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

Manipulating quantum interference of dressed photon fields

Open Access Open Access

Abstract

Through quantum electrodynamics (QED) we investigate the interactions between a three-level atom and two photon fields under perturbation limit. The dispersion relation and (relative) transmission of the probe photons are obtained by calculating the corresponding Feynman diagrams. The quantum interference in this three-level system such as Fano resonance and electromagnetically induced transparency (EIT) can be tuned by varying the intensities of the control and probe beams. Moreover, by considering that the control beam with periodic modulation, that is, the so-called Landau-Zener-Stückelberg (LZS) type source, the accumulated phase after Landau-Zener transitions is found to show the alternating Fano (EIT) lineshapes in the transmission of the probe photons. We further find that the transmissions can become almost stationary in addition to a wide EIT window in time even though the control beam is a LZS-type oscillating source.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Since the discovery of quantum mechanics, the interactions between (artificial) atoms and photons have become an important research area in both fundamental physics and practical applications [16]. Quantum interference [710,14,15] especially plays a substantial role for the diversified manifestations of the light-matter interactions [1618]. In principle, enhancement, reduction, or no change of transition probability may happen due to the interference between amplitudes from alternative pathways. The photon fields can thus control the optical response of the matter and the quantum interference effect can occur as phenomena of the electromagnetically induced transparency (EIT) [813] and the Fano resonance [14,15]. These abnormal optical responses resulting from quantum interference have been observed in atomic and molecular systems [19,20], photonic crystals [21], solid-state systems [22], microresonators [23], metamaterials [24] and electronic systems [25]. How to tune the quantum coherence has then become a very important issue. In the past few decades, the Landau–Zener–Stückelberg (LZS) [26] interference has received a lot of attention for serving as a feasible way to control the quantum interference by periodically changing the relative phase between quantum states [27].

In this paper, we investigate the interaction between a three-level atom and two photon fields with different frequencies as depicted in Fig. 1. We assume that the intensity of one light field (control beam, $\Omega _c$) is much stronger than the other (probe beam, $\Omega _p$). Due to interactions between the strong control beam and the atom, the interacting atom becomes a $``\rm {dressed}"$ one. Then it would have great impact on the quantum states of the probe beam which becomes a $``\rm {dressed}"$ field through interacting with the dressed atom. Henceforth, the dispersion relation and associated physical behaviors of the dressed probe photon would be quite different from those of free photons [28]. Through the well-known Feynman’s diagrammatic method in field theory of QED, we calculate the dispersion relation and transmission of the dressed probe field. We find significant tunable quantum interferences when the frequency of the control beam is around the resonant regimes of the atom. For example, the asymmetric patterns and locations of Fano resonances and the widths of EIT windows ($\propto \Omega _c^2$) can be adjusted by tuning the intensities of the two photon fields, and the frequency of the control field. With the amplitude modulated control beam, i.e., the so called Landau-Zener-Stückelberg (LZS) type source, we find that the signatures of quantum coherence shows oscillating behavior in the transmission of the probe photon. We further find that the destructive interference from a sequence of Landau-Zener transitions can lead to a wide EIT window in time. In this paper, we use natural units by putting $\hbar = c= 1$ as is widely adopted in field theory literature, $e.g.$, Ref. [29].

 figure: Fig. 1.

Fig. 1. A schematic diagram of the bare-state $\Lambda$-type three-level EIT/Fano system: ground state $|1\rangle$; metastable state $|2\rangle$; excited state $|3\rangle$. The probe field of frequency $\omega _p$ governs the $|1\rangle \rightarrow |3\rangle$ transition with the electric current density $J_p$, and the control field of frequency $\omega _c$ drives the $|2\rangle \leftrightarrow |3\rangle$ transition with the electric current density $J_c$.

Download Full Size | PDF

2. Model formulation

Here we assume the EM waves to propagate in $x$-axis and they are uniform in the transverse. Therefore, in radiation gauge ($\nabla \cdot \vec {A}=0$), the vector potential $\vec {A}$ can be written as $\vec {A}(x)=(0,A_y(x),A_z(x))$. The photonic annihilation operator $A$, and the photonic creation operator ${A}^{\dagger }$ are defined as,

$$A=(A_y+iA_z)/\sqrt{2}, \,\, {A}^{{\dagger}}=(A_y-iA_z)/\sqrt{2},$$
and the field energy density $[\dot {\vec {A}}^2+(\nabla \times \vec {A})^2]/2$ can be written as $(\dot {{A}^{\dagger }}\dot {{A}}+\nabla {A}^{\dagger }\cdot \nabla {A})$. Then the Hamiltonian for the free photon is
$$H_{em} = \int dx\, (\dot{{A}^{{\dagger}}}\dot{{A}}+\nabla {A}^{{\dagger}}\cdot\nabla {A} +J^*\cdot A+J\cdot A^\dagger) ,$$
where $J$ $(J^*)$ is the amplitude of photon-source (sink) corresponding to the external electric current density. Note that in the following discussions, $J$ contains two components, $J=J_p+J_c$, while $J_c$ stands for the control beam, and $J_p$ stands for the probe beam, respectively. The propagator of free photon is
$$\widetilde{G}_{(0)}( k ,\omega ) = \frac{i}{\omega^{2}- k^{2}+i\epsilon }$$
with $\epsilon$ an infinitesimal positive number. For a three-level atomic (3LA) system, the Hamiltonian is
$${\mathcal{H}}_{3LA}(\{c \}) = \sum_{\alpha=1,2,3} (\varepsilon_\alpha-i\delta) c_{\alpha}^{{\dagger}}c_{\alpha} +\frac{U}{2}\sum_{\alpha>\beta}c_{\alpha}^{{\dagger}}c_{\beta}^{{\dagger}}c_{\alpha}c_{\beta},$$
where $c_{\alpha }$ is the field operator of electron in the $\alpha$-th energy level ($\varepsilon _{\alpha }$), and $\delta$ is a small and positive number with its inverse corresponding to the relaxation time. The $bare$ propagator of the electron field $c_\alpha$ is
$$\widetilde{\Delta}^{({\pm})}_{(0),\alpha} (\omega) = \frac{i}{\omega-\varepsilon_{\alpha} \pm i\delta}, \;\alpha= 1,2,3,$$
where + (-) sign is for electron moving forward (backward) in time. The four-body term in Eq. (4) describes the on-site hard-core interaction between electrons of two different energy levels to avoid any two electrons occupying the atom at the same time. Here U is the strength of the strong repulsive interaction between particles (electrons). Initially, it is treated as a finite positive number in the calculations of any physical quantity involving the four-body interaction between particles; and it will be set as infinity at the end of calculations for the hard-core repulsion. The interaction between the atom and photons is
$${\mathcal{H}}_{int}(\{c \},A) = \,g\,[A( x_a)c_{\alpha}^{{\dagger}}c_{\beta}+ \,\, h.c.],$$
where $x_a$ is the location of the atom. The Hamiltonian ${\mathcal {H}}_{int}(\{c \},A)$ is the light-matter interaction adopted in QED in which a field with lower energy can absorb a photon to become a field with higher energy, and vice versa; the coupling constant $g\sim \sqrt {e^2/\hbar c} \sim 1/\sqrt {137}$ is the coupling between $bare$ electrons and $bare$ photons as is widely adopted in field theory literatures, $e.g.$, Ref. [29]. Please note that the collective behavior of couplings between bare electrons and photons in different geometry or confinements results in different (effective) coupling strength in quantum optics.

