Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Time reflection of light from a quantum perspective and vacuum entanglement

Open Access Open Access

Abstract

If a boundary between two static media is moving with a constant superluminal velocity, or there is a sudden change of the refractive index with time, this yields generation of entangled pairs of photons out of vacuum propagating in the opposite directions. Here we show that during this process, entanglement of Minkowski vacuum is transferred to the entanglement of the generated photon pairs. If initially an electromagnetic pulse is present in the medium the photon generation is stimulated into the pulse mode, and since photons are created as entangled pairs the counter-propagating photon partners produce a pulse moving in the opposite direction, which is known as time reflection. Thus, time reflection occurs due to stimulated generation of the entangled photon pairs out of entangled vacuum and no photons in the original pulse are in fact being reflected. This is different from the mechanism of light reflection from spatial inhomogeneities for which no photons are generated.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Interaction of a light pulse with an atom is determined by the frequency content of the pulse in the atom’s reference frame. If one of the frequencies coincides with the frequency of the atomic transition the atom can become excited by absorbing a photon. A positive-frequency pulse can have negative-frequency content in the frame of atoms moving along certain worldlines leading to emission of a photon by ground-state atoms together with atom’s excitation. This is, e.g., the case if the atom is accelerated through an empty space which yields Unruh acceleration radiation [13].

Negative frequency photons describe field in a medium in a moving frame which leads to Cherenkov radiation of neutral atoms which is a quantum effect. The latter can be interpreted as emission of the negative frequency photons by a uniformly moving atom [46]. For description of the Cherenkov effect it is convenient to choose mode functions of the field as plane waves which in the medium with refractive index $n$ read (in the lab frame)

$$\phi _{\mathbf{k}}(t,\mathbf{r})=e^{{-}i\frac{ck}{n}t+i\mathbf{kr}},$$
where $\mathbf {k}$ is the photon wave vector. In the frame moving with velocity $\mathbf {V}$ the photon frequency is Doppler shifted and becomes
$$\nu =\frac{\frac{ck}{n}-\mathbf{V}\cdot \mathbf{k}}{\sqrt{1-\frac{V^{2}}{ c^{2}}}}.$$

In this frame, if $V>c/n$, photons propagating inside the Cherenkov cone $\cos \theta >c/Vn$, where $\theta$ is the angle between $\mathbf {V}$ and $\mathbf {k}$, have negative frequency [7]. This leads to Cherenkov radiation. Namely, an atom which is at rest in the moving frame can become excited by emitting a negative-frequency photon inside the Cherenkov cone. In this process the total energy is conserved because the emitted photon has negative energy in the atom’s frame. Other ground-state atoms which are at rest in the moving frame can not absorb the emitted negative-energy photon, but can emit identical negative-energy photons by the stimulated emission mechanism [6]. Such ensemble of ground-state atoms acts as an inverted medium in a laser which yields light amplification.

An excited atom decays back to the ground state by spontaneous emission of a positive-frequency photon outside the Cherenkov cone. This yields generation of entangled photon pairs, one of the photons in the pair is emitted inside the Cherenkov cone and the other one is emitted outside the cone [8].

Negative frequency is associated with many other physical effects. For example, quantum friction between relatively moving dielectrics stems from the mixing of positive and negative frequency waves in the two materials [9,10]. The behavior of the negative-energy quanta is essential to understanding the kinematics of amplification of waves which can explain the physics of traveling-wave-tube-type amplifiers [11], a resistive-wall amplifier [12], and amplification of ultrasound in semiconductors [13]. A ground-state atom moving above a metal surface can become excited by emitting a surface plasmon which, from the atom’s perspective, has negative frequency [14]. The role of negative frequencies in nonlinear fiber optics was discussed in [15] .

Frequency content of a pulse can change, e.g., due to interaction with an accelerating mirror [16] or due to a sudden change of the medium properties with time. These can lead to appearance of negative frequencies in the pulse spectrum in the lab frame and, as a consequence, to it’s unusual interaction with atoms. For example, a fixed ground-state atom cannot become excited by absorbing a negative-frequency photon because this process would violate energy conservation. The atom can, however, become excited by emitting the negative-frequency photon and this process is stimulated by the photons present in the field.

By changing the basis set of the field modes one can describe the same state of the field in a picture in which photons have only positive frequencies and, thus, interact with atoms in a usual fashion. In this representation the number of photons in the pulse would be different, implying particle creation or annihilation. In particular, in the new representation, a vacuum state is filled with photons. Thus, if a physical process yields appearance of negative frequencies in the pulse spectrum this means generation of particles in a conventional description of the field in terms of the positive-frequency photons.

Let us consider a massless scalar field that obeys the one-dimensional wave equation in Minkowski spacetime

$$\frac{1}{c^{2}}\frac{\partial ^{2}\phi }{\partial t^{2}}-\frac{\partial ^{2}\phi }{\partial x^{2}}=0.$$

The inner product of two field modes $\phi _{1}(t,x)$ and $\phi _{2}(t,x)$ obeying Eq. (1) is defined as the Klein–Gordon inner product which is a generalization of the Wronksian [16]

$$\left\langle \phi _{1},\phi _{2}\right\rangle =\frac{i}{2c}\int_{-\infty }^{\infty }\left( \phi _{1}^{{\ast} }\frac{\partial \phi _{2}}{\partial t}- \frac{\partial \phi _{1}^{{\ast} }}{\partial t}\phi _{2}\right) dx.$$

The inner product is independent of time and has the following properties

$$\left\langle \phi _{1}^{{\ast} },\phi _{1}^{{\ast} }\right\rangle ={-}\left\langle \phi _{1},\phi _{1}\right\rangle ,\quad \left\langle \phi _{1},\phi _{1}^{{\ast} }\right\rangle =\left\langle \phi _{1}^{{\ast} },\phi _{1}\right\rangle =0.$$

For example, for a plane-wave mode

$$\phi (t,x)=e^{{-}i\nu t+ikx}$$
the norm $\left \langle \phi,\phi \right \rangle \propto \nu$ and, thus, sign of the norm is determined by the sign of the plane-wave frequency $\nu$. Equation (4) shows that the change $\nu \rightarrow -\nu$ is equivalent to the time reversal transformation $t\rightarrow -t$, that is a negative frequency mode can be viewed as a positive frequency mode propagating backward in time. The classical Eq. (1) is, however, symmetric under the time reversal and the mode function $\phi (-t,x)$ also describes a physical solution.

The situation is different in quantum description because Schrodinger equation is not symmetric under time reversal. For simplicity, let us consider a single-mode field with a mode function $\phi (t,x)$ which has positive norm, $\left \langle \phi,\phi \right \rangle >0$. In quantum description, the positive norm modes are associated with the photon annihilation operators $\hat {a}$. The complex conjugate $\phi ^{\ast }(t,x)$ is a negative-norm mode corresponds to the photon creation operators $\hat {a} ^{\dagger }$. Expansion of the field operator $\hat {\Phi }$ in terms of the modes $\phi$ and $\phi ^{\ast }$ reads

$$\hat{\Phi}(t,x)=\phi (t,x)\hat{a}+\phi ^{{\ast} }(t,x)\hat{a}^{{\dagger} }.$$

The field quantization procedure breaks the time reversal symmetry of the Eq. (1). Namely, if we make transformation $t\rightarrow -t$ in Eq. (5), the transformed field operator

$$\hat{\Phi}({-}t,x)=\phi ({-}t,x)\hat{a}+\phi ^{{\ast} }({-}t,x)\hat{a}^{{\dagger} },$$
in which $\hat {a}$ is the annihilation operator for the photon in the mode $\phi (-t,x),$ is forbidden. This is the case because $\phi (-t,x)$ has the opposite sign of the norm and, hence, according to the quantization procedure, the mode function $\phi (-t,x)$ must be associated with the photon creation operator. The operator (6) describes a pulse propagating backward in time. Thus, quantum mechanics constrains the backward in time evolution.

However, a positive-norm mode function $\phi (t,x)$ can have negative frequency components. For example, a superposition of the right and left-moving Gaussian pulses with opposite carrier frequencies $\nu >0$ and $-\nu$

$$\phi (t,x)=e^{{-}i\nu \left( t-x/c\right) }e^{-\Omega ^{2}(t-x/c)^{2}}+Ae^{i\nu \left( t+x/c\right) }e^{-\Omega ^{2}(t+x/c)^{2}}$$
has a positive norm if $|A|<1$. The last term in Eq. (7) describes a pulse propagating backward in time (the pulse has negative carrier frequency). Such a pulse can, e.g., be generated if propagation of light through a uniform medium is accompanied by an abrupt change of the refractive index. This is known as time reflection.

Classical Maxwell’s equations which do not take into account creation of photons out of vacuum yield that during time reflection a pulse undergoes a frequency shift and splits into two pulses propagating in the opposite directions [17]. The pulse propagating in the direction opposite to that of the incident wave has negative frequency. The time boundary problem conserves the wave momentum but not the energy, in contrast with the usual space boundary problem where the wave energy is conserved but the momentum is not. The difference between reflected and time-reflected pulses in classical description is illustrated in Fig. 1.

 figure: Fig. 1.

Fig. 1. Classical electrodynamics picture of reflection and refraction of light from a space boundary at $x=0$ between two media with refractive indices $n_{1}$ and $n_{2}$ in 1+1 dimensions (a) and a time boundary at $t=0$ (b). During reflection and refraction from the space (time) boundary the light energy (momentum) is conserved. When a pulse of frequency $\nu$ passes through a time boundary it splits into two pulses propagating in the opposite directions with frequencies $\pm n_{1}\nu /n_{2}$.

Download Full Size | PDF

Pulses propagating backward in time (having negative frequency) can be produced in the lab frame using superluminal boundaries between two media or if properties of the host material depend on time [17,18]. Such boundaries have been demonstrated experimentally. For example, superluminal ionization fronts have been used to generate electromagnetic radiation [19]. In a nonlinear medium, one can make refractive index inhomogeneities propagate faster than light by creating interference between two laser beams with differing frequencies (flying mirrors) [20]. Recently temporal reflection at photonic time interfaces has been observed experimentally [21,22]. Negative frequency pulses have been produced experimentally by optical solitons through a resonant emission process [23].

As we mentioned above, appearance of the negative frequency pulses can be interpreted as particle generation or annihilation. Namely, by changing the basis set of the photon mode functions one can describe the state of the field in terms of the positive frequency photons but with a different particle content. In the latter description, the superluminal interface or an abrupt change of the refractive index yield generation of positive-frequency entangled photon pairs propagating in the opposite directions [24]. This is analogous to the creation of entangled particle pairs in the dynamical Casimir effect and Hawking radiation. In Section 5 we show that during the process of particle pair production, entanglement of vacuum is transferred to the entanglement of the generated photon pairs.

If initially an electromagnetic pulse is present in the medium the photon generation is stimulated into the pulse mode, and since photons are created as entangled pairs the counter-propagating photon partners produce a positive-frequency pulse moving in the opposite direction. Thus, pulse splitting (time reflection) under an abrupt change of the refractive index occurs due to stimulated generation of particle pairs and no photons in the original pulse are in fact reflected back.

The moving boundary mimics an event horizon of a black hole (BH) if it’s speed is subluminal for the medium 1 (analog of BH exterior) and superluminal for the medium 2 (BH interior). In Section 6 we show that in this case the moving boundary generates entangled photon pairs, with one photon propagating into the medium 1 and the other into the medium 2, which is analogous to Hawking radiation. Production of photon pairs by the event horizon can be stimulated by electromagnetic field present in the system.

2. Interaction of light pulses with a superluminal boundary: classical description

Let us assume that a Gaussian pulse with electric field

$$E(t,x)=e^{{-}i\nu \left( t\pm \frac{n_{1}}{c}x\right) }e^{-\Omega ^{2}\left( t\pm \frac{n_{1}}{c}x\right) ^{2}}$$
falls perpendicularly at the interface between two dielectric media with refractive indices $n_{1}$ and $n_{2}$. The field is linearly polarized parallel to the interface. The media are at rest, but the interface is moving with a superluminal speed for both media $V>c/n_{1,2}$. The upper (lower) sign in Eq. (8) describes the left (right) moving pulse. Interaction of the right-moving pulse with the interface is illustrated in Fig. 2.

 figure: Fig. 2.

Fig. 2. (a) A right-moving Gaussian pulse passes through a superluminal boundary between two dielectric media. The latter are at rest and only the interface is moving. (b) The interface splits the pulse into two transmitted Gaussian pulses moving in the opposite directions, and there is no reflected pulse (that is there is no pulse on the left side of the boundary after interaction). In the classical description the pulse $\nu _{3}$ has negative frequency and propagates backward in time.

