation, the switching function κ(τ)\kappa(\tau) reduces to unity.

III.1.1 Response in the Minkowski vacuum

Let us now evaluate the response of the rotating detector in the Minkowski vacuum. The Wightman function G+(x~,x~)G^{+}(\tilde{x},\tilde{x}^{\prime}) associated with the massless scalar field in the Minkowski vacuum can be expressed as (in this regard, also see App. A)

G+(x,x)\displaystyle G^{+}(x,x^{\prime}) =\displaystyle= 0M|Φ^(x~)Φ^(x~)|0M\displaystyle\langle 0_{{}_{\mathrm{M}}}|\hat{\Phi}(\tilde{x})\,\hat{\Phi}(\tilde{x}^{\prime})|0_{{}_{\mathrm{M}}}\rangle (19)
=\displaystyle= 0dq4πm=Jm(qρ)Jm(qρ)\displaystyle\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\sum_{m=-\infty}^{\infty}J_{m}(q\,\rho)\,J_{m}(q\,\rho^{\prime})
×eiq(tt)eim(ϕϕ),\displaystyle\times\,\mathrm{e}^{-i\,q\,(t-t^{\prime})}\,\mathrm{e}^{i\,m\,(\phi-\phi^{\prime})},

where we have made use of the fact that the Bessel functions Jn(x)J_{n}(x) are real for integer values of nn and real arguments. In order to arrive at the response of the detector, we need to calculate the above Wightman function along the trajectory of the detector. Let us assume that the detector is moving along a circular trajectory with radius σ\sigma, at a constant angular velocity Ω\Omega. If τ\tau is the proper time in the frame of the detector, then the trajectory of the detector is given by (see, for example, Refs. Gutti et al. (2011); Picanço et al. (2020))

t=γτ,ρ=σ,ϕ=γΩτ,t=\gamma\,\tau,\quad\rho=\sigma,\quad\phi=\gamma\,\Omega\,\tau, (20)

where γ=1/1v2\gamma=1/\sqrt{1-v^{2}} is the Lorentz factor associated with the linear velocity v=σΩv=\sigma\,\Omega of the detector. Along such a trajectory, the Wightman function above reduces the following form:

G+(u)=0dq4πm=Jm2(qσ)eiγ(qmΩ)u,G^{+}(u)=\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\sum_{m=-\infty}^{\infty}J_{m}^{2}(q\,\sigma)\,\mathrm{e}^{-i\,\gamma\,(q-m\,\Omega)\,u}, (21)

where, as we mentioned above, u=(ττ)u=(\tau-\tau^{\prime}). In such a situation, we can define the transition probability rate of the detector to be

R()=dueiuG+(u).R(\mathcal{E})=\int_{-\infty}^{\infty}{\rm d}u\,\mathrm{e}^{-i\,\mathcal{E}\,u}\,G^{+}(u). (22)

On substituting the Wightman function (21) along the trajectory of the rotating detector in the above expression, we find that the transition probability rate can be expressed as

R(¯)\displaystyle R(\bar{\mathcal{E}}) =\displaystyle= m=0dq2γJm2(qσ)δ(1)[q(m¯)Ω]\displaystyle\sum_{m=-\infty}^{\infty}\,\int_{0}^{\infty}\frac{{\rm d}q}{2\,\gamma}\,J^{2}_{m}(q\,\sigma)\,\delta^{(1)}[q-(m-\bar{\mathcal{E}})\,\Omega] (23)
=\displaystyle= 12γm¯Jm2[(m¯)v],\displaystyle\frac{1}{2\,\gamma}\,\sum_{m\geq\bar{\mathcal{E}}}^{\infty}J^{2}_{m}\left[(m-\bar{\mathcal{E}})\,v\right],

where we have introduced the dimensionless energy gap ¯=/(γΩ)\bar{\mathcal{E}}=\mathcal{E}/(\gamma\,\Omega) and, recall that, v=σΩv=\sigma\,\Omega is the linear velocity of the detector. Note that, since q[0,)q\in[0,\infty), the Dirac delta function in the above expression leads to non-zero contributions only when m¯m\geq\bar{\mathcal{E}}, which is reflected in the lower limit of the sum in the final expression.

Evidently, the transition probability rate (23) of the detector in circular motion depends only on its linear velocity vv. The sum in the expression for the transition probability rate proves to be difficult to evaluate analytically. But, it converges quickly enough to be computed numerically. For a given value of ¯\bar{\mathcal{E}}, we find that the sum converges exponentially beyond a certain value of mm. For instance, we find that, when ¯\bar{\mathcal{E}} is chosen to be unity, for v=(0.25,0.5,0.75)v=(0.25,0.5,0.75), the quantity Jm2[(m¯)v]J^{2}_{m}\left[(m-\bar{\mathcal{E}})\,v\right] that appears in the sum characterizing the transition probability rate of the detector decreases at least as fast as (e2.0m,e0.9m,e0.25m)(\mathrm{e}^{-2.0\,m},\mathrm{e}^{-0.9\,m},\mathrm{e}^{-0.25\,m}), respectively, at suitably large mm. We should clarify that we have confirmed this behavior for adequately large values of mm, much beyond the values we sum over to arrive at the result. The rapid convergence of the sum allows us to easily calculate the transition probability rate of the rotating detector numerically by summing up to a finite value of mm (in this regard, also see, App. B). We have also ensured that, over the domain of ¯\bar{\mathcal{E}} we focus on, the contributions beyond the maximum value of mm we have worked with are insignificant. In Fig. 2, we have illustrated the transition probability rate of the detector as a function of the dimensionless energy gap ¯\bar{\mathcal{E}} for a few different values of the velocity.

Refer to caption
Refer to caption
Figure 2: The transition probability rate R(¯)R(\bar{\mathcal{E}}) of the Unruh-DeWitt detector that is moving on a circular trajectory and remains switched on forever has been plotted as a function of the dimensionless energy gap ¯\bar{\mathcal{E}}. We have plotted the response of the detector in the Minkowski vacuum (on the left) and in a thermal bath (on the right). In the case of the Minkowski vacuum, we have plotted the results for three different values of the velocity of the detector, viz. v=(0.25,0.5,0.75)v=(0.25,0.5,0.75) (in red, blue and green) and, in the case of the thermal bath, we have fixed the velocity to be v=0.5v=0.5 and have plotted the results for three different values of the dimensionless inverse temperature, viz. β¯=(0.1,1,10)\bar{\beta}=(0.1,1,10) (in red, blue and green). We have arrived at these results by summing until (m¯)=50(m-\bar{\mathcal{E}})=50 [in Eqs. (23) and (26)] and we have confirmed that summing up to higher values of mm does not significantly alter the results. Note that the transition probability rate of the rotating detector is larger at smaller energies for a given velocity and temperature. Also, we find that, for a given energy and temperature, the rate is higher at a higher velocity of the detector. Moreover, for a given energy and velocity, the transition probability rate is higher when the temperature of the bath is higher. Clearly, this can be attributed to the fact that there are more quanta available to excite the detector at higher temperatures.

The figure suggests that, higher the velocity of the detector, higher is its transition probability rate. Moreover, for a given velocity, the transition probability of the detector is larger at smaller values of the energy gap such that ¯1\bar{\mathcal{E}}\ll 1.

III.1.2 Response in a thermal bath

Let us now turn to evaluate the response of the detector in a thermal bath. We shall assume that the massless scalar field Φ\Phi of our interest is in equilibrium with a thermal bath maintained at the inverse temperature β\beta. We can utilize the decomposition (16) of the scalar field in terms of the normal modes (15) to arrive at the following expression for the Wightman function at a finite temperature (in this context, also see App. A):

Gβ+(x~,x~)\displaystyle G^{+}_{\beta}(\tilde{x},\tilde{x}^{\prime}) =\displaystyle= 0dq4πm=Jm(qρ)Jm(qρ)\displaystyle\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\sum_{m=-\infty}^{\infty}J_{m}(q\,\rho)\,J_{m}(q\,\rho^{\prime}) (24)
×{ei[q(tt)m(ϕϕ)]1eβq\displaystyle\times\,\biggl{\{}\frac{\mathrm{e}^{-i\,[q\,(t-t^{\prime})-m\,(\phi-\phi^{\prime})]}}{1-\mathrm{e}^{-\beta\,q}}
+ei[q(tt)m(ϕϕ)]eβq1}.\displaystyle+\,\frac{\mathrm{e}^{i\,[q\,(t-t^{\prime})-m\,(\phi-\phi^{\prime})]}}{\mathrm{e}^{\beta\,q}-1}\biggr{\}}.

Along the trajectory (20) of the rotating detector, this finite temperature Wightman function too turns out to be invariant under translations in the proper time of the detector as in the Minkowski vacuum. We find that, in the frame of the rotating detector, the finite temperature Wightman function reduces to

Gβ+(u)\displaystyle G^{+}_{\beta}(u) =\displaystyle= 0dq4πm=Jm2(qσ)\displaystyle\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\sum_{m=-\infty}^{\infty}J^{2}_{m}(q\,\sigma) (25)
×[eiγ(qmΩ)u1eβq+eiγ(qmΩ)ueβq1].\displaystyle\times\,\left[\frac{\mathrm{e}^{-i\,\gamma\,(q-m\,\Omega)\,u}}{1-\mathrm{e}^{-\beta\,q}}+\frac{\mathrm{e}^{i\,\gamma\,(q-m\,\Omega)\,u}}{\mathrm{e}^{\beta\,q}-1}\right].\qquad

Since the Wightman function depends only on the quantity uu, we can define a transition probability rate for the detector as in the Minkowski vacuum [cf. Eq. (22)]. The transition probability rate of the rotating detector at a finite temperature can be easily evaluated to be

R(¯)\displaystyle R(\bar{\mathcal{E}}) =\displaystyle= m=0dq2γJm2(qσ){δ(1)[q(m¯)Ω]1eβq\displaystyle\sum_{m=-\infty}^{\infty}\,\int_{0}^{\infty}\frac{{\rm d}q}{2\,\gamma}\,J^{2}_{m}(q\,\sigma)\,\biggl{\{}\frac{\delta^{(1)}[q-(m-\bar{\mathcal{E}})\,\Omega]}{1-\mathrm{e}^{-\beta\,q}} (26)
+δ(1)[q(m+¯)Ω]eβq1}\displaystyle+\,\frac{\delta^{(1)}[q-(m+\bar{\mathcal{E}})\,\Omega]}{\mathrm{e}^{\beta\,q}-1}\biggr{\}}
=\displaystyle= 12γ{m¯Jm2[(m¯)v]1eβ¯(m¯)\displaystyle\frac{1}{2\,\gamma}\,\biggl{\{}\sum_{m\geq\bar{\mathcal{E}}}^{\infty}\,\frac{J^{2}_{m}\left[(m-\bar{\mathcal{E}})\,v\right]}{1-\mathrm{e}^{-\bar{\beta}\,(m-\bar{\mathcal{E}})}}
+m¯Jm2[(m+¯)v]eβ¯(m+¯)1},\displaystyle+\,\sum_{m\geq-\bar{\mathcal{E}}}^{\infty}\frac{J^{2}_{m}\left[(m+\bar{\mathcal{E}})\,v\right]}{\mathrm{e}^{\bar{\beta}\,(m+\bar{\mathcal{E}})}-1}\biggr{\}},

where β¯=βΩ\bar{\beta}=\beta\,\Omega denotes the dimensionless inverse temperature of the bath. We had noted earlier that, the quantity Jm2[(m¯)v]J^{2}_{m}\left[(m-\bar{\mathcal{E}})\,v\right] which appears in the sum characterizing the transition probability rate decreases exponentially at large mm. Though the sums in the above expression are again difficult to evaluate analytically, such a rapid convergence allows us to compute them numerically rather easily (again, in this regard, see, App. B). Actually, in the case of the second term, the exponential in the denominator also aids in a faster convergence of the sum. In Fig. 2, we have plotted the above transition probability rate of the detector for a fixed value of the velocity vv and a few different values of the dimensionless inverse temperature β¯\bar{\beta}. It should be clear from the figure that, larger the temperature (or, equivalently, smaller the value of β¯\bar{\beta}), larger is the transition probability rate of the rotating detector.

At this stage, there is a technical point that we need to discuss. Actually, the finite temperature Wightman function (25) contains an infrared divergence. As q0q\to 0, the functions Jm2(qσ)J_{m}^{2}(q\,\sigma) behave as q2mq^{2\,m} and, hence, the m=0m=0 term in the Wightman function diverges logarithmically in this limit even for finite separation of the spacetime points. This behavior is a surprising and less known peculiarity of the thermal Green’s function in (2+1)(2+1)-dimensional Minkowski spacetime and, in fact, the divergence is absent at zero temperature [as can be easily checked with the Wightman function (21)]. Also, it can be readily shown that such an infrared divergence is not encountered in (3+1)(3+1)-dimensional Minkowski spacetime. We should point out that the infrared divergence occurs in addition to the ultraviolet divergence which arises at large qq. The ultraviolet divergence can, as usual, be regulated using the (iϵ)(i\,\epsilon)-prescription. Note that, in arriving at the rate R(¯)R(\bar{\mathcal{E}}) in Eq. (26), we chose to calculate the integral over uu first before evaluating the integral over qq. In the process, the infrared divergence is transferred to the m=±¯m=\pm\bar{\mathcal{E}} term in the sum, and it manifests itself only in the ¯0\bar{\mathcal{E}}\to 0 limit. But, since we have assumed that >0\mathcal{E}>0, we do not actually encounter the divergence when evaluating the sum. In the following section, when we consider detectors which are switched on for a finite duration, we shall find that the divergence at the finite temperature cannot be circumvented in a similar manner. To handle the divergence, we shall adopt a procedure which allows us to reproduce the results in the different limits, viz. at zero temperature and when the detector is switched on for infinite duration.

There is another point that we need to clarify regarding the results illustrated in Fig. 2. Note that, in the case of results plotted at a finite temperature, the transition probability rate R(¯)R(\bar{\mathcal{E}}) turns out to be more than unity for small values of ¯\bar{\mathcal{E}}. This may cause concern. But, it occurs due to the fact that we have dropped an overall factor of |μ|2|\mu|^{2}, where μ\mu is the coupling constant that determines the strength of the interaction between the detector and the field [cf. Eq. (6)], when evaluating the transition probability rate. In order for the perturbative expansion of the time evolution operator in Eq. (7) to be valid, we require μ\mu to be much smaller than unity. Evidently, for a suitably small value of |μ|2|\mu|^{2}, the transition probability rate will reduce to a value less than unity for all energies ¯\bar{\mathcal{E}}.

We will now show that the transition probability rate of the rotating detector in a thermal bath we have obtained above corresponds to the accumulation of different types of radiative processes that occur in the system. Let us introduce the quantities

𝒩r(q,¯)=12γm=Jm2(qσ)δ(1)[q(m¯)Ω]\mathcal{N}_{r}(q,\bar{\mathcal{E}})=\frac{1}{2\,\gamma}\sum_{m=-\infty}^{\infty}J^{2}_{m}(q\,\sigma)\,\delta^{(1)}[q-(m-\bar{\mathcal{E}})\,\Omega] (27)

and

𝒩β(q)=1eβq1.\mathcal{N}_{\beta}(q)=\frac{1}{\mathrm{e}^{\beta\,q}-1}. (28)

Since (1ex)1=1+(ex1)1(1-\mathrm{e}^{-x})^{-1}=1+(\mathrm{e}^{x}-1)^{-1}, in terms of the above quantities, we can re-express the first equality of Eq. (26) in the following form:

R(¯)\displaystyle R(\bar{\mathcal{E}}) =\displaystyle= 0dq𝒩r(q,¯)R1(¯)\displaystyle\underbrace{\int_{0}^{\infty}{\rm d}q\,\mathcal{N}_{r}(q,\bar{\mathcal{E}})}_{R_{1}(\bar{\mathcal{E}})} (29)
+0dq𝒩β(q)[𝒩r(q,¯)+𝒩r(q,¯)]R2(¯,β).\displaystyle+\,\underbrace{\int_{0}^{\infty}{\rm d}q\,\mathcal{N}_{\beta}(q)\,\left[\mathcal{N}_{r}(q,\bar{\mathcal{E}})+\mathcal{N}_{r}(q,-\bar{\mathcal{E}})\right]}_{R_{2}(\bar{\mathcal{E}},\beta)}.\qquad

In other words, we can express the transition probability rate of the rotating detector in a thermal bath as a sum of the two contributions R1(¯)R_{1}(\bar{\mathcal{E}}) and R2(¯,β)R_{2}(\bar{\mathcal{E}},\beta), which, as we shall soon discuss, can be attributed to different types of radiative processes. Note that the first term R1(¯)R_{1}(\bar{\mathcal{E}}) is only a function of ¯\bar{\mathcal{E}} and is independent of β\beta, whereas the second term R2(¯,β)R_{2}(\bar{\mathcal{E}},\beta) is a function of both ¯\bar{\mathcal{E}} and β\beta. The fact that the term R1(¯)R_{1}(\bar{\mathcal{E}}) is the same as the first equality in Eq. (23) clearly suggests that it corresponds to the response of the rotating detector in the Minkowski vacuum. The contribution arises due to modes with the magnetic quantum numbers m1=(q/Ω)+¯m_{1}=(q/\Omega)+\bar{\mathcal{E}} for a given value of momentum qq. A contribution from the Minkowski vacuum can always be expected to occur and the contribution can be interpreted as arising due to the spontaneous excitation of the detector. The interesting aspect of the term R2(¯,β)R_{2}(\bar{\mathcal{E}},\beta) is the appearance of the factor 𝒩β(q)\mathcal{N}_{\beta}(q). The factor represents the distribution of scalar field modes with momentum qq and it reflects the fact that the scalar field is immersed in a thermal bath. Actually, the contribution R2(¯,β)R_{2}(\bar{\mathcal{E}},\beta) consists of two parts. The first part involving 𝒩r(q,¯)\mathcal{N}_{r}(q,\bar{\mathcal{E}}) can be interpreted as the excitation of the rotating detector due to the thermal nature of the scalar field, with modes corresponding to the magnetic quantum number m1m_{1} contributing to the transition probability rate of the detector. The second part involving 𝒩r(q,¯)\mathcal{N}_{r}(q,-\bar{\mathcal{E}}) too signifies the excitation of detector by the thermal character of the field, but with the contributions arising from modes with a different set of magnetic quantum number, viz. m¯1=(q/Ω)¯\bar{m}_{1}=(q/\Omega)-\bar{\mathcal{E}}. Evidently, these latter two contributions can be attributed to the combined effects of both the circular motion of the detector as well as the thermal bath. Since these contributions are influenced by the presence of the thermal bath and vanish in its absence (i.e. when β\beta\to\infty), these contributions can be interpreted as arising due to the stimulated excitation of the detector. Therefore, the overall response of the rotating detector in the thermal bath can be interpreted as arising due to the accumulation of three types of radiative process—two processes (spontaneous and stimulated excitation) to which the scalar field modes with the magnetic quantum number m1m_{1} contribute and another process (stimulated excitation) which arises due to the contributions by the modes with quantum number m¯1\bar{m}_{1}.

In this regard, it may be pointed out that a similar interpretation has also been suggested earlier for the response of a uniformly accelerated detector in a thermal bath (see Refs. Kolekar and Padmanabhan (2015); Kolekar (2014); Kolekar and Padmanabhan (2014)). However, we observe a noticeable difference between the transition probability rates of accelerated detectors and our present system. In the accelerated case, the second term in Eq. (29), which we had interpreted as due to stimulated excitation, has a simpler structure. It is composed of a purely thermal contribution and an excitation due to the acceleration effects stimulated by the thermal bath. Therefore, for an accelerated detector in a thermal bath, there are two independent, spontaneous excitations—one is due to the acceleration (known as the Unruh effect), and the other is purely due to the thermal bath; apart from stimulated excitation, which is influenced by the thermal background. Whereas, in the present study, we do not find any contribution due to the thermal bath that is independent of rotation. This aspect of the transition probability rates of rotating detectors in a thermal bath provides a signature distinct from the case of accelerated detectors.

III.2 Detector switched on for a finite duration

Let us now consider the case wherein the detectors are assumed to be switched on for a finite time interval, say, TT. We shall consider the switching functions κ(τ)\kappa(\tau) to be of the following Gaussian form Sriramkumar and Padmanabhan (1996):

κ(τ)=exp(τ2T2),\kappa(\tau)=\mathrm{exp}\,\left(-\frac{\tau^{2}}{T^{2}}\right), (30)

where, evidently, TT denotes the duration for which the detector remains effectively switched on. In the presence of such a switching function, the transition probability of the detector is given by [cf. Eq. (11)]

FT()\displaystyle F_{T}(\mathcal{E}) =\displaystyle= dτdτei(ττ)G+(τ,τ)\displaystyle\int_{-\infty}^{\infty}{\rm d}\tau^{\prime}\int_{-\infty}^{\infty}{\rm d}\tau\,\mathrm{e}^{-i\,\mathcal{E}\,(\tau-\tau^{\prime})}\,G^{+}(\tau,\tau^{\prime}) (31)
×exp[(τ2+τ2)/T2].\displaystyle\times\,\mathrm{exp}\,\left[-(\tau^{2}+{\tau^{\prime}}^{2})/T^{2}\right].