t-matrix. To handle the hard-core interaction in Eq. (4), we can follow the calculations done in Ref. [30] by using the method of binary collision [31,32]. We add up all the repeated scatterings between different fields $c_\alpha$ and $c_\beta$ (ladder diagrams in Fig. 2(a)) to get a finite effective coupling ($u_0$) between them at low energy [30]. For brevity, we shall not repeat calculations parallel to those shown in Ref. [30]; and similar results are obtained: The effective coupling $u_0=\frac {1}{2}(\varepsilon _{\alpha }+\varepsilon _{\beta }-P_0)$ with the assumption that the total collision energy $P_0\ll (\varepsilon _{\alpha }+\varepsilon _{\beta })$ for long wavelength scatterings. The modification to the self-mass is $\Sigma _\alpha (\omega )=\frac {1}{4}(\varepsilon _{\alpha }-\omega )+$ constant $C_1$. It follows that all three (bare) energy levels $\varepsilon _{\alpha }$’s are shifted by the same amount of constant $\frac {4}{5} C_1$ due to the hard-core interaction, and this can be incorporated into the theory by redefining the shifted energy levels as the renormalized (physical) energy levels $\nu _{\alpha }$ as what is obtained in Ref. [30].

 figure: Fig. 2.

Fig. 2. (a) Diagrammatic expansion of the effective coupling $u_0$ which is the sum of the repeated and continuous scatterings between two different fields $c_\alpha$ and $c_\beta$. The external legs are only for the eyes, the black dot $``\bullet "$ is $U$ in Eq. (4), and the internal double-line or single line are only used to represent different index of the electronic bare propagator of field $c_\gamma$, $\widetilde \Delta ^{(\pm )}_{(0),\gamma }$’s, $\gamma$=1,2,3 (Eq. (5)). (b) Diagrammatic expansions of the self-mass $\Sigma ^{(2)}_{\alpha }(\omega )$ (Eq. (9)) for the renormalized propagators $\widetilde \Delta ^{(\pm )}_{(2),\alpha }$ (Eq. (8)) of the dressed electronic field $c_\alpha$. The internal straight line represents electron propagator $\widetilde \Delta ^{(\pm )}_{(1),\alpha }$ (Eq. (7)). The curvy line represents the propagator of free photon $\widetilde {G}_{(0)}$ (Eq. (3)). (c), (d) Diagrammatic representations of self-masses $\Pi ^{(2)}(\omega )$ (Eq. (11)) and $\Pi ^{(4)}(\omega )$ (Eq. (12)) of the renormalized propagator of the dressed probe photon, $\widetilde {G}(k,\omega )$ (Eq. (10)) are shown in (c) and (d), respectively. The solid black lines represent the renormalized electronic propagators $\widetilde \Delta ^{(\pm )}_{(2),\alpha }$, and the (dashed) curvy external legs are only for the eyes.

Download Full Size | PDF

Electron propagator. It follows that the energy differences between atomic energy levels are unchanged. In the mean time, the renormalized propagators of the electron field $c_\alpha$ become

$$\widetilde{\Delta}^{({\pm})}_{(1),\alpha} (\omega) = \frac{i}{\omega-\nu_\alpha \pm i\delta},\; \alpha= 1,2,3,$$
with $\nu _\alpha$ the physical energy of the $\alpha$-th level without the light-matter interaction. After including the interaction between the atom and the strong control beam with frequency $\omega _c$ and amplitude $J_c$, the further renormalized propagator of the electron field $c_\alpha$ satisfies
$$\widetilde\Delta^{({\pm})}_{(2),\alpha} (\omega;J_c,\omega_c) ^{{-}1} =\widetilde\Delta^{({\pm})}_{(1),\alpha} (\omega) ^{{-}1}- \frac{\Sigma^{(2)}_{\alpha} (\omega;J_c,\omega_c)}{i},$$
where
$$\begin{aligned} &\Sigma^{(2)}_{\alpha}(\omega;J_c,\omega_c)\\ = &\frac{g^2}{i}|\langle\tilde{ A}(\omega_c,J_c) \rangle|^2\, \sum_{\beta\neq\alpha} \widetilde\Delta^+_{(1),\beta} ( \omega-\textrm{sgn}(\alpha-\beta)\,{\omega_c} )\\ = &\frac{g^2|J_c|^2 }{4 \omega_c^2 }\sum_{\beta\neq\alpha} \frac{1}{\omega-\textrm{sgn}(\alpha-\beta)\,\omega_c-\nu_\alpha +i\delta}, \end{aligned}$$
is the self-mass of the $\alpha$-th field (Fig. 2(b)) responsible for the interaction with the control beam, and $|\langle \tilde { A}(\omega _c, J_c) \rangle |^2 \equiv |\int \frac {d k^{\prime }}{ 2\pi } \frac {d \omega ^{\prime }}{ 2\pi } \, \widetilde {G}_{(0)}(k^{\prime },\omega ^{\prime }) \,\tilde {J}(k^{\prime },\omega ^{\prime })|^2 = |\frac {J_c }{2 \omega _c }|^2$.