Download Full Size | PDF

At the interface, the pulse splits into two transmitted Gaussian pulses moving in the opposite directions with the speed $\pm c/n_{2}$, and there is no reflected pulse, that is there is no pulse on the left side of the boundary after interaction (see Fig. 2(b)). As we show in Appendix A, classical Maxwell’s equations yield for the field in the medium $2$

$$E(t,x)=A_{2}e^{{-}i\nu _{2}\left( t-\frac{n_{2}}{c}x\right) }e^{-\Omega _{2}^{2}\left( t-\frac{n_{2}}{c}x\right) ^{2}}+A_{3}e^{{-}i\nu _{3}\left( t+ \frac{n_{2}}{c}x\right) }e^{-\Omega _{3}^{2}\left( t+\frac{n_{2}}{c}x\right) ^{2}},$$
where carrier frequencies of the transmitted pulses are
$$\nu _{2}=\frac{1\mp n_{1}\frac{V}{c}}{1+n_{2}\frac{V}{c}}\nu ,\quad \nu _{3}= \frac{1\mp n_{1}\frac{V}{c}}{1-n_{2}\frac{V}{c}}\nu ,$$
while the pulse amplitudes $A_{2,3}$ and widths $\Omega _{2,3}$ are given by Eqs. (63), (64) and (65). In the medium $2$, the mode function (9) consists of two pulses and the total mode function has positive norm.

If the pulse in the medium $1$ is moving to the right (lower sign in Eq. (10)), and velocity of the interface $V>c/n_{1,2}$, the carrier frequency of the left-moving pulse $A_{3}$ is negative $(\nu _{3}<0)$, while $\nu _{2}>0$. Thus, pulse $A_{3}$ propagates backward in time and it’s mode function is obtained from the pulse moving forward in time

$$A_{3}e^{{-}i|\nu _{3}|\left( t-\frac{n_{2}}{c}x\right) }e^{-\Omega _{3}^{2}\left( t-\frac{n_{2}}{c}x\right) ^{2}}$$
by the time-reversal transformation $t\rightarrow -t$. If the pulse in the medium $1$ is moving to the left (upper sign in Eq. (10)), the carrier frequency of the transmitted right-moving pulse $A_{2}$ is negative ( $\nu _{2}<0$), while $\nu _{3}>0$.

Thus, superluminal boundary produces pulses with negative frequency (time reflection). In quantum description this implies particle creation which we discuss next.

3. Stimulated generation of entangled photon pairs by a superluminal boundary

As we mentioned above, one can change the basis set of the field mode functions such that photons in the medium 2 have only positive frequencies in their Fourier spectrum. To do this we consider two independent orthogonal Gaussian modes of the field propagating perpendicularly to the moving boundary (we assume that motion of the interface is superluminal for both media, $V>c/n_{1,2}$)

$$E(t,x)=\left\{ \begin{array}{c} f(t,x),\quad x<{-}Vt \\ C_{2}f_{2}(t,x)+C_{3}f_{3}^{{\ast} }(t,x),\quad x>{-}Vt \end{array} \right. ,$$
$$E_{1}(t,x)=\left\{ \begin{array}{c} f_{1}(t,x),\quad x<{-}Vt \\ C_{2}f_{3}(t,x)+C_{3}f_{2}^{{\ast} }(t,x),\quad x>{-}Vt \end{array} \right. ,$$
where
$$f(t,x)=\sqrt{\frac{\Omega \nu }{n_{1}}}e^{{-}i\nu \left( t-\frac{n_{1}}{c} x\right) }e^{-\Omega ^{2}\left( t-\frac{n_{1}}{c}x\right) ^{2}},$$
$$f_{1}(t,x)={-}\sqrt{\frac{\Omega _{1}\nu _{1}}{n_{1}}}e^{{-}i\nu _{1}\left( t+ \frac{n_{1}}{c}x\right) }e^{-\Omega _{1}^{2}\left( t+\frac{n_{1}}{c}x\right) ^{2}},$$
$$f_{2}(t,x)=\sqrt{\frac{\Omega _{2}\nu _{2}}{n_{2}}}e^{{-}i\nu _{2}\left( t- \frac{n_{2}}{c}x\right) }e^{-\Omega _{2}^{2}\left( t-\frac{n_{2}}{c}x\right) ^{2}},$$
$$f_{3}(t,x)=\sqrt{\frac{\Omega _{3}\nu _{3}}{n_{2}}}e^{{-}i\nu _{3}\left( t+ \frac{n_{2}}{c}x\right) }e^{-\Omega _{3}^{2}\left( t+\frac{n_{2}}{c}x\right) ^{2}}$$
are properly normalized Gaussian mode functions with positive carrier frequency,
$$\frac{\nu _{1}}{\nu }=\frac{\Omega _{1}}{\Omega }=\left\vert \frac{n_{1} \frac{V}{c}+1}{n_{1}\frac{V}{c}-1}\right\vert ,$$
$$\frac{\nu _{2}}{\nu }=\frac{\Omega _{2}}{\Omega }=\left\vert \frac{n_{1} \frac{V}{c}+1}{n_{2}\frac{V}{c}+1}\right\vert ,$$
$$\frac{\nu _{3}}{\nu }=\frac{\Omega _{3}}{\Omega }=\left\vert \frac{n_{1} \frac{V}{c}+1}{n_{2}\frac{V}{c}-1}\right\vert ,$$
and the pulse amplitudes in the medium 2
$$C_{2}=\frac{n_{1}+n_{2}}{2\sqrt{n_{1}n_{2}}}, \qquad C_{3}=\frac{n_{2}-n_{1} }{2\sqrt{n_{1}n_{2}}}$$
obey the property
$$C_{2}^{2}-C_{3}^{2}=1.$$

Since $\nu$, $\nu _{1}$, $\nu _{2}$ and $\nu _{3}$ are positive the mode functions (11) and (12) have a positive norm. In the medium $2$ ($x>-Vt$) the mode functions (11) and (12) are linear combinations of Gaussian pulses with the opposite sign of the carrier frequencies. Photon operators in the modes (11) and (12) we denote as $\hat {a}$ and $\hat {a}_{1}$, while operators of photons in the modes

$$E_{2}(t,x)=\left\{ \begin{array}{c} C_{2}f(t,x)-C_{3}f_{1}^{{\ast} }(t,x),\quad x<{-}Vt \\ f_{2}(t,x),\quad x>{-}Vt \end{array} \right. ,$$
$$E_{3}(t,x)=\left\{ \begin{array}{c} C_{2}f_{1}(t,x)-C_{3}f^{{\ast} }(t,x),\quad x<{-}Vt \\ f_{3}(t,x),\quad x>{-}Vt \end{array} \right. ,$$
we denote as $\hat {b}_{2}$ and $\hat {b}_{3}$ respectively. In the medium 2 the mode functions (21) and (22) are single Gaussian pulses with the positive carrier frequency. Equations (11), (12), (21) and (22) yield the following Bogoliubov transformations between the photon operators
$$\hat{a}=C_{2}\hat{b}_{2}+C_{3}\hat{b}_{3}^{\dagger},\quad \hat{a}^{\dagger}=C_{2} \hat{b}_{2}^{\dagger}+C_{3}\hat{b}_{3},$$
$$\hat{a}_{1}=C_{2}\hat{b}_{3}+C_{3}\hat{b}_{2}^{\dagger},\quad \hat{a} _{1}^{\dagger}=C_{2}\hat{b}_{3}^{\dagger}+C_{3}\hat{b}_{2},$$
while inverse transformations are given by
$$\hat{b}_{2}=C_{2}\hat{a}-C_{3}\hat{a}_{1}^{\dagger},\quad \hat{b}_{2}^{\dagger}=C_{2}\hat{a}^{\dagger}-C_{3}\hat{a}_{1},$$
$$\hat{b}_{3}=C_{2}\hat{a}_{1}-C_{3}\hat{a}^{\dagger},\quad \hat{b}_{3}^{\dagger}=C_{2}\hat{a}_{1}^{\dagger}-C_{3}\hat{a}.$$

Equation (20) guarantees that operators $\hat {a}$, $\hat {a}_{1}$, and $\hat {b}_{2}$, $\hat {b}_{3}$ obey the same bosonic commutation relations.

In the medium 1, the mode functions (11) and (12) have only positive frequency term. Therefore, if state of the field is such that a fixed atom doesn’t become excited in the medium 1 (Minkowski vacuum) then it is a vacuum state for the photons $\hat {a}$ and $\hat {a}_{1}$. Number of photons in the modes $\hat {a}$ and $\hat {a}_{1}$ does not change with time and, thus, no $\hat {a}$ and $\hat {a}_{1}$ photons are present in both media. However, according to Eqs. (11) and (12), in the medium 2 the fixed atom can become excited by emitting photons into the modes $\hat {a}$ and $\hat {a}_{1}$. This is the case because in the medium 2 the modes (11) and (12) have negative frequency components.

If we look at the same state of the field from the perspective of photons $\hat {b}_{2}$ and $\hat {b}_{3}$, then media 1 and 2 are filled with particles. Using Bogoliubov transformations (25) and (26) we obtain for the average number of photons in the modes $\hat {b}_{2}$ and $\hat {b}_{3}$

$$N_{2}=\left\langle \hat{b}_{2}^{\dagger}\hat{b}_{2}\right\rangle =N+C_{3}^{2}(N+N_{1}+1),$$
$$N_{3}=\left\langle \hat{b}_{3}^{\dagger}\hat{b}_{3}\right\rangle =N_{1}+C_{3}^{2}(N+N_{1}+1),$$
where $N$ and $N_{1}$ are the average number of photons in the modes $\hat {a}$ and $\hat {a}_{1}$ respectively and we assumed that $\left \langle \hat {a} \hat {a}_{1}\right \rangle =0$. If state of the field is vacuum for the photons $\hat {a}$ and $\hat {a}_{1}$ ($N=N_{1}=0$), Eqs. (27) and (28) yield
$$N_{2}=N_{3}=\frac{\left( n_{2}-n_{1}\right) ^{2}}{4n_{1}n_{2}}.$$

According to Eqs. (21) and (22), in the medium 2 the mode functions of photons $\hat {b}_{2}$ and $\hat {b}_{3}$ have only positive frequency components, and, thus, a fixed atom can become excited only by absorbing photons $\hat {b}_{2}$ and $\hat {b}_{3}$. If we describe the state of the field by positive-frequency photons in both media (that is by $\hat {a}$ and $\hat {a}_{1}$ in the medium 1 and by $\hat {b}_{2}$ and $\hat {b}_{3}$ in the medium 2), in this description the superluminal interface generates photons out of vacuum with the average photon number per mode given by Eq. (29).

Moreover, Bogoliubov transformations (25) and (26) yield that in terms of photons $\hat {b}_{2}$ and $\hat {b}_{3}$ the state of the field is a two-mode squeezed state, namely,

$$\left\vert 0_{a}0_{a_{1}}\right\rangle =\sqrt{1-\gamma ^{2}}e^{-\gamma \hat{b }_{2}^{{\dagger} }\hat{b}_{3}^{{\dagger} }}\left\vert 0_{b_{2}}0_{b_{3}}\right\rangle ,$$
where
$$\gamma =\frac{C_{3}}{C_{2}}=\frac{n_{2}-n_{1}}{n_{2}+n_{1}},$$
$\left \vert 0_{a}0_{a_{1}}\right \rangle$ and $\left \vert 0_{b_{2}}0_{b_{3}}\right \rangle$ refer to the state with no photons in the modes $\hat {a}$, $\hat {a}_{1}$ and $\hat {b}_{2}$, $\hat {b}_{3}$ respectively. Thus, the superluminal interface generates entangled pairs of positive-frequency photons with the frequency ratio
$$\frac{\nu _{3}}{\nu _{2}}=\left\vert \frac{1+n_{2}\frac{V}{c}}{1-n_{2}\frac{V }{c}}\right\vert$$
propagating in the opposite directions.

If modes $\hat {b}_{2}$ and $\hat {b}_{3}$ are considered separately, then tracing over or absorbing one of the modes leaves the remaining mode in a thermal state. Namely, if we trace over the mode $\hat {b}_{3}$, the reduced density operator for the photons $\hat {b}_{2}$ is thermal

$$\hat{\rho}_{b_{2}}=\text{Tr}_{b_{3}}\left( \left\vert 0_{a}0_{a_{1}}\right\rangle \left\langle 0_{a}0_{a_{1}}\right\vert \right) =\left( 1-\gamma ^{2}\right) \sum_{m=0}^{\infty }\gamma ^{2m}\left\vert m\right\rangle \left\langle m\right\vert ,$$
where $\left \vert m\right \rangle$ is a state with $m$ photons in the mode $\hat {b}_{2}$.