On changing variables to u=(ττ)u=(\tau-\tau^{\prime}) and v=(τ+τ)v=(\tau+\tau^{\prime}), the integral can be expressed as

FT()\displaystyle F_{T}(\mathcal{E}) =\displaystyle= dv2ev2/(2T2)\displaystyle\int_{-\infty}^{\infty}\frac{{\rm d}v}{2}\,\mathrm{e}^{-v^{2}/(2\,T^{2})} (32)
×dueiuG+(u,v)eu2/(2T2).\displaystyle\times\,\int_{-\infty}^{\infty}{\rm d}u\,\mathrm{e}^{-i\,\mathcal{E}\,u}\,G^{+}(u,v)\,\mathrm{e}^{-u^{2}/(2\,T^{2})}.\qquad

In cases wherein the Wightman function is invariant under time translations in the frame of the detector, i.e. when G+(τ,τ)=G+(u)G^{+}(\tau,\tau^{\prime})=G^{+}(u)—as in the case of the rotating detector [cf. Eqs. (21) and (25)]—we can carry out the Gaussian integral over vv to define the transition probability rate of the detector as follows:

RT()\displaystyle R_{T}(\mathcal{E}) =\displaystyle= FT()(π/2)T\displaystyle\frac{F_{T}(\mathcal{E})}{\sqrt{(\pi/2)}\;T} (33)
=\displaystyle= dueiuG+(u)eu2/(2T2).\displaystyle\int_{-\infty}^{\infty}{\rm d}u\,\mathrm{e}^{-i\,\mathcal{E}\,u}\,G^{+}(u)\,\mathrm{e}^{-u^{2}/(2\,T^{2})}.\qquad

Note that, as required, RT()R()R_{T}(\mathcal{E})\to R(\mathcal{E}) [cf. Eq. (22)] when TT\to\infty. We shall now utilize the above expression to evaluate the finite time response of the rotating detector in the Minkowski vacuum and in a thermal bath.

III.2.1 Response in the Minkowski vacuum

On utilizing the expression (21) for the Wightman function in the Minkowski vacuum along the circular trajectory, the transition probability rate of the detector that is switched on for a finite time interval TT can be expressed as

RT()\displaystyle R_{T}(\mathcal{E}) =\displaystyle= 0dq4πm=Jm2(qσ)\displaystyle\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\sum_{m=-\infty}^{\infty}J^{2}_{m}(q\,\sigma) (34)
×duei[+γ(qmΩ)]ueu2/(2T2).\displaystyle\times\,\int_{-\infty}^{\infty}{\rm d}u\,\mathrm{e}^{-i\,\left[\mathcal{E}+\gamma\,(q-m\,\Omega)\right]\,u}\,\mathrm{e}^{-u^{2}/(2\,T^{2})}.\qquad\;\;

The Gaussian integral over uu can be calculated easily to arrive at

RT()\displaystyle R_{T}(\mathcal{E}) =\displaystyle= 2πTm=0dq4πJm2(qσ)\displaystyle\sqrt{2\,\pi}\;T\,\sum_{m=-\infty}^{\infty}\,\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,J^{2}_{m}(q\,\sigma) (35)
×e[+γ(qmΩ)]2T2/2.\displaystyle\times\,\mathrm{e}^{-\left[\mathcal{E}+\gamma\,(q-m\,\Omega)\right]^{2}\,T^{2}/2}.\qquad

Recall that the Dirac delta function can be represented in terms of the Gaussian function as follows:

δ(1)(z)=limαα2πeα2z2/2.\delta^{(1)}(z)=\lim_{\alpha\to\infty}\frac{\alpha}{\sqrt{2\,\pi}}\,\mathrm{e}^{-\alpha^{2}\,z^{2}/2}. (36)

Therefore, in the limit TT\to\infty, the expression (35) for the finite time response rate of the detector reduces to the result (23) for detectors that remain switched on for infinite time (with the identification of α=γT\alpha=\gamma\,T), as required. Upon noting that Jn2(x)=Jn2(x)J_{n}^{2}(x)=J_{-n}^{2}(x) for integer nn and real xx and setting q=xΩq=x\,\Omega, the integral (35) can be expressed as

RT(¯)\displaystyle R_{T}(\bar{\mathcal{E}}) =\displaystyle= 2πT¯4πγe¯2T¯2/2{0dxJ02(xv)e[(x2/2)+x¯]T¯2\displaystyle\frac{\sqrt{2\,\pi}\;\bar{T}}{4\,\pi\gamma}\,\mathrm{e}^{-\bar{\mathcal{E}}^{2}\,\bar{T}^{2}/2}\,\Biggl{\{}\int_{0}^{\infty}{\rm d}x\,J^{2}_{0}(x\,v)\,\mathrm{e}^{-\left[(x^{2}/2)+x\,\bar{\mathcal{E}}\right]\,\bar{T}^{2}}\, (37)
+ 2m=1em2T¯2/20dxJm2(xv)e[(x2/2)+x¯]T¯2cosh[m(x+¯)T¯2]},\displaystyle+\,2\,\sum_{m=1}^{\infty}\,\mathrm{e}^{-m^{2}\,\bar{T}^{2}/2}\int_{0}^{\infty}{\rm d}x\,J^{2}_{m}(x\,v)\,\mathrm{e}^{-\left[(x^{2}/2)+x\,\bar{\mathcal{E}}\right]\,\bar{T}^{2}}\,\mathrm{cosh}\left[m\,(x+\bar{\mathcal{E}})\,\bar{T}^{2}\right]\Biggr{\}},

where we have defined the dimensionless time interval T¯=γΩT\bar{T}=\gamma\,\Omega\,T. The above integrals and sum seem difficult to evaluate analytically, but they can be computed numerically. We find that the functions Jm2(xv)J_{m}^{2}(x\,v) that appear in the integrands behave as x2mx^{2\,m} when x0x\to 0, and as x1x^{-1} when xx\to\infty (see, for instance, Refs. Watson (2015); Olver ). Note that the factors e[(x2/2)+x¯]T¯2\mathrm{e}^{-[(x^{2}/2)+x\,\bar{\mathcal{E}}]\,\bar{T}^{2}} and cosh[m(x+¯)T¯2]\mathrm{cosh}\left[m\,(x+\bar{\mathcal{E}})\,\bar{T}^{2}\right] reduce to constants as x0x\to 0. As a result, the integrals are well behaved at small xx. Moreover, at large xx, the integrals over xx are dominated by the factor ex2T¯2/2\mathrm{e}^{-x^{2}\,\bar{T}^{2}/2}, and hence converge quickly (in this regard, also see App. C). We evaluate the integrals up to a suitably large value of xx and then carry out the sum involved. In fact, we find that, because of the factor em2T¯2/2\mathrm{e}^{-m^{2}\,\bar{T}^{2}/2}, the sum too converges extremely quickly. In Fig. 3, we have plotted the transition probability rate of the detector in circular motion for a given value of the velocity vv and a few different values of the dimensionless time interval T¯\bar{T}.

Refer to caption
Refer to caption
Figure 3: The transition probability rate RT(¯)R_{T}(\bar{\mathcal{E}}) of the detector in motion on a circular trajectory that has been switched on for a finite time TT has been plotted as a function of the dimensionless energy gap ¯\bar{\mathcal{E}}. We have plotted the results for T¯=(1,5,10)\bar{T}=(1,5,10) (in red, blue and green), assuming the field is in the Minkowski vacuum (on the left) and in a thermal bath (on the right). We have set v=0.5v=0.5 and β¯=1\bar{\beta}=1 in the plotting of these figures. We have arrived at the results for a finite time by integrating over xx from zero up to 10210^{2}, and carrying the sum over mm until (m¯)=10(m-\bar{\mathcal{E}})=10. We have checked that increasing the upper limits of the integral and the sum does not significantly change the results we have obtained. In the figures, we have also indicated the results in the case wherein the detector is switched on forever, i.e. when T¯\bar{T}\to\infty (in gray). Interestingly, we find that the transition probability rate of the detector in the Minkowski vacuum is higher when it is switched on for a shorter duration. We find that, in a thermal bath, the transition probability rate of the detector also exhibits a similar behavior at high ¯\bar{\mathcal{E}} (when ¯0.5\bar{\mathcal{E}}\gtrsim 0.5, for the values of the parameters we have worked with), while at low ¯\bar{\mathcal{E}} the behavior is reversed. At a sufficiently low temperature (say, β¯150\bar{\beta}\gtrsim 150), when T¯\bar{T} is decreased, we observe that the transition probability rate increases over the whole domain of ¯\bar{\mathcal{E}}, just like in the case of the Minkowski vacuum (in this context, see Fig. 4).

Note that, when T¯\bar{T} is made larger, as expected, the transition probability rate approaches the response rate of the detector that remains switched on forever. Interestingly, for a given energy ¯\bar{\mathcal{E}} and velocity vv, the transition probability rate of a detector that is switched on for a shorter duration is higher. However, this seems to occur up to a specific value of ¯\bar{\mathcal{E}}. At sufficiently large values of ¯\bar{\mathcal{E}} and T¯\bar{T}, the transition rate RT(¯)R_{T}(\bar{\mathcal{E}}) decreases, and all the curves corresponding to different values of T¯\bar{T} merge with the transition probability rate of the detector that remains switched on forever.

III.2.2 Response in a thermal bath

Let us now evaluate the finite time response of the detector in circular motion when it is immersed in a thermal bath. As earlier, we shall consider Gaussian switching functions [cf. Eq. (30)]. On substituting the Wightman function at a finite temperature along the trajectory of the rotating detector [cf. Eq. (25)] in the expression (33) that governs the transition probability rate of the detector, we obtain that

RT()\displaystyle R_{T}(\mathcal{E}) =\displaystyle= 0dq4πm=Jm2(qσ)\displaystyle\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\sum_{m=-\infty}^{\infty}J^{2}_{m}(q\,\sigma) (38)
×du{ei[+γ(qmΩ)]ueu2/(2T2)1eβq\displaystyle\times\int_{-\infty}^{\infty}{\rm d}u\,\biggl{\{}\frac{\mathrm{e}^{-i\,\left[\mathcal{E}+\gamma\,(q-m\,\Omega)\right]\,u}\,\mathrm{e}^{-u^{2}/(2\,T^{2})}}{1-\mathrm{e}^{-\beta\,q}}
+ei[γ(qmΩ)]ueu2/(2T2)eβq1}.\displaystyle+\,\frac{\mathrm{e}^{-i\,\left[\mathcal{E}-\gamma\,(q-m\,\Omega)\right]\,u}\,\mathrm{e}^{-u^{2}/(2\,T^{2})}}{\mathrm{e}^{\beta\,q}-1}\biggr{\}}.

Upon carrying out the Gaussian integral over uu, we arrive at

RT()\displaystyle R_{T}(\mathcal{E}) =\displaystyle= 2πT0dq4πm=Jm2(qσ)\displaystyle\sqrt{2\,\pi}\,T\,\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\sum_{m=-\infty}^{\infty}J^{2}_{m}(q\,\sigma) (39)
×{e[+γ(qmΩ)]2T2/21eβq\displaystyle\times\,\biggl{\{}\frac{\mathrm{e}^{-\left[\mathcal{E}+\gamma\,(q-m\,\Omega)\right]^{2}\,T^{2}/2}}{1-\mathrm{e}^{-\beta\,q}}
+e[γ(qmΩ)]2T2/2eβq1}.\displaystyle+\,\frac{\mathrm{e}^{-\left[\mathcal{E}-\gamma\,(q-m\,\Omega)\right]^{2}\,T^{2}/2}}{\mathrm{e}^{\beta\,q}-1}\biggr{\}}.

One can further simplify this expression to eventually obtain that

RT(¯)\displaystyle R_{T}(\bar{\mathcal{E}}) =\displaystyle= 2πT¯4πγe¯2T¯2/2{0dxJ02(xv)[e[(x2/2)+x¯]T¯21eβ¯x+e[(x2/2)x¯]T¯2eβ¯x1]\displaystyle\frac{\sqrt{2\,\pi}\,\bar{T}}{4\,\pi\gamma}\,\mathrm{e}^{-\bar{\mathcal{E}}^{2}\,\bar{T}^{2}/2}\,\Biggl{\{}\int_{0}^{\infty}{\rm d}x\,J^{2}_{0}(x\,v)\,\Biggl{[}\frac{\mathrm{e}^{-\left[(x^{2}/2)+x\,\bar{\mathcal{E}}\right]\,\bar{T}^{2}}}{1-\mathrm{e}^{-\bar{\beta}\,x}}+\frac{\mathrm{e}^{-\left[(x^{2}/2)-x\,\bar{\mathcal{E}}\right]\,\bar{T}^{2}}}{\mathrm{e}^{\bar{\beta}\,x}-1}\Biggr{]} (40)
+ 2m=1em2T¯2/20dxJm2(xv)[e[(x2/2)+x¯]T¯21eβ¯xcosh[m(x+¯)T¯2]\displaystyle+\,2\,\sum_{m=1}^{\infty}\,\mathrm{e}^{-m^{2}\,\bar{T}^{2}/2}\int_{0}^{\infty}{\rm d}x\,J^{2}_{m}(x\,v)\,\Biggl{[}\frac{\mathrm{e}^{-\left[(x^{2}/2)+x\,\bar{\mathcal{E}}\right]\,\bar{T}^{2}}}{1-\mathrm{e}^{-\bar{\beta}\,x}}\,\mathrm{cosh}\left[m\,(x+\bar{\mathcal{E}})\,\bar{T}^{2}\right]
+e[(x2/2)x¯]T¯2eβ¯x1cosh[m(x¯)T¯2]]}.\displaystyle+\,\frac{\mathrm{e}^{-\left[(x^{2}/2)-x\,\bar{\mathcal{E}}\right]\,\bar{T}^{2}}}{\mathrm{e}^{\bar{\beta}\,x}-1}\,\mathrm{cosh}\left[m\,(x-\bar{\mathcal{E}})\,\bar{T}^{2}\right]\Biggr{]}\Biggr{\}}.

Recall that the functions Jm2(xv)J_{m}^{2}(x\,v) behave as x2mx^{2\,m} when x0x\to 0. Also, as we pointed out, the factors e[(x2/2)+x¯]T¯2\mathrm{e}^{-[(x^{2}/2)+x\,\bar{\mathcal{E}}]\,\bar{T}^{2}} and cosh[m(x+¯)T¯2]\mathrm{cosh}\left[m\,(x+\bar{\mathcal{E}})\,\bar{T}^{2}\right] reduce to constants when x0x\to 0. Moreover, note that the functions (1eβ¯x)(1-\mathrm{e}^{-\bar{\beta}\,x}) and (eβ¯x1)(\mathrm{e}^{\bar{\beta}\,x}-1) behave as xx when x0x\to 0. Therefore, in the m=0m=0 term, as x0x\to 0, the integrand in the above expression for RT(¯)R_{T}(\bar{\mathcal{E}}) behaves as x1x^{-1} and the integration over xx leads to a logarithmic divergence as x0x\to 0. This is the infrared divergence at the finite temperature which we had discussed earlier. In contrast to the case wherein the detectors are switched on forever, it proves to be more involved to handle the divergence when the detectors are switched on for a finite duration. We shall regulate the divergence using a procedure which ensures that we recover the results we have already obtained in the limits T¯\bar{T}\to\infty and β¯\bar{\beta}\to\infty. In order to avoid a long digression, we have discussed the procedure in App. E. Once we have tackled the divergence, the remaining terms can be evaluated without any difficulty. As in the earlier cases, we have evaluated the integrals and sum in Eq. (40) numerically to arrive at the response of the detector. We had pointed out that, even in the Minkowski vacuum, the integrals at large xx and the sum at large mm are dominated by the factors ex2T¯2/2\mathrm{e}^{-x^{2}\,\bar{T}^{2}/2} and em2T¯2/2\mathrm{e}^{-m^{2}\,\bar{T}^{2}/2}, respectively. Hence, they converge very quickly, making it convenient to compute them numerically. In the finite temperature case of our interest, additionally, the contribution due to the exponential factor in the denominator in the final term leads to a more rapid convergence of the integral. We have checked that the results are robust against increasing the upper limits of the integral over xx and the sum over mm.

In Fig. 3, we have plotted the transition probability rate of the detector in circular motion and is immersed in a thermal bath for different values of the time interval T¯\bar{T} for which the detector remains effectively switched on, assuming a given velocity vv and inverse temperature β\beta.

Refer to caption
Refer to caption
Figure 4: The transition probability rate RT(¯,β¯)R_{T}(\bar{\mathcal{E}},\bar{\beta}) of the detector in circular motion that is immersed in a thermal bath and has been switched on for a finite time TT has been plotted as a function of the dimensionless inverse temperature β¯\bar{\beta}. We have set v=0.5v=0.5, and have plotted the results for two different values of the dimensionless energy gap, viz. ¯=(0.02,0.6)\bar{\mathcal{E}}=(0.02,0.6) (on the left and right), and three different values of switching time, viz. for T¯=(1,5,10)\bar{T}=(1,5,10) (in red, blue and green, respectively). These plots clearly indicate that the transition probability rate is larger for larger temperatures of the thermal bath (or, equivalently, for smaller values of β¯\bar{\beta}). We also observe that, when ¯\bar{\mathcal{E}} is suitably large, the transition probability rate increases with decreasing T¯\bar{T}, as in the case of the Minkowski vacuum. However, for smaller values of ¯\bar{\mathcal{E}}, we find that this behavior can be reversed when β¯\bar{\beta} is small.

In contrast to the response in the Minkowski vacuum, we find that, at a finite temperature, the transition probability rate of the detector is lower at lower energies, when the detector is switched on for a shorter duration. It is only at high enough energies ¯\bar{\mathcal{E}} that the response is higher when T¯\bar{T} is made smaller. This point should also be clear from Fig. 4 wherein we have plotted the transition probability rate of the detector as a function of the dimensionless inverse temperature β¯\bar{\beta} for different values of ¯\bar{\mathcal{E}} and T¯\bar{T}.

We should point out here that, in the limit β¯\bar{\beta}\to\infty, the quantities [1exp(β¯x)]1[1-\mathrm{exp}\,({-\bar{\beta}\,x})]^{-1} and [exp(β¯x)1]1[\mathrm{exp}\,({\bar{\beta}\,x})-1]^{-1} in Eq. (40) reduce to unity and zero, respectively. Note that the limit β¯\bar{\beta}\to\infty implies a vanishing temperature for the thermal bath and hence corresponds to the Minkowski vacuum. In such a case, the transition probability rates (26) and (40) reduce to the corresponding results for the Minkowski vacuum, viz. Eqs. (23) and (37), as required. Moreover, recall that, earlier, in Eq. (29), we had expressed the transition probability rate of the rotating detector as a sum of different contributions arising due to spontaneous and stimulated excitations. We should point out here that the transition probability rate of the rotating detector which has been switched on for a finite time interval can also be expressed in a similar fashion.

IV Radiative processes of entangled detectors in circular motion

Having discussed the response of individual Unruh-DeWitt detectors, let us now turn to examine the responses of two entangled detectors that are moving on circular trajectories.

IV.1 Detectors switched on for infinite duration

As we have done earlier, let us first discuss the case wherein the detectors are switched on forever, before going to study the situations wherein the detectors are switched on for a finite time interval.

IV.1.1 Response in the Minkowski vacuum

Consider two detectors that are moving along the circular trajectories x~j=(γjτj,σj,γjΩjτj)\tilde{x}_{j}=(\gamma_{j}\,\tau_{j},\sigma_{j},\gamma_{j}\,\Omega_{j}\,\tau_{j}) and x~l=(γlτl,σl,γlΩlτl)\tilde{x}_{l}=(\gamma_{l}\,\tau_{l},\sigma_{l},\gamma_{l}\,\Omega_{l}\,\tau_{l}), with vj(l)=σj(l)Ωj(l))v_{j(l)}=\sigma_{j(l)}\,\Omega_{j(l))} and γj(l)=1/1vj(l)2\gamma_{j(l)}=1/\sqrt{1-v_{j(l)}^{2}}. The circular trajectories of the two detectors are, in general, assumed to be independent. On utilizing the mode decomposition (16) of the scalar field, the positive frequency Wightman function in the Minkowski vacuum which connects two spacetime points corresponding to two such detectors in circular motion can be obtained to be

Gjl+[xj(τj),xl(τl)]\displaystyle G^{+}_{jl}\left[x_{j}(\tau_{j}),x_{l}(\tau^{\prime}_{l})\right] =\displaystyle= 0dq4πm=Jm(qσj)Jm(qσl)\displaystyle\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\sum_{m=-\infty}^{\infty}J_{m}(q\,\sigma_{j})\,J_{m}(q\,\sigma_{l}) (41)
×eiq(γjτjγlτl)\displaystyle\times\,\mathrm{e}^{-i\,q\,(\gamma_{j}\,\tau_{j}-\gamma_{l}\,\tau^{\prime}_{l})}
×eim(γjΩjτjγlΩlτl).\displaystyle\times\,\mathrm{e}^{i\,m\,(\gamma_{j}\,\Omega_{j}\,\tau_{j}-\gamma_{l}\,\Omega_{l}\,\tau^{\prime}_{l})}.