Photon propagator. Under the illumination of the control beam with frequency $\omega _c$ and amplitude $J_c$, the propagator of the dressed photon $\widetilde {G}( k,\omega )$ from the probe beam can be calculated from the Feynman diagrams (Figs. 2(c) and 2(d)), and it satisfies the following Dyson’s equation,

$$\widetilde{G}(k,\omega )^{{-}1} =\widetilde{G}_{(0)}( k,\omega )^{{-}1} -i^{{-}1}[\Pi^{(2)}(\omega)+\Pi^{(4)}(\omega)], \\$$
and
$$\begin{aligned} &\Pi^{(2)}(\omega)\\ = &\frac{-g^2}{2} \sum_{\alpha>\beta} \int \frac{d\omega'}{2\pi}\, \widetilde\Delta^{(+)}_{(2),\,\alpha}(\omega+\omega';J_c,\omega_c)\,\widetilde\Delta^{(-)}_{(2),\,\beta}(\omega';J_c,\omega_c), \end{aligned}$$
$$\begin{aligned} &\Pi^{(4)}(\omega)\\ = &\frac{g^4}{2} |\langle\tilde{ A}(\omega,J_p) \rangle|^2 \int \frac{d\omega'}{2\pi}\, \widetilde\Delta^{(+)}_{(2),\,3}(\omega+\omega'+\omega;J_c,\omega_c)\cdot\\ &\widetilde\Delta^{(+)}_{(2),\,2}(\omega+\omega';J_c,\omega_c) \,\widetilde\Delta^{(-)}_{(2),\,2} (\omega'+\omega;J_c,\omega_c) \widetilde\Delta^{(-)}_{(2),\,1} (\omega';J_c,\omega_c), \end{aligned}$$
where $\Pi ^{(2)}(\omega )$ and $\Pi ^{(4)}(\omega )$ represent the two leading order contributions of the self-mass of the dressed probe photon from the light-matter interaction, and they are diagrammatically depicted in Fig. 2(c) and Fig. 2(d), respectively. The famous $\Lambda$-type coherent processes [8,14] is embedded in $\Pi ^{(2)}(\omega )$. Here $J_p$ is the electric current density of the probe beam. The dispersion relation of the dressed probe photon $K_\omega$ can be obtained by taking the pole of the propagator $\widetilde {G}(k,\omega )$ into Eq. (10),
$$K_\omega^2=\omega^2-[\Pi^{(2)}(\omega)+\Pi^{(4)}(\omega)],$$
as shown in Fig. 3. It is worth noting that $\partial K_{\omega }/\partial \omega$ takes its local maxima, or the group velocity takes its local minima, around the flips of the dispersion curve. For example, when $\omega$ is around $0.7\nu _0$, the group velocity is roughly $0.1c$ (slow photon).

 figure: Fig. 3.

Fig. 3. The dispersion relation of free photon, $K_{\omega }c$ versus $\omega /\nu _0$, shown as the straight green dotted line. The real and imaginary parts of $K_{\omega }c$ of the dressed probe photon are represented as the blue solid and red dashed lines, respectively. Here we assign the three energy levels $\nu _1$, $\nu _2$, $\nu _3$, to be $\nu _0$, 3$\nu _0$, 4$\nu _0$, respectively, with $\nu _0$ an appropriate energy scale denoting the transition frequency between $|2\rangle$ and $|3\rangle$. In plotting this figure, we choose $g^2=1/137$ and $\delta =8\times 10^{-3}$, while the frequency and intensity of the control beam are $\omega _c/\nu _0=2.9$ and $J_c=9/\nu _0$, respectively. The intensity of the probe beam is $J_p=3/\nu _0$. The inset shows the real part of the index of refraction $n$. All the above-mentioned parameters are shown in arbitrary unit.

Download Full Size | PDF

The index of refraction $n(\omega )$ can also be obtained [33] (here we put $c$ back),

$$n(\omega)=\frac{c}{\omega}\,\textrm{Re} \sqrt{[\textrm{Re}(K_\omega)]^2-[\textrm{Im}(K_\omega)]^2+2i\, \textrm{Re}(K_\omega)\,\textrm{Im}(K_\omega)},$$
which is depicted in the inset of Fig. 3. The relative rate of transmission of a dressed probe photon to that of a free photon (with no interaction and $c=1$) observed at a distance $l$ ($l$ is set to be unity throughout this paper) from the location of the atom is
$$\begin{aligned} T(\omega)&=\left|\int \frac{dk}{2\pi}\,\widetilde{G}( k,\omega)\,e^{i(kl-\omega t)}\,\right|^2/\left|\int \frac{dk}{2\pi}\,\widetilde{G}_{(0)}( k,\omega)\,e^{i(kl-\omega t)}\,\right|^2\\ &=\left| \widetilde{G}(K_\omega,\omega) \,e^{i(K_\omega\,l-\omega t)}\,\right|^2/ \left|\widetilde{G}_{(0)}(\frac{\omega}{c},\omega)\,e^{i(\frac{\omega}{c}l-\omega t)} \,\right|^2\\ &=\frac{\omega^2}{|K_\omega|^2}\exp{({-}2\, \textrm{Im} (K_\omega) \,l}), \hspace{-0.0in}\end{aligned}$$
as shown in Fig. 4.

 figure: Fig. 4.

Fig. 4. The relative transmission $T(\omega )$ of dressed probe photon. Here, we use the same parameters as those in Fig. 3 except the $J_c$: $J_c= 17/\nu _0$ (red dashed line), $J_c= 9/\nu _0$ (black solid line), and $J_c= 0$ (blue dotted line).