Equation (29) is obtained under the assumption that media have no dispersion. If this is the case the number of generated photons in each mode is independent of frequency as well. In reality, however, the medium has dispersion and $n_{1,2}\rightarrow 1$ at high frequencies. In this limit the interface is no longer superluminal and, thus, there is no generation of high frequency photons. This is analogous to Cherenkov radiation for which high-frequency photons are not emitted because particle motion is not superluminal relative to the speed of high-frequency photons in the medium.

One can use Eqs. (27) and (28) to calculate pulse amplification by the superluminal interface. A right-moving initial pulse described by the mode $\hat {a}$ carries energy $W_{0}=\hslash \nu N$. The interface splits the pulse into two pulses into the modes $\hat {b}_{2}$ and $\hat {b}_{3}$ with the net energy $W=\hslash \nu _{2}N_{2}+\hslash \nu _{3}N_{3}$. Plug here Eqs. (27) and (28) with $N_{1}=0$ and, using Eqs. (18 ) and (19), in the limit $N\gg 1$ we obtain

$$\frac{W}{W_{0}}=\left\vert \frac{1+n_{1}\frac{V}{c}}{1+n_{2}\frac{V}{c}} \right\vert \frac{\left( n_{1}+n_{2}\right) ^{2}}{4n_{1}n_{2}}+\left\vert \frac{1+n_{1}\frac{V}{c}}{1-n_{2}\frac{V}{c}}\right\vert \frac{\left( n_{1}-n_{2}\right) ^{2}}{4n_{1}n_{2}}.$$

Equation (33) shows that pulse amplification is large if velocity of the interface is close to the speed of light in the medium 2, $V\approx c/n_{2}$. In this case the frequency and the duration of the counter-propagating pulse $A_{3}$ go as

$$\nu _{3}\approx \frac{1+\frac{n_{1}}{n_{2}}}{\left\vert 1-n_{2}\frac{V}{c} \right\vert }\nu ,\quad \tau _{3}\approx \frac{\left\vert 1-n_{2}\frac{V}{c} \right\vert }{1+\frac{n_{1}}{n_{2}}}\tau .$$

That is pulse undergoes substantial frequency up-conversion and shortening. This property can be used for generation of intense short pulses with high frequency.

Equations (27) and (28) provide an interesting insight on the mechanism of the counter-propagating pulse generation (time reflection). If initially there is only right-moving pulse with the photon number $N$, Eqs. (27) and (28) reduce to

$$N_{2}=N+C_{3}^{2}(N+1),$$
$$N_{3}=C_{3}^{2}(N+1).$$

Equations (34) and (35) show that pair production is stimulated by the pulse present in the medium. Namely, spontaneous generation term $C_{3}^{2}$ in Eq. (34) is enhanced by the factor $N+1$. That is generation of the right-moving photon is stimulated by the pulse. However, since photons are generated as entangled pairs, the counter-propagating entangled photon partners yield a pulse moving in the opposite direction. Thus, the counter-propagating pulse appears due to stimulated generation of the entangled photon pairs by the moving boundary. This is different from the mechanism of pulse splitting by a subluminal interface or by spatial inhomogeneities for which no photons are generated.

4. Photon generation and time reflection by a sudden change of the medium refractive index

According to Eq. (29), the number of generated photons in each mode is independent of the velocity of the superluminal interface. This can be understood from the following argument. By making Lorentz transformation one can choose an inertial frame in which the superluminal boundary has the worldline $t=t_{0}$. In this frame, the boundary corresponds to a sudden change of the medium refractive index.

Such Lorentz transformation exists if the boundary moves at a speed greater than speed of light in vacuum $c$. For example, a worldline

$$x+ut=0,\quad u>c,$$
under Lorentz transformation
$$t=\frac{t^{\prime }-\beta x^{\prime }/c}{\sqrt{1-\beta ^{2}}},\quad x=\frac{ x^{\prime }-\beta ct^{\prime }}{\sqrt{1-\beta ^{2}}},$$
where $\beta =c/u<1$, transforms into the worldline $t^{\prime }=0$. Superluminal motion of the boundary (with speed greater than speed of light in vacuum) does not violate causality because motion of the interface between regions with different refractive indices does not necessarily correspond to motion of particles faster than light.

In this Section we consider a non-moving dielectric medium for which the refractive index suddenly changes from $n_{1}$ to $n_{2}$ (see Fig. 3). As a result of such change, the Gaussian pulse (8) splits into two Gaussian pulses (9) moving in the opposite directions. The carrier frequencies of these pulses and the pulse widths can be obtained from Eqs. (10) and (65) by taking the limit $V\rightarrow \infty$. This is the case because previous calculations are valid for any speed of the interface, including those faster than light in vacuum. As a verification we performed calculations for the case of a sudden change of the medium refractive index separately. In either way we obtain

$$\nu _{2}={-}\nu _{3}={\mp} \frac{n_{1}}{n_{2}}\nu ,\quad \Omega _{2}=\Omega _{3}= \frac{n_{1}}{n_{2}}\Omega ,$$
where $\nu$ and $\Omega$ are the carrier frequency and the width of the initial pulse. Thus, the generated counter-propagating (time reflected) pulse has negative frequency and moves backward in time.

 figure: Fig. 3.

Fig. 3. A pulse is moving through a medium whose refractive index suddenly changes from $n_{1}$ to $n_{2}$. This change splits the pulse into two pulses propagating in the opposite directions.

Download Full Size | PDF

However, in the conventional quantum treatment of the problem we choose the positive-frequency modes (15) and (16) to quantize the field. In this picture, both generated pulses have the same positive frequency $\nu n_{1}/n_{2}$ but different particle content. Namely, the average number of photons in the two pulses is given by Eq. (34) for the co-moving pulse, and Eq. (35) for the counter-propagating pulse.

The last term in Eqs. (34) and (35) describes stimulated generation of particles by the sudden change of the refractive index. As we mentioned in the previous Section, appearance of the counter-propagating pulse (35) is caused by the stimulated generation of the entangled photon pairs. That is no photons in the original pulse are in fact being reflected and the energy of the backward-propagating pulse comes from the particle generation. This is the quantum mechanism of time reflection of light.

It has been shown that time refraction/reflection can be described by a squeezing transformation [24]. The total momentum of the field is conserved during time reflection since the medium is spatially uniform (there is space translation symmetry). However, the energy is not conserved because there is no time translation symmetry. In the limit $N\gg 1$ Eqs. (34) and (35) give that sudden change of the refractive index from $n_{1}$ to $n_{2}$ changes energy of the field by a factor $(1+n_{1}^{2}/n_{2}^{2})/2$. This is different from the ordinary reflection from stationary spatial inhomogeneities of the refractive index for which the field energy is conserved but the momentum does not (there is time but no space translation symmetry in the problem).

If initial state of the field is vacuum the sudden change of the refractive index yields generation of entangled pairs of photons with equal frequencies moving in the opposite directions. The photons in the pair are in the two-mode squeezed state given by Eq. (30). Tracing over the entangled partner leaves the remaining photon in a thermal state (32) with the average number of photons per mode given by Eq. (29).

Particle generation by superluminal boundaries or by a sudden change of the medium properties is a quantum effect. Quantum vacuum emission from a moving refractive-index step in 1+1 dimensions for realistic dispersive media has been developed in [25,26]. Hawking-like radiation can be produced in this setup [26]. Time-varying media have inspired the discussion of new quantum optical phenomena. For example, a sudden change in the medium anisotropy allows us to control the angular distribution of the photons generated via vacuum amplification effects [27]. By matching temporal layers one can suppress generation of backward waves in temporal boundaries (antireflection temporal coatings) [28,29]. Transluminal gratings can also yield spontaneous photon generation [30].

In the next section we will make a connection between generation of entangled photon pairs by a sudden change of the medium properties and vacuum entanglement.

5. Transfer of Minkowski vacuum entanglement to the entanglement of the generated photon pairs by a sudden change of refractive index

Minkowski vacuum is a ground state of free field in Minkowski spacetime. However, particle content of Minkowski vacuum depends on the choice of mode functions to quantize the field. For example, if we choose plane-wave modes as a basis set the ground-state of the free-field Hamiltonian

$$\hat{H}=\sum_{\mathbf{k}}\hbar \nu _{k}\hat{a}_{\mathbf{k}}^{\dagger}\hat{a}_{ \mathbf{k}},$$
where $\hat {a}_{\mathbf {k}}$ are annihilation operators of the plane-wave photons, has no particles. That is there are no plane-wave photons in Minkowski vacuum.

However, if, e.g., we choose the basis set as a combination of two plane-waves with opposite frequencies and equal momenta

$$\phi _{\mathbf{k}}(t,\mathbf{r})=e^{{-}i\nu _{k}t+i\mathbf{kr}}+\gamma e^{i\nu _{k}t+i\mathbf{kr}},$$
where $\nu _{k}=ck/n$, $n$ is the medium refractive index and $\gamma <1$ is a parameter, the photons in the modes $\phi _{\mathbf {k}}$ and $\phi _{- \mathbf {k}}$ interact with each similar to the particle interaction in a weakly interacting Bose-Einstein condensate. Annihilation operators $\hat {b} _{\mathbf {k}}$ of photons (37) are related to $\hat {a}_{\mathbf {k}}$ by the Bogoliubov transformation
$$\hat{a}_{\mathbf{k}}=\frac{\hat{b}_{\mathbf{k}}-\gamma \hat{b}_{-\mathbf{k} }^{\dagger}}{\sqrt{1-\gamma ^{2}}},\quad \hat{a}_{\mathbf{k}}^{\dagger}=\frac{ \hat{b}_{\mathbf{k}}^{\dagger}-\gamma \hat{b}_{-\mathbf{k}}}{\sqrt{1-\gamma ^{2}}}.$$

Plug Eq. (38) into Eq. (36) yields the free-field Hamiltonian in terms of operators $\hat {b}_{\mathbf {k}}$ which upto an irrelevant constant reads

$$\hat{H}=\sum_{\mathbf{k}}\hbar \nu _{k}\left( \frac{1+\gamma ^{2}}{1-\gamma ^{2}}\hat{b}_{\mathbf{k}}^{\dagger}\hat{b}_{\mathbf{k}}+\frac{2\gamma }{ 1-\gamma ^{2}}\left( \hat{b}_{\mathbf{k}}^{\dagger}\hat{b}_{-\mathbf{k}}^{\dagger}+\hat{b}_{\mathbf{k}}\hat{b}_{-\mathbf{k}}\right) \right) .$$

The last term in the Hamiltonian (39) shows that photons $\hat {b}_{ \mathbf {k}}$ are created and annihilated out of vacuum in pairs with opposite momenta. Similarly to the interacting Bose-Einstein condensate the ground state of the Hamiltonian (39) is filled with excitations. Namely, in terms of photons $\hat {b}_{\mathbf {k}}$, Minkowski vacuum is a product of two-mode squeezed states (see Appendix E in [31])

$$\left\vert 0_{M}\right\rangle =\prod_{\mathbf{k}}\sqrt{1-\gamma ^{2}} e^{\gamma \hat{b}_{\mathbf{k}}^{\dagger}\hat{b}_{-\mathbf{k}}^{\dagger}}\left\vert 0_{b}\right\rangle ,$$
where $\left \vert 0_{b}\right \rangle$ is the state with no photons in the modes $\hat {b}_{\mathbf {k}}$. Thus, Minkowski vacuum is filled with entangled pairs of photons $\hat {b}_{\mathbf {k}}$ and $\hat {b}_{-\mathbf {k}}$.

However, photons $\hat {b}_{\mathbf {k}}$ are “invisible” for stationary atoms. This is clear because stationary atoms cannot become excited in Minkowski vacuum. One can also show this directly by calculating the excitation probability of a stationary atom interacting with photons $\hat {b}_{\mathbf {k}}$. Since mode functions (37) have both positive and negative frequency components the atom can become excited by absorbing or emitting photons $\hat {b}_{\mathbf {k}}$. Calculations yield that the process in which the atom becomes excited by absorbing (emitting) photon $\hat {b}_{\mathbf {k}}$ destructively interferes with the process in which the atom becomes excited by emitting (absorbing) photon $\hat {b}_{-\mathbf {k}}$, and the corresponding probability amplitudes cancel each other. Entanglement between photons $\hat {b}_{\mathbf {k}}$ and $\hat {b}_{-\mathbf {k}}$, described by Eq. (40), is crucial for such cancellation.