We can arrive at the total transition probability (9) corresponding to the entangled detectors by evaluating the auto and cross transition probabilities Fjl()F_{jl}(\mathcal{E}) defined in Eq. (11).

To evaluate the auto and cross transition probabilities, let us consider the change of variables u¯=τjτl\bar{u}=\tau_{j}-\tau^{\prime}_{l} and v¯=τj+τl\bar{v}=\tau_{j}+\tau^{\prime}_{l}. In terms of these new variables, for the case of detectors that are switched on forever, i.e. when κj(τj)=κl(τl)=1\kappa_{j}(\tau_{j})=\kappa_{l}(\tau_{l})=1, the transition probabilities Fjl()F_{jl}(\mathcal{E}) can be expressed as

Fjl()=dv¯2du¯eiu¯Gjl+(u¯,v¯).F_{jl}(\mathcal{E})=\int_{-\infty}^{\infty}\frac{{\rm d}\bar{v}}{2}\,\int_{-\infty}^{\infty}{\rm d}\bar{u}\,\mathrm{e}^{-i\,\mathcal{E}\,\bar{u}}\,G^{+}_{jl}(\bar{u},\bar{v}). (42)

Upon using the inverse transformations τj=(v¯+u¯)/2\tau_{j}=(\bar{v}+\bar{u})/2 and τl=(v¯u¯)/2\tau^{\prime}_{l}=(\bar{v}-\bar{u})/2 in the expression (41), we obtain the Wightman function Gjl+(u¯,v¯)G^{+}_{jl}(\bar{u},\bar{v}) for the two detectors moving on circular trajectories to be

Gjl+(u¯,v¯)\displaystyle G^{+}_{jl}(\bar{u},\bar{v}) =\displaystyle= 0dq4πm=Jm(qσj)Jm(qσl)\displaystyle\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\sum_{m=-\infty}^{\infty}J_{m}(q\,\sigma_{j})\,J_{m}(q\,\sigma_{l}) (43)
×ei[α1(q)v¯+α2(q)u¯]/2,\displaystyle\times\,\mathrm{e}^{-i\,[\alpha_{1}(q)\,\bar{v}+\alpha_{2}(q)\,\bar{u}]/2},

where the quantities α1(q)\alpha_{1}(q) and α2(q)\alpha_{2}(q) are given by

α1(q)\displaystyle\alpha_{1}(q) =\displaystyle= q(γjγl)m(γjΩjγlΩl),\displaystyle q\,(\gamma_{j}-\gamma_{l})-m\,(\gamma_{j}\,\Omega_{j}-\gamma_{l}\,\Omega_{l}), (44a)
α2(q)\displaystyle\alpha_{2}(q) =\displaystyle= q(γj+γl)m(γjΩj+γlΩl).\displaystyle q\,(\gamma_{j}+\gamma_{l})-m\,(\gamma_{j}\,\Omega_{j}+\gamma_{l}\,\Omega_{l}). (44b)

On carrying the integral over u¯\bar{u}, we obtain the transition probability Fjl()F_{jl}(\mathcal{E}) to be

Fjl()\displaystyle F_{jl}(\mathcal{E}) =\displaystyle= 0dq2m=Jm(qσj)Jm(qσl)\displaystyle\int_{0}^{\infty}\frac{{\rm d}q}{2}\sum_{m=-\infty}^{\infty}J_{m}(q\,\sigma_{j})\,J_{m}(q\,\sigma_{l}) (45)
×δ(1)[+α2(q)2]dv¯2eiα1(q)v¯/2\displaystyle\times\,\delta^{(1)}\left[\mathcal{E}+\frac{\alpha_{2}(q)}{2}\right]\,\int_{-\infty}^{\infty}\frac{{\rm d}\bar{v}}{2}\,\mathrm{e}^{-i\,\alpha_{1}(q)\,\bar{v}/2}
=\displaystyle= 1γj+γlm^Jm(q0σj)Jm(q0σl)\displaystyle\frac{1}{\gamma_{j}+\gamma_{l}}\,\sum_{m\geq\hat{\mathcal{E}}}^{\infty}J_{m}(q_{0}\,\sigma_{j})\,J_{m}(q_{0}\,\sigma_{l})~{}
×dv¯2eiα1(q0)v¯/2,\displaystyle\times\,\int_{-\infty}^{\infty}\frac{{\rm d}\bar{v}}{2}\,\mathrm{e}^{-i\,\alpha_{1}(q_{0})\,\bar{v}/2},

where the quantity q0q_{0} is defined as

q0=(m^)γj+γl(γjΩj+γlΩl)q_{0}=\frac{(m-\hat{\mathcal{E}})}{\gamma_{j}+\gamma_{l}}\,(\gamma_{j}\,\Omega_{j}+\gamma_{l}\,\Omega_{l}) (46)

with ^\hat{\mathcal{E}} being given by

^=2γjΩj+γlΩl.\hat{\mathcal{E}}=\frac{2\,\mathcal{E}}{\gamma_{j}\,\Omega_{j}+\gamma_{l}\,\Omega_{l}}. (47)

We should point out that the condition m^m\geq\hat{\mathcal{E}} arises since q00q_{0}\geq 0. Let us now explicitly evaluate the transition probabilities Fjl()F_{jl}(\mathcal{E}) for the cases j=lj=l and jlj\neq l.

When j=lj=l, the trajectories correspond to the same detector so that we have γj=γl\gamma_{j}=\gamma_{l} and Ωj=Ωl\Omega_{j}=\Omega_{l}, which lead to α1(q0)=0\alpha_{1}(q_{0})=0. In such a case, the expression (45) reduces to the transition probability of a single detector. We can also define the corresponding transition probability rate, say, Rjj()R_{jj}(\mathcal{E}), by dividing the quantity Fjj()F_{jj}(\mathcal{E}) in Eq. (45) by the integral over v¯\bar{v} [cf. Eq. (22)]. Since, ^=/(γjΩj)=¯j\hat{\mathcal{E}}=\mathcal{E}/(\gamma_{j}\,\Omega_{j})=\bar{\mathcal{E}}_{j} and q0=(m¯j)Ωjq_{0}=(m-\bar{\mathcal{E}}_{j})\,\Omega_{j} when j=lj=l, we obtain the transition probability rate of the detector to be

Rjj(¯j)=12γjm¯jJm2[(m¯j)vj],R_{jj}(\bar{\mathcal{E}}_{j})=\frac{1}{2\,\gamma_{j}}\,\sum_{m\geq\bar{\mathcal{E}}_{j}}^{\infty}J^{2}_{m}\left[(m-\bar{\mathcal{E}}_{j})\,v_{j}\right], (48)

where, as we mentioned before, the condition m¯jm\geq\bar{\mathcal{E}}_{j} arises because q00q_{0}\geq 0. This is exactly the result we had obtained earlier when we had considered the response of a single detector [cf. Eq. (23)]. When j=1j=1, the transition probability rate R11(¯1)R_{11}(\bar{\mathcal{E}}_{1}) is given by the expression (23), with (v,γ,¯(v,\gamma,\bar{\mathcal{E}}) replaced by (v1,γ1,¯1)(v_{1},\gamma_{1},\bar{\mathcal{E}}_{1}). If we define the dimensionless parameters Ω¯=Ω2/Ω1\bar{\Omega}=\Omega_{2}/\Omega_{1} and γ¯=γ2/γ1\bar{\gamma}=\gamma_{2}/\gamma_{1}, then the transition probability R22(¯1)R_{22}(\bar{\mathcal{E}}_{1}) can be expressed as

R22(¯1)=12γ2m¯1/(γ¯Ω¯)Jm2[(m¯1γ¯Ω¯)v2].R_{22}(\bar{\mathcal{E}}_{1})=\frac{1}{2\,\gamma_{2}}\,\sum_{m\geq\bar{\mathcal{E}}_{1}/(\bar{\gamma}\,\bar{\Omega})}^{\infty}J^{2}_{m}\left[\left(m-\frac{\bar{\mathcal{E}}_{1}}{\bar{\gamma}\,\bar{\Omega}}\right)\,v_{2}\right]. (49)

Let us now consider the case wherein jlj\neq l. When jlj\neq l, the integral over v¯\bar{v} in Eq. (45) results in a delta function of the form δ(1)[α1(q0)]\delta^{(1)}[\alpha_{1}(q_{0})], which can be utilized to define the transition probability rate. Also, the delta function leads to an additional constraint on mm. For the transition probability rate Rjl()R_{jl}(\mathcal{E}) to be non-zero, other than the condition m^m\geq\hat{\mathcal{E}}, we also require that

m=m0=(γlγj)γjγl(ΩjΩl).m=m_{0}=\frac{(\gamma_{l}-\gamma_{j})\,\mathcal{E}}{\gamma_{j}\,\gamma_{l}\,(\Omega_{j}-\Omega_{l})}. (50)

In other words, the contribution to the transition probability rate arises due to only one term in the sum over mm, leading to

Rjl(^)=1(γj+γl)Jm0(q0σj)Jm0(q0σl),\displaystyle R_{jl}(\hat{\mathcal{E}})=\frac{1}{(\gamma_{j}+\gamma_{l})}\,J_{m_{0}}(q_{0}\,\sigma_{j})\,J_{m_{0}}(q_{0}\,\sigma_{l}), (51)

and we should stress that this result is true only when jlj\neq l. However, since m0m_{0} has to be an integer, the relation (50) implies that it is only for some specific values of the parameters of the system that the transition probability rates R12()R_{12}(\mathcal{E}) and R21()R_{21}(\mathcal{E}) contribute to the radiative process. In terms of the dimensionless parameters γ¯\bar{\gamma} and Ω¯\bar{\Omega}, we can express the transition probability rate R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}) as follows:

R12(¯1)\displaystyle R_{12}(\bar{\mathcal{E}}_{1}) =\displaystyle= 1(γ1+γ2)Jm0[(m0^)(1+γ¯Ω¯)(1+γ¯)v1]\displaystyle\frac{1}{(\gamma_{1}+\gamma_{2})}\,J_{m_{0}}\left[(m_{0}-\hat{\mathcal{E}})\,\frac{(1+\bar{\gamma}\,\bar{\Omega})}{(1+\bar{\gamma})}\,v_{1}\right] (52)
×Jm0[(m0^)(1+γ¯Ω¯)(1+γ¯)Ω¯v2]\displaystyle\times\,J_{m_{0}}\left[(m_{0}-\hat{\mathcal{E}})\,\frac{(1+\bar{\gamma}\,\bar{\Omega})}{(1+\bar{\gamma})\,\bar{\Omega}}\,v_{2}\right]
=\displaystyle= R21(¯1)\displaystyle R_{21}(\bar{\mathcal{E}}_{1})

with ^\hat{\mathcal{E}} and m0m_{0} being given by

^\displaystyle\hat{\mathcal{E}} =\displaystyle= 2γ1Ω1+γ2Ω2=2¯11+γ¯Ω¯,\displaystyle\frac{2\,\mathcal{E}}{\gamma_{1}\,\Omega_{1}+\gamma_{2}\,\Omega_{2}}=\frac{2\,\bar{\mathcal{E}}_{1}}{1+\bar{\gamma}\,\bar{\Omega}},
m0\displaystyle m_{0} =\displaystyle= (γ2γ1)γ1γ2(Ω1Ω2)=(11/γ¯)¯11Ω¯.\displaystyle\frac{(\gamma_{2}-\gamma_{1})\,\mathcal{E}}{\gamma_{1}\,\gamma_{2}\,(\Omega_{1}\,-\Omega_{2})}=\frac{(1-1/\bar{\gamma})\,\bar{\mathcal{E}}_{1}}{1-\bar{\Omega}}. (53)

Let us now understand if the transition probability rate R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}) can be non-zero for values of the parameters describing the trajectories of the two detectors, viz. (Ω1,γ1)(\Omega_{1},\gamma_{1}) and (Ω2,γ2)(\Omega_{2},\gamma_{2}), that we shall focus on. First, consider the case wherein γ1γ2\gamma_{1}\neq\gamma_{2}, while (Ω1Ω2)0+(\Omega_{1}-\Omega_{2})\to 0^{+}. In such a situation, m0m_{0}\to\infty as Ω1Ω2\Omega_{1}\to\Omega_{2}, and the required condition m0^m_{0}\geq\hat{\mathcal{E}} will indeed be satisfied. However, we find that, as mm\to\infty, the function Jm(z)J_{m}(z) goes to zero (in this context, see Refs. Watson (2015); Olver ). This implies that the transition probability rate R12()R_{12}(\mathcal{E}) vanishes. Second, when γ1=γ2\gamma_{1}=\gamma_{2}, we require Ω1Ω2\Omega_{1}\neq\Omega_{2} so that they correspond to different trajectories for the two detectors. In such a situation, m0=0m_{0}=0. However, since ^>0\hat{\mathcal{E}}>0, the condition m0^m_{0}\geq\hat{\mathcal{E}} cannot be satisfied leading to a vanishing R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}). Hence, in these situations, the complete transition probability rate of the two entangled detectors will be solely determined by the rates R11(¯1)R_{11}(\bar{\mathcal{E}}_{1}) and R22(¯1)R_{22}(\bar{\mathcal{E}}_{1}) of the individual detectors. Naturally, constructive or destructive effects due to the cross transition probability rates R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}) and R21(¯1)R_{21}(\bar{\mathcal{E}}_{1}) will be absent in the corresponding total transition rate. However, in general, non-zero contributions due to R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}) and R21(¯1)R_{21}(\bar{\mathcal{E}}_{1}) can be expected to arise when one considers, say, transitions from the symmetric and anti-symmetric Bell states to the collective excited state. We shall discuss these points further in the concluding section.

IV.1.2 Response in a thermal bath

Let us now evaluate the response of the entangled detectors in a thermal bath. In a thermal bath, upon using the decomposition (16) of the scalar field, one can obtain the positive frequency Wightman function connecting two spacetime points corresponding to two differently rotating detectors, denoted by the subscripts jj and ll, to be

Gβjl+[x~j(τj),x~l(τl)]\displaystyle G^{+}_{\beta_{jl}}\left[\tilde{x}_{j}(\tau_{j}),\tilde{x}_{l}(\tau^{\prime}_{l})\right] =\displaystyle= 0dq4πm=Jm(qσj)Jm(qσl)\displaystyle\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\sum_{m=-\infty}^{\infty}J_{m}(q\,\sigma_{j})\,J_{m}(q\,\sigma_{l}) (54)
×[ei[q(γjτjγlτl)m(γjΩjτjγlΩlτl)]1eβq+ei[q(γjτjγlτl)m(γjΩjτjγlΩlτl)]eβq1].\displaystyle\times\,\biggl{[}\frac{\mathrm{e}^{-i\,[q\,(\gamma_{j}\,\tau_{j}-\gamma_{l}\,\tau^{\prime}_{l})-m\,(\gamma_{j}\,\Omega_{j}\,\tau_{j}-\gamma_{l}\,\Omega_{l}\,\tau^{\prime}_{l})]}}{1-\mathrm{e}^{-\beta\,q}}+\frac{\mathrm{e}^{i\,[q\,(\gamma_{j}\,\tau_{j}-\gamma_{l}\,\tau^{\prime}_{l})-m\,(\gamma_{j}\,\Omega_{j}\,\tau_{j}-\gamma_{l}\,\Omega_{l}\,\tau^{\prime}_{l})]}}{\mathrm{e}^{\beta\,q}-1}\biggr{]}.

To calculate the corresponding transition probabilities Fjl()F_{jl}(\mathcal{E}) [cf. Eq. (11)], we proceed as in the case of the Minkowski vacuum and consider the change of variables u¯=τjτl\bar{u}=\tau_{j}-\tau^{\prime}_{l} and v¯=τj+τl\bar{v}=\tau_{j}+\tau^{\prime}_{l}. With the change of variables, the above Wightman function simplifies to be

Gβjl+(u¯,v¯)=0dq4πm=Jm(qσj)Jm(qσl)[ei[α1(q)v¯+α2(q)u¯]/21eβq+ei[α1(q)v¯+α2(q)u¯]/2eβq1],\displaystyle G^{+}_{\beta_{jl}}(\bar{u},\bar{v})=\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\sum_{m=-\infty}^{\infty}J_{m}(q\,\sigma_{j})\,J_{m}(q\,\sigma_{l})\,\left[\frac{\mathrm{e}^{-i\,\left[\alpha_{1}(q)\,\bar{v}+\alpha_{2}(q)\,\bar{u}\right]/2}}{1-\mathrm{e}^{-\beta\,q}}+\frac{\mathrm{e}^{i\,\left[\alpha_{1}(q)\,\bar{v}+\alpha_{2}(q)\,\bar{u}\right]/2}}{\mathrm{e}^{\beta\,q}-1}\right], (55)

where the quantities α1(q)\alpha_{1}(q) and α2(q)\alpha_{2}(q) are given by Eq. (44). Upon substituting this expression in Eq. (11) and integrating over u¯\bar{u}, we obtain the transition probabilities Fjl()F_{jl}(\mathcal{E}) to be

Fjl()\displaystyle F_{jl}(\mathcal{E}) =\displaystyle= 0dq2m=Jm(qσj)Jm(qσl){δ(1)[+α2(q)/2]1eβq+δ(1)[+α2(q)/2]eβq1}dv¯2eiα1(q)v¯/2\displaystyle\int_{0}^{\infty}\frac{{\rm d}q}{2}\,\sum_{m=-\infty}^{\infty}J_{m}(q\,\sigma_{j})\,J_{m}(q\,\sigma_{l})\,\biggl{\{}\frac{\delta^{(1)}\left[\mathcal{E}+\alpha_{2}(q)/2\right]}{1-\mathrm{e}^{-\beta\,q}}+\frac{\delta^{(1)}\left[-\mathcal{E}+\alpha_{2}(q)/2\right]}{\mathrm{e}^{\beta\,q}-1}\biggr{\}}\,\int_{-\infty}^{\infty}\frac{{\rm d}\bar{v}}{2}\,\mathrm{e}^{-i\,\alpha_{1}(q)\,\bar{v}/2} (56)
=\displaystyle= 1(γj+γl)[m^Jm(q0σj)Jm(q0σl)1eβq0dv¯2eiα1(q0)v¯/2\displaystyle\frac{1}{(\gamma_{j}+\gamma_{l})}\,\biggl{[}\sum_{m\geq\hat{\mathcal{E}}}^{\infty}\frac{J_{m}(q_{0}\,\sigma_{j})\,J_{m}(q_{0}\,\sigma_{l})}{1-\mathrm{e}^{-\beta\,q_{0}}}\int_{-\infty}^{\infty}\frac{{\rm d}\bar{v}}{2}\,\mathrm{e}^{-i\,\alpha_{1}(q_{0})\,\bar{v}/2}
+m^Jm(q¯0σj)Jm(q¯0σl)eβq¯01dv¯2eiα1(q¯0)v¯/2],\displaystyle+\,\sum_{m\geq-\hat{\mathcal{E}}}^{\infty}\frac{J_{m}(\bar{q}_{0}\,\sigma_{j})\,J_{m}(\bar{q}_{0}\,\sigma_{l})}{\mathrm{e}^{\beta\,\bar{q}_{0}}-1}\,\int_{-\infty}^{\infty}\frac{{\rm d}\bar{v}}{2}\,\mathrm{e}^{-i\,\alpha_{1}(\bar{q}_{0})\,\bar{v}/2}\biggr{]},

where q0q_{0} is given by Eq. (46), while q¯0\bar{q}_{0} is defined to be

q¯0=(m+^)γj+γl(γjΩj+γlΩl)\bar{q}_{0}=\frac{(m+\hat{\mathcal{E}})}{\gamma_{j}+\gamma_{l}}\,(\gamma_{j}\,\Omega_{j}+\gamma_{l}\,\Omega_{l}) (57)

and, as before, ^\hat{\mathcal{E}} is given by Eq. (47).