Download Full Size | PDF

As can be seen in Fig. 3, the dispersion relation of the dressed probe photon $K_\omega$ is nonlinear with energy gaps around which the momenta are complex corresponding to attenuated waves of absorbed photons. It should be noticed that the shape of $K_\omega$ and especially the locations of gaps depend largely on the frequency $\omega _c$ and the electric current densities of the control and probe beams ($J_c$ and $J_p$). Therefore, the relative rate of transmission of the dressed probe photon $T(\omega )$ shows different performances in Fig. 4. Taking the black solid line in Fig. 4 as an example that the corresponding electric current density and frequency of the control beam are $J_c=9/\nu _0$ and $\omega _c/\nu _0=2.9$. Note that $\nu _0$ is an appropriate energy scale denoting the transition frequency between $|2\rangle$ and $|3\rangle$. The results show an apparent Fano resonance around probe photon frequency $\omega /\nu _0=0.7$, a less apparent one is around $\omega /\nu _0=1.22$, and a window of EIT roughly located between $\omega /\nu _0=0.8$ and $\omega /\nu _0=1.1$. Comparing the dispersion relation in Fig. 3 with the transmission curve shown as the black solid line in Fig. 4, it can be found that slow photons appear around the Fano resonance frequencies. Moreover, the differences between the red dashed line ($J_c=17/\nu _0$) and the blue dotted line ($J_c=0$) (same $\omega _c$ and $J_p$) shown in Fig. 4 indicate that the asymmetric patterns and locations of Fano resonances, and the widths of EIT windows are modified by tuning the electric current density of the control beam $J_c$. When the control beam is absent, there usually exists Lorentzian lineshapes in the scattering/absorption spectra. However, the blue-dotted curve in Fig. 4 shows a non-Lorentzian behavior around $\omega =0.9 \nu _0$. This is because the amplitude of the next-to-leading order diagram [Fig. 2(d)] cannot be ignored compared with that of the leading-order diagram [Fig. 2(c)] therein, the net effect of both leading-order and next-to-leading order scattering processes therefore leads to this non-Lorentzian lineshape in transmission. This has been demonstrated more clearly in Fig. 5 which shows the influences of adjusting the intensities (amplitudes) of the control and probe beams on the transmission. For example, for fixed frequencies of the control and probe beams, the (relative) transmission $T(\omega )$ can be varied from $T(\omega )<1$ to $T(\omega )>1$ when we tune the amplitude of the control beam (probe beam) and keep the amplitude of the probe beam (control beam). The varying quantum interference can also be clearly seen in Fig. 5. For example, when $5J_p=10/{\nu _0}$ in the upper x-scale (which coincides with $J_c= 10/{\nu _0}$ in the lower x-scale ), the lineshape of $T_{p1} (T_{p2})$ shows the constructive (destructive) interference.

 figure: Fig. 5.

Fig. 5. (a) (Lower x-scale) The relative transmission $T(\omega )$ of dressed probe photon versus the electric current density of the control beam $J_c$. The probe frequency $\omega _p/\nu _0=0.5$ ($T_{c1}$, the black solid line) and $\omega _p/\nu _0=0.6$ ($T_{c2}$, the red solid line), and the amplitude of the probe beam is fixed as $J_p=1/\nu _0$. (b) (Upper x-scale) The relative transmission $T(\omega )$ of dressed probe photon versus 5 times of amplitude of probe beam ($5J_p$). The probe frequency $\omega _p/\nu _0=0.5$ ($T_{p1}$, the black dashed line) and $\omega _p/\nu _0=0.6$ ($T_{p2}$, the red dashed line), and the amplitude of the control beam is fixed as $J_c=14.6/\nu _0$. Other parameters like $\omega _c$, $g$, and $\delta$ are the same as those in Fig. 3.

Download Full Size | PDF

3. Landau-Zener-Stückelberg interference

We now study the effect of periodic driving to this three-level system, that is, to consider the control beams with auxiliary waves of oscillatory amplitudes. The resulting quantum interference is called the Landau-Zener-Stückelberg (LZS) interferometry [26]. The amplitude of the oscillatory control beam can be written as

$$\begin{aligned} &J(t) = J_c\,[1+2\gamma \sin(\Omega t)]\,e^{{-}i\omega_c t}\\ = &J_c(t)+ J_+(t)+J_-(t),\\ \textrm{with}\; &J_c(t)=J_c e^{{-}i\omega_ct}~\textrm{and}~ J_{{\pm}}=\gamma J_c e^{{-}i(\omega_c\pm\Omega)t}, \end{aligned}$$
where $\gamma \ll 1$ and $\Omega \ll \omega _c$. Note that the amplitude can be easily modulated by applying an electro-optic modulator in the experiment. The above LZS control beam is in fact composed of three component waves, including one carrier frequency $\omega _c$ with the corresponding electric current density $J_c$ as before, and two sideband frequencies $\omega _c\pm \Omega$ with the corresponding electric current density $\gamma J_c$. For the multi-photon scattering process involving the LZS control beam, the field theoretic method is especially useful.

The leading order contributions to the modifications of transmission amplitudes of dressed probe photons (frequency $\omega$) with the two sideband waves $J_\pm (t)$ on the control beam are represented diagrammatically in Fig. 6. It involves the processes of wave $J_c(t)$ as the source (sink), and the sidebands $J_\pm (t)$ as the sink (source). Note that higher order terms of $\gamma$ would be ignored for a sufficiently small $\gamma$. The analytical expression of the probability amplitude represented by Fig. 6 is

$$\begin{aligned} &{\cal A}_{i,j}(\omega) =\frac{g^4}{2} \langle\tilde{ A}(\omega_i,J_i)^* \rangle \langle\tilde{ A}(\omega_j,J_j)\rangle \cdot\\ &\int \frac{d\omega'}{2\pi}\widetilde\Delta^{(+)}_{(2),\,3}(\omega+\omega'-\omega_i+\omega_j;J_c,\omega_c)\cdot \widetilde\Delta^{(+)}_{(2),\,2}(\omega+\omega'-\omega_i;J_c,\omega_c)\cdot\\ . &\qquad\widetilde\Delta^{(+)}_{(2),\,3} (\omega'+\omega;J_c,\omega_c) \widetilde\Delta^{(-)}_{(2),\,1} (\omega';J_c,\omega_c), \end{aligned}$$
where $i,j\in \{c,\pm \}$, with $\omega _\pm =\omega _c\pm \Omega$ and $J_\pm =\gamma J_c$. This may end up to be a nonlinear process for the transmission of dressed probe photon. For a probe photon carrying frequency $\omega$, as $i=c$, $j=\pm$, for example, the transmitted photon can carry frequency $\omega \pm \Omega$. By adding up the contributions from all possible ${\cal A}_{i,j}(\omega )$’s to the previous monochromatic Green function (propagator) in Eq. (10), we can obtain the new Green function for the dressed probe photon with the LZS component in the control beam
$$\widetilde{G}(K_\omega,\omega) \,e^{i(K_\omega\,x-\omega t)}+ \sum_{i,j\in\{c,\pm\}} {\cal A}_{i,j}(\omega)e^{i[\frac{\omega}{c}x-(\omega-\omega_i+\omega_j )t]}.$$

Therefore, similar to Eq. (15), we can derive the relative transmission of probe beam with frequency $\omega$ at time t, $T(\omega ;t)$.

 figure: Fig. 6.