Next we show that sudden change of the medium refractive index transfers entanglement of Minkowski vacuum into entanglement of the generated photon pairs. According to Eq. (21) taken in the limit $V\rightarrow \infty$, under sudden change of the refractive index from $n$ to $n_{2}$ the mode function

$$e^{{-}i\nu \left( t-\frac{n}{c}x\right) }+\frac{n_{2}-n}{n_{2}+n}e^{i\nu \left( t+\frac{n}{c}x\right) }$$
transforms into
$$\frac{2n^{2}}{n_{2}\left( n+n_{2}\right) }e^{{-}i\nu \left( \frac{n}{n_{2}}t- \frac{n}{c}x\right) }.$$

Thus, if we choose $n_{2}$ such that $(n_{2}-n)/(n_{2}+n)=\gamma$, that is

$$n_{2}=\frac{1+\gamma }{1-\gamma }n,$$
the sudden change of the refractive index $n\rightarrow n_{2}$ transforms the mode functions (37) into single plane-waves with positive frequency
$$\phi _{\mathbf{k}}(t,\mathbf{r})\rightarrow \frac{\left( 1-\gamma \right) ^{2}}{1+\gamma }e^{{-}i\nu _{k}^{\prime }t+i\mathbf{kr}},$$
where $\nu _{k}^{\prime }=ck/n_{2}$. Under this change the state of the field is given by the same formula (40), but now the operators $\hat {b }_{\mathbf {k}}$ describe photons in the modes (41) which are “visible”, that is they can excite stationary atoms. Thus, entanglement of the photon pairs generated under the change of the refractive index originates from the entanglement of the “invisible” photons (37) present in Minkowski vacuum.

6. Optical analog of Hawking radiation

General relativity predicts existence of black holes (BHs). These hypothetical objects possess an event horizon - the boundary under which no particles, at least if they are treated classically and moving forward in time, can escape. In BH spacetime, particles under the event horizon always move away from the horizon toward the BH center. In 1974 S. Hawking showed that entangled particle-antiparticle pairs can be generated near the event horizon of a BH, with one carrying positive energy to infinity and the other carrying negative energy into the BH [32,33]. This is known as Hawking radiation which leads to BH evaporation.

Hawking predicted that BH radiation is thermal. However, if nonunitary absorption of the negative energy photons near the BH center is taken into account, this alters the state of the outgoing entangled photon partner. As a result, radiation of the evaporating BH is not thermal; it carries information about BH interior, and entropy is preserved during evaporation [34].

In this Section we discuss an optical analog of Hawking radiation in which a superluminal boundary between two dielectric media plays the role of the event horizon. In Sections 2 and 3 we assumed that the boundary moves faster than the speed of light in both media 1 and 2, and showed that the moving boundary generates entangled photon pairs out of vacuum in the medium 2 (see Fig. 2). To make connection with the BH event horizon the speed of the boundary must be subluminal for the medium 1 and superluminal for the medium 2 ($c/n_{1}>V>c/n_{2}$). In this case, medium 2 is analogous to the BH interior region - light from the medium 2 can not escape into the medium 1 (exterior region). Light always move away from the moving boundary in the medium 2. In contrast, in the medium 1, light can move both toward and away from the boundary which is analogous to the BH exterior region.

Next we show that under these conditions the moving boundary generates entangled photon pairs, with one photon propagating into the medium 1 and the other photon into the medium 2. In addition, the boundary generates pairs of photons propagating in the medium 2 in the opposite directions.

In the present case, the total field consists of the incident and reflected waves in the medium 1, and two refracted waves in the medium 2. Thus, the boundary-value problem is degenerate and can be solved uniquely only if we make additional assumptions about the structure of the interface, nonlinearities, etc. [35]. Keeping in mind analogy with BH radiation, we assume that the interface structure yields no reflected wave, as shown in Fig. 4(a). The corresponding positive-norm mode function is given by

$$E(t,x)=\left\{ \begin{array}{c} f(t,x),\quad x<{-}Vt \\ C_{2}f_{2}(t,x)-C_{3}f_{3}^{{\ast} }(t,x),\quad x>{-}Vt \end{array} \right. ,$$
where notations are the same as in Section 3. The mode function (42) differs from Eq. (11) by change $C_{3}\rightarrow -C_{3}$. Photon operator for the mode (42) we denote as $\hat {a}$. Equation (42 ) yields the following Bogoliubov transformation between the photon operators in the medium 2
$$\hat{a}=C_{2}\hat{b}_{2}-C_{3}\hat{b}_{3}^{\dagger},$$
where $\hat {b}_{2}$ and $\hat {b}_{3}$ are operators of photons which in the medium 2 have the positive-frequency mode functions $f_{2}(t,x)$ and $f_{3}(t,x)$ respectively. Namely, properly normalized orthogonal mode functions of the photons $\hat {b}_{2}$ and $\hat {b}_{3}$ read
$$E_{2}(t,x)=\left\{ \begin{array}{c} C_{2}f(t,x)+C_{3}f_{1}(t,x),\quad x<{-}Vt \\ f_{2}(t,x)+\sqrt{2}C_{3}f_{4}^{{\ast} }(t,x),\quad x>{-}Vt \end{array} \right. ,$$
$$E_{3}(t,x)=\left\{ \begin{array}{c} C_{2}f_{1}^{{\ast} }(t,x)+C_{3}f^{{\ast} }(t,x),\quad x<{-}Vt \\ f_{3}(t,x)+\sqrt{2}C_{2}f_{4}(t,x),\quad x>{-}Vt \end{array} \right. ,$$
where $f_{4}(t,x)$ is a left-moving pulse in the medium 2 propagating far away from the interface (at $x\rightarrow \infty$). The pulse $f_{4}$ has the same norm as $f,$ $f_{1}$, $f_{2}$ and $f_{3}$, but can have a mixture of positive and negative frequencies. Pulse $f_{4}$ does not interact with the interface and is introduced to assure proper normalization of the mode functions (42), (44), (45) and (47).

 figure: Fig. 4.

Fig. 4. Interface between two static media moving with a constant velocity that is subluminal (superluminal) for the medium 1 (2) mimics a BH event horizon. Orthogonal mode functions of the field obeying the proper boundary conditions at the interface with one ingoing and two transmitted waves (Eq. (refk1a)), and three outgoing waves (Eq. (refk1d)) are sketched in (a) and (b) respectively.

Download Full Size | PDF

We assume that state of the field is vacuum for photons $\hat {a}$, that is there are no photons falling on the interface. Then Eq. (43) gives

$$\left\langle \left( C_{2}\hat{b}_{2}^{\dagger}+C_{3}\hat{b}_{3}\right) \left( C_{2}\hat{b}_{2}-C_{3}\hat{b}_{3}^{\dagger}\right) \right\rangle =0,$$
or
$$N_{2}=\frac{C_{3}^{2}}{C_{2}^{2}}(N_{3}+1),$$
where $N_{2}=\left \langle \hat {b}_{2}^{\dagger}\hat {b}_{2}\right \rangle$ and $N_{3}=\left \langle \hat {b}_{3}^{\dagger}\hat {b}_{3}\right \rangle$ are number of photons in the outgoing modes $\hat {b}_{2}$ and $\hat {b}_{3}$, and we took into account that $\left \langle \hat {b}_{2}^{\dagger}\hat {b}_{3}^{\dagger}- \hat {b}_{3}\hat {b}_{2}\right \rangle =0$. The latter identity follows from the fact that state of the field is vacuum for the operator (43) and both $C_{2}$ and $C_{3}$ are real numbers. Indeed, taking the expectation value from the operator $\hat {b}_{2}^{\dagger}\left ( C_{2}\hat {b}_{2}-C_{3} \hat {b}_{3}^{\dagger}\right )$ we obtain $\left \langle \hat {b}_{2}^{\dagger}\hat { b}_{3}^{\dagger}\right \rangle =N_{2}C_{2}/C_{3}$. Thus, if both $C_{2}$ and $C_{3}$ are real then $\left \langle \hat {b}_{2}^{\dagger}\hat {b}_{3}^{\dagger}\right \rangle$ is equal to its complex conjugate $\left \langle \hat {b}_{3} \hat {b}_{2}\right \rangle$.

A mode function orthogonal to (42) reads

$$E_{1}(t,x)=\left\{ \begin{array}{c} f_{1}^{{\ast} }(t,x),\quad x<{-}Vt \\ C_{2}f_{3}(t,x)-C_{3}f_{2}^{{\ast} }(t,x)+\sqrt{2}f_{4}(t,x),\text{ }x>{-}Vt \end{array} \right. .$$

It is sketched in Fig, 4(b). The mode function (47) consists of three outgoing waves and the left-moving pulse $f_{4}$ propagating at $x\rightarrow \infty$ which is not shown in the figure. Photon operator of the mode $E_{1}(t,x)$ we denote as $\hat {a}_{1}$. It can be expressed in terms of $\hat {b}_{2}$ and $\hat {b}_{3}$ as

$$\hat{a}_{1}=C_{2}\hat{b}_{3}-C_{3}\hat{b}_{2}^{\dagger}.$$

There is no reason to believe that state of the field is vacuum for photons $\hat {a}_{1}$. The latter assumption would yield that there are photons in the outgoing mode $f_{1}$ even if there is no interface ($n_{1}=n_{2}$).

To find the average number of photons $N_{1,2,3}$ in the outgoing modes we will use physical arguments. Consider outgoing plane waves moving in the media 1 and 2 that have positive frequencies given by Eqs. (17)–(19). In the reference frame of the moving interface all three outgoing waves have the same absolute value of frequency, however, $\nu _{3}<0$, while $\nu _{1,2}>0$. Thus, the interface degrees of freedom can become excited by emitting a photon into the mode $\nu _{3}$ and then spontaneously decay back to the ground state by emitting the positive-frequency photon $\nu _{1}$ or $\nu _{2}$. This process generates photons in pairs, namely, a pair of $\nu _{3}$ and $\nu _{1}$ photons, and $\nu _{3}$ and $\nu _{2}$ photons. For a small change of the dielectric constant across the interface ($C_{3}^{2}\ll 1$) the probabilities of emission of the $\nu _{1}$ and $\nu _{2}$ photons are approximately the same which gives

$$N_{1}\approx N_{2}.$$

Energy conservation in the moving frame yields an additional equation

$$N_{1}+N_{2}=N_{3}.$$

Combining Eqs. (48) and (49) with Eq. (46), we obtain in the limit $C_{3}^{2}\ll 1$

$$N_{1}=N_{2}=\frac{N_{3}}{2}=C_{3}^{2}=\frac{\left( n_{2}-n_{1}\right) ^{2}}{ 4n_{1}n_{2}}.$$

Equation (50) yields $N_{1}=(N_{2}+N_{3})/3$, that is number of generated photons propagating into the exterior region (medium 1) is three times smaller than that generated in the interior region (medium 2). This is the case because there is an additional channel of generation the particle pairs in the medium 2.

7. Discussion

In 1970, G. Moore predicted that an accelerated mirror can excite the quantum vacuum, generating photons [16]. This effect is nowadays known as the dynamical Casimir effect and was investigated, during the 1970s, in other pioneering articles by DeWitt [36], Fulling and Davies [37,38], Candelas and Deutsch [39], among others. In the dynamical Casimir effect the particle creation from the vacuum occurs when a quantized field is submitted to time-dependent boundary conditions, with moving mirrors being a particular case. A fundamental property of the effect is that particles are created out of vacuum as entangled pairs.

The mechanism of the dynamical Casimir effect is somewhat similar to that of Cherenkov radiation of a neutral atom if both are viewed from the negative frequency perspective [6]. As we discussed in Introduction, if an atom is moving through a medium faster than light in the medium, in the reference frame of the atom the frequencies of the plane-wave photons propagating inside the Cherenkov cone are negative. Thus, the moving atom can become excited by emitting the negative-frequency photon inside the Cherenkov cone followed by spontaneous decay back to the ground state by emitting a positive-frequency photon outside the Cherenkov cone. As a result, the superluminal motion of the atom through a medium is accompanied by generation of entangled photon pairs [8]. In the lab frame, frequencies of both generated photons are positive.

One should mention that Cherenkov radiation of a neutral atom is a quantum effect. It is different from the classical Vavilov-Cherenkov radiation of an electric charge moving at a superluminal speed in a medium in the same way as spontaneous emission of an atom differs from radiation of a classical electric dipole.