Let us first discuss the results in the cases wherein j=lj=l and jlj\neq l. When j=lj=l, we have α1(q0)=α1(q¯0)=0\alpha_{1}(q_{0})=\alpha_{1}(\bar{q}_{0})=0 [cf. Eq. (44a)], and one can readily determine the corresponding transition probability rate to be

Rjj(¯j)\displaystyle R_{jj}(\bar{\mathcal{E}}_{j}) =\displaystyle= 12γj{m¯jJm2[(m¯j)vj]1eβ¯j(m¯j)\displaystyle\frac{1}{2\,\gamma_{j}}\,\biggl{\{}\sum_{m\geq\bar{\mathcal{E}}_{j}}^{\infty}\frac{J^{2}_{m}\left[(m-\bar{\mathcal{E}}_{j})\,v_{j}\right]}{1-\mathrm{e}^{-\bar{\beta}_{j}\,(m-\bar{\mathcal{E}}_{j})}} (58)
+m¯jJm2[(m+¯j)vj]eβ¯j(m+¯j)1},\displaystyle+\,\sum_{m\geq-\bar{\mathcal{E}}_{j}}^{\infty}\frac{J^{2}_{m}\left[(m+\bar{\mathcal{E}}_{j})\,v_{j}\right]}{\mathrm{e}^{\bar{\beta}_{j}\,(m+\bar{\mathcal{E}}_{j})}-1}\biggr{\}},

where we have set β¯j=βΩj\bar{\beta}_{j}=\beta\,\Omega_{j}. As expected, this result is same as the transition probability rate of a single detector in a thermal bath that we had obtained earlier. For instance, the transition probability rate R11(¯1)R_{11}(\bar{\mathcal{E}}_{1}) is given by the expression (26), with (v,γ,¯,β¯)(v,\gamma,\bar{\mathcal{E}},\bar{\beta}) replaced by (v1,γ1,¯1,β¯1)(v_{1},\gamma_{1},\bar{\mathcal{E}}_{1},\bar{\beta}_{1}). Also, we can express the transition probability rate R22(¯1)R_{22}(\bar{\mathcal{E}}_{1}) in terms of the dimensionless parameters Ω¯\bar{\Omega} and γ¯\bar{\gamma} as follows:

R22(¯1)\displaystyle R_{22}(\bar{\mathcal{E}}_{1}) =\displaystyle= 12γ2{m¯1/(γ¯Ω¯)Jm2[(m¯1γ¯Ω¯)v2]\displaystyle\frac{1}{2\,\gamma_{2}}\,\biggl{\{}\sum_{m\geq\bar{\mathcal{E}}_{1}/(\bar{\gamma}\,\bar{\Omega})}^{\infty}J^{2}_{m}\left[\left(m-\frac{\bar{\mathcal{E}}_{1}}{\bar{\gamma}\,\bar{\Omega}}\right)\,v_{2}\right] (59)
×11eβ¯1Ω¯[m(¯1/γ¯Ω¯)]\displaystyle\times\,\frac{1}{1-\mathrm{e}^{-\bar{\beta}_{1}\,\bar{\Omega}\,[m-(\bar{\mathcal{E}}_{1}/\bar{\gamma}\,\bar{\Omega})]}}
+m¯1/(γ¯Ω¯)Jm2[(m+¯1γ¯Ω¯)v2]\displaystyle+\,\sum_{m\geq-\bar{\mathcal{E}}_{1}/(\bar{\gamma}\,\bar{\Omega})}^{\infty}J^{2}_{m}\left[\left(m+\frac{\bar{\mathcal{E}}_{1}}{\bar{\gamma}\,\bar{\Omega}}\right)\,v_{2}\right]
×1eβ¯1Ω¯[m+(¯1/γ¯Ω¯)]1}.\displaystyle\times\,\frac{1}{\mathrm{e}^{\bar{\beta}_{1}\,\bar{\Omega}\,[m+(\bar{\mathcal{E}}_{1}/\bar{\gamma}\,\bar{\Omega})]}-1}\biggr{\}}.

Let us now turn to the case wherein jlj\neq l. In this case, we can notice from Eq. (56) that the integral over v¯\bar{v} leads to δ(1)[α1(q0)]\delta^{(1)}[\alpha_{1}(q_{0})] and δ(1)[α1(q¯0)]\delta^{(1)}[\alpha_{1}(\bar{q}_{0})] in the first and the second sums, respectively. This implies that non-trivial contributions arise only when α1(q0)\alpha_{1}(q_{0}) and α1(q¯0)\alpha_{1}(\bar{q}_{0}) vanish in these sums. We find that, α1(q0)=0\alpha_{1}(q_{0})=0 corresponds to m=m0m=m_{0} as have encountered earlier in the Minkowski vacuum, whereas α1(q¯0)=0\alpha_{1}(\bar{q}_{0})=0 corresponds to m=m0m=-m_{0}. Note that, in order to lead to non-zero contributions to the transition probability rate Rjl()R_{jl}(\mathcal{E}) (with jlj\neq l), while the constraint m0^m_{0}\geq\hat{\mathcal{E}} must be fulfilled in the first sum, the condition m0^-m_{0}\geq-\hat{\mathcal{E}} must be satisfied in the second sum. It is easy to observe that these two conditions cannot be satisfied simultaneously as the second condition, which corresponds to m0^m_{0}\leq\hat{\mathcal{E}}, is mathematically the opposite of the first one.

In particular, if the first condition is fulfilled, i.e. when m0^m_{0}\geq\hat{\mathcal{E}}, then the transition probability rate is given by

R12(¯1)\displaystyle R_{12}(\bar{\mathcal{E}}_{1}) =\displaystyle= 1(γ1+γ2)Jm0[(m0^)(1+γ¯Ω¯)(1+γ¯)v1]\displaystyle\frac{1}{(\gamma_{1}+\gamma_{2})}\,J_{m_{0}}\left[(m_{0}-\hat{\mathcal{E}})\,\frac{(1+\bar{\gamma}\,\bar{\Omega})}{(1+\bar{\gamma})}\,v_{1}\right] (60)
×Jm0[(m0^)(1+γ¯Ω¯)(1+γ¯)Ω¯v2]\displaystyle\times\,J_{m_{0}}\left[(m_{0}-\hat{\mathcal{E}})\,\frac{(1+\bar{\gamma}\,\bar{\Omega})}{(1+\bar{\gamma})\,\bar{\Omega}}\,v_{2}\right]
×11eβ¯1(m0^)(1+γ¯Ω¯)/(1+γ¯)\displaystyle\times\,\frac{1}{1-\mathrm{e}^{-\bar{\beta}_{1}\,(m_{0}-\hat{\mathcal{E}})\,(1+\bar{\gamma}\,\bar{\Omega})/(1+\bar{\gamma})}}
=\displaystyle= R21(¯1),\displaystyle R_{21}(\bar{\mathcal{E}}_{1}),

where β¯1=βΩ1\bar{\beta}_{1}=\beta\,\Omega_{1}. As we had discussed before, we have m0=0m_{0}=0, when γ1=γ2\gamma_{1}=\gamma_{2} and Ω1Ω2\Omega_{1}\neq\Omega_{2}. In such a case, the condition m0^m_{0}\geq\hat{\mathcal{E}} will not be satisfied and hence R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}) vanishes. Moreover, when γ1γ2\gamma_{1}\neq\gamma_{2} and (Ω1Ω2)0+(\Omega_{1}-\Omega_{2})\to 0^{+}, m0m_{0}\to\infty. Since ^\hat{\mathcal{E}} is finite, the condition m0^m_{0}\geq\hat{\mathcal{E}} will indeed be satisfied. However, as we had pointed out earlier, when mm\to\infty, the Bessel functions Jm(z)J_{m}(z) go to zero Watson (2015); Olver . As a result, the transition probability rate R12()R_{12}(\mathcal{E}) vanishes in this case as well.

On the other hand, when the second condition m0^m_{0}\leq\hat{\mathcal{E}} is satisfied, the transition probability rate is given by

R12(¯1)\displaystyle R_{12}(\bar{\mathcal{E}}_{1}) =\displaystyle= 1(γ1+γ2)Jm0[(m0+^)(1+γ¯Ω¯)(1+γ¯)v1]\displaystyle\frac{1}{(\gamma_{1}+\gamma_{2})}\,J_{m_{0}}\left[(-m_{0}+\hat{\mathcal{E}})\,\frac{(1+\bar{\gamma}\,\bar{\Omega})}{(1+\bar{\gamma})}\,v_{1}\right] (61)
×Jm0[(m0+^)(1+γ¯Ω¯)(1+γ¯)Ω¯v2]\displaystyle\times\,J_{m_{0}}\left[(-m_{0}+\hat{\mathcal{E}})\,\frac{(1+\bar{\gamma}\,\bar{\Omega})}{(1+\bar{\gamma})\,\bar{\Omega}}\,v_{2}\right]
×1eβ¯1(m0+^)(1+γ¯Ω¯)/(1+γ¯)1\displaystyle\times\,\frac{1}{\mathrm{e}^{\bar{\beta}_{1}\,(-m_{0}+\hat{\mathcal{E}})\,(1+\bar{\gamma}\,\bar{\Omega})/(1+\bar{\gamma})}-1}
=\displaystyle= R21(¯1).\displaystyle R_{21}(\bar{\mathcal{E}}_{1}).

Recall that, when γ1γ2\gamma_{1}\neq\gamma_{2} and Ω1Ω2\Omega_{1}\to\Omega_{2}, m0m_{0}\to\infty. In such a case, clearly, the condition m0^m_{0}\leq\hat{\mathcal{E}} cannot be met and the cross transition probability rate R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}) between the two entangled detectors will be zero. But, for the case wherein γ1=γ2\gamma_{1}=\gamma_{2} and Ω1Ω2\Omega_{1}\neq\Omega_{2}, since m0=0m_{0}=0, clearly, the condition m0^m_{0}\leq\hat{\mathcal{E}} is satisfied and the cross transition probability rate will be given by the above expression for R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}). Evidently, in contrast to the response in the Minkowski vacuum wherein the cross transition probability rates were always zero (for the parameters we focus on), these rates can contribute non-trivially in a thermal bath for certain sets of the parameters involved. In Fig. 5, we have plotted the cross transition probability rate R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}) for a set of parameters that satisfy the condition m0^m_{0}\leq\hat{\mathcal{E}} and also correspond to an integer value for m0m_{0} (in fact, for m0=0)m_{0}=0).

Refer to caption
Figure 5: The cross transition probability rate R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}) of the two entangled detectors moving on circular trajectories and immersed in a thermal bath has been plotted as a function of ¯1\bar{\mathcal{E}}_{1}, for the case wherein the detectors remain switched on forever. We have set v1=v2=0.5v_{1}=v_{2}=0.5 so that γ¯=1\bar{\gamma}=1, which corresponds to m0=0m_{0}=0, and have chosen Ω¯=5\bar{\Omega}=5. We have plotted the results for the cases wherein β¯1=(0.1,1,10)\bar{\beta}_{1}=(0.1,1,10) (in red, blue, and green, respectively). Clearly, the cross transition probability rate is higher at a higher temperature.

We have plotted the rate for a few different values of the dimensionless inverse temperature β¯1\bar{\beta}_{1}. The figure suggests that the cross transition probability rate decreases with increasing β¯1\bar{\beta}_{1}, a behavior we had encountered earlier when we had discussed the results for the case of a single detector, which corresponds to the auto transition probability rate. In the next section, we will discuss the complete transition probability of the rotating detector, including the auto and the cross transition probability rates. As we shall see, the non-zero cross transition probability rates contribute constructively and destructively for the transition from the symmetric and anti-symmetric Bell states to the collective excited state.

Lastly, it should be noted that, in the limit β\beta\to\infty, i.e. when the temperature of the thermal bath vanishes, the transition probability rate R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}) as given by Eq. (61) above reduces to zero. In the same limit, the last factor in the transition probability rate R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}) as given by Eq. (60) simplifies to unity and the expression reduces to the result in the Minkowski vacuum [cf. Eq. (52)], as required.

IV.2 Detectors switched on for a finite duration

Let us now turn to discuss the responses of entangled detectors that have been switched on for a finite time interval.

IV.2.1 Response in the Minkowski vacuum

As we have done earlier in the case of single detectors, let us now introduce Gaussian switching functions [cf. Eq. (30)] to examine the response of entangled detectors that are switched on for a finite time interval. In such a case, the transition probabilities (11) for two entangled detectors is given by

FjlT()\displaystyle F_{jl}^{T}(\mathcal{E}) =\displaystyle= dτldτjei(τjτl)G+(τj,τl)\displaystyle\int_{-\infty}^{\infty}{\rm d}\tau^{\prime}_{l}\int_{-\infty}^{\infty}{\rm d}\tau_{j}\,\mathrm{e}^{-i\,\mathcal{E}\,(\tau_{j}-\tau^{\prime}_{l})}\,G^{+}\left(\tau_{j},\tau^{\prime}_{l}\right) (62)
×exp[(τj2+τl2)/T2]\displaystyle\times\,\exp\left[-(\tau^{2}_{j}+\tau^{\prime 2}_{l})/T^{2}\right]

which, in terms of the variables u¯=(τjτl)\bar{u}=(\tau_{j}-\tau^{\prime}_{l}) and v¯=(τj+τl)\bar{v}=(\tau_{j}+\tau^{\prime}_{l}), can be expressed as

FjlT()\displaystyle F_{jl}^{T}(\mathcal{E}) =\displaystyle= dv¯2ev¯2/(2T2)\displaystyle\int_{-\infty}^{\infty}\frac{{\rm d}\bar{v}}{2}\,\mathrm{e}^{-\bar{v}^{2}/(2\,T^{2})} (63)
×du¯eiu¯G+(u¯,v¯)eu¯2/(2T2).\displaystyle\times\,\int_{-\infty}^{\infty}{\rm d}\bar{u}\,\mathrm{e}^{-i\,\mathcal{E}\,\bar{u}}\,G^{+}\left(\bar{u},\bar{v}\right)\,\mathrm{e}^{-\bar{u}^{2}/(2\,T^{2})}.\qquad

Recall that, in the Minkowski vacuum, the Wightman function associated with the two entangled detectors that are in circular motion can be expressed as in Eq. (43). Also, as in the case of the single detector, we can define the transition probability rate of the detectors to be

RjlT()=FjlT()(π/2)T.R_{jl}^{T}(\mathcal{E})=\frac{F_{jl}^{T}(\mathcal{E})}{\sqrt{(\pi/2)}\,T}. (64)

Upon substituting the Wightman function (43) in Eq. (63), we find that the corresponding transition probability rate can be expressed as

RjlT()\displaystyle R_{jl}^{T}(\mathcal{E}) =\displaystyle= 1(π/2)T0dq4πm=Jm(qσj)Jm(qσl)\displaystyle\frac{1}{\sqrt{(\pi/2)}\,T}\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\sum_{m=-\infty}^{\infty}J_{m}(q\,\sigma_{j})\,J_{m}(q\,\sigma_{l}) (65)
×dv¯2e[(v¯2/T2)+iα1(q)v¯]/2\displaystyle\times\,\int_{-\infty}^{\infty}\frac{{\rm d}\bar{v}}{2}\,\mathrm{e}^{-\left[(\bar{v}^{2}/T^{2})+i\,\alpha_{1}(q)\,\bar{v}\right]/2}
×du¯eiu¯e[(u¯2/T2)+iα2(q)u¯]/2.\displaystyle\times\,\int_{-\infty}^{\infty}{\rm d}\bar{u}\,\mathrm{e}^{-i\,\mathcal{E}\,\bar{u}}\,\mathrm{e}^{-\left[(\bar{u}^{2}/T^{2})+i\,\alpha_{2}(q)\,\bar{u}\right]/2}.

After carrying out the Gaussian integrals, we obtain that

RjlT()\displaystyle R_{jl}^{T}(\mathcal{E}) =\displaystyle= 2πT0dq4πm=Jm(qσj)Jm(qσl)\displaystyle\sqrt{2\,\pi}\,T\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\sum_{m=-\infty}^{\infty}J_{m}(q\,\sigma_{j})\,J_{m}(q\,\sigma_{l}) (66)
×e{α12(q)+[α2(q)+2]2}T2/8.\displaystyle\times\,\mathrm{e}^{-\left\{\alpha_{1}^{2}(q)+\left[\alpha_{2}(q)+2\,\mathcal{E}\right]^{2}\right\}\,T^{2}/8}.

Since the integral over qq and the sum over mm do not seem to be analytically tractable, we need to compute them numerically as in the case of the single detector.

Let us first consider the auto transition probability rates of the two detectors. As we had discussed, when j=lj=l, we have α1(q)=0\alpha_{1}(q)=0 and α2(q)=2γj(qmΩj)\alpha_{2}(q)=2\,\gamma_{j}(q-m\,\Omega_{j}) and, in such a situation, the above transition probability rate reduces to

RjjT()\displaystyle R_{jj}^{T}(\mathcal{E}) =\displaystyle= 2πT0dq4πm=Jm2(qσj)\displaystyle\sqrt{2\,\pi}\,T\,\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\sum_{m=-\infty}^{\infty}\,J^{2}_{m}(q\,\sigma_{j}) (67)
×e[+γj(qmΩj)]2T2/2.\displaystyle\times\,\mathrm{e}^{-\left[\mathcal{E}+\gamma_{j}\,(q-m\,\Omega_{j})\right]^{2}\,T^{2}/2}.

Note that, as expected, this result exactly matches the transition probability rate of a single detector [cf. Eq. (35)]. We can also introduce the dimensionless variable xj=q/Ωjx_{j}=q/\Omega_{j} and the dimensionless parameter T¯j=γjΩjT\bar{T}_{j}=\gamma_{j}\,\Omega_{j}\,T to rewrite the above integral and sum, as we have done earlier. If we do so, we find that the result for R11T(¯1)R_{11}^{T}(\bar{\mathcal{E}}_{1}) is given by the expression in Eq. (37) with (v,γ,¯,x,T¯)(v,\gamma,\bar{\mathcal{E}},x,\bar{T}) replaced by (v1,γ1,¯1,x1,T¯1)(v_{1},\gamma_{1},\bar{\mathcal{E}}_{1},x_{1},\bar{T}_{1}). With respect to the same set of dimensionless quantities as well as the dimensionless parameters γ¯\bar{\gamma} and Ω¯\bar{\Omega}, we find that the auto transition probability rate R22T(¯1)R_{22}^{T}(\bar{\mathcal{E}}_{1}) can be expressed as follows:

R22T(¯1)\displaystyle R_{22}^{T}(\bar{\mathcal{E}}_{1}) =\displaystyle= 2πT¯14πγ10dx1m=Jm2(x1v2/Ω¯)\displaystyle\frac{\sqrt{2\,\pi}\;\bar{T}_{1}}{4\,\pi\,\gamma_{1}}\,\int_{0}^{\infty}{\rm d}x_{1}\,\sum_{m=-\infty}^{\infty}\,J^{2}_{m}(x_{1}\,v_{2}/\bar{\Omega}) (68)
×e[¯1+γ¯(x1mΩ¯)]2T¯12/2.\displaystyle\times\,\mathrm{e}^{-\left[\bar{\mathcal{E}}_{1}+\bar{\gamma}\,(x_{1}-\,m\,\bar{\Omega})\right]^{2}\,\bar{T}_{1}^{2}/2}.

We should point out that, when γ¯=1=Ω¯\bar{\gamma}=1=\bar{\Omega} and v2=v1v_{2}=v_{1}, this expression reduces to R11T(¯1)R_{11}^{T}(\bar{\mathcal{E}}_{1}).

When j=1j=1 and l=2l=2, we can express the cross transition probability rate R12T(¯1)R_{12}^{T}(\bar{\mathcal{E}}_{1}) as

R12T(¯1)\displaystyle R_{12}^{T}(\bar{\mathcal{E}}_{1})\!\! =\displaystyle= 2πT¯14πγ1m=0dx1Jm(x1v1)Jm(x1v2/Ω¯)\displaystyle\!\!\frac{\sqrt{2\,\pi}\,\bar{T}_{1}}{4\,\pi\,\gamma_{1}}\!\sum_{m=-\infty}^{\infty}\int_{0}^{\infty}{\rm d}x_{1}\,J_{m}(x_{1}\,v_{1})\,J_{m}(x_{1}v_{2}/\bar{\Omega}) (69)
×exp({[(1γ¯)x1m(1γ¯Ω¯)]2\displaystyle\times\,\exp\,\biggl{(}-\biggl{\{}\left[(1-\bar{\gamma})\,x_{1}-m\,(1-\bar{\gamma}\,\bar{\Omega})\right]^{2}
+[2¯1+(1+γ¯)x1m(1+γ¯Ω¯)]2}T¯12/8)\displaystyle+\left[2\,\bar{\mathcal{E}}_{1}+(1+\bar{\gamma})\,x_{1}-m\,(1+\bar{\gamma}\,\bar{\Omega})\right]^{2}\biggr{\}}\,\bar{T}_{1}^{2}/8\biggr{)}
=\displaystyle= R21T(¯1).\displaystyle R_{21}^{T}(\bar{\mathcal{E}}_{1}).

This expression too reduces to that of R11T()R_{11}^{T}(\mathcal{E}) when γ¯=1=Ω¯\bar{\gamma}=1=\bar{\Omega} and v2=v1v_{2}=v_{1}. In App. D, for convenience, we have explicitly listed the complete expressions for the finite time auto and cross transition probability rates R11T(¯1)R_{11}^{T}(\bar{\mathcal{E}}_{1}), R22T(¯1)R_{22}^{T}(\bar{\mathcal{E}}_{1}) and R12T(¯1)R_{12}^{T}(\bar{\mathcal{E}}_{1}). It is these expressions that we actually utilize to numerically compute the rates. While computing the rates, we work with the limits for the variables xx and mm for the integral and the sum that we had considered in the case of a single detector (in this context, see the caption of Fig. 3). In Fig. 6 we have plotted the above cross transition probability rate R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}) for a given set of parameters describing the circular motion of the detector and for different values of the dimensionless time parameter T¯1\bar{T}_{1}.

Refer to caption
Figure 6: The cross transition probability rate R12T(¯1)R_{12}^{T}(\bar{\mathcal{E}}_{1}) of the two entangled detectors in the Minkowski vacuum has been plotted as a function of ¯1\bar{\mathcal{E}}_{1} when the detectors are in motion on circular trajectories and are switched on for a finite time interval TT. As in the previous figure, we have set v1=v2=0.5v_{1}=v_{2}=0.5 so that γ¯=1\bar{\gamma}=1, and have chosen Ω¯=5\bar{\Omega}=5. We have plotted the results for three different values of the dimensionless time interval, viz. T¯1=(1,5,10)\bar{T}_{1}=(1,5,10) (in red, blue, and green, respectively). Note that, as in the case of the auto transition probability rate of a single detector (see Fig. 3), the cross transition probability rate R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}) is higher when the detector is switched on for a shorter time interval.