Fig. 6. Diagrammatic representation of the processes for the leading orders of transmission of dressed photon with the LZS control beam. Here $J^*_i=J^*_c$, $J_j=J_\pm$, or $J^*_i=J^*_\pm$, $J_j=J_c$, and $\omega _c\approx \nu _3-\nu _1$, $\omega \approx \nu _3-\nu _2$ will be adopted in the numerical calculations shown in Figs. 7 and 8.

Download Full Size | PDF

Numerical results of $T(\omega ;t)$ are depicted in Figs. 7 and 8. In Fig. 7, the (relative) transmission $T(\omega ;t)$ shows more cycles of Fano resonance and EIT due to the LZS interference. For most ranges of frequencies, the (relative) transmission is modified to be oscillatory in time due to the oscillatory LZS control beam (black solid line and blue dashed line in Fig. 8). It is interesting to note that when the frequency of the probe beam is larger than a certain value, e.g., $\omega _p/\nu _0=1.3$ in Fig. 8, $T(\omega ;t)$ maintains almost a constant in time. Specifically, the (relative) transmission rate $T(\omega ;t) \approx 1$ as $\omega _p/\nu _0=1.3$ at which EIT occurs (red dotted line in Fig. 8). Figures 7 and 8 indicate that with the interference of the Stückelberg phase accumulated between Landau-Zener transitions, the Fano resonance and EIT can occur in an alternating way. More importantly, with appropriate frequencies of the probe photons, the transmissions become almost stationary in addition to a wide EIT window in time.

 figure: Fig. 7.

Fig. 7. The relative transmission $T(\omega ;t)$ of dressed probe photon with LZS type control beam observed at different times $t=0$ (red solid line) and $t=7.5~(\nu _0^{-1})$ (blue dashed line). The parameters of the oscillatory part of the control beam defined in Eq. (16) are $\gamma =0.05$, $\Omega /\nu _0=0.2$. The black dotted line is used for comparison, and it indicates the relative transmission when the control beam does not contain the LZS oscillation component ($\gamma =\Omega =0$). Here, $J_c=13/\nu _0$, $J_p=3/\nu _0$, and other parameters like $\omega _c$, $g$, and $\delta$ are the same as those in Fig. 3.

Download Full Size | PDF

 figure: Fig. 8.

Fig. 8. The relative transmission $T(\omega ;t)$ of dressed probe photon with LZS type control beam versus observation time $t$/$\pi$. Here, the control beam frequency $\omega _p/\nu _0=1$ (black solid line), $\omega _p/\nu _0=1.225$ (blue dashed line), $\omega _p/\nu _0=1.3$ (red dotted line), and $\omega _p/\nu _0=1.35$ (green dotted dashed line). Other parameters like $\omega _c$, $J_c$, $J_p$, $g$, and $\delta$ are the same as in Fig. 7.

Download Full Size | PDF

As shown in Figs. (4)–(8), the (relative) transmission can go beyond unity. This is because in this work, we adopted a field theoretical approach [Eq. (1)–Eq. (6)] which deals with a many-body system. As bosonic fields, photon fields can have particle numbers more than one, and we study multi-photon scattering processes such that the transmission of the dressed probe beam can be greater than unity under some conditions.

4. Conclusions

Fano resonance and EIT are well-known signatures of interference in multi-level quantum systems. However, in this paper, we show that the quantum interference in a three-level atom is not only frequency-dependent but also tunable through adjusting the intensities of control and probe fields. We also present that the alternating Fano/EIT lineshapes appear when the LZS interferometry is introduced to the control fields. Besides, under the condition that the frequency of the probe beam is larger than a certain value, the system turns out to be transparent for a relatively long time.

The theoretical scheme on tuning quantum interference we studied in this work can be experimentally realized in many three-level quantum systems, such as superconducting qubits, gate-control quantum dots, and trapped atoms, and microcavity systems. Since the scheme has the potential tuning the quantum interference, the results of tunable dispersion and EIT (Fano)-window can therefore be applied to the studies of lossless light-transmission, light delay and storage [3436], quantum sensing, the optical quantum memory [3741] in the nanostructures, the optical nonlinearity [4244], the high-dimension quantum entanglement, and the adiabatic quantum computing [45,46].

Funding

Ministry of Science and Technology, Taiwan (108-2627-E-006-001, 109-2112-M-110-008-MY3, 110-2112-M-005-002, 110-2112-M-005-006, 110-2123-M-006-001).

Acknowledgments

C. C. Chang acknowledges the support of the Ministry of Science and Technology, Taiwan [Grants No. MOST 110-2112-M-005-006]. Y. H. Chen acknowledges the support of the Ministry of Science and Technology, Taiwan [Grants No. MOST 109-2112-M-110-008-MY3 and MOST 110-2123-M-006-001]. G. Y. Chen is supported partially by the National Center for Theoretical Sciences, and acknowledges the support of the Ministry of Science and Technology, Taiwan [Grants No. MOST 110-2123-M-006-001 and MOST 110-2112-M-005-002].

Disclosures

The authors declare no conflicts of interest.

Data availability

No data were generated or analyzed in the presented research.

References

1. I. Buluta, S. Ashhab, and F. Nori, “Natural and artificial atoms for quantum computation,” Rep. Prog. Phys. 74(10), 104401 (2011). [CrossRef]  

2. Z. L. Xiang, S. Ashhab, J. You, and F. Nori, “Hybrid quantum circuits: Superconducting circuits interacting with other quantum systems,” Rev. Mod. Phys. 85(2), 623–653 (2013). [CrossRef]  

3. I. Georgescu, S. Ashhab, and F. Nori, “Quantum Simulation,” Rev. Mod. Phys. 86(1), 153–185 (2014). [CrossRef]  

4. N. Rivera and I. Kaminer, “Light–matter interactions with photonic quasiparticles,” Nat. Rev. Phys. 2(10), 538–561 (2020). [CrossRef]  

5. C. Maschler and H. Ritsch, “Cold Atom Dynamics in a Quantum Optical Lattice Potential,” Phys. Rev. Lett. 95(26), 260401 (2005). [CrossRef]  