Now let us assume that a mirror is uniformly accelerated through Minkowski vacuum in 1+1 dimension along the trajectory

$$t(\tau )=\frac{c}{a}\sinh \left( \frac{a\tau }{c}\right) ,\quad x(\tau )= \frac{c^{2}}{a}\cosh \left( \frac{a\tau }{c}\right) .$$

In Eq. (51), $\tau$ is the proper time of the mirror and $a>0$ is the proper acceleration. The left-moving positive-norm Unruh-Minkowski modes of a scalar field are defined as [40]

$$F_{1\Omega }(t,x)=\frac{|t+x/c|^{i\Omega }}{\sqrt{2\Omega \sinh (\pi \Omega ) }}\left\{ \begin{array}{c} e^{-\frac{\pi \Omega }{2}},\quad t+x/c>0 \\ e^{\frac{\pi \Omega }{2}},\quad t+x/c<0 \end{array} \right. ,$$
$$F_{2\Omega }(t,x)=\frac{|t+x/c|^{{-}i\Omega }}{\sqrt{2\Omega \sinh (\pi \Omega )}}\left\{ \begin{array}{c} e^{\frac{\pi \Omega }{2}},\quad t+x/c>0 \\ e^{-\frac{\pi \Omega }{2}},\quad t+x/c<0 \end{array} \right. ,$$
where $\Omega >0$. In Minkowski vacuum there are no Unruh-Minkowski photons. The coordinate transformation
$$t(\tau ,\bar{x})=\frac{c}{a}e^{a\bar{x}/c^{2}}\sinh \left( \frac{a\tau }{c} \right) , \qquad x(\tau ,\bar{x})=\frac{c^{2}}{a}e^{a\bar{x}/c^{2}}\cosh \left( \frac{a\tau }{c}\right) ,$$
where $a$ is a constant, converts Minkowski spacetime to the Rindler space $\tau,$ $\bar {x}$ [41]. A mirror moving along the trajectory $\bar {x }=0$ in the Rindler space is uniformly accelerating in the Minkowski space along the trajectory (51).

In Rindler space, the Unruh-Minkowski mode functions are plane waves

$$F_{1\Omega }(\tau ,\bar{x})\propto e^{i\frac{a\Omega }{c}\left( \tau +\frac{ \bar{x}}{c}\right) }, \qquad F_{2\Omega }(\tau ,\bar{x})\propto e^{{-}i\frac{ a\Omega }{c}\left( \tau +\frac{\bar{x}}{c}\right) }.$$

Equations (52) show that in the Rindler space the mode function $F_{1\Omega }$ has a negative frequency $\nu =-a\Omega /c$, while $F_{2\Omega }$ has a positive frequency $a\Omega /c$.

Since the mirror is composed of particles, particles at the mirror’s surface can become excited by emitting Unruh-Minkowski photon with the negative frequency into the mode $F_{1\Omega }$ and then decay back to the ground state by emitting a photon into the positive-frequency mode $F_{2\Omega }$. Thus, mirror acceleration yields generation of entangled pairs of Unruh-Minkowski photons, which is the dynamical Casimir effect.

One can also explain the effect by making a connection with Glauber’s coupled oscillators one of which is inverted, that is it has negative frequency (a negative potential and kinetic energy). Such a system works as an amplifier which leads to excitation of the oscillators out of the vacuum fluctuations [42]. Indeed, in the rotating wave approximation, two coupled harmonic oscillators with frequencies $\omega$ and $-\omega$ are described by the Hamiltonian

$$\hat{H}=\hbar \omega \left( \hat{b}_{1}^{\dagger}\hat{b}_{1}-\hat{b}_{2}^{\dagger}\hat{b}_{2}\right) +\hbar g\left( \hat{b}_{1}\hat{b}_{2}+\hat{b}_{1}^{\dagger} \hat{b}_{2}^{\dagger}\right) ,$$
which yields that the number of the oscillator excitations exponentially grow with time from zero
$$\left\langle \hat{b}_{1}^{\dagger}\hat{b}_{1}\right\rangle =\left\langle \hat{b }_{2}^{\dagger}\hat{b}_{2}\right\rangle =\sinh ^{2}(gt),$$
and the evolution of the system’s state vector is given by
$$\left\vert \psi (t)\right\rangle =\frac{1}{\cosh (gt)}e^{{-}i\tanh (gt)\hat{b} _{1}^{\dagger}\hat{b}_{2}^{\dagger}}\left\vert 0\right\rangle ,$$
which is a two-mode squeezed state of the oscillators. Please note that in the quantum description the oscillators become excited out of vacuum without initial kick, which is similar to the spontaneous emission of atoms. An accelerated mirror couples the positive and the negative frequency Unruh-Minkowski modes which leads to their excitation out of the vacuum in a two-mode squeezed state, similarly to the Glauber’s coupled oscillators.

In this paper we consider two static media separated by a boundary moving with a constant velocity $\mathbf {V}$ (Secs. 2 and 3). If the boundary moves from the medium 2 to the medium 1 with a speed exceeding the speed of light in both media this yields generation of entangled pairs of photons out of vacuum in the medium 2 in a two-mode squeezed state. Such effect is quantum and is analogous to the creation of entangled particle pairs in the dynamical Casimir effects. Indeed, in the reference frame of the moving boundary the frequency of plane-wave photons propagating inside the Cherenkov cone in the medium 2 is negative, while photons propagating outside the Cherenkov cone have positive frequency. The moving boundary couples these positive and negative frequency modes which leads to their excitation as in the case of Glauber oscillators.

The limit $V\rightarrow \infty$ corresponds to a sudden change of the medium properties with time. In the absence of the medium dispersion, the number of generated photons in each mode is independent of the frequency and the boundary velocity. That is in the limit $V\rightarrow \infty$ the number of generated photons is given by the same formula (29) as in the case of finite $V$ (see Sec. 4.). In Section 5 we show that during the process of particle pair production by a sudden change of the refractive index, entanglement of vacuum is transferred to the entanglement of the generated photon pairs.

If an electromagnetic pulse crosses the superluminal boundary or there is an abrupt change of the refractive index with time the pulse stimulates photon generation into the pulse mode. Since photons are created as entangled pairs the counter-propagating photon partners produce a pulse moving in the opposite direction. Thus, pulse splitting by the superluminal boundary or by a sudden change of the refractive index (time reflection) occurs due to stimulated generation of the entangled photon pairs. That is during time reflection no photons in the original pulse are actually being reflected and the energy of the backward-propagating pulse comes from the particle generation. This is different from the mechanism of pulse splitting by a subluminal interface or spatial inhomogeneities for which no photons are generated.

The moving boundary mimics an event horizon of a black hole if it’s speed is subluminal for the medium 1 (analog of BH exterior) and superluminal for the medium 2 (BH interior). In this case, the boundary couples three outgoing waves $\nu _{1,2,3}$. One of them (mode $\nu _{3}$) propagates inside the Cherenkov cone in the medium 2 and has negative frequency in the reference frame of the moving boundary. This is analogous to three coupled oscillators one of which is inverted, and yields generation of photons out of vacuum in these three modes. Photons are produced in pairs, namely, a pair of $\nu _{3}$ and $\nu _{1}$ photons, and $\nu _{3}$ and $\nu _{2}$ photons. One of the generated photons ($\nu _{1}$) propagates into the medium 1 and photons $\nu _{2}$ and $\nu _{3}$ propagate into the medium 2. Thus, the optical event horizon created by the moving boundary generates photons into the exterior region (medium 1) similarly to Hawking radiation of a BH (see Sec. 6.). The radiation mechanism is analogous to that of the dynamical Casimir effect which, in turn, can be explained by means of the Glauber oscillators [43].

One should mention that analogs of gravitational effects in various systems, including quantum emission of particles, have been previously discussed in the literature in different settings, see, e.g., [4456] and references therein. Stimulated Hawking-like radiation was measured using light pulses in nonlinear fiber optics to establish artificial event horizon [57]. Our paper is an attempt to provide an additional insight on Hawking radiation and make a connection with the Cherenkov and the dynamical Casimir effects.

It is worth mentioning that the latter non necessarily requires an accelerated mirror. For instance, particles can be also generated by a motionless mirror in vacuum whose internal properties rapidly vary in time [58]. In this context, Wilson et al. [59] observed experimentally the particle creation from vacuum in a transmission line with a superconducting quantum interference device (SQUID) at one end. The SQUID forms a tunable inductance, and by changing the magnetic flux threading it, it can be tuned from almost a short to a highly inductive state. A time-dependent magnetic flux applied to the SQUID changes the SQUID inductance, yielding a time-dependent boundary condition. Schneider et al. demonstrated broadband entanglement of microwave photon pairs generated by the dynamical Casimir effect in a superconducting circuit which consists of a semi-infinite transmission line, terminated by two parallel SQUIDs [60].

Physics behind generation of photon pairs in these experiments is the sum combination resonance which occurs when system’s parameters are modulated with frequency $\nu _{d}$ equal to the sum of two normal mode frequencies $\nu _{1}$ and $\nu _{2}$

$$\nu _{d}=\nu _{1}+\nu _{2}.$$

In both experiments mentioned above, the boundary condition was modulated harmonically with a pump frequency $\nu _{d}$ in the GHz range. This creates a pair of microwave photons, whose frequencies add up to $\nu _{d}$. Other experiments have also been done in metamaterials [61] and fibers [62].

Optical parametric amplification (OPA) [6366] which transfers energy from the driving field to the signal and idler waves is another example of the sum combination resonance caused by the nonlinear light-matter interaction. OPA led to development of tunable radiation sources throughout the infrared, visible, and ultraviolet spectral regions [67]. Such sources yield amplification of coherent radiation at lower-frequencies $\nu _{1}$ and $\nu _{2}$ by pumping nonlinear optical crystals with intense laser light at the sum frequency $\nu _{d}$.

One should mention that Glauber’s oscillators, described by the Hamiltonian (53), are a special case of the sum combination resonance for which $\nu _{1}=-\nu _{2}=\omega$. In this case the modulation frequency $\nu _{d}=0$.

Particles can also be created in entangled pairs with zero net momentum from the vacuum by the curved space-time of an expanding universe [68,69]. The process of cosmological particle creation is accompanied by the generation of quantum entanglement at large distances and played an important role in the early Universe. Together with other effects, it is considered to be responsible for the reheating of our Universe after the inflationary period. Furthermore, a very similar process based on the tearing apart of quantum vacuum fluctuations by an expanding space-time during cosmic inflation explains the creation of the seeds for structure formation.

The expansion of the universe stretches all length scales, including the wavelengths of the particle modes. Thus, modes evolve with time leading to appearance of the negative frequency components. This implies that the modes at early and late times are related by a Bogoliubov transformation [70]. If this stretching process becomes too fast, the cosmic expansion tears apart the initial quantum vacuum fluctuations and turns them into pairs of particles with opposite momenta. This yields spontaneous production of cosmological particles, including the primordial density fluctuations which led to the acoustic peaks in the cosmic microwave background radiation spectrum [7173].

A. Interaction of light with superluminal interface between two media

Here we consider a plane interface between two media moving with a constant velocity $\mathbf {V}$, while both media are at rest (see Fig. 5). In the absence of surface charges and currents, the boundary conditions for the electromagnetic field at the interface read

$$\hat{n}\times (\mathbf{E}_{2}-\mathbf{E}_{1})=\frac{V}{c}(\mathbf{B}_{2}- \mathbf{B}_{1}),$$
$$\hat{n}\times (\mathbf{H}_{2}-\mathbf{H}_{1})={-}\frac{V}{c}(\mathbf{D}_{2}- \mathbf{D}_{1}),$$
where $V$ is the projection of the velocity of the interface along the normal $\hat {n}$ to it, and the index $1$ ($2$) refers to the field at the left (right) side of the interface. Speeds of light in the two media are given by
$$c_{1}=\frac{c}{\sqrt{\varepsilon _{1}\mu _{1}}},\quad c_{2}=\frac{c}{\sqrt{ \varepsilon _{2}\mu _{2}}}$$
respectively. In the absence of currents, the Maxwell’s equations and the constitutive relations for linear materials read
$$\nabla \times \mathbf{E}={-}\frac{1}{c}\frac{\partial \mathbf{B}}{\partial t}, \qquad \nabla \times \mathbf{H}=\frac{1}{c}\frac{\partial \mathbf{D}}{ \partial t},$$
$$\mathbf{D}=\varepsilon \mathbf{E},\quad \mathbf{B}=\mu \mathbf{H}.$$

We will assume that the media are not magnetic, that is $\mu _{1}=\mu _{2}=1$. Then, using Eqs. (54), (55) and (57), we obtain

$$\hat{n}\times (\hat{n}\times (\mathbf{E}_{2}-\mathbf{E}_{1}))={-}\frac{V^{2}}{ c^{2}}(\mathbf{D}_{2}-\mathbf{D}_{1}),$$
or
$$\hat{n}\left( \hat{n}\cdot (\mathbf{E}_{2}-\mathbf{E}_{1})\right) -(\mathbf{E }_{2}-\mathbf{E}_{1})={-}\frac{V^{2}}{c^{2}}(\mathbf{D}_{2}-\mathbf{D}_{1}).$$

 figure: Fig. 5.