We observe that the results for the cross transition probability rate are broadly similar to the auto transition probability rate we had obtained earlier in the case of a single detector (see Fig. 3). These plots also suggest that the transition probability rate is higher when the detectors interact with the field for a smaller time interval.

IV.2.2 Response in a thermal bath

We can repeat the procedure that we have adopted earlier to determine the response of two entangled detectors in circular motion and are immersed in a thermal bath. For Gaussian switching functions, we can make use of the expression (63) for the transition probability of the detectors, with the Wightman function in the thermal bath being given by Eq. (55). Upon carrying out the resulting Gaussian integrals over u¯\bar{u} and v¯\bar{v}, we find that the transition probability rate of the detectors [as defined in Eq. (64)] can be expressed as

RjlT()\displaystyle R_{jl}^{T}(\mathcal{E}) =\displaystyle= 2πT0dq4πm=Jm(qσj)Jm(qσl)\displaystyle\sqrt{2\,\pi}\,T\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\sum_{m=-\infty}^{\infty}\,J_{m}(q\,\sigma_{j})\,J_{m}(q\,\sigma_{l}) (70)
×{e{α12(q)+[α2(q)+2]2}T2/81eβq\displaystyle\times\,\biggl{\{}\frac{\mathrm{e}^{-\left\{\alpha_{1}^{2}(q)+[\alpha_{2}(q)+2\,\mathcal{E}]^{2}\right\}\,T^{2}/8}}{1-\mathrm{e}^{-\beta\,q}}
+e{α12(q)+[α2(q)2]2}T2/8eβq1}.\displaystyle+\,\frac{\mathrm{e}^{-\left\{\alpha_{1}^{2}(q)+[\alpha_{2}(q)-2\,\mathcal{E}]^{2}\right\}\,T^{2}/8}}{\mathrm{e}^{\beta\,q}-1}\biggr{\}}.

When j=lj=l, since α1(q)=0\alpha_{1}(q)=0 and α2(q)=2γj(qmΩj)\alpha_{2}(q)=2\,\gamma_{j}\,(q-m\,\Omega_{j}), the above transition probability rate can be expressed as

RjjT()\displaystyle R_{jj}^{T}(\mathcal{E}) =\displaystyle= 2πT0dq4πm=Jm2(qσj)\displaystyle\sqrt{2\,\pi}\,T\,\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\sum_{m=-\infty}^{\infty}J^{2}_{m}(q\,\sigma_{j}) (71)
×{e[+γj(qmΩj)]2T2/21eβq\displaystyle\times\,\bigg{\{}\frac{\mathrm{e}^{-\left[\mathcal{E}+\gamma_{j}\,(q-m\,\Omega_{j})\right]^{2}\,T^{2}/2}}{1-\mathrm{e}^{-\beta\,q}}
+e[γj(qmΩj)]2T2/2eβq1},\displaystyle+\,\frac{\mathrm{e}^{-\left[\mathcal{E}-\gamma_{j}\,(q-m\,\Omega_{j})\right]^{2}\,T^{2}/2}}{\mathrm{e}^{\beta\,q}-1}\biggr{\}},

which is essentially the transition probability rate of a single detector that we encountered earlier [cf. Eq (40)]. For instance, the transition probability rate R11T(¯1)R_{11}^{T}(\bar{\mathcal{E}}_{1}) is given by the expression (40) with (v,γ,¯,β¯,T¯)(v,\gamma,\bar{\mathcal{E}},\bar{\beta},\bar{T}) replaced by (v1,γ1,¯1,β¯1,T¯1)(v_{1},\gamma_{1},\bar{\mathcal{E}}_{1},\bar{\beta}_{1},\bar{T}_{1}). In terms of the dimensionless variables γ¯\bar{\gamma} and Ω¯\bar{\Omega} we had introduced, we find that the transition probability rate R22T(¯1)R_{22}^{T}(\bar{\mathcal{E}}_{1}) can be written as

R22T(¯1)\displaystyle R_{22}^{T}(\bar{\mathcal{E}}_{1}) =\displaystyle= 2πT¯14πγ10dx1m=Jm2(x1v2/Ω¯)\displaystyle\frac{\sqrt{2\,\pi}\;\bar{T}_{1}}{4\,\pi\,\gamma_{1}}\,\int_{0}^{\infty}{\rm d}x_{1}\,\sum_{m=-\infty}^{\infty}\,J^{2}_{m}(x_{1}\,v_{2}/\bar{\Omega}) (72)
×{e[¯1+γ¯(x1mΩ¯)]2T¯12/21eβ¯1x1\displaystyle\times\,\biggl{\{}\frac{\mathrm{e}^{-\left[\bar{\mathcal{E}}_{1}+\bar{\gamma}\,(x_{1}-m\,\bar{\Omega})\right]^{2}\,\bar{T}_{1}^{2}/2}}{1-\mathrm{e}^{-\bar{\beta}_{1}\,x_{1}}}
+e[¯1γ¯(x1mΩ¯)]2T¯12/2eβ¯1x11}.\displaystyle+\frac{\mathrm{e}^{-\left[\bar{\mathcal{E}}_{1}-\bar{\gamma}\,(x_{1}-m\,\bar{\Omega})\right]^{2}\,\bar{T}_{1}^{2}/2}}{\mathrm{e}^{\bar{\beta}_{1}\,x_{1}}-1}\biggr{\}}.

Again, we can numerically compute the integral over the variable x1x_{1} and carry out the sum over mm.

When j=1j=1 and l=2l=2, the transition probability rate R12T(¯1)R_{12}^{T}(\bar{\mathcal{E}}_{1}) can be expressed as follows:

R12T(¯1)\displaystyle R_{12}^{T}(\bar{\mathcal{E}}_{1}) =\displaystyle= 2πT¯14πγ10dx1m=Jm(x1v1)Jm(x1v2/Ω¯)\displaystyle\frac{\sqrt{2\,\pi}\,\bar{T}_{1}}{4\,\pi\,\gamma_{1}}\,\int_{0}^{\infty}{\rm d}x_{1}\sum_{m=-\infty}^{\infty}J_{m}(x_{1}\,v_{1})\,J_{m}(x_{1}\,v_{2}/\bar{\Omega}) (73)
×(11eβ¯1x1e{[(1γ¯)x1m(1γ¯Ω¯)]2+[(1+γ¯)x1m(1+γ¯Ω¯)+2¯1]2}T¯12/8\displaystyle\times\,\biggl{(}\frac{1}{1-\mathrm{e}^{-\bar{\beta}_{1}\,x_{1}}}\,\mathrm{e}^{-\left\{\left[(1-\bar{\gamma})\,x_{1}-m\,(1-\bar{\gamma}\,\bar{\Omega})\right]^{2}+\left[(1+\bar{\gamma})\,x_{1}-m\,(1+\bar{\gamma}\,\bar{\Omega})+2\,\bar{\mathcal{E}}_{1}\right]^{2}\right\}\bar{T}_{1}^{2}/8}
+1eβ¯1x11e{[(1γ¯)x1m(1γ¯Ω¯)]2+[(1+γ¯)x1m(1+γ¯Ω¯)2¯1]2}T¯12/8)=R21T(¯1).\displaystyle+\frac{1}{\mathrm{e}^{\bar{\beta}_{1}\,x_{1}}-1}\,\mathrm{e}^{-\left\{\left[(1-\bar{\gamma})\,x_{1}-m\,(1-\bar{\gamma}\,\bar{\Omega})\right]^{2}+\left[(1+\bar{\gamma})\,x_{1}-m\,(1+\bar{\gamma}\,\bar{\Omega})-2\,\bar{\mathcal{E}}_{1}\right]^{2}\right\}\bar{T}_{1}^{2}/8}\biggr{)}=R_{21}^{T}(\bar{\mathcal{E}}_{1}).

This expression also reduces to R11T()R_{11}^{T}(\mathcal{E}) when γ¯=1=Ω¯\bar{\gamma}=1=\bar{\Omega} and v2=v1v_{2}=v_{1}. As in the case of the response in the Minkowski vacuum, in App. D, we have provided the complete expressions for the auto and cross transition probability rates of the rotating detectors in a thermal bath. Moreover, we carry out the integral and the sum over the domain we had indicated earlier (in the caption of Fig. 3). In Fig. 7 we have plotted the cross transition probability rate R12(¯1)R_{12}(\bar{\mathcal{E}}_{1}) for different values of the dimensionless time T¯1\bar{T}_{1} and a fixed value of the dimensionless inverse temperature β¯1\bar{\beta}_{1}.

Refer to caption
Figure 7: The cross transition probability rate R12T(¯1)R^{T}_{12}(\bar{\mathcal{E}}_{1}) of the two entangled detectors moving on circular trajectories and immersed in a thermal bath has been plotted as a function of the dimensionless energy gap ¯1\bar{\mathcal{E}}_{1} for the case wherein the detectors remain switched on for a finite time interval. As in the previous two figures, we have set v1=v2=0.5v_{1}=v_{2}=0.5 (so that γ¯=1\bar{\gamma}=1) and Ω¯=5\bar{\Omega}=5. We have plotted the results for T¯1=(1,5,10)\bar{T}_{1}=(1,5,10) (in red, blue and green) and β¯1=1\bar{\beta}_{1}=1. Note that the cross transition probability rate behaves in a manner similar to the auto transition probability rate we had plotted earlier (see Fig. 3).

Broadly, the cross transition probability rate exhibits the same behavior as the auto transition probability rate we had encountered earlier [cf. Fig. 3].

V Summary and discussion

In this section, we shall summarize the results we have obtained and conclude with a broader discussion.

V.1 Summary

In the previous section, we had calculated the auto and the cross transition probability rates of the two entangled detectors that are moving on circular trajectories. The auto transition probability rates of the detectors are evidently the same as the response of the single detectors we had discussed initially (in Sec. III). Note that the complete transition probability of the entangled detectors is given by the expression (9), which involves contributions from the auto and cross transition probabilities. Let us now discuss the complete probability rates for transitions from the symmetric and anti-symmetric Bell states to the excited state of the two entangled detectors.

Recall that the transition amplitude of the monopole operator is given by mjωω¯=ω¯|m^j(0)|ωm_{j}^{\omega\bar{\omega}}=\langle\bar{\omega}|\hat{m}_{j}(0)|\omega\rangle, with m^j(0)\hat{m}_{j}(0) being defined in Eq. (10). As we had mentioned, for a transition from the symmetric or the anti-symmetric Bell states (i.e. from |s|s\rangle or |a)|a\rangle) to the collective excited state (i.e. |e)|e\rangle), the transition amplitudes of the monopole operator are found to be m1se=m2se=1/2m_{1}^{se}=m_{2}^{se}=1/\sqrt{2} and m1ae=m2ae=1/2m_{1}^{ae}=-m_{2}^{ae}=-1/\sqrt{2}. Due to this reason, the corresponding transition probability (9) will contain an overall factor of 1/21/2, apart from the factor of μ2\mu^{2} that arises due to the strength of the coupling between the detectors and the scalar field. Since the overall factor μ2/2\mu^{2}/2 does not depend on either the trajectory of the detector or the state of the field, we shall drop the quantity or, equivalently, consider the total transition probability rate, say, ωω¯T~()\mathcal{R}^{\tilde{T}}_{\omega\bar{\omega}}(\mathcal{E}), to be given by

ωω¯T~()=2μ2Γ|ω|ω¯()T~,\displaystyle\mathcal{R}^{\tilde{T}}_{\omega\bar{\omega}}(\mathcal{E})=\frac{2}{\mu^{2}}\,\frac{\Gamma_{|\omega\rangle\to|\bar{\omega}\rangle}(\mathcal{E})}{\tilde{T}}, (74)

where T~=(π/2)T\tilde{T}=\sqrt{(\pi/2)}\,T in the case of detectors that are switched on for a finite duration through the Gaussian switching functions and T~=limTT\tilde{T}=\lim_{T\to\infty}T in the case of detectors that remain switched on forever. Therefore, the total transition probability rates from the symmetric and anti-symmetric Bell states to the collective excited state, referred to by the subscripts ‘sese’ and ‘aeae’, respectively, can be expressed as

seT~()\displaystyle\mathcal{R}_{se}^{\tilde{T}}(\mathcal{E})\!\! =\displaystyle= R11T~()+R22T~()+[R12T~()+R21T~()],\displaystyle\!\!R_{11}^{\tilde{T}}(\mathcal{E})+R_{22}^{\tilde{T}}(\mathcal{E})+\left[R_{12}^{\tilde{T}}(\mathcal{E})+R_{21}^{\tilde{T}}(\mathcal{E})\right], (75a)
aeT~()\displaystyle\mathcal{R}_{ae}^{\tilde{T}}(\mathcal{E})\!\! =\displaystyle= R11T~()+R22T~()[R12T~()+R21T~()].\displaystyle\!\!R_{11}^{\tilde{T}}(\mathcal{E})+R_{22}^{\tilde{T}}(\mathcal{E})-\left[R_{12}^{\tilde{T}}(\mathcal{E})+R_{21}^{\tilde{T}}(\mathcal{E})\right].\qquad\quad (75b)

Note that, in these expressions, for convenience, we have used the notation introduced above, viz. that RjlT~()R_{jl}^{\tilde{T}}(\mathcal{E}) denotes the auto or cross transition probability rate of the detectors switched on for a finite or infinite time interval.

Let us first consider the case wherein the detectors are switched on for infinite duration. When the scalar field is assumed to be in the Minkowski vacuum, in the situations wherein γ1=γ2\gamma_{1}=\gamma_{2} and Ω1Ω2\Omega_{1}\neq\Omega_{2} that we had focused on, the cross transition probability rates R12()R_{12}(\mathcal{E}) and R21()R_{21}(\mathcal{E}) vanish. This implies that the total transition probability rates seT~()\mathcal{R}_{se}^{\tilde{T}}(\mathcal{E}) and aeT~()\mathcal{R}_{ae}^{\tilde{T}}(\mathcal{E}) will be equal and both the rates can be entirely expressed in terms of the auto transition probability rates R11()R_{11}(\mathcal{E}) and R22()R_{22}(\mathcal{E}). As a result, the rates seT~()\mathcal{R}_{se}^{\tilde{T}}(\mathcal{E}) and aeT~(¯)\mathcal{R}_{ae}^{\tilde{T}}(\bar{\mathcal{E}}) can be expected to be similar to that of, say, R11()R_{11}(\mathcal{E}) (in this regard, see Fig. 2). When the detectors are assumed to be immersed in a thermal bath of quanta associated with the scalar field, we had found that, for γ1=γ2\gamma_{1}=\gamma_{2} and Ω1Ω2\Omega_{1}\neq\Omega_{2}, the cross transition probability rates R12()R_{12}(\mathcal{E}) and R21()R_{21}(\mathcal{E}) prove to be non-zero. Consequently, the total transition probability rates seT~()\mathcal{R}_{se}^{\tilde{T}}(\mathcal{E}) and aeT~(¯)\mathcal{R}_{ae}^{\tilde{T}}(\bar{\mathcal{E}}) can be expected to be different. This is evident from Fig. 8 where we have presented these total transition probability rates.

Refer to caption
Refer to caption
Figure 8: The total transition probability rates seT~(¯1)\mathcal{R}^{\tilde{T}}_{se}(\bar{\mathcal{E}}_{1}) (on the left) and aeT~(¯1)\mathcal{R}^{\tilde{T}}_{ae}(\bar{\mathcal{E}}_{1}) (on the right) of the two entangled, rotating detectors in a thermal bath have been plotted as functions of ¯1\bar{\mathcal{E}}_{1} for the case wherein the detectors remain switched on forever. We have set v1=v2=0.5v_{1}=v_{2}=0.5, corresponding to γ¯=1\bar{\gamma}=1, and have chosen Ω¯=5\bar{\Omega}=5, as we have done in the earlier figures. We have plotted the results for three different values of the dimensionless inverse temperature, viz. β¯1=(0.1,1,10)\bar{\beta}_{1}=(0.1,1,10) (in red, blue and green). Note that, for the values of the parameters we have worked with, the total rate seT~(¯1)\mathcal{R}^{\tilde{T}}_{se}(\bar{\mathcal{E}}_{1}) is a factor of 10310^{3} higher at small energies than the rate aeT~(¯1)\mathcal{R}^{\tilde{T}}_{ae}(\bar{\mathcal{E}}_{1}).

Note that the total rate seT~(¯1)\mathcal{R}_{se}^{\tilde{T}}(\bar{\mathcal{E}}_{1}) is about 10310^{3} times larger in magnitude than the rate seT~(¯1)\mathcal{R}_{se}^{\tilde{T}}(\bar{\mathcal{E}}_{1}) for suitably small energies (in fact, for ¯10.01)\bar{\mathcal{E}}_{1}\lesssim 0.01). These findings can provide, in principle, observable distinction in the radiative processes of entangled detectors between the Minkowski vacuum and a thermal bath.

Let us now discuss the cases wherein the detectors are switched on for a finite time interval using the Gaussian switching functions. In Fig. 9, we have plotted the total transition probability rates seT~()\mathcal{R}_{se}^{\tilde{T}}(\mathcal{E}) and aeT~(¯)\mathcal{R}_{ae}^{\tilde{T}}(\bar{\mathcal{E}}) of the entangled detectors in the Minkowski vacuum as well as the thermal bath.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 9: The total transition probability rates seT~(¯1)\mathcal{R}^{\tilde{T}}_{se}(\bar{\mathcal{E}}_{1}) (on the left) and aeT~(¯1)\mathcal{R}^{\tilde{T}}_{ae}(\bar{\mathcal{E}}_{1}) (on the right) of two entangled detectors moving on circular trajectories have been plotted as functions of ¯1\bar{\mathcal{E}}_{1}, when the detectors have been switched on for a finite time interval. We have plotted the results in the Minkowski vacuum (on top) as well as in a thermal bath (at the bottom). As before, we have set v1=v2=0.5v_{1}=v_{2}=0.5 (corresponding to γ¯=1\bar{\gamma}=1) and have chosen Ω¯=5\bar{\Omega}=5. We have set the dimensionless inverse temperature of the thermal bath to be β¯1=1\bar{\beta}_{1}=1. Moreover, as earlier, we have plotted the total rates for three different values of the dimensionless time interval, viz. T¯1=(1,5,10)\bar{T}_{1}=(1,5,10) (in red, blue and green). Clearly, the total rate seT~(¯1)\mathcal{R}^{\tilde{T}}_{se}(\bar{\mathcal{E}}_{1}) is significantly higher than the rate aeT~(¯1)\mathcal{R}^{\tilde{T}}_{ae}(\bar{\mathcal{E}}_{1}) due to the interference effects.

We find that the total rate seT~()\mathcal{R}_{se}^{\tilde{T}}(\mathcal{E}) for a transition from the symmetric Bell state to the excited state is nearly 1010 times than the total rate aeT~(¯)\mathcal{R}_{ae}^{\tilde{T}}(\bar{\mathcal{E}}) from the asymmetric Bell state, due to the constructive and destructive interference we mentioned above. Apart from this aspect, the total rates broadly exhibit characteristics that are similar to what we had encountered in the auto and cross transition probability rates.

V.2 Discussion

In this work, we have examined the response of detectors that are moving on circular trajectories in (2+1)(2+1) dimensional flat spacetime. As has been pointed out before (in this context, see, for example, Ref. Bell and Leinaas (1983)), it seems more realistic and practical to consider detectors that are in motion on circular trajectories than detectors that are moving on uniformly accelerated trajectories. We should mention that certain aspects of the response of entangled detectors that are in motion on circular trajectories have been studied earlier in the literature (see, for instance, Ref. Picanço et al. (2020)). We believe that there are many interesting aspects of the rotating detectors that we have uncovered. To begin with, we find that, in the case of two entangled, rotating detectors, the cross transition probability rates can be comparable to the auto transition probability rates of the individual detectors (cf. Figs. 5, 6, 7). Second, when the detectors are switched on for infinite duration (in both single and entangled cases), the transition probability rate of the rotating detectors in the Minkowski vacuum and a thermal bath are higher at smaller values of the energy gap of the detectors, higher values of their velocity and higher values of the temperature (cf. Fig. 2). Third, in the Minkowski vacuum, interestingly, we find that the transition probability rates of the detectors are higher when they are switched on for a shorter duration (cf. Fig. 3). Though, at first, this result may seem counter-intuitive, it can be interpreted as a manifestation of the energy-time uncertainty principle. The shorter the interval of time that the detector remains switched on, the larger can be the energy of the virtual quanta that are available to excite the detector. Fourth, in a thermal bath, when the detectors are switched on for a finite time interval, we observe that the transition probability rate is higher for smaller intervals of time only when the temperature of the thermal bath is lower or the energy gap of the detectors is higher. In fact, we observe that the behavior can be reversed at higher temperatures and smaller energy gaps (cf. Figs. 3 and 4). Fifth, from Eq. (29) and the related discussions in Sec. III, we identified a specific difference in the nature of the response of single Unruh-DeWitt detectors in a thermal bath, while they are on circular trajectories, when compared to the accelerated case. There is a single spontaneous excitation in the case of circular trajectories due to the motion, while the thermal bath contributes to the stimulated excitations. On the other hand, there are two independent, spontaneous excitations due to the motion and the thermal bath in the accelerated case, in addition to the stimulated excitation due to the thermal bath. Finally, as we had discussed in the previous subsection, due to constructive or destructive interference, the total transition probability rates from the symmetric and anti-symmetric Bell states to the collective excited state can be substantially different in a thermal bath or when they are switched on for a finite time interval in the Minkowski vacuum. We should mention that a similar behavior is also observed when one considers the de-excitation of the detector from the symmetric or the anti-symmetric Bell states to the ground state (in this regard, see the discussion in Refs. Ficek and Tanaś (2002); Menezes and Svaiter (2015)).