6. J. Liu and Z.-Y. Li, “Interaction of a two-level atom with single-mode optical field beyond the rotating wave approximation,” Opt. Express 22(23), 28671–28682 (2014). [CrossRef]  

7. Z. Ficek and S. Swain, Quantum Interference and Coherence, 1st ed. (Springer, 2005).

8. M. Fleischhauer, A. Imamoglu, and J. P. Marangos, “Electromagnetically induced transparency: Optics in coherent media,” Rev. Mod. Phys. 77(2), 633–673 (2005). [CrossRef]  

9. B. Peng, Ş. K. Özdemir, W. Chen, F. Nori, and L. Yang, “What is and what is not electromagnetically induced transparency in whispering-gallery microcavities,” Nat. Commun. 5(1), 5082 (2014). [CrossRef]  

10. Y. C. Liu, B. B. Li, and Y. F. Xiao, “Electromagnetically induced transparency in optical microcavities,” Nanophotonics 6(5), 789–811 (2017). [CrossRef]  

11. H. Ian, Y. X. Liu, and F. Nori, “Tunable electromagnetically induced transparency and absorption with dressed superconducting qubits,” Phys. Rev. A 81(6), 063823 (2010). [CrossRef]  

12. H. C. Sun, Y. X. Liu, H. Ian, J. Q. You, E. Il’ichev, and F. Nori, “Electromagnetically induced transparency and Autler-Townes splitting in superconducting flux quantum circuits,” Phys. Rev. A 89(6), 063822 (2014). [CrossRef]  

13. Q. C. Liu, T. F. Li, X. Q. Luo, H. Zhao, W. Xiong, Y. S. Zhang, Z. Chen, J. S. Liu, W. Chen, F. Nori, J. S. Tsai, and J. Q. You, “Method for identifying electromagnetically induced transparency in a tunable circuit quantum electrodynamics system,” Phys. Rev. A 93(5), 053838 (2016). [CrossRef]  

14. M. F. Limonov, M. V. Rybin, A. N. Poddubny, and Y. S. Kivshar, “Fano resonances in photonics,” Nat. Photonics 11(9), 543–554 (2017). [CrossRef]  

15. A. E. Miroshnichenko, S. Flach, and Y. S. Kivshar, “Fano resonances in nanoscale structures,” Rev. Mod. Phys. 82(3), 2257–2298 (2010). [CrossRef]  

16. R. G. Mani, J. H. Smet, K. von Klitzing, V. Narayanamurti, W. B. Johnson, and V. Umansky, “Zero-resistance states induced by electromagnetic-wave excitation in GaAs/AlGaAs heterostructures,” Nature 420(6916), 646–650 (2002). [CrossRef]  

17. R. L. Samaraweera, H.-C. Liu, Z. Wang, C. Reichl, W. Wegscheider, and R. G. Mani, “Mutual influence between currentinduced giant magnetoresistance and radiation-induced magnetoresistance oscillations in the GaAs/AlGaAs 2DES,” Sci. Rep. 7(1), 5074 (2017). [CrossRef]  

18. R. L. Samaraweera, H.-C. Liu, B. Gunawardana, A. Kriisa, C. Reichl, W. Wegscheider, and R. G. Mani, “Coherent backscattering in quasiballistic ultra-high mobility GaAs/AlGaAs 2DES,” Sci. Rep. 8(1), 10061 (2018). [CrossRef]  

19. M. D. Lukin, M. Fleischhauer, A. S. Zibrov, H. G. Robinson, V. L. Velichansky, L. Hollberg, and M. O. Scully, “Spectroscopy in dense coherent media: line narrowing and interference effects,” Phys. Rev. Lett. 79(16), 2959–2962 (1997). [CrossRef]  

20. M. Mücke, E. Figueroa, J. Bochmann, C. Hahn, K. Murr, S. Ritter, C. J. Villas-Boas, and G. Rempe, “Electromagnetically induced transparency with single atoms in a cavity,” Nature 465(7299), 755–758 (2010). [CrossRef]  

21. Y. Xiaodong, M. Yu, D. -L. Kwong, and C. H. Wong, “All-optical analog to electromagnetically induced transparency in multiple coupled photonic crystal cavities,” Phys. Rev. Lett. 102(17), 173902 (2009). [CrossRef]  

22. J. J. Longdell, E. Fraval, M. J. Sellars, and N. B. Manson, “Stopped light with storage times greater than one second using electromagnetically induced transparency in a solid,” Phys. Rev. Lett. 95(6), 063601 (2005). [CrossRef]  

23. L. Maleki, A. B. Matsko, A. A. Savchenkov, and V. S. Ilchenko, “Tunable delay line with interacting whispering-gallery-mode resonators,” Opt. Lett. 29(6), 626–628 (2004). [CrossRef]  

24. N. Papasimakis, V. A. Fedotov, N. I. Zheludev, and S. L. Prosvirnin, “Metamaterial analog of electromagnetically induced transparency,” Phys. Rev. Lett. 101(25), 253903 (2008). [CrossRef]  

25. C. G. Alzar, M. A. G. Martinez, and P. Nussenzveig, “Classical analog of electromagnetically induced transparency,” Am. J. Phys. 70(1), 37–41 (2002). [CrossRef]  

26. S. N. Shevchenko, S. Ashhab, and F. Nori, “Landau–Zener–Stückelberg interferometry,” Phys. Rep. 492(1), 1–30 (2010). [CrossRef]  

27. J. R. Petta, H. Lu, and A. C. Gossard, “A coherent beam splitter for electronic spin states,” Science 327(5966), 669–672 (2010). [CrossRef]  

28. O. Di Stefano, R. Stassi, L. Garziano, A. Frisk Kockum, S. Savasta, and F. Nori, “Feynman-diagrams approach to the quantum Rabi model for ultrastrong cavity QED: stimulated emission and reabsorption of virtual particles dressing a physical excitation,” New J. Phys. 19(5), 053010 (2017). [CrossRef]  

29. F. Mandl and G. Shaw, Quantum Field Theory, 2nd ed. (Wiley, 2010), pp. 88.

30. C.-C. Chang, L. Lin, and G. Y. Chen, “Photon-Assisted Perfect Conductivity Between Arrays of Two-Level Atoms,” Sci. Rep. 9(1), 13033 (2019). [CrossRef]  

31. A. A. Abrisokov, L. P. Gor’kov, and I. E. Dzyaloshinsky, Quantum Field Theoretical Methods in Statistical Physics, 2nd ed. (Pergamon, 1965).