Fig. 5. Ingoing wave $A_{1}$ propagates through a superluminal interface between two dielectric media and splits into two transmitted waves $A_{2}$ and $A_{3}$ without reflection.

Download Full Size | PDF

If a plane electromagnetic wave propagates through the interface at the normal incidence, the boundary condition (58) reduces to

$$\mathbf{E}_{2}-\mathbf{E}_{1}=\frac{V^{2}}{c^{2}}(\mathbf{D}_{2}-\mathbf{D} _{1}).$$

Using the constitutive relations (57), we find

$$\left( 1-\frac{V^{2}}{c^{2}}\varepsilon _{1}\right) \mathbf{E}_{1}=\left( 1- \frac{V^{2}}{c^{2}}\varepsilon _{2}\right) \mathbf{E}_{2}.$$

Assuming that the $x-$axis is normal to the surface, $\mathbf {E}$ is polarized along the $z-$axis, and $\mathbf {B}$ is along the $y-$axis (see Fig. 5), Eq. (54) yields the second boundary condition in the form

$$E_{2}-E_{1}+\frac{V}{c}\left( B_{2}-B_{1}\right) =0.$$

The wave equation for the field in a dielectric medium

$$\left( \frac{\partial ^{2}}{\partial t^{2}}-\frac{c^{2}}{\varepsilon }\frac{ \partial ^{2}}{\partial x^{2}}\right) \mathbf{E}=0$$
has the following plane wave solutions
$$\mathbf{E}=\hat{z}e^{i\nu \left( t\pm \frac{\sqrt{\varepsilon }}{c}x\right) }$$
describing right and left moving waves. Maxwell’s equations yield for the magnetic field
$$\mathbf{B}={\mp} \sqrt{\varepsilon }\hat{y}e^{i\nu \left( t\pm \frac{\sqrt{ \varepsilon }}{c}x\right) }.$$

We will assume that a right or left-moving plane wave falls into the interface, that is

$$E_{1}=A_{1}e^{i\nu \left( t\pm \sqrt{\varepsilon _{1}}\frac{x}{c}\right) },\quad B_{1}={\mp} \sqrt{\varepsilon _{1}}E_{1},$$
there is no reflected wave, and there are two waves $A_{2}$ and $A_{3}$ moving in the other side of the interface (see Fig. 5)
$$E_{2}=A_{2}e^{i\nu _{2}\left( t-\frac{\sqrt{\varepsilon _{2}}}{c}x\right) }+A_{3}e^{i\nu _{3}\left( t+\frac{\sqrt{\varepsilon _{2}}}{c}x\right) }, \qquad B_{2}=\sqrt{\varepsilon _{2}}A_{2}e^{i\nu _{2}\left( t-\frac{\sqrt{ \varepsilon _{2}}}{c}x\right) }-\sqrt{\varepsilon _{2}}A_{3}e^{i\nu _{3}\left( t+\frac{\sqrt{\varepsilon _{2}}}{c}x\right) }.$$

The wave $A_{2}$ is right-moving and $A_{3}$ is left-moving. If the interface is moving along the trajectory $x=-Vt$, the boundary conditions (59) and (60) yield the following equations

$$\left( 1-\varepsilon _{1}\frac{V^{2}}{c^{2}}\right) A_{1}e^{i\nu t\left( 1\mp \sqrt{\varepsilon _{1}}\frac{V}{c}\right) }= \left( 1-\varepsilon _{2} \frac{V^{2}}{c^{2}}\right) \left( A_{2}e^{i\nu _{2}t\left( 1+\sqrt{ \varepsilon _{2}}\frac{V}{c}\right) }+A_{3}e^{i\nu _{3}t\left( 1-\sqrt{ \varepsilon _{2}}\frac{V}{c}\right) }\right) ,$$
$$\left( 1\mp \sqrt{\varepsilon _{1}}\frac{V}{c}\right) A_{1}e^{i\nu t\left( 1\mp \sqrt{\varepsilon _{1}}\frac{V}{c}\right) }= \left( 1+\sqrt{\varepsilon _{2}}\frac{V}{c}\right) A_{2}e^{i\nu _{2}t\left( 1+\sqrt{\varepsilon _{2}} \frac{V}{c}\right) } +\left( 1-\sqrt{\varepsilon _{2}}\frac{V}{c}\right) A_{3}e^{i\nu _{3}t\left( 1-\sqrt{\varepsilon _{2}}\frac{V}{c}\right) }.$$

Solution of these equations gives the following result for the wave frequencies and amplitudes

$$\nu _{2}=\frac{1\mp n_{1}\frac{V}{c}}{1+n_{2}\frac{V}{c}}\nu ,\quad \nu _{3}= \frac{1\mp n_{1}\frac{V}{c}}{1-n_{2}\frac{V}{c}}\nu ,$$
$$A_{2}=\frac{1}{2}\left( 1\mp \frac{n_{1}}{n_{2}}\right) \frac{\nu _{2}}{\nu } A_{1},$$
$$A_{3}=\frac{1}{2}\left( 1\pm \frac{n_{1}}{n_{2}}\right) \frac{\nu _{3}}{\nu } A_{1},$$
where $n_{1,2}=\sqrt {\varepsilon _{1,2}}$ are the media refractive indices.

Equations (62), (63) and (64) show that the boundary conditions for the electromagnetic field at the moving interface can be satisfied assuming no reflected wave and two waves propagating at the other side of the interface in the opposite directions. This is the case for any speed of the interface (superluminal or subluminal). For the subluminal motion, the wave $A_{3}$ is interpreted as ingoing, while for the superluminal interface, $A_{3}$ is interpreted as transmitted.

Next we assume that the ingoing field is a Gaussian pulse with a carrier frequency $\nu$, that is

$$E_{1}=A_{1}e^{{-}i\nu \left( t\pm \frac{n_{1}}{c}x\right) }e^{-\Omega ^{2}\left( t\pm \frac{n_{1}}{c}x\right) ^{2}},\quad B_{1}={\mp} n_{1}E_{1},$$
and there are two transmitted pulses
$$E_{2}=A_{2}e^{{-}i\nu _{2}\left( t-\frac{n_{2}}{c}x\right) }e^{-\Omega _{2}^{2}\left( t-\frac{n_{2}}{c}x\right) ^{2}} +A_{3}e^{{-}i\nu _{3}\left( t+ \frac{n_{2}}{c}x\right) }e^{-\Omega _{3}^{2}\left( t+\frac{n_{2}}{c}x\right) ^{2}},$$
$$B_{2}=\sqrt{\varepsilon _{2}}A_{2}e^{{-}i\nu _{2}\left( t-\frac{n_{2}}{c} x\right) }e^{-\Omega _{2}^{2}\left( t-\frac{n_{2}}{c}x\right) ^{2}} -\sqrt{ \varepsilon _{2}}A_{3}e^{{-}i\nu _{3}\left( t+\frac{n_{2}}{c}x\right) }e^{-\Omega _{3}^{2}\left( t+\frac{n_{2}}{c}x\right) ^{2}}.$$

The boundary conditions (59) and (60) yield the same carrier frequencies (62) and the amplitudes (63), (64) for the transmitted Gaussian pulses as for the plane waves, and the following relations between the pulse widths

$$\Omega _{2}=\frac{1\mp n_{1}\frac{V}{c}}{1+n_{2}\frac{V}{c}}\Omega ,\quad \Omega _{3}=\frac{1\mp n_{1}\frac{V}{c}}{1-n_{2}\frac{V}{c}}\Omega .$$

Reflection and transmission of a light pulse when passing through superluminal inhomogeneities of the refractive index have been investigated theoretically using classical Maxwell’s equations in [35,7477].

Funding

U.S. Department of Energy (DE-SC-0023103, FWP-ERW7011, DE-SC0024882); Welch Foundation (A-1261); National Science Foundation (PHY-2013771); Air Force Office of Scientific Research (FA9550-20-1-0366).

Acknowledgments

I am very grateful to Marlan Scully and Bill Unruh for stimulating discussions.

Disclosures

The author declares no conflicts of interest.

Data availability

No data were generated or analyzed in the presented research.

References

1. S. A. Fulling, “Nonuniqueness of canonical field quantization in riemannian space time,” Phys. Rev. D 7(10), 2850–2862 (1973). [CrossRef]  

2. P. C. W. Davies, “Scalar production in Schwarzschild and Rindler metrics,” J. Phys. A: Math. Gen. 8(4), 609–616 (1975). [CrossRef]  

3. W. G. Unruh, “Notes on black hole evaporation,” Phys. Rev. D 14(4), 870–892 (1976). [CrossRef]  

4. V. L. Ginzburg and V. P. Frolov, “Excitation and emission of a “detector” in accelerated motion in a vacuum or in uniform motion at a velocity above the velocity of light in a medium,” Pis’Ma Zh. Eksp. Teor. Fiz. 43, 265 (1986); [JETP Lett. 43, 339 (1986)].

5. V. P. Frolov and V. L. Ginzburg, “Excitation and radiation of an accelerated detector and anomalous Doppler effect,” Phys. Lett. A 116(9), 423–426 (1986). [CrossRef]  

6. A. A. Svidzinsky, A. Azizi, J. S. Ben-Benjamin, et al., “Unruh and Cherenkov radiation from a negative frequency perspective,” Phys. Rev. Lett. 126(6), 063603 (2021). [CrossRef]  

7. C. Luo, M. Ibanescu, S. G. Johnson, et al., “Cherenkov radiation in photonic crystals,” Science 299(5605), 368–371 (2003). [CrossRef]  

8. M. O. Scully, A. A. Svidzinsky, and W. Unruh, “Entanglement in Unruh, Hawking, and Cherenkov radiation from a quantum optical perspective,” Phys. Rev. Res. 4(3), 033010 (2022). [CrossRef]  

9. S. A. R. Horsley, “Canonical quantization of the electromagnetic field interacting with a moving dielectric medium,” Phys. Rev. A 86(2), 023830 (2012). [CrossRef]  

10. Y. Guo and Z. Jacob, “Singular evanescent wave resonances in moving media,” Opt. Express 22(21), 26193 (2014). [CrossRef]  

11. A. H. W. Beck, Space-charge waves and slow electromagnetic waves (Pergamon Press. Inc., New York, 1958).

12. C. K. Birdsall, G. R. Brewer, and A. V. Haeff, “The resistive-wall amplifier,” Proc. IRE 41(7), 865–875 (1953). [CrossRef]  

13. A. R. Hutson, J. H. McFee, and D. L. White, “Ultrasonic Amplification in CdS,” Phys. Rev. Lett. 7(6), 237–239 (1961). [CrossRef]  

14. A. A. Svidzinsky, “Excitation of a uniformly moving atom through vacuum fluctuations,” Phys. Rev. Res. 1(3), 033027 (2019). [CrossRef]  

15. R. Aguero-Santacruz and D. Bermudez, “Negative frequencies in pulse propagation equations and the double analytic signal,” New J. Phys. 25(10), 103045 (2023). [CrossRef]  

16. G. Moore, “Quantum theory of the electromagnetic field in a variable-length one-dimensional cavity,” J. Math. Phys. 11(9), 2679–2691 (1970). [CrossRef]  

17. J. T. Mendonça and P. K. Shukla, “Time refraction and time reflection: two basic concepts,” Phys. Scr. 65(2), 160–163 (2002). [CrossRef]  

18. D. L. Sounas and A. Alù, “Non-reciprocal photonics based on time modulation,” Nat. Photonics 11(12), 774–783 (2017). [CrossRef]  

19. D. Hashimshony, A. Zigler, and K. Papadopoulos, “Conversion of Electrostatic to Electromagnetic Waves by Superluminous Ionization Fronts,” Phys. Rev. Lett. 86(13), 2806–2809 (2001). [CrossRef]  

20. N. N. Rozanov, “Superluminal parametric Doppler effect in insulators and in an electron-positron vacuum,” JETP Lett. 95(12), 609–612 (2012). [CrossRef]  

21. H. Moussa, G. Xu, S. Yin, et al., “Observation of temporal reflection and broadband frequency translation at photonic time interfaces,” Nat. Phys. 19(6), 863–868 (2023). [CrossRef]  