There are many further aspects of the rotating and entangled detectors that remain to be explored. We need to urgently extend all our analysis to (3+1)(3+1)-spacetime dimensions. In (3+1)(3+1)-spacetime dimensions, while we expect the results to be qualitatively similar to the (2+1)(2+1)-dimensional case we have considered here, we can expect some quantitative differences. Also, we have to examine whether two initially uncorrelated atomic detectors moving on circular trajectories can get entangled over time, a phenomenon that has been referred to as entanglement harvesting (in this regard, see Refs. Reznik (2003); Martin-Martinez et al. (2016); Koga et al. (2018, 2019); Barman et al. (2021a, b)). In particular, we need to investigate entanglement harvesting in the presence of a thermal bath, with detectors switched on for infinite as well as finite intervals of time (for previous studies in this context involving static and non-inertial detectors in a thermal bath, see Refs. Brown (2013); Simidzija and Martín-Martínez (2018) and Ref. Barman et al. (2021a), respectively). Moreover, it will be interesting to study the effects due to the presence of boundaries Davies et al. (1996); Picanço et al. (2020). Further, on the practical front, it is easier to set charged particles in motion on circular trajectories, using, say, with the help of an external magnetic field. If a charged particle (say, an ion) is to be used as a detector, then it may emit classical synchrotron radiation as it moves along the circular trajectories. We need to understand the implications or effects of the synchrotron radiation for the detection of quanta emitted or absorbed by the detector due to the quantum phenomena we are investigating. We are presently working on these issues.

Acknowledgements.
SB would like to thank the Science and Engineering Research Board, Government of India, for supporting this work through the National Postdoctoral Fellowship PDF/2022/000428.

Appendix A Wightman function in polar coordinates

In (2+1)(2+1)-dimensional flat spacetime, when working in the Cartesian coordinates 𝒙=(x,y){\bm{x}}=(x,y), the Wightman function associated with a massive, minimally coupled, scalar field can be written as (in this regard, see, for instance, Ref. Birrell and Davies (1984b))

G+(x~,x~)=d2𝒌(2π)2(2ω)eiω(tt)+i𝒌(𝒙𝒙),\displaystyle G^{+}(\tilde{x},\tilde{x}^{\prime})=\int\frac{{\rm d}^{2}{\bm{k}}}{(2\,\pi)^{2}\,(2\,\omega)}\,\mathrm{e}^{-i\,\omega\,(t-t^{\prime})+i\,{\bm{k}}\cdot({\bm{x}}-{\bm{x}}^{\prime})}, (76)

where ω=(|𝒌|2+μ2)1/2\omega=(|{\bm{k}}|^{2}+\mu^{2})^{1/2}, and μ\mu denotes the mass of the field. Let us write both the wave vector 𝒌{\bm{k}} and the position vector 𝒙{\bm{x}} in terms of the corresponding polar coordinates, say, (q,α)(q,\alpha) and (ρ,ϕ)(\rho,\phi), as follows:

kx\displaystyle k_{x} =\displaystyle= qcosα,ky=qsinα,\displaystyle q\,\cos{\alpha},\quad\;k_{y}=q\,\sin{\alpha}, (77a)
x\displaystyle x =\displaystyle= ρcosϕ,y=ρsinϕ,\displaystyle\rho\,\cos{\phi},\quad\;\;\,y=\rho\,\sin{\phi}, (77b)
x\displaystyle x^{\prime} =\displaystyle= ρcosϕ,y=ρsinϕ.\displaystyle\rho^{\prime}\,\cos{\phi^{\prime}},\quad y^{\prime}=\rho^{\prime}\,\sin{\phi^{\prime}}. (77c)

In such a case, the Wightman function (76) can be expressed as

G+(x~,x~)\displaystyle G^{+}(\tilde{x},\tilde{x}^{\prime}) =\displaystyle= 0dqq(2π)(2ω)02πdα2πeiω(tt)\displaystyle\int_{0}^{\infty}\frac{{\rm d}q\,q}{(2\,\pi)\,(2\,\omega)}\int_{0}^{2\,\pi}\frac{{\rm d}\alpha}{2\,\pi}\,\mathrm{e}^{-i\,\omega\,(t-t^{\prime})} (78)
×eiq[ρcos(ϕα)ρcos(ϕα)].\displaystyle\times\,\mathrm{e}^{i\,q\,\left[\rho\,\cos(\phi-\alpha)-\rho^{\prime}\,\cos(\phi^{\prime}-\alpha)\right]}.

If we now use the following identity (known as the Jacobi–Anger identity; in this context, see, for instance, Ref. Arfken (1985))

eizcosϕ=m=imJm(z)eimϕ,\displaystyle\mathrm{e}^{i\,z\,\cos{\phi}}=\sum_{m=-\infty}^{\infty}i^{m}\,J_{m}(z)\,\mathrm{e}^{i\,m\,\phi}, (79)

where Jm(z)J_{m}(z) are the Bessel functions, then the Wightman function can be written as

G+(x~,x~)\displaystyle G^{+}(\tilde{x},\tilde{x}^{\prime}) =\displaystyle= 0dqq(2π)(2ω)02πdα2πeiω(tt)\displaystyle\int_{0}^{\infty}\frac{{\rm d}q\,q}{(2\,\pi)\,(2\,\omega)}\,\int_{0}^{2\pi}\,\frac{{\rm d}\alpha}{2\,\pi}\,\mathrm{e}^{-i\,\omega\,(t-t^{\prime})}
×m=imJm(qρ)eim(ϕα)\displaystyle\times\,\sum_{m=-\infty}^{\infty}i^{m}\,J_{m}(q\,\rho)\,\mathrm{e}^{i\,m\,(\phi-\alpha)}\,
×m=imJm(qρ)eim(ϕα)\displaystyle\times\,\sum_{m^{\prime}=-\infty}^{\infty}i^{m^{\prime}}\,J_{m^{\prime}}(-q\,\rho^{\prime})\,\mathrm{e}^{i\,m^{\prime}\,(\phi^{\prime}-\alpha)}
=\displaystyle= 0dqq4πωeiω(tt)\displaystyle\int_{0}^{\infty}\frac{{\rm d}q\,q}{4\,\pi\,\omega}\,\mathrm{e}^{-i\,\omega\,(t-t^{\prime})}
×m=Jm(qρ)Jm(qρ)eim(ϕϕ).\displaystyle\times\,\sum_{m=-\infty}^{\infty}J_{m}(q\,\rho)\,J_{-m}(-q\,\rho^{\prime})\,\mathrm{e}^{i\,m\,(\phi-\phi^{\prime})}.

We should mention that, to arrive at the final equality, we have used the relation

02πdαei(m+m)α=(2π)δm,m,\int_{0}^{2\,\pi}{\rm d}\alpha\,\mathrm{e}^{-i\,(m+m^{\prime})\,\alpha}=(2\,\pi)\,\delta_{m,-m^{\prime}}, (81)

where δn,n\delta_{n,n^{\prime}} denotes the Kronecker delta. On using the identity Jm(qρ)=Jm(qρ)J_{-m}(-q\,\rho^{\prime})=J_{m}(q\,\rho^{\prime}), in the case of a massless field (i.e. when μ=0\mu=0 so that ω=q\omega=q), we can arrive at the expression (19) for the Wightman function we have mentioned earlier.

Similarly, when working in the Cartesian coordinates, the Wightman function for the scalar field at a finite temperature in (2+1)(2+1) spacetime dimensions can be easily obtained to be (see, for instance, Ref. Birrell and Davies (1984b); Kolekar and Padmanabhan (2014); Chowdhury et al. (2019))

Gβ+(x~,x~)\displaystyle G_{\beta}^{+}(\tilde{x},\tilde{x}^{\prime}) =\displaystyle= d2𝒌(2π)2(2ω){ei[ω(tt)𝒌(𝒙𝒙)]1eβω\displaystyle\int\frac{{\rm d}^{2}{\bm{k}}}{(2\,\pi)^{2}\,(2\,\omega)}\,\biggl{\{}\frac{\mathrm{e}^{-i\,\left[\omega\,(t-t^{\prime})-{\bm{k}}\cdot({\bm{x}}-{\bm{x}}^{\prime})\right]}}{1-\mathrm{e}^{-\beta\,\omega}} (82)
+ei[ω(tt)𝒌(𝒙𝒙)]eβω1}.\displaystyle+\frac{\mathrm{e}^{i\,\left[\omega\,(t-t^{\prime})-{\bm{k}}\cdot({\bm{x}}-{\bm{x}}^{\prime})\right]}}{\mathrm{e}^{\beta\,\omega-1}}\biggr{\}}.

In the case of a massless field, upon carrying out the transformations (77), we can arrive at the expression for the above Wightman function in terms of the polar coordinates (in both real and momentum space), which is the result (24) we have quoted earlier.

Appendix B Behavior of the transition probability rate as a function of velocity of the detector

In our discussion, barring in Fig. 4, we have been primarily interested in computing the auto and cross transition probability rates of the detectors as a function of the energy gap \mathcal{E}, for given angular and linear velocities Ω\Omega and vv of the detector, inverse temperature β\beta of the thermal bath and the time interval TT for which the detector is switched on. We had pointed out that the sums over mm which appear in the transition probability rates converge fairly quickly. Specifically, we had mentioned that, for the parameters we have considered, it is adequate to evaluate the sum until (m¯)=50(m-\bar{\mathcal{E}})=50 and (m¯)=10(m-\bar{\mathcal{E}})=10 when the detectors are switched on for infinite or a finite duration (in this regard, see the captions of Figs. 2 and 3). However, when the detector is switched on for infinite duration, as we had discussed in Sec. III [see our discussion following Eq. (23)], the convergence of the sum depends on the velocity of the detector. We find that, for detector velocities very close to the velocity of light (say, for v0.9v\gtrsim 0.9), it becomes necessary to evaluate the sum to larger values of mm. To illustrate this point, in Fig. 10, we have plotted the transition probability rate of the rotating detector in the Minkowski vacuum as a function of the linear velocity vv of the detector.

Refer to caption
Figure 10: The transition probability rate of the detector R(¯,v)R(\bar{\mathcal{E}},v) in the Minkowski vacuum, when it is moving on a circular trajectory and has been switched on for infinite duration, has been plotted as a function of the velocity vv, when the sum over mm is carried out to larger and larger values, viz. until (m¯)=(50,100,150)(m-\bar{\mathcal{E}})=(50,100,150) (in red, blue and green, respectively). We have set ¯=0.1\bar{\mathcal{E}}=0.1 in arriving at these plots. Note that, for velocities close to unity, the peak in the transition probability rates shifts towards higher values of the velocity as we sum until larger values of mm. This illustrates that care needs to be exercised when the sum over mm is carried out. In the plots we have presented earlier, at every stage, we have checked and confirmed that the terms we have ignored do not contribute significantly to the transition probability rate of the detector.

We have fixed the value of the energy gap \mathcal{E} in plotting the results and have assumed that the detector remains switched on forever. Note that, for v0.9v\gtrsim 0.9, the peak in the transition probability rate of the detector shifts towards higher velocities as we sum to larger and larger values of mm. In the results we have presented in all the earlier figures, we have ensured that, for the parameters we have worked with, summing to larger values of mm does not significantly change the results we obtain.

Appendix C Behavior of the integrand in the finite time transition probability rate

Recall that, in the case of detectors that are switched on for a finite time interval, we have to carry out an integral over xx, apart from summing over mm. Immediately after Eq. (37), we had discussed the behavior of the integrals at large and small values of xx. In this appendix, we shall briefly illustrate the behavior of the integrands that are encountered when evaluating the response of the detector in the Minkowski vacuum. Note that the integrands in this case are of the following form:

T(x)\displaystyle\mathcal{I}_{T}(x) =\displaystyle= Jm2(xv)e[(x2/2)+x¯]T¯2em2T¯2/2\displaystyle J^{2}_{m}(x\,v)\,\mathrm{e}^{-\left[(x^{2}/2)+x\,\bar{\mathcal{E}}\right]\,\bar{T}^{2}}\,\mathrm{e}^{-m^{2}\,\bar{T}^{2}/2} (83)
×cosh[m(x+¯)T¯2].\displaystyle\times\,\mathrm{cosh}\left[m\,(x+\bar{\mathcal{E}})\,\bar{T}^{2}\right].

In Fig. 11, we have plotted this integrand for m=0m=0 and m=1m=1 and a few different values of the dimensionless time interval T¯\bar{T}, assuming fixed values for the velocity vv and the dimensionless energy gap ¯\bar{\mathcal{E}}.

Refer to caption
Refer to caption
Figure 11: The integrand T(x)\mathcal{I}_{T}(x) [cf. Eq. (83) has been plotted as a function of the dimensionless variable xx for m=0m=0 (on the left) and m=1m=1 (on the right). We have chosen v=0.5v=0.5 and ¯=0.01\bar{\mathcal{E}}=0.01, and have plotted the integrand for T¯=(1,5,10)\bar{T}=(1,5,10) (in red, blue, and green). Clearly, the integrand is well behaved at small xx and quickly dies down at large xx, allowing us to efficiently compute the integrals.

It is evident that the integrands are well behaved and, in particular, they decrease rapidly at large xx due to the ex2T¯2/2\mathrm{e}^{-x^{2}\,\bar{T}^{2}/2} factor. Moreover, for m>0m>0, the factor em2T¯2/2\mathrm{e}^{-m^{2}\,\bar{T}^{2}/2} also suppresses the overall amplitude of the integrand. Such a rapid decrease allows us to quickly compute the integrals involved. As we had mentioned, for the values of the parameters we work with, we find that it is adequate to integrate up to x=102x=10^{2} (see the caption of Fig. 3) and we have also checked that the results we obtain are not altered if we increase the upper limit.

Appendix D Transition probability rates of entangled detectors for finite duration

In this appendix, we shall provide explicit expressions for the transition probability rates of the entangled detectors that are moving on circular trajectories and are interacting with the scalar field for a finite time interval. It is these expressions that we eventually use to numerically compute the transition probability rates.

Let us first consider the case of the Minkowski vacuum. We find that the expressions (67) and (68) can be utilized to write R11T(¯1)R_{11}^{T}(\bar{\mathcal{E}}_{1}) and R22T(¯1)R_{22}^{T}(\bar{\mathcal{E}}_{1}) as

R11T(¯1)\displaystyle R_{11}^{T}(\bar{\mathcal{E}}_{1}) =\displaystyle= 2πT¯14πγ1e¯12T¯12/2{0dx1J02(x1v1)e[(x12/2)+x1¯1]T¯12\displaystyle\frac{\sqrt{2\,\pi}\;\bar{T}_{1}}{4\,\pi\,\gamma_{1}}\,\mathrm{e}^{-\bar{\mathcal{E}}_{1}^{2}\,\bar{T}_{1}^{2}/2}\,\Biggl{\{}\int_{0}^{\infty}{\rm d}x_{1}\,J^{2}_{0}(x_{1}\,v_{1})\,\mathrm{e}^{-\left[(x_{1}^{2}/2)+x_{1}\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}} (84a)
+ 2m=1em2T¯12/20dx1Jm2(x1v1)e[(x12/2)+x1¯1]T¯12cosh[m(x1+¯1)T¯12]},\displaystyle+\,2\,\sum_{m=1}^{\infty}\,\mathrm{e}^{-m^{2}\,\bar{T}_{1}^{2}/2}\int_{0}^{\infty}{\rm d}x_{1}\,J^{2}_{m}(x_{1}\,v_{1})\,\mathrm{e}^{-\left[(x_{1}^{2}/2)+x_{1}\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}}\,\mathrm{cosh}\left[m\,(x_{1}+\bar{\mathcal{E}}_{1})\,\bar{T}_{1}^{2}\right]\Biggr{\}}\,,
R22T()\displaystyle R_{22}^{T}(\mathcal{E}) =\displaystyle= 2πT¯14πγ1e¯12T¯12/2{0dx1J02(x1v2/Ω¯)e[(x12/2)+x1¯1/γ¯]γ¯2T¯12\displaystyle\frac{\sqrt{2\,\pi}\;\bar{T}_{1}}{4\,\pi\,\gamma_{1}}\,\mathrm{e}^{-\bar{\mathcal{E}}_{1}^{2}\,\bar{T}_{1}^{2}/2}\,\Biggl{\{}\int_{0}^{\infty}{\rm d}x_{1}\,J^{2}_{0}\left(x_{1}\,v_{2}/\bar{\Omega}\right)\,\mathrm{e}^{-\left[(x_{1}^{2}/2)+x_{1}\,\bar{\mathcal{E}}_{1}/\bar{\gamma}\right]\,\bar{\gamma}^{2}\,\bar{T}_{1}^{2}} (84b)
+ 2m=1em2γ¯2Ω¯2T¯12/20dx1Jm2(x1v2/Ω¯)e[(x12/2)+x1¯1/γ¯]γ¯2T¯12\displaystyle+\,2\,\sum_{m=1}^{\infty}\,\mathrm{e}^{-m^{2}\,\bar{\gamma}^{2}\,\bar{\Omega}^{2}\,\bar{T}_{1}^{2}/2}\,\int_{0}^{\infty}{\rm d}x_{1}\,J^{2}_{m}\left(x_{1}v_{2}/\bar{\Omega}\right)\,\mathrm{e}^{-\left[(x_{1}^{2}/2)+x_{1}\,\bar{\mathcal{E}}_{1}/\bar{\gamma}\right]\,\bar{\gamma}^{2}\,\bar{T}_{1}^{2}}
×cosh[m(x1+¯1/γ¯)γ¯2Ω¯T¯12]}.\displaystyle\times\,\mathrm{cosh}\left[m\,\left(x_{1}+\bar{\mathcal{E}}_{1}/\bar{\gamma}\right)\,\bar{\gamma}^{2}\,\bar{\Omega}\,\bar{T}_{1}^{2}\right]\Biggr{\}}.\qquad\quad

Similarly, the expression (69) for the cross transition probability rate R12T(¯1)R_{12}^{T}(\bar{\mathcal{E}}_{1}) in the Minkowski vacuum can be written as

R12T(¯1)\displaystyle R_{12}^{T}(\bar{\mathcal{E}}_{1}) =\displaystyle= 2πT¯14πγ1e¯12T¯12/2(0dx1J0(x1v1)J0(x1v2/Ω¯)e[(1+γ¯2)x12+2(1+γ¯)x1¯1]T¯12/4\displaystyle\frac{\sqrt{2\,\pi}\,\bar{T}_{1}}{4\,\pi\,\gamma_{1}}\,\mathrm{e}^{-\bar{\mathcal{E}}_{1}^{2}\,\bar{T}_{1}^{2}/2}\,\Biggl{(}\int_{0}^{\infty}{\rm d}x_{1}\,J_{0}(x_{1}\,v_{1})\,J_{0}(x_{1}\,v_{2}/\bar{\Omega})\,\mathrm{e}^{-\left[(1+\bar{\gamma}^{2})\,x_{1}^{2}+2\,(1+\bar{\gamma})\,x_{1}\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}/4} (85)
+ 2m=1em2(1+γ¯2Ω¯2)T¯12/40dx1Jm(x1v1)Jm(x1v2/Ω¯)e[(1+γ¯2)x12+2(1+γ¯)x1¯1]T¯12/4\displaystyle+\,2\,\sum_{m=1}^{\infty}\,\mathrm{e}^{-m^{2}\,(1+\bar{\gamma}^{2}\,\bar{\Omega}^{2})\,\bar{T}_{1}^{2}/4}\int_{0}^{\infty}{\rm d}x_{1}\,J_{m}(x_{1}\,v_{1})\,J_{m}(x_{1}\,v_{2}/\bar{\Omega})\,\mathrm{e}^{-\left[(1+\bar{\gamma}^{2})\,x_{1}^{2}+2\,(1+\bar{\gamma})\,x_{1}\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}/4}
×cosh{m[(1+γ¯2Ω¯)x1+(1+γ¯Ω¯)¯1]T¯12/2})=R21T(¯1).\displaystyle\times\,\mathrm{cosh}\left\{m\,\left[(1+\bar{\gamma}^{2}\,\bar{\Omega})\,x_{1}+(1+\bar{\gamma}\,\bar{\Omega})\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}/2\right\}\Biggr{)}=R_{21}^{T}(\bar{\mathcal{E}}_{1}).