32. T. D. Lee and C. N. Yang, “Many-Body Problem in Quantum Statistical Mechanics I. General Formulation,” Phys. Rev. 113(5), 1165–1177 (1959). [CrossRef]  

33. J. Jackson, Classical Electrodynamics, 3rd ed. (Wiley, 1998), pp. 296, 310.

34. C. Liu, Z. Dutton, C. H. Behroozi, and L. V. Hau, “Observation of coherent optical information storage in an atomic medium using halted light pulses,” Nature 409(6819), 490–493 (2001). [CrossRef]  

35. D. F. Phillips, A. Fleischhauer, A. Mair, R. L. Walsworth, and M. D. Lukin, “Storage of light in atomic vapor,” Phys. Rev. Lett. 86(5), 783–786 (2001). [CrossRef]  

36. M. F. Yanik, W. Suh, Z. Wang, and S. Fan, “Stopping light in a waveguide with an all-optical analog of electromagnetically induced transparency,” Phys. Rev. Lett. 93(23), 233903 (2004). [CrossRef]  

37. A. I. Lvovsky, B. C. Sanders, and W. Tittel, “Optical quantum memory,” Nat. Photonics 3(12), 706–714 (2009). [CrossRef]  

38. H. J. Kimble, “The quantum internet,” Nature 453(7198), 1023–1030 (2008). [CrossRef]  

39. L. Wang, J. Li, S. Zhang, K. Su, Y. Zhou, K. Liao, S. Du, H. Yan, and S.-L. Zhu, “Efficient quantum memory for single-photon polarization qubits,” Nat. Photonics 13(5), 346–351 (2019). [CrossRef]  

40. Y.-F. Hsiao, P. J. Tsai, H. S. Chen, S. X. Lin, C. C. Hung, C. H. Lee, Y. H. Chen, Y. F. Chen, I. A. Yu, and Y. C. Chen, “Highly efficient coherent optical memory based on electromagnetically induced transparency,” Phys. Rev. Lett. 120(18), 183602 (2018). [CrossRef]  

41. P. Vernaz-Gris, K. Huang, M. Cao, S. S. Alexandra, and L. Julien, “Highly-efficient quantum memory for polarization qubits in a spatially-multiplexed cold atomic ensemble,” Nat. Commun. 9(1), 363 (2018). [CrossRef]  

42. D. Budker, D. F. Kimball, S. M. Rochester, and V. V. Yashchuk, “Nonlinear magneto-optics and reduced group velocity of light in atomic vapor with slow ground state relaxation,” Phys. Rev. Lett. 83(9), 1767–1770 (1999). [CrossRef]  

43. T. Peyronel, O. Firstenberg, Q. Y. Liang, S. Hofferberth, A. V. Gorshkov, T. Pohl, M. D. Lukin, and V. Vuletić, “Quantum nonlinear optics with single photons enabled by strongly interacting atoms,” Nature 488(7409), 57–60 (2012). [CrossRef]  

44. A. F. Kockum, A. Miranowicz, V. Macri, S. Savasta, and F. Nori, “Deterministic quantum nonlinear optics with single atoms and virtual photons,” Phys. Rev. A 95(6), 063849 (2017). [CrossRef]  

45. G. Sun, X. Wen, B. Mao, J. Chen, Y. Yu, P. Wu, and S. Han, “Tunable quantum beam splitters for coherent manipulation of a solid-state tripartite qubit system,” Nat. Commun. 1(1), 51 (2010). [CrossRef]  

46. G. Cao, H. Li, T. Tu, L. Wang, C. Zhou, M. Xiao, G. Guo, H. Jiang, and G. Guo, “Ultrafast universal quantum control of a quantum-dot charge qubit using Landau–Zener–Stückelberg interference,” Nat. Commun. 4(1), 1401 (2013). [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 (8)