22. E. Lustig, O. Segal, S. Saha, et al., “Time-refraction optics with single cycle modulation,” Nanophotonics 12(12), 2221–2230 (2023). [CrossRef]  

23. E. Rubino, J. McLenaghan, S. C. Kehr, et al., “Negative-Frequency Resonant Radiation,” Phys. Rev. Lett. 108(25), 253901 (2012). [CrossRef]  

24. J. T. Mendonca, A. Guerreiro, and A. M. Martins, “Quantum theory of time refraction,” Phys. Rev. A 62(3), 033805 (2000). [CrossRef]  

25. S. Finazzi and I. Carusotto, “Quantum vacuum emission in a nonlinear optical medium illuminated by a strong laser pulse,” Phys. Rev. A 87(2), 023803 (2013). [CrossRef]  

26. M. Jacquet and F. König, “Quantum vacuum emission from a refractive-index front,” Phys. Rev. A 92(2), 023851 (2015). [CrossRef]  

27. J. E. Vázquez-Lozano and I. Liberal, “Shaping the quantum vacuum with anisotropic temporal boundaries,” Nanophotonics 12(3), 539–548 (2023). [CrossRef]  

28. V. Pacheco-Pena and N. Engheta, “Antireflection temporal coatings,” Optica 7(4), 323 (2020). [CrossRef]  

29. I. Liberal, J. E. Vázquez-Lozano, and V. Pacheco-Pe na, “Quantum antireflection temporal coatings: Quantum state frequency shifting and inhibited thermal noise amplification,” Laser Photonics Rev. 17(9), 2200720 (2023). [CrossRef]  

30. S. A. Horsley and J. B. Pendry, “Quantum electrodynamics of time-varying gratings,” Proc. Natl. Acad. Sci. U.S.A. 120(36), e2302652120 (2023). [CrossRef]  

31. A. A. Svidzinsky, A. Azizi, J. S. Ben-Benjamin, et al., “Causality in quantum optics and entanglement of Minkowski vacuum,” Phys. Rev. Res. 3(1), 013202 (2021). [CrossRef]  

32. S. W. Hawking, “BH explosions?” Nature (London) 248(5443), 30–31 (1974). [CrossRef]  

33. S. W. Hawking, “Particle creation by black holes,” Commun. Math. Phys. 43(3), 199–220 (1975). [CrossRef]  

34. A. A. Svidzinsky, “Nonthermal radiation of evaporating black holes,” Phys. Rev. D 107(2), 024038 (2023). [CrossRef]  

35. L. A. Ostrovskii, “Some “moving boundaries paradoxes” in electrodynamics,” Usp. Fiz. Nauk 116(6), 315 (1975) [Sov. Phys. Usp. 18, 452 (1975)]. [CrossRef]  

36. B. S. DeWitt, “Quantum field theory in curved spacetime,” Phys. Rep. 19(6), 295–357 (1975). [CrossRef]  

37. S. A. Fulling and P. C. W. Davies, “Radiation from a moving mirror in two dimensional space-time: Conformal anomaly,” Proc. R. Soc. Lond. A 348(1654), 393–414 (1976). [CrossRef]  

38. P. C. W. Davies and S. A. Fulling, “Radiation from moving mirrors and from black holes,” Proc. R. Soc. A 356(1685), 237–257 (1977).

39. P. Candelas and D. Deutsch, “On the vacuum stress induced by uniform acceleration or supporting the ether,” Proc. R. Soc. A 354(1676), 79–99 (1977). [CrossRef]  

40. W. G. Unruh and R. M. Wald, “What happens when an accelerating observer detects a Rindler particle,” Phys. Rev. D 29(6), 1047–1056 (1984). [CrossRef]  

41. W. Rindler, “Kruskal space and the uniformly accelerated frame,” Am. J. Phys. 34(12), 1174–1178 (1966). [CrossRef]  

42. R. J. Glauber, “Amplifiers, Attenuators and Schrödinger’s Cat,” Ann. N. Y. Acad. Sci. 480(1), 336–372 (1986). [CrossRef]  

43. F. Ullinger, M. Zimmermann, and W. P. Schleich, “The logarithmic phase singularity in the inverted harmonic oscillator,” AVS Quantum Sci. 4(2), 024402 (2022). [CrossRef]  

44. W. G. Unruh, “Experimental Black-Hole Evaporation?” Phys. Rev. Lett. 46(21), 1351–1353 (1981). [CrossRef]  

45. G. E. Volovik, “Simulation of a Panlevé -Gullstrand black hole in a thin 3He-A film,” JETP Lett. 69(9), 705–713 (1999). [CrossRef]  

46. U. Leonhardt and P. Piwnicki, “Relativistic Effects of Light in Moving Media with Extremely Low Group Velocity,” Phys. Rev. Lett. 84(5), 822–825 (2000). [CrossRef]  

47. I. Brevik and G. Halnes, “Light rays at optical black holes in moving media,” Phys. Rev. D 65(2), 024005 (2001). [CrossRef]  

48. U. Leonhardt, “A laboratory analogue of the event horizon using slow light in an atomic medium,” Nature (London) 415(6870), 406–409 (2002). [CrossRef]  

49. R. Schützhold, G. Plunien, and G. Soff, “Dielectric Black Hole Analogs,” Phys. Rev. Lett. 88(6), 061101 (2002). [CrossRef]  

50. V. A. De Lorenci, R. Klippert, and Yu. N. Obukhov, “Optical black holes in moving dielectrics,” Phys. Rev. D 68(6), 061502 (2003). [CrossRef]  

51. R. Schützhold and W. G. Unruh, “Hawking Radiation in an Electromagnetic Waveguide?” Phys. Rev. Lett. 95(3), 031301 (2005). [CrossRef]  

52. T. G. Philbin, C. Kuklewicz, S. Robertson, et al., “Fiber-Optical Analog of the Event Horizon,” Science 319(5868), 1367–1370 (2008). [CrossRef]  

53. S. L. Cacciatori, F. Belgiorno, V. Gorini, et al., “Spacetime geometries and light trapping in travelling refractive index perturbations,” New J. Phys. 12(9), 095021 (2010). [CrossRef]  

54. F. Belgiorno, S. L. Cacciatori, G. Ortenzi, et al., “Dielectric black holes induced by a refractive index perturbation and the Hawking effect,” Phys. Rev. D 83(2), 024015 (2011). [CrossRef]  

55. M. F. Linder, R. Schützhold, and W. G. Unruh, “Derivation of Hawking radiation in dispersive dielectric media,” Phys. Rev. D 93(10), 104010 (2016). [CrossRef]  

56. M. J. Jacquet and F. König, “Analytical description of quantum emission in optical analogs to gravity,” Phys. Rev. A 102(1), 013725 (2020). [CrossRef]  

57. J. Drori, Y. Rosenberg, D. Bermudez, et al., “Observation of stimulated Hawking radiation in an optical analogue,” Phys. Rev. Lett. 122(1), 010404 (2019). [CrossRef]  

58. Y. E. Lozovik, V. G. Tsvetus, and E. A. Vinogradov, “Femtosecond parametric excitation of electromagnetic field in a cavity,” Pis’ma Zh. Éksp. Teor. Fiz. 61(9), 711–716 (1995).

59. C. M. Wilson, G. Johansson, A. Pourkabirian, et al., “Observation of the dynamical Casimir effect in a superconducting circuit,” Nature (London) 479(7373), 376–379 (2011). [CrossRef]  

60. B. H. Schneider, A. Bengtsson, I. M. Svensson, et al., “Observation of Broadband Entanglement in Microwave Radiation from a Single Time-Varying Boundary Condition,” Phys. Rev. Lett. 124(14), 140503 (2020). [CrossRef]  

61. P. Lähteenmäki, G. S. Paraoanu, J. Hassel, et al., “Dynamical Casimir effect in a Josephson metamaterial,” Proc. Natl. Acad. Sci. U.S.A. 110(11), 4234–4238 (2013). [CrossRef]  

62. S. Vezzoli, A. Mussot, N. Westerberg, et al., “Optical analogue of the dynamical Casimir effect in a dispersion-oscillating fibre,” Commun. Phys. 2(1), 84 (2019). [CrossRef]  

63. R. H. Kingston, “Parametric amplification and oscillation at optical frequencies,” Proc. IRE 50, 472 (1962).

64. N. M. Kroll, “Parametric Amplification in Spatially Extended Media and Application to the Design of Tuneable Oscillators at Optical Frequencies,” Phys. Rev. 127(4), 1207–1211 (1962). [CrossRef]  

65. S. A. Akhmanov and R. V. Khokhlov, “Concerning one possibility of amplitication of light waves,” Zh. Eksp. Teor. Fiz. 4(3), 351 (1962) [Sov. Phys. JETP 16, 252 (1963)].

66. J. A. Giordmaine and R. C. Miller, “Tunable Coherent Parametric Oscillation in LiNbO3 at Optical Frequencies,” Phys. Rev. Lett. 14(24), 973–976 (1965). [CrossRef]  

67. W. R. Bosenberg, W. S. Pelouch, and C. L. Tang, “High-efficiency and narrow-linewidth operation of a two-crystal β-BaB2O4 optical parametric oscillator,” Appl. Phys. Lett. 55(19), 1952–1954 (1989). [CrossRef]  

68. L. Parker, “Particle creation in expanding universes,” Phys. Rev. Lett. 21(8), 562–564 (1968). [CrossRef]  

69. L. Parker, “Particle creation in isotropic cosmologies,” Phys. Rev. Lett. 28(11), 705–708 (1972). [CrossRef]  

70. L. Parker, “Particle creation and particle number in an expanding universe,” J. Phys. A: Math. Theor. 45(37), 374023 (2012). [CrossRef]  

71. P. J. E. Peebles and J. T. Yu, “Primeval adiabatic perturbation in an expanding universe,” Astrophys. J. 162, 815 (1970). [CrossRef]  

72. L. P. Grishchuk and Y. V. Sidorov, “Squeezed quantum states of relic gravitons and primordial density fluctuations,” Phys. Rev. D 42(10), 3413–3421 (1990). [CrossRef]  

73. V. F. Mukhanov, H. A. Feldman, and R. H. Brandenberger, “Theory of cosmological perturbations,” Phys. Rep. 215(5-6), 203–333 (1992). [CrossRef]  

74. M. Lampe, E. Ott, and J. H. Waler, “Interaction of electromagnetic waves with a moving ionization front,” Phys. Fluids 21(1), 42–54 (1978). [CrossRef]  

75. B. M. Bolotovskii and S. N. Stolyarov, “Reflection of light from a moving mirror and related problems,” Sov. Phys. Usp. 32(9), 813–827 (1989). [CrossRef]  

76. N. N. Rozanov, “Subluminal and superluminal parametric doppler effects in the case of light reflection from a moving smooth medium inhomogeneity,” J. Exp. Theor. Phys. 115(6), 962–968 (2012). [CrossRef]  

77. Z. Bu, B. Shen, S. Huang, et al., “Light reversing and folding based on a superluminal flying mirror in a plasma with increasing density,” Plasma Phys. Control. Fusion 58(7), 075008 (2016). [CrossRef]  

Data availability

No data were generated or analyzed in the presented research.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. Classical electrodynamics picture of reflection and refraction of light from a space boundary at $x=0$ between two media with refractive indices $n_{1}$ and $n_{2}$ in 1+1 dimensions (a) and a time boundary at $t=0$ (b). During reflection and refraction from the space (time) boundary the light energy (momentum) is conserved. When a pulse of frequency $\nu$ passes through a time boundary it splits into two pulses propagating in the opposite directions with frequencies $\pm n_{1}\nu /n_{2}$.
Fig. 2.
Fig. 2. (a) A right-moving Gaussian pulse passes through a superluminal boundary between two dielectric media. The latter are at rest and only the interface is moving. (b) The interface splits the pulse into two transmitted Gaussian pulses moving in the opposite directions, and there is no reflected pulse (that is there is no pulse on the left side of the boundary after interaction). In the classical description the pulse $\nu _{3}$ has negative frequency and propagates backward in time.
Fig. 3.
Fig. 3. A pulse is moving through a medium whose refractive index suddenly changes from $n_{1}$ to $n_{2}$. This change splits the pulse into two pulses propagating in the opposite directions.
Fig. 4.
Fig. 4. Interface between two static media moving with a constant velocity that is subluminal (superluminal) for the medium 1 (2) mimics a BH event horizon. Orthogonal mode functions of the field obeying the proper boundary conditions at the interface with one ingoing and two transmitted waves (Eq. (refk1a)), and three outgoing waves (Eq. (refk1d)) are sketched in (a) and (b) respectively.
Fig. 5.
Fig. 5. Ingoing wave $A_{1}$ propagates through a superluminal interface between two dielectric media and splits into two transmitted waves $A_{2}$ and $A_{3}$ without reflection.