The expressions (71) and (72) can be used to write the auto transition probability rates of the rotating detectors in a thermal bath as

R11T(¯1)\displaystyle R_{11}^{T}(\bar{\mathcal{E}}_{1}) =\displaystyle= 2πT¯14πγ1e¯12T¯12/2{0dx1J02(x1v1)[e[(x12/2)+x1¯1]T¯121eβ¯1x1+e[(x12/2)x1¯1]T¯12eβ¯1x11]\displaystyle\frac{\sqrt{2\,\pi}\;\bar{T}_{1}}{4\,\pi\,\gamma_{1}}\,\mathrm{e}^{-\bar{\mathcal{E}}_{1}^{2}\,\bar{T}_{1}^{2}/2}\,\Biggl{\{}\int_{0}^{\infty}{\rm d}x_{1}\,J^{2}_{0}(x_{1}\,v_{1})\,\left[\frac{\mathrm{e}^{-\left[(x_{1}^{2}/2)+x_{1}\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}}}{1-\mathrm{e}^{-\bar{\beta}_{1}\,x_{1}}}+\frac{\mathrm{e}^{-\left[(x_{1}^{2}/2)-x_{1}\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}}}{\mathrm{e}^{\bar{\beta}_{1}\,x_{1}}-1}\,\right] (86a)
+ 2m=1em2T¯12/20dx1Jm2(x1v1)[e[(x12/2)+x1¯1]T¯121eβ¯1x1cosh[m(x1+¯1)T¯12]\displaystyle+\,2\,\sum_{m=1}^{\infty}\,\mathrm{e}^{-m^{2}\,\bar{T}_{1}^{2}/2}\int_{0}^{\infty}{\rm d}x_{1}\,J^{2}_{m}(x_{1}\,v_{1})\,\biggl{[}\frac{\mathrm{e}^{-\left[(x_{1}^{2}/2)+x_{1}\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}}}{1-\mathrm{e}^{-\bar{\beta}_{1}\,x_{1}}}\,\mathrm{cosh}\left[m\,(x_{1}+\bar{\mathcal{E}}_{1})\,\bar{T}_{1}^{2}\right]
+e[(x12/2)x1¯1]T¯12eβ¯1x11cosh[m(x1¯1)T¯12]},\displaystyle+\,\frac{\mathrm{e}^{-\left[(x_{1}^{2}/2)-x_{1}\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}}}{\mathrm{e}^{\bar{\beta}_{1}\,x_{1}}-1}\,\mathrm{cosh}\left[m\,(x_{1}-\bar{\mathcal{E}}_{1})\,\bar{T}_{1}^{2}\right]\Biggr{\}},
R22T(¯1)\displaystyle R_{22}^{T}(\bar{\mathcal{E}}_{1}) =\displaystyle= 2πT¯14πγ1e¯12T¯12/2(0dx1J02(x1v2Ω¯)[e[(x12/2)+x1¯1/γ¯]γ¯2T¯121eβ¯1x1+e[(x12/2)x1¯1/γ¯]γ¯2T¯12eβ¯1x11]\displaystyle\frac{\sqrt{2\,\pi}\;\bar{T}_{1}}{4\,\pi\gamma_{1}}\,\mathrm{e}^{-\bar{\mathcal{E}}_{1}^{2}\,\bar{T}_{1}^{2}/2}\,\Biggl{(}\int_{0}^{\infty}{\rm d}x_{1}\,J^{2}_{0}\left(\frac{x_{1}\,v_{2}}{\bar{\Omega}}\right)\,\left[\frac{\mathrm{e}^{-\left[(x_{1}^{2}/2)+x_{1}\,\bar{\mathcal{E}}_{1}/\bar{\gamma}\right]\,\bar{\gamma}^{2}\,\bar{T}_{1}^{2}}}{1-\mathrm{e}^{-\bar{\beta}_{1}\,x_{1}}}+\frac{\mathrm{e}^{-\left[(x_{1}^{2}/2)-x_{1}\,\bar{\mathcal{E}}_{1}/\bar{\gamma}\right]\,\bar{\gamma}^{2}\,\bar{T}_{1}^{2}}}{\mathrm{e}^{\bar{\beta}_{1}\,x_{1}}-1}\right] (86b)
+ 2m=1em2γ¯2Ω¯2T¯12/20dx1Jm2(x1v2/Ω¯){e[(x12/2)+x1¯1/γ¯]γ¯2T¯121eβ¯1x1cosh[m(x1+¯1γ¯)γ¯2Ω¯T¯12]\displaystyle+\,2\,\sum_{m=1}^{\infty}\,\mathrm{e}^{-m^{2}\,\bar{\gamma}^{2}\,\bar{\Omega}^{2}\,\bar{T}_{1}^{2}/2}\,\int_{0}^{\infty}{\rm d}x_{1}\,J^{2}_{m}\left(x_{1}v_{2}/\bar{\Omega}\right)\,\biggl{\{}\frac{\mathrm{e}^{-\left[(x_{1}^{2}/2)+x_{1}\,\bar{\mathcal{E}}_{1}/\bar{\gamma}\right]\,\bar{\gamma}^{2}\,\bar{T}_{1}^{2}}}{1-\mathrm{e}^{-\bar{\beta}_{1}\,x_{1}}}\,\mathrm{cosh}\left[m\,\left(x_{1}+\frac{\bar{\mathcal{E}}_{1}}{\bar{\gamma}}\right)\,\bar{\gamma}^{2}\,\bar{\Omega}\,\bar{T}_{1}^{2}\right]
+e[(x12/2)x1¯1/γ¯]γ¯2T¯12eβ¯1x11cosh[m(x1¯1γ¯)γ¯2Ω¯T¯12]}).\displaystyle+\,\frac{\mathrm{e}^{-\left[(x_{1}^{2}/2)-x_{1}\,\bar{\mathcal{E}}_{1}/\bar{\gamma}\right]\,\bar{\gamma}^{2}\,\bar{T}_{1}^{2}}}{\mathrm{e}^{\bar{\beta}_{1}\,x_{1}}-1}\mathrm{cosh}\left[m\left(x_{1}-\frac{\bar{\mathcal{E}}_{1}}{\bar{\gamma}}\right)\,\bar{\gamma}^{2}\,\bar{\Omega}\,\bar{T}_{1}^{2}\right]\biggr{\}}\Biggr{)}.

Similarly, upon using the expression (73), the corresponding cross transition transition probability rate R12T(¯1)R_{12}^{T}(\bar{\mathcal{E}}_{1}) can be written as

R12T(¯1)\displaystyle R_{12}^{T}(\bar{\mathcal{E}}_{1}) =\displaystyle= 2πT¯14πγ1e¯12T¯12/2[0dx1J0(x1v1)J0(x1v2/Ω¯)[e[(1+γ¯2)x12+2(1+γ¯)x1¯1]T¯12/41eβ¯1x1\displaystyle\frac{\sqrt{2\,\pi}\;\bar{T}_{1}}{4\,\pi\,\gamma_{1}}\,\mathrm{e}^{-\bar{\mathcal{E}}_{1}^{2}\,\bar{T}_{1}^{2}/2}\,\Biggl{[}\int_{0}^{\infty}{\rm d}x_{1}\,J_{0}(x_{1}\,v_{1})\,J_{0}(x_{1}\,v_{2}/\bar{\Omega})\,\biggl{[}\frac{\mathrm{e}^{-\left[(1+\bar{\gamma}^{2})\,x_{1}^{2}+2\,(1+\bar{\gamma})\,x_{1}\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}/4}}{1-\mathrm{e}^{-\bar{\beta}_{1}\,x_{1}}}\, (87)
+e[(1+γ¯2)x122(1+γ¯)x1¯1]T¯12/4eβ¯1x11]\displaystyle+\,\frac{\mathrm{e}^{-\left[(1+\bar{\gamma}^{2})\,x_{1}^{2}-2\,(1+\bar{\gamma})\,x_{1}\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}/4}}{\mathrm{e}^{\bar{\beta}_{1}\,x_{1}}-1}\biggr{]}
+ 2m=1em2(1+γ¯2Ω¯2)T¯12/40dx1Jm(x1v1)Jm(x1v2/Ω¯)\displaystyle+\,2\,\sum_{m=1}^{\infty}\,\mathrm{e}^{-m^{2}\,(1+\bar{\gamma}^{2}\,\bar{\Omega}^{2})\,\bar{T}_{1}^{2}/4}\,\int_{0}^{\infty}{\rm d}x_{1}\,J_{m}(x_{1}\,v_{1})\,J_{m}(x_{1}\,v_{2}/\bar{\Omega})
×(e[(1+γ¯2)x12+2(1+γ¯)x1¯1]T¯12/41eβ¯1x1cosh{m[(1+γ¯2Ω¯)x1+(1+γ¯Ω¯)¯1]T¯12/2}\displaystyle\times\,\bigg{(}\frac{\mathrm{e}^{-\left[(1+\bar{\gamma}^{2})\,x_{1}^{2}+2\,(1+\bar{\gamma})\,x_{1}\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}/4}}{1-\mathrm{e}^{-\bar{\beta}_{1}\,x_{1}}}\,\mathrm{cosh}\left\{m\,\left[(1+\bar{\gamma}^{2}\,\bar{\Omega})\,x_{1}+(1+\bar{\gamma}\,\bar{\Omega})\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}/2\right\}
+e[(1+γ¯2)x122(1+γ¯)x1¯1]T¯12/4eβ¯1x11cosh{m[(1+γ¯2Ω¯)x1(1+γ¯Ω¯)¯1]T¯12/2})]=R21T(¯1).\displaystyle+\,\frac{\mathrm{e}^{-\left[(1+\bar{\gamma}^{2})\,x_{1}^{2}-2\,(1+\bar{\gamma})\,x_{1}\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}/4}}{\mathrm{e}^{\bar{\beta}_{1}\,x_{1}}-1}\,\mathrm{cosh}\left\{m\,\left[(1+\bar{\gamma}^{2}\,\bar{\Omega})\,x_{1}-(1+\bar{\gamma}\,\bar{\Omega})\,\bar{\mathcal{E}}_{1}\right]\,\bar{T}_{1}^{2}/2\right\}\biggr{)}\Biggr{]}=R_{21}^{T}(\bar{\mathcal{E}}_{1}).

Appendix E A remedy for the infrared divergence

In our analysis, we had encountered an infrared divergence when calculating the Wightman function at a finite temperature. The occurrence of infrared divergences in Green’s functions in spacetime dimensions less than (3+1)(3+1) is not uncommon. We can turn to the calculation of the Green’s functions in (1+1)(1+1)-spacetime dimensions to identify possible remedies to regulate the divergence (in this context, see, for example, Ref. Chowdhury et al. (2019)). Note that, in the case of (2+1)(2+1)-spacetime dimensions, we encounter the divergence only when calculating the Green’s function at a finite temperature. (We should clarify that such a divergence does not arise in (3+1)(3+1)-spacetime dimensions.) Also, the divergence occurs only in the m=0m=0 term in the sum in Eq. (25). In the case wherein the detectors are switched on forever, the divergence in the Wightman function does not affect the transition probability rate of the detector, as the m=0m=0 term does not contribute [cf. Eq. (26)]. We should mention here that such a behavior has also been noticed earlier in a related work Barman et al. (2021b). However, when we consider detectors that are switched on for a finite duration, the m=0m=0 term in Eq. (40) leads to a non-zero contribution and we need to formally regulate the divergence. We need to do so in such a way that we recover the result in the limit of β¯\bar{\beta}\to\infty [viz. Eq. (37)], i.e. when the temperature of the thermal bath vanishes. Needless to add, we also need to reproduce our earlier result (26) at a finite temperature for the case of detectors that remain switched on forever (i.e in the limit T¯\bar{T}\to\infty).

Let us first single out the term containing the infrared divergence in the finite temperature Wightman function (25). Using the following identity (cf. Ref. Gradshteyn and Ryzhik (2007), 8.531.1):

J02(z)+2m=1Jm2(z)=1,J_{0}^{2}(z)+2\sum_{m=1}^{\infty}\,J_{m}^{2}(z)=1, (88)

we can express the Wightman function (25) as

Gβ+(u)=𝒜0(u)+𝒜1(u),\displaystyle G^{+}_{\beta}(u)=\mathcal{A}_{0}(u)+\mathcal{A}_{1}(u), (89)

where the quantities 𝒜0(u)\mathcal{A}_{0}(u) and 𝒜1(u)\mathcal{A}_{1}(u) are given by

𝒜0(u)\displaystyle\mathcal{A}_{0}(u) =\displaystyle= 0dq4π(eiγqu1eβq+eiγqueβq1),\displaystyle\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\left(\frac{\mathrm{e}^{-i\,\gamma\,q\,u}}{1-\mathrm{e}^{-\beta\,q}}+\frac{\mathrm{e}^{i\,\gamma\,q\,u}}{\mathrm{e}^{\beta\,q}-1}\right), (90a)
𝒜1(u)\displaystyle\mathcal{A}_{1}(u) =\displaystyle= 0dq4π{m=m0Jm2(qσ)\displaystyle\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\biggl{\{}\sum_{\begin{subarray}{c}m=-\infty\\ m\neq 0\end{subarray}}^{\infty}J^{2}_{m}(q\,\sigma) (90b)
×[eiγ(qmΩ)u1eβq+eiγ(qmΩ)ueβq1]\displaystyle\times\,\left[\frac{\mathrm{e}^{-i\,\gamma\,(q-m\,\Omega)\,u}}{1-\mathrm{e}^{-\beta\,q}}+\frac{\mathrm{e}^{i\,\gamma\,(q-m\,\Omega)\,u}}{\mathrm{e}^{\beta\,q}-1}\right]~{}
 2(eiγqu1eβq+eiγqueβq1)\displaystyle-\,2\,\left(\frac{\mathrm{e}^{-i\,\gamma\,q\,u}}{1-\mathrm{e}^{-\beta\,q}}+\frac{\mathrm{e}^{i\,\gamma\,q\,u}}{\mathrm{e}^{\beta\,q}-1}\right)
×m=1Jm2(qσ)}.\displaystyle\times\,\sum_{m=1}^{\infty}J^{2}_{m}(q\,\sigma)\biggr{\}}.

Therefore, the transition probability rate of a detector switched on for a finite time through the Gaussian window function can be expressed as

RT()\displaystyle R_{T}(\mathcal{E}) =\displaystyle= RT0()+RT1(),\displaystyle R_{T}^{0}(\mathcal{E})+R_{T}^{1}(\mathcal{E})~{}, (91)

where, evidently, RT0,1()R_{T}^{0,1}(\mathcal{E}) are given by

RT0,1()=dueiu𝒜0,1(u)eu2/(2T2).R_{T}^{0,1}(\mathcal{E})=\int_{-\infty}^{\infty}{\rm d}u\,\mathrm{e}^{-i\,\mathcal{E}\,u}\,\mathcal{A}_{0,1}(u)\,\mathrm{e}^{-u^{2}/(2\,T^{2})}. (92)

Since the term 𝒜1(u)\mathcal{A}_{1}(u) in the finite temperature Wightman function does not contain the infrared divergence, the corresponding transition probability rate RT1()R_{T}^{1}(\mathcal{E}) can be evaluated as we have done earlier in the other cases. Therefore, let us turn to the calculation of the rate RT0()R_{T}^{0}(\mathcal{E}) that depends on the term 𝒜0(u)\mathcal{A}_{0}(u). To do so, let us first explicitly evaluate 𝒜0(u)\mathcal{A}_{0}(u). As is often done in the case of (1+1)(1+1)-spacetime dimensions, in order to avoid the divergence, we shall take the derivative of 𝒜0(u)\mathcal{A}_{0}(u) with respect to the variable uu, thus rendering it safe from the infrared divergence. We can then evaluate the integral over qq as usual, by introducing an ultraviolet regulator of the form of eϵq\mathrm{e}^{-\epsilon\,q} to obtain that

𝒜0(u)u\displaystyle\frac{\partial\mathcal{A}_{0}(u)}{\partial u} =\displaystyle= iγ4πβ2[ψ(1)(iγu+ϵβ)\displaystyle-\frac{i\,\gamma}{4\,\pi\,\beta^{2}}\,\biggl{[}\psi^{(1)}\left(\frac{i\,\gamma\,u+\epsilon}{\beta}\right) (93)
ψ(1)(iγu+β+ϵβ)],\displaystyle-\,\psi^{(1)}\left(\frac{-i\,\gamma\,u+\beta+\epsilon}{\beta}\right)\biggr{]},

where ψ(n)(z)\psi^{(n)}(z) denotes the polygamma function of order nn. Upon integrating over uu, we arrive at the expression

𝒜0(u)\displaystyle\mathcal{A}_{0}(u) =\displaystyle= 14πβ[ψ(0)(iγu+ϵβ)\displaystyle-\frac{1}{4\pi\beta}\,\biggl{[}\psi^{(0)}\left(\frac{i\,\gamma\,u+\epsilon}{\beta}\right) (94)
+ψ(0)(iγu+β+ϵβ)].\displaystyle+\,\psi^{(0)}\left(\frac{-i\,\gamma\,u+\beta+\epsilon}{\beta}\right)\biggr{]}.

In the limit of zero temperature (i.e. as β\beta\to\infty), this expression reduces to

limβ𝒜0(u)\displaystyle\lim_{\beta\to\infty}\mathcal{A}_{0}(u) =\displaystyle= i4π1γuiϵ\displaystyle-\frac{i}{4\,\pi}\,\frac{1}{\gamma\,u-i\,\epsilon} (95)

which is the result we would have obtained had we taken the limit β\beta\to\infty in Eq. (90a) (i.e. before taking the derivative and carrying out the integration with respect to variable uu).

We can now evaluate the transition probability rate RT0()R_{T}^{0}(\mathcal{E}) of the detector using the expression (94) for 𝒜0(u)\mathcal{A}_{0}(u). To do so, let us define the dimensionless variable u~=γΩu\tilde{u}=\gamma\,\Omega\,u and parameter ϵ¯=ϵΩ\bar{\epsilon}=\epsilon\,\Omega. Let us also make use of the Fourier transform

eu~2/(2T¯2)=T¯2πdξeiu~ξξ2T¯2/2\mathrm{e}^{-\tilde{u}^{2}/(2\,\bar{T}^{2})}=\frac{\bar{T}}{\sqrt{2\,\pi}}\,\int_{-\infty}^{\infty}{\rm d}\xi\,\mathrm{e}^{i\,\tilde{u}\,\xi-\xi^{2}\,\bar{T}^{2}/2} (96)

and the following series expansion of the polygamma function:

ψ(0)(z)=k=01(z+k).\psi^{(0)}(z)=-\sum_{k=0}^{\infty}\frac{1}{(z+k)}. (97)

From this last expression, it would be clear that the first and the second polygamma functions in Eq. (94) would have poles at u~=i(kβ¯+ϵ¯)=ip1(k)\tilde{u}=i\,(k\,\bar{\beta}+\bar{\epsilon})=i\,p_{1}(k) (i.e. in the upper half of the complex u~\tilde{u}-plane) and at u~=i[(k+1)β¯+ϵ¯]=ip2(k)\tilde{u}=-i\,[(k+1)\,\bar{\beta}+\bar{\epsilon}]=-i\,p_{2}(k) (i.e. in the lower half of the complex u~\tilde{u}-plane), respectively. The transition probability rate RT0()R_{T}^{0}(\mathcal{E}) can be written as

RT0()\displaystyle R_{T}^{0}(\mathcal{E}) =\displaystyle= T¯4πiγ2πk=0dξeξ2T¯2/2du~ei(ξ¯)u~[1u~ip1(k)1u~+ip2(k)].\displaystyle\frac{\bar{T}}{4\,\pi\,i\,\,\gamma\,\sqrt{2\,\pi}}\,\sum_{k=0}^{\infty}\int_{-\infty}^{\infty}{\rm d}\xi\,\mathrm{e}^{-\xi^{2}\,\bar{T}^{2}/2}\int_{-\infty}^{\infty}{\rm d}\tilde{u}\,\mathrm{e}^{i\,(\xi-\bar{\mathcal{E}})\,\tilde{u}}\,\left[\frac{1}{\tilde{u}-i\,p_{1}(k)}-\frac{1}{\tilde{u}+i\,p_{2}(k)}\right]. (98)

After carrying out the integration over u~\tilde{u} and imposing the appropriate conditions for non-vanishing residues (such as when ξ>¯\xi>\bar{\mathcal{E}} and ξ<¯\xi<\bar{\mathcal{E}} for the first and the second terms in the square brackets), we obtain that

RT0()=T¯2γ2πk=0[¯dξeξ2T¯2/2e(ξ¯)p1(k)+¯dξeξ2T¯2/2e(ξ¯)p2(k)].R_{T}^{0}(\mathcal{E})=\frac{\bar{T}}{2\,\gamma\sqrt{2\,\pi}}\,\sum_{k=0}^{\infty}\biggl{[}\int_{\bar{\mathcal{E}}}^{\infty}~{}{\rm d}\xi\,\mathrm{e}^{-\xi^{2}\,\bar{T}^{2}/2}\,\mathrm{e}^{-(\xi-\bar{\mathcal{E}})\,p_{1}(k)}+\int_{-\infty}^{\bar{\mathcal{E}}}{\rm d}\xi\,\mathrm{e}^{-\xi^{2}\,\bar{T}^{2}/2}\,\mathrm{e}^{(\xi-\bar{\mathcal{E}})\,p_{2}(k)}\biggr{]}. (99)

On further calculating the integral over ξ\xi and then taking the limit ϵ¯0\bar{\epsilon}\to 0, we arrive at the expression

RT0()\displaystyle R_{T}^{0}(\mathcal{E}) =\displaystyle= 14γk=0{eβ¯k(2¯T¯2+β¯k)/(2T¯2)rfc[(β¯k+¯T¯2)/(2T¯)]\displaystyle\frac{1}{4\,\gamma}\sum_{k=0}^{\infty}\biggl{\{}\mathrm{e}^{\bar{\beta}\,k\,\left(2\,\bar{\mathcal{E}}\,\bar{T}^{2}+\bar{\beta}\,k\right)/(2\,\bar{T}^{2})}\,\mathcal{E}\text{rfc}\left[(\bar{\beta}\,k+\bar{\mathcal{E}}\,\bar{T}^{2})/(\sqrt{2}\,\bar{T})\right] (100)
+eβ¯(k+1)(β¯+β¯k2¯T¯2)/(2T¯2)rfc[(β¯+β¯k¯T¯2)/(2T¯)]}.\displaystyle+\,\mathrm{e}^{\bar{\beta}\,(k+1)\left(\bar{\beta}+\bar{\beta}\,k-2\,\bar{\mathcal{E}}\,\bar{T}^{2}\right)/(2\,\bar{T}^{2})}\,\mathcal{E}\text{rfc}\left[(\bar{\beta}+\bar{\beta}\,k-\bar{\mathcal{E}}\,\bar{T}^{2})/(\sqrt{2}\,\bar{T})\right]\biggr{\}}.

We find that, when the temperature of the thermal bath vanishes (i.e. when β¯\bar{\beta}\to\infty), in Eq. (99), the quantity p2(k)p_{2}(k)\to\infty for all accessible values of kk. In the same limit, we have p1(k)=ϵ¯p_{1}(k)=\bar{\epsilon} for k=0k=0, whereas, for the all other values of kk, we have p1(k)p_{1}(k)\to\infty. Therefore, when β\beta\to\infty, in Eq. (99), we are left with only one term of the sum, and it can be expressed as

limβϵ¯0RT0()\displaystyle\lim_{\begin{subarray}{c}\beta\to\infty\\ \bar{\epsilon}\to 0\end{subarray}}R_{T}^{0}(\mathcal{E}) =\displaystyle= T¯2γ2π¯dξeξ2T¯2/2\displaystyle\frac{\bar{T}}{2\,\gamma\,\sqrt{2\,\pi}}\,\int_{\bar{\mathcal{E}}}^{\infty}{\rm d}\xi\,\mathrm{e}^{-\xi^{2}\,\bar{T}^{2}/2}\, (101)
=\displaystyle= 14γrfc(¯T¯2).\displaystyle\frac{1}{4\,\gamma}\,\mathcal{E}\text{rfc}\left(\frac{\bar{\mathcal{E}}\,\bar{T}}{\sqrt{2}}\right).

In a similar manner, in the β¯\bar{\beta}\to\infty limit, the expression (100) is non-zero only when k=0k=0. Also, in this limit, the result reduces to same expression as in Eq. (101).

Furthermore, we can take the limit of infinite interaction time, i.e. T¯\bar{T}\to\infty) in Eq. (100) and observe that the quantity RT0()R_{T}^{0}(\mathcal{E}) reduces to

RT0()\displaystyle R_{T}^{0}(\mathcal{E}) =\displaystyle= 12γk=0e(k+1)β¯¯=12γ1eβ¯¯1.\displaystyle\frac{1}{2\,\gamma}\sum_{k=0}^{\infty}\mathrm{e}^{-(k+1)\bar{\beta}\,\bar{\mathcal{E}}}=\frac{1}{2\,\gamma}\frac{1}{\mathrm{e}^{\bar{\beta}\,\bar{\mathcal{E}}}-1}.\quad (102)

The same quantity can also be obtained from the second sum of Eq. (26) with m=0m=0 and by utilizing the relation J02(z)=12m=1Jm2(z)J_{0}^{2}(z)=1-2\sum_{m=1}^{\infty}\,J_{m}^{2}(z). We also observe that in the zero temperature limit and for infinite interaction time, i.e. when T¯\bar{T}\to\infty, the quantity RT0()R_{T}^{0}(\mathcal{E}) vanishes, which is evident from Eq. (101). (For a different approach to handle this infrared divergence, we would refer the reader to Ref. Bunney and Louko (2023).)

Let us now provide a similar regularization procedure for the thermal Green’s function in Eqs. (54) and (55) that, in general, connect two detector events. In the same manner as in Eq. (89), we can express the Green’s function (55) as

Gβjl+(u¯,v¯)=𝒜¯0jl(u¯,v¯)+𝒜¯1jl(u¯,v¯).\displaystyle G^{+}_{\beta_{jl}}(\bar{u},\bar{v})=\bar{\mathcal{A}}_{0{jl}}(\bar{u},\bar{v})+\bar{\mathcal{A}}_{1{jl}}(\bar{u},\bar{v}). (103)

As earlier, in the quantity 𝒜0¯(u¯,v¯)\bar{\mathcal{A}_{0}}(\bar{u},\bar{v}), we have singled out the contribution containing the infrared divergence. The quantity 𝒜¯1(u¯,v¯)\bar{\mathcal{A}}_{1}(\bar{u},\bar{v}) contains all the other contributions and it does not diverge in the infrared limit. These quantities are given by

𝒜¯0jl(u¯,v¯)\displaystyle\bar{\mathcal{A}}_{0{jl}}(\bar{u},\bar{v}) =\displaystyle= 0dq4π[ei[α¯1(q)v¯+α¯2(q)u¯]/21eβq+ei[α¯1(q)v¯+α¯2(q)u¯]/2eβq1],\displaystyle\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\left[\frac{\mathrm{e}^{-i\,\left[\bar{\alpha}_{1}(q)\,\bar{v}+\bar{\alpha}_{2}(q)\,\bar{u}\right]/2}}{1-\mathrm{e}^{-\beta\,q}}+\frac{\mathrm{e}^{i\,\left[\bar{\alpha}_{1}(q)\,\bar{v}+\bar{\alpha}_{2}(q)\,\bar{u}\right]/2}}{\mathrm{e}^{\beta\,q}-1}\right]~{},
𝒜¯1jl(u¯,v¯)\displaystyle\bar{\mathcal{A}}_{1{jl}}(\bar{u},\bar{v}) =\displaystyle= 0dq4π{m=m0Jm(qσj)Jm(qσl)[ei[α1(q)v¯+α2(q)u¯]/21eβq+ei[α1(q)v¯+α2(q)u¯]/2eβq1]\displaystyle\int_{0}^{\infty}\frac{{\rm d}q}{4\,\pi}\,\Bigg{\{}\sum_{\begin{subarray}{c}m=-\infty\\ m\neq 0\end{subarray}}^{\infty}J_{m}(q\,\sigma_{j})\,J_{m}(q\,\sigma_{l})\,\left[\frac{\mathrm{e}^{-i\,\left[\alpha_{1}(q)\,\bar{v}+\alpha_{2}(q)\,\bar{u}\right]/2}}{1-\mathrm{e}^{-\beta\,q}}+\frac{\mathrm{e}^{i\,\left[\alpha_{1}(q)\,\bar{v}+\alpha_{2}(q)\,\bar{u}\right]/2}}{\mathrm{e}^{\beta\,q}-1}\right]~{} (104)
[ei[α¯1(q)v¯+α¯2(q)u¯]/21eβq+ei[α¯1(q)v¯+α¯2(q)u¯]/2eβq1][1J0(qσj)J0(qσl)]},\displaystyle-\,\left[\frac{\mathrm{e}^{-i\,\left[\bar{\alpha}_{1}(q)\,\bar{v}+\bar{\alpha}_{2}(q)\,\bar{u}\right]/2}}{1-\mathrm{e}^{-\beta\,q}}+\frac{\mathrm{e}^{i\,\left[\bar{\alpha}_{1}(q)\,\bar{v}+\bar{\alpha}_{2}(q)\,\bar{u}\right]/2}}{\mathrm{e}^{\beta\,q}-1}\right]\left[1-J_{0}(q\,\sigma_{j})\,J_{0}(q\,\sigma_{l})\right]\Bigg{\}},

where α¯1(q)=q(γjγl)\bar{\alpha}_{1}(q)=q\,(\gamma_{j}-\gamma_{l}) and α¯2(q)=q(γj+γl)\bar{\alpha}_{2}(q)=q\,(\gamma_{j}+\gamma_{l}), i.e. they correspond to α1(q)\alpha_{1}(q) and α2(q)\alpha_{2}(q) when m=0m=0. We can define the corresponding transition probability rates as

RjlT()\displaystyle R_{jl}^{T}(\mathcal{E}) =\displaystyle= R0jlT()+R1jlT(),\displaystyle R^{T}_{0{jl}}(\mathcal{E})+R^{T}_{1{jl}}(\mathcal{E})~{}, (105)

where R0jlT()R^{T}_{0{jl}}(\mathcal{E}) depends exclusively on 𝒜¯0jl(u¯,v¯)\bar{\mathcal{A}}_{0{jl}}(\bar{u},\bar{v}) and R1jlT()R^{T}_{1{jl}}(\mathcal{E}) on 𝒜¯1jl(u¯,v¯)\bar{\mathcal{A}}_{1_{jl}}(\bar{u},\bar{v}). We can evaluate the contribution R1jlT()R^{T}_{1{jl}}(\mathcal{E}) using the same procedure we had adopted in Sec. IV.2.2. Therefore, we shall now focus only the evaluation of R0jlT()R^{T}_{0{jl}}(\mathcal{E}). In particular, one can consider general γj\gamma_{j} and γl\gamma_{l} for the evaluation of 𝒜¯0jl(u¯,v¯)\bar{\mathcal{A}}_{0{jl}}(\bar{u},\bar{v}). However, recall that, we had considered the same velocities for the two different detectors, with different radial distances and angular velocities, to estimate total transition probabilities (see Figs. 7 and 9). Therefore, for simplicity, we shall set γj=γl=γ\gamma_{j}=\gamma_{l}=\gamma, so that 𝒜¯0jl(u¯,v¯)=𝒜0(u¯)\bar{\mathcal{A}}_{0{jl}}(\bar{u},\bar{v})=\mathcal{A}_{0}(\bar{u}) [cf. Eq. (90a)]. In such a case, the expression R0jlT()R^{T}_{0{jl}}(\mathcal{E}) will be given exactly by Eq. (100) for all jj and ll. We can use this result to plot the different transition probability rates as in Figs. 7 and 9.

References

  • Fulling (1973) S. A. Fulling, Phys.Rev. D7, 2850 (1973).
  • Birrell and Davies (1984a) N. D. Birrell and P. C. W. Davies, Quantum Fields in Curved Space (Cambridge University Press, Cambridge, England, 1984a).
  • Fulling (1989) S. A. Fulling, Aspects of Quantum Field Theory in Curved Spacetime (Cambridge University Press, Cambridge, England, 1989).
  • Mukhanov and Winitzki (2007) V. Mukhanov and S. Winitzki, Introduction to Quantum Effects in Gravity (Cambridge University Press, Cambridge, England, 2007).
  • Parker and Toms (2009) L. E. Parker and D. Toms, Quantum Field Theory in Curved Spacetime: Quantized Field and Gravity (Cambridge University Press, Cambridge, England, 2009).
  • Letaw (1981) J. R. Letaw, Phys. Rev. D 23, 1709 (1981).
  • Letaw and Pfautsch (1981) J. R. Letaw and J. D. Pfautsch, Phys. Rev. D 24, 1491 (1981).
  • Sriramkumar and Padmanabhan (2002) L. Sriramkumar and T. Padmanabhan, Int. J. Mod. Phys. D 11, 1 (2002), eprint arXiv:gr-qc/9903054.
  • Unruh (1976) W. Unruh, Phys.Rev. D14, 870 (1976).
  • DeWitt (1980) B. S. DeWitt, QUANTUM GRAVITY: THE NEW SYNTHESIS (1980), pp. 680–745.
  • Crispino et al. (2008) L. C. Crispino, A. Higuchi, and G. E. Matsas, Rev.Mod.Phys. 80, 787 (2008), eprint arXiv:0710.5373.
  • Bell and Leinaas (1983) J. S. Bell and J. M. Leinaas, Nucl. Phys. B 212, 131 (1983).
  • Svaiter and Svaiter (1992) B. F. Svaiter and N. F. Svaiter, Phys. Rev. D 46, 5267 (1992), [Erratum: Phys.Rev.D 47, 4802 (1993)].
  • Higuchi et al. (1993) A. Higuchi, G. E. A. Matsas, and C. B. Peres, Phys. Rev. D 48, 3731 (1993).
  • Sriramkumar and Padmanabhan (1996) L. Sriramkumar and T. Padmanabhan, Class. Quant. Grav. 13, 2061 (1996), eprint arXiv:gr-qc/9408037.
  • Davies et al. (1996) P. C. W. Davies, T. Dray, and C. A. Manogue, Phys. Rev. D 53, 4382 (1996), eprint arXiv:gr-qc/9601034.
  • Korsbakken and Leinaas (2004) J. I. Korsbakken and J. M. Leinaas, Phys. Rev. D 70, 084016 (2004), eprint arXiv:hep-th/0406080.
  • Gutti et al. (2011) S. Gutti, S. Kulkarni, and L. Sriramkumar, Phys. Rev. D 83, 064011 (2011), eprint arXiv:1005.1807.
  • Jaffino Stargen et al. (2017) D. Jaffino Stargen, N. Kajuri, and L. Sriramkumar, Phys. Rev. D 96, 066002 (2017), eprint arXiv:1706.05834.
  • Louko and Upton (2018) J. Louko and S. D. Upton, Phys. Rev. D 97, 025008 (2018), eprint arXiv:1710.06954.
  • Picanço et al. (2020) G. Picanço, N. F. Svaiter, and C. A. Zarro, JHEP 08, 025 (2020), eprint arXiv:2002.06085.
  • Biermann et al. (2020) S. Biermann, S. Erne, C. Gooding, J. Louko, J. Schmiedmayer, W. G. Unruh, and S. Weinfurtner, Phys. Rev. D 102, 085006 (2020), eprint arXiv:2007.09523.
  • Sriramkumar (2017) L. Sriramkumar, Fundam. Theor. Phys. 187, 451 (2017), eprint arXiv:1612.08579.
  • Schutzhold et al. (2006) R. Schutzhold, G. Schaller, and D. Habs, Phys.Rev.Lett. 97, 121302 (2006), eprint arXiv:quant-ph/0604065.
  • Sudhir et al. (2021) V. Sudhir, N. Stritzelberger, and A. Kempf, Phys. Rev. D 103, 105023 (2021), eprint arXiv:2102.03367.
  • Sánchez-Kuntz et al. (2022) N. Sánchez-Kuntz, A. Parra-López, M. Tolosa-Simeón, T. Haas, and S. Floerchinger, Phys. Rev. D 105, 105020 (2022), eprint arXiv:2202.10440.
  • Hu and Yu (2015) J. Hu and H. Yu, Phys. Rev. A 91, 012327 (2015), eprint arXiv:1501.03321.
  • Menezes and Svaiter (2015) G. Menezes and N. Svaiter, Phys. Rev. A 92, 062131 (2015), eprint arXiv:1508.04513.
  • Arias et al. (2016) E. Arias, J. Dueñas, G. Menezes, and N. Svaiter, JHEP 07, 147 (2016), eprint arXiv:1510.00047.
  • Flores-Hidalgo et al. (2015) G. Flores-Hidalgo, M. Rojas, and O. Rojas (2015), eprint arXiv:1511.01416.
  • Menezes and Svaiter (2016) G. Menezes and N. Svaiter, Phys. Rev. A 93, 052117 (2016), eprint arXiv:1512.02886.
  • Menezes (2016) G. Menezes, Phys. Rev. D94, 105008 (2016), eprint arXiv:1512.03636.
  • Rizzuto et al. (2016) L. Rizzuto, M. Lattuca, J. Marino, A. Noto, S. Spagnolo, R. Passante, and W. Zhou, Phys. Rev. A 94, 012121 (2016), eprint arXiv:1601.04502.
  • Rodríguez-Camargo et al. (2018) C. Rodríguez-Camargo, N. Svaiter, and G. Menezes, Annals Phys. 396, 266 (2018), eprint arXiv:1608.03365.
  • Zhou et al. (2016) W. Zhou, R. Passante, and L. Rizzuto, Phys. Rev. D 94, 105025 (2016), eprint arXiv:1609.06931.
  • Cai and Ren (2018) H. Cai and Z. Ren, Class. Quant. Grav. 35, 025016 (2018).
  • Liu et al. (2018) X. Liu, Z. Tian, J. Wang, and J. Jing, Phys. Rev. D 97, 105030 (2018), eprint arXiv:1805.04470.
  • Zhou and Yu (2020) W. Zhou and H. Yu, Phys. Rev. D 101, 025009 (2020), eprint arXiv:2001.00750.
  • Barman and Majhi (2021) S. Barman and B. R. Majhi, JHEP 03, 245 (2021), eprint arXiv:2101.08186.
  • Zhang and Yu (2020) J. Zhang and H. Yu, Phys. Rev. D 102, 065013 (2020), eprint arXiv:2008.07980.
  • Louko and Satz (2006) J. Louko and A. Satz, Class. Quant. Grav. 23, 6321 (2006), eprint arXiv:gr-qc/0606067.
  • Satz (2007) A. Satz, Class. Quant. Grav. 24, 1719 (2007), eprint arXiv:gr-qc/0611067.
  • Brown et al. (2013) E. G. Brown, E. Martin-Martinez, N. C. Menicucci, and R. B. Mann, Phys. Rev. D 87, 084062 (2013), eprint arXiv:1212.1973.
  • Dicke (1954) R. H. Dicke, Phys. Rev. 93, 99 (1954).
  • Kolekar and Padmanabhan (2015) S. Kolekar and T. Padmanabhan, Class. Quant. Grav. 32, 202001 (2015), eprint arXiv:1308.6289.
  • Kolekar (2014) S. Kolekar, Phys. Rev. D 89, 044036 (2014), eprint arXiv:1309.3261.
  • Kolekar and Padmanabhan (2014) S. Kolekar and T. Padmanabhan, Phys. Rev. D 89, 064055 (2014), eprint arXiv:1309.4424.
  • Watson (2015) G. Watson, A Treatise on the Theory of Bessel Functions - Scholar’s Choice Edition (Creative Media Partners, LLC, 2015), ISBN 9781294994701.
  • (49) F. W. J. Olver, The asymptotic expansion of bessel functions of large order. Philosophical Transactions of the Royal Society of London. Series A, Mathematical and Physical Sciences 247: 328–368.
  • Ficek and Tanaś (2002) Z. Ficek and R. Tanaś, Physics Reports 372, 369 (2002), ISSN 0370-1573.
  • Reznik (2003) B. Reznik, Found. Phys. 33, 167 (2003), eprint arXiv:quant-ph/0212044.
  • Martin-Martinez et al. (2016) E. Martin-Martinez, A. R. H. Smith, and D. R. Terno, Phys. Rev. D 93, 044001 (2016), eprint arXiv:1507.02688.
  • Koga et al. (2018) J.-I. Koga, G. Kimura, and K. Maeda, Phys. Rev. A 97, 062338 (2018), eprint arXiv:1804.01183.
  • Koga et al. (2019) J.-i. Koga, K. Maeda, and G. Kimura, Phys. Rev. D 100, 065013 (2019), eprint arXiv:1906.02843.
  • Barman et al. (2021a) D. Barman, S. Barman, and B. R. Majhi, JHEP 07, 124 (2021a), eprint arXiv:2104.11269.
  • Barman et al. (2021b) S. Barman, D. Barman, and B. R. Majhi (2021b), eprint arXiv:2112.01308.
  • Brown (2013) E. G. Brown, Phys. Rev. A 88, 062336 (2013), eprint arXiv:1309.1425.
  • Simidzija and Martín-Martínez (2018) P. Simidzija and E. Martín-Martínez, Phys. Rev. D 98, 085007 (2018), eprint arXiv:1809.05547.
  • Birrell and Davies (1984b) N. D. Birrell and P. C. W. Davies, Quantum fields in curved space, Cambridge Monographs on Mathematical Physics (Cambridge University Press, 1984b).
  • Arfken (1985) G. Arfken, Mathematical Methods for Physicists (Academic Press, Inc., San Diego, 1985), 3rd ed.
  • Chowdhury et al. (2019) C. Chowdhury, S. Das, S. Dalui, and B. R. Majhi, Phys. Rev. D 99, 045021 (2019), eprint arXiv:1902.06900.
  • Gradshteyn and Ryzhik (2007) I. S. Gradshteyn and I. M. Ryzhik, Table of integrals, series, and products (Elsevier/Academic Press, Amsterdam, 2007), seventh ed., translation edited and with a preface by Alan Jeffrey and Daniel Zwillinger.
  • Bunney and Louko (2023) C. R. D. Bunney and J. Louko, Class. Quant. Grav. 40, 155001 (2023), eprint arXiv:2303.12690.