Equations (97)

Equations on this page are rendered with MathJax. Learn more.

ϕ k ( t , r ) = e i c k n t + i k r ,
ν = c k n V k 1 V 2 c 2 .
1 c 2 2 ϕ t 2 2 ϕ x 2 = 0.
ϕ 1 , ϕ 2 = i 2 c ( ϕ 1 ϕ 2 t ϕ 1 t ϕ 2 ) d x .
ϕ 1 , ϕ 1 = ϕ 1 , ϕ 1 , ϕ 1 , ϕ 1 = ϕ 1 , ϕ 1 = 0.
ϕ ( t , x ) = e i ν t + i k x
Φ ^ ( t , x ) = ϕ ( t , x ) a ^ + ϕ ( t , x ) a ^ .
Φ ^ ( t , x ) = ϕ ( t , x ) a ^ + ϕ ( t , x ) a ^ ,
ϕ ( t , x ) = e i ν ( t x / c ) e Ω 2 ( t x / c ) 2 + A e i ν ( t + x / c ) e Ω 2 ( t + x / c ) 2
E ( t , x ) = e i ν ( t ± n 1 c x ) e Ω 2 ( t ± n 1 c x ) 2
E ( t , x ) = A 2 e i ν 2 ( t n 2 c x ) e Ω 2 2 ( t n 2 c x ) 2 + A 3 e i ν 3 ( t + n 2 c x ) e Ω 3 2 ( t + n 2 c x ) 2 ,
ν 2 = 1 n 1 V c 1 + n 2 V c ν , ν 3 = 1 n 1 V c 1 n 2 V c ν ,
A 3 e i | ν 3 | ( t n 2 c x ) e Ω 3 2 ( t n 2 c x ) 2
E ( t , x ) = { f ( t , x ) , x < V t C 2 f 2 ( t , x ) + C 3 f 3 ( t , x ) , x > V t ,
E 1 ( t , x ) = { f 1 ( t , x ) , x < V t C 2 f 3 ( t , x ) + C 3 f 2 ( t , x ) , x > V t ,
f ( t , x ) = Ω ν n 1 e i ν ( t n 1 c x ) e Ω 2 ( t n 1 c x ) 2 ,
f 1 ( t , x ) = Ω 1 ν 1 n 1 e i ν 1 ( t + n 1 c x ) e Ω 1 2 ( t + n 1 c x ) 2 ,
f 2 ( t , x ) = Ω 2 ν 2 n 2 e i ν 2 ( t n 2 c x ) e Ω 2 2 ( t n 2 c x ) 2 ,
f 3 ( t , x ) = Ω 3 ν 3 n 2 e i ν 3 ( t + n 2 c x ) e Ω 3 2 ( t + n 2 c x ) 2
ν 1 ν = Ω 1 Ω = | n 1 V c + 1 n 1 V c 1 | ,
ν 2 ν = Ω 2 Ω = | n 1 V c + 1 n 2 V c + 1 | ,
ν 3 ν = Ω 3 Ω = | n 1 V c + 1 n 2 V c 1 | ,
C 2 = n 1 + n 2 2 n 1 n 2 , C 3 = n 2 n 1 2 n 1 n 2
C 2 2 C 3 2 = 1.
E 2 ( t , x ) = { C 2 f ( t , x ) C 3 f 1 ( t , x ) , x < V t f 2 ( t , x ) , x > V t ,
E 3 ( t , x ) = { C 2 f 1 ( t , x ) C 3 f ( t , x ) , x < V t f 3 ( t , x ) , x > V t ,
a ^ = C 2 b ^ 2 + C 3 b ^ 3 , a ^ = C 2 b ^ 2 + C 3 b ^ 3 ,
a ^ 1 = C 2 b ^ 3 + C 3 b ^ 2 , a ^ 1 = C 2 b ^ 3 + C 3 b ^ 2 ,
b ^ 2 = C 2 a ^ C 3 a ^ 1 , b ^ 2 = C 2 a ^ C 3 a ^ 1 ,
b ^ 3 = C 2 a ^ 1 C 3 a ^ , b ^ 3 = C 2 a ^ 1 C 3 a ^ .
N 2 = b ^ 2 b ^ 2 = N + C 3 2 ( N + N 1 + 1 ) ,
N 3 = b ^ 3 b ^ 3 = N 1 + C 3 2 ( N + N 1 + 1 ) ,
N 2 = N 3 = ( n 2 n 1 ) 2 4 n 1 n 2 .
| 0 a 0 a 1 = 1 γ 2 e γ b ^ 2 b ^ 3 | 0 b 2 0 b 3 ,
γ = C 3 C 2 = n 2 n 1 n 2 + n 1 ,
ν 3 ν 2 = | 1 + n 2 V c 1 n 2 V c |
ρ ^ b 2 = Tr b 3 ( | 0 a 0 a 1 0 a 0 a 1 | ) = ( 1 γ 2 ) m = 0 γ 2 m | m m | ,
W W 0 = | 1 + n 1 V c 1 + n 2 V c | ( n 1 + n 2 ) 2 4 n 1 n 2 + | 1 + n 1 V c 1 n 2 V c | ( n 1 n 2 ) 2 4 n 1 n 2 .
ν 3 1 + n 1 n 2 | 1 n 2 V c | ν , τ 3 | 1 n 2 V c | 1 + n 1 n 2 τ .
N 2 = N + C 3 2 ( N + 1 ) ,
N 3 = C 3 2 ( N + 1 ) .
x + u t = 0 , u > c ,
t = t β x / c 1 β 2 , x = x β c t 1 β 2 ,
ν 2 = ν 3 = n 1 n 2 ν , Ω 2 = Ω 3 = n 1 n 2 Ω ,
H ^ = k ν k a ^ k a ^ k ,
ϕ k ( t , r ) = e i ν k t + i k r + γ e i ν k t + i k r ,
a ^ k = b ^ k γ b ^ k 1 γ 2 , a ^ k = b ^ k γ b ^ k 1 γ 2 .
H ^ = k ν k ( 1 + γ 2 1 γ 2 b ^ k b ^ k + 2 γ 1 γ 2 ( b ^ k b ^ k + b ^ k b ^ k ) ) .
| 0 M = k 1 γ 2 e γ b ^ k b ^ k | 0 b ,
e i ν ( t n c x ) + n 2 n n 2 + n e i ν ( t + n c x )
2 n 2 n 2 ( n + n 2 ) e i ν ( n n 2 t n c x ) .
n 2 = 1 + γ 1 γ n ,
ϕ k ( t , r ) ( 1 γ ) 2 1 + γ e i ν k t + i k r ,
E ( t , x ) = { f ( t , x ) , x < V t C 2 f 2 ( t , x ) C 3 f 3 ( t , x ) , x > V t ,
a ^ = C 2 b ^ 2 C 3 b ^ 3 ,
E 2 ( t , x ) = { C 2 f ( t , x ) + C 3 f 1 ( t , x ) , x < V t f 2 ( t , x ) + 2 C 3 f 4 ( t , x ) , x > V t ,
E 3 ( t , x ) = { C 2 f 1 ( t , x ) + C 3 f ( t , x ) , x < V t f 3 ( t , x ) + 2 C 2 f 4 ( t , x ) , x > V t ,
( C 2 b ^ 2 + C 3 b ^ 3 ) ( C 2 b ^ 2 C 3 b ^ 3 ) = 0 ,
N 2 = C 3 2 C 2 2 ( N 3 + 1 ) ,
E 1 ( t , x ) = { f 1 ( t , x ) , x < V t C 2 f 3 ( t , x ) C 3 f 2 ( t , x ) + 2 f 4 ( t , x ) ,   x > V t .
a ^ 1 = C 2 b ^ 3 C 3 b ^ 2 .
N 1 N 2 .
N 1 + N 2 = N 3 .
N 1 = N 2 = N 3 2 = C 3 2 = ( n 2 n 1 ) 2 4 n 1 n 2 .
t ( τ ) = c a sinh ( a τ c ) , x ( τ ) = c 2 a cosh ( a τ c ) .
F 1 Ω ( t , x ) = | t + x / c | i Ω 2 Ω sinh ( π Ω ) { e π Ω 2 , t + x / c > 0 e π Ω 2 , t + x / c < 0 ,
F 2 Ω ( t , x ) = | t + x / c | i Ω 2 Ω sinh ( π Ω ) { e π Ω 2 , t + x / c > 0 e π Ω 2 , t + x / c < 0 ,
t ( τ , x ¯ ) = c a e a x ¯ / c 2 sinh ( a τ c ) , x ( τ , x ¯ ) = c 2 a e a x ¯ / c 2 cosh ( a τ c ) ,
F 1 Ω ( τ , x ¯ ) e i a Ω c ( τ + x ¯ c ) , F 2 Ω ( τ , x ¯ ) e i a Ω c ( τ + x ¯ c ) .
H ^ = ω ( b ^ 1 b ^ 1 b ^ 2 b ^ 2 ) + g ( b ^ 1 b ^ 2 + b ^ 1 b ^ 2 ) ,
b ^ 1 b ^ 1 = b ^ 2 b ^ 2 = sinh 2 ( g t ) ,
| ψ ( t ) = 1 cosh ( g t ) e i tanh ( g t ) b ^ 1 b ^ 2 | 0 ,
ν d = ν 1 + ν 2 .
n ^ × ( E 2 E 1 ) = V c ( B 2 B 1 ) ,
n ^ × ( H 2 H 1 ) = V c ( D 2 D 1 ) ,
c 1 = c ε 1 μ 1 , c 2 = c ε 2 μ 2
× E = 1 c B t , × H = 1 c D t ,
D = ε E , B = μ H .
n ^ × ( n ^ × ( E 2 E 1 ) ) = V 2 c 2 ( D 2 D 1 ) ,
n ^ ( n ^ ( E 2 E 1 ) ) ( E 2 E 1 ) = V 2 c 2 ( D 2 D 1 ) .
E 2 E 1 = V 2 c 2 ( D 2 D 1 ) .
( 1 V 2 c 2 ε 1 ) E 1 = ( 1 V 2 c 2 ε 2 ) E 2 .
E 2 E 1 + V c ( B 2 B 1 ) = 0.
( 2 t 2 c 2 ε 2 x 2 ) E = 0
E = z ^ e i ν ( t ± ε c x )
B = ε y ^ e i ν ( t ± ε c x ) .
E 1 = A 1 e i ν ( t ± ε 1 x c ) , B 1 = ε 1 E 1 ,
E 2 = A 2 e i ν 2 ( t ε 2 c x ) + A 3 e i ν 3 ( t + ε 2 c x ) , B 2 = ε 2 A 2 e i ν 2 ( t ε 2 c x ) ε 2 A 3 e i ν 3 ( t + ε 2 c x ) .
( 1 ε 1 V 2 c 2 ) A 1 e i ν t ( 1 ε 1 V c ) = ( 1 ε 2 V 2 c 2 ) ( A 2 e i ν 2 t ( 1 + ε 2 V c ) + A 3 e i ν 3 t ( 1 ε 2 V c ) ) ,
( 1 ε 1 V c ) A 1 e i ν t ( 1 ε 1 V c ) = ( 1 + ε 2 V c ) A 2 e i ν 2 t ( 1 + ε 2 V c ) + ( 1 ε 2 V c ) A 3 e i ν 3 t ( 1 ε 2 V c ) .
ν 2 = 1 n 1 V c 1 + n 2 V c ν , ν 3 = 1 n 1 V c 1 n 2 V c ν ,
A 2 = 1 2 ( 1 n 1 n 2 ) ν 2 ν A 1 ,
A 3 = 1 2 ( 1 ± n 1 n 2 ) ν 3 ν A 1 ,
E 1 = A 1 e i ν ( t ± n 1 c x ) e Ω 2 ( t ± n 1 c x ) 2 , B 1 = n 1 E 1 ,
E 2 = A 2 e i ν 2 ( t n 2 c x ) e Ω 2 2 ( t n 2 c x ) 2 + A 3 e i ν 3 ( t + n 2 c x ) e Ω 3 2 ( t + n 2 c x ) 2 ,
B 2 = ε 2 A 2 e i ν 2 ( t n 2 c x ) e Ω 2 2 ( t n 2 c x ) 2 ε 2 A 3 e i ν 3 ( t + n 2 c x ) e Ω 3 2 ( t + n 2 c x ) 2 .
Ω 2 = 1 n 1 V c 1 + n 2 V c Ω , Ω 3 = 1 n 1 V c 1 n 2 V c Ω .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.