Fig. 1.
Fig. 1. A schematic diagram of the bare-state $\Lambda$-type three-level EIT/Fano system: ground state $|1\rangle$; metastable state $|2\rangle$; excited state $|3\rangle$. The probe field of frequency $\omega _p$ governs the $|1\rangle \rightarrow |3\rangle$ transition with the electric current density $J_p$, and the control field of frequency $\omega _c$ drives the $|2\rangle \leftrightarrow |3\rangle$ transition with the electric current density $J_c$.
Fig. 2.
Fig. 2. (a) Diagrammatic expansion of the effective coupling $u_0$ which is the sum of the repeated and continuous scatterings between two different fields $c_\alpha$ and $c_\beta$. The external legs are only for the eyes, the black dot $``\bullet "$ is $U$ in Eq. (4), and the internal double-line or single line are only used to represent different index of the electronic bare propagator of field $c_\gamma$, $\widetilde \Delta ^{(\pm )}_{(0),\gamma }$’s, $\gamma$=1,2,3 (Eq. (5)). (b) Diagrammatic expansions of the self-mass $\Sigma ^{(2)}_{\alpha }(\omega )$ (Eq. (9)) for the renormalized propagators $\widetilde \Delta ^{(\pm )}_{(2),\alpha }$ (Eq. (8)) of the dressed electronic field $c_\alpha$. The internal straight line represents electron propagator $\widetilde \Delta ^{(\pm )}_{(1),\alpha }$ (Eq. (7)). The curvy line represents the propagator of free photon $\widetilde {G}_{(0)}$ (Eq. (3)). (c), (d) Diagrammatic representations of self-masses $\Pi ^{(2)}(\omega )$ (Eq. (11)) and $\Pi ^{(4)}(\omega )$ (Eq. (12)) of the renormalized propagator of the dressed probe photon, $\widetilde {G}(k,\omega )$ (Eq. (10)) are shown in (c) and (d), respectively. The solid black lines represent the renormalized electronic propagators $\widetilde \Delta ^{(\pm )}_{(2),\alpha }$, and the (dashed) curvy external legs are only for the eyes.
Fig. 3.
Fig. 3. The dispersion relation of free photon, $K_{\omega }c$ versus $\omega /\nu _0$, shown as the straight green dotted line. The real and imaginary parts of $K_{\omega }c$ of the dressed probe photon are represented as the blue solid and red dashed lines, respectively. Here we assign the three energy levels $\nu _1$, $\nu _2$, $\nu _3$, to be $\nu _0$, 3$\nu _0$, 4$\nu _0$, respectively, with $\nu _0$ an appropriate energy scale denoting the transition frequency between $|2\rangle$ and $|3\rangle$. In plotting this figure, we choose $g^2=1/137$ and $\delta =8\times 10^{-3}$, while the frequency and intensity of the control beam are $\omega _c/\nu _0=2.9$ and $J_c=9/\nu _0$, respectively. The intensity of the probe beam is $J_p=3/\nu _0$. The inset shows the real part of the index of refraction $n$. All the above-mentioned parameters are shown in arbitrary unit.
Fig. 4.
Fig. 4. The relative transmission $T(\omega )$ of dressed probe photon. Here, we use the same parameters as those in Fig. 3 except the $J_c$: $J_c= 17/\nu _0$ (red dashed line), $J_c= 9/\nu _0$ (black solid line), and $J_c= 0$ (blue dotted line).
Fig. 5.
Fig. 5. (a) (Lower x-scale) The relative transmission $T(\omega )$ of dressed probe photon versus the electric current density of the control beam $J_c$. The probe frequency $\omega _p/\nu _0=0.5$ ($T_{c1}$, the black solid line) and $\omega _p/\nu _0=0.6$ ($T_{c2}$, the red solid line), and the amplitude of the probe beam is fixed as $J_p=1/\nu _0$. (b) (Upper x-scale) The relative transmission $T(\omega )$ of dressed probe photon versus 5 times of amplitude of probe beam ($5J_p$). The probe frequency $\omega _p/\nu _0=0.5$ ($T_{p1}$, the black dashed line) and $\omega _p/\nu _0=0.6$ ($T_{p2}$, the red dashed line), and the amplitude of the control beam is fixed as $J_c=14.6/\nu _0$. Other parameters like $\omega _c$, $g$, and $\delta$ are the same as those in Fig. 3.
Fig. 6.
Fig. 6. Diagrammatic representation of the processes for the leading orders of transmission of dressed photon with the LZS control beam. Here $J^*_i=J^*_c$, $J_j=J_\pm$, or $J^*_i=J^*_\pm$, $J_j=J_c$, and $\omega _c\approx \nu _3-\nu _1$, $\omega \approx \nu _3-\nu _2$ will be adopted in the numerical calculations shown in Figs. 7 and 8.
Fig. 7.
Fig. 7. The relative transmission $T(\omega ;t)$ of dressed probe photon with LZS type control beam observed at different times $t=0$ (red solid line) and $t=7.5~(\nu _0^{-1})$ (blue dashed line). The parameters of the oscillatory part of the control beam defined in Eq. (16) are $\gamma =0.05$, $\Omega /\nu _0=0.2$. The black dotted line is used for comparison, and it indicates the relative transmission when the control beam does not contain the LZS oscillation component ($\gamma =\Omega =0$). Here, $J_c=13/\nu _0$, $J_p=3/\nu _0$, and other parameters like $\omega _c$, $g$, and $\delta$ are the same as those in Fig. 3.
Fig. 8.
Fig. 8. The relative transmission $T(\omega ;t)$ of dressed probe photon with LZS type control beam versus observation time $t$/$\pi$. Here, the control beam frequency $\omega _p/\nu _0=1$ (black solid line), $\omega _p/\nu _0=1.225$ (blue dashed line), $\omega _p/\nu _0=1.3$ (red dotted line), and $\omega _p/\nu _0=1.35$ (green dotted dashed line). Other parameters like $\omega _c$, $J_c$, $J_p$, $g$, and $\delta$ are the same as in Fig. 7.

Equations (18)

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

A = ( A y + i A z ) / 2 , A = ( A y i A z ) / 2 ,
H e m = d x ( A ˙ A ˙ + A A + J A + J A ) ,
G ~ ( 0 ) ( k , ω ) = i ω 2 k 2 + i ϵ
H 3 L A ( { c } ) = α = 1 , 2 , 3 ( ε α i δ ) c α c α + U 2 α > β c α c β c α c β ,
Δ ~ ( 0 ) , α ( ± ) ( ω ) = i ω ε α ± i δ , α = 1 , 2 , 3 ,
H i n t ( { c } , A ) = g [ A ( x a ) c α c β + h . c . ] ,
Δ ~ ( 1 ) , α ( ± ) ( ω ) = i ω ν α ± i δ , α = 1 , 2 , 3 ,
Δ ~ ( 2 ) , α ( ± ) ( ω ; J c , ω c ) 1 = Δ ~ ( 1 ) , α ( ± ) ( ω ) 1 Σ α ( 2 ) ( ω ; J c , ω c ) i ,
Σ α ( 2 ) ( ω ; J c , ω c ) = g 2 i | A ~ ( ω c , J c ) | 2 β α Δ ~ ( 1 ) , β + ( ω sgn ( α β ) ω c ) = g 2 | J c | 2 4 ω c 2 β α 1 ω sgn ( α β ) ω c ν α + i δ ,
G ~ ( k , ω ) 1 = G ~ ( 0 ) ( k , ω ) 1 i 1 [ Π ( 2 ) ( ω ) + Π ( 4 ) ( ω ) ] ,
Π ( 2 ) ( ω ) = g 2 2 α > β d ω 2 π Δ ~ ( 2 ) , α ( + ) ( ω + ω ; J c , ω c ) Δ ~ ( 2 ) , β ( ) ( ω ; J c , ω c ) ,
Π ( 4 ) ( ω ) = g 4 2 | A ~ ( ω , J p ) | 2 d ω 2 π Δ ~ ( 2 ) , 3 ( + ) ( ω + ω + ω ; J c , ω c ) Δ ~ ( 2 ) , 2 ( + ) ( ω + ω ; J c , ω c ) Δ ~ ( 2 ) , 2 ( ) ( ω + ω ; J c , ω c ) Δ ~ ( 2 ) , 1 ( ) ( ω ; J c , ω c ) ,
K ω 2 = ω 2 [ Π ( 2 ) ( ω ) + Π ( 4 ) ( ω ) ] ,
n ( ω ) = c ω Re [ Re ( K ω ) ] 2 [ Im ( K ω ) ] 2 + 2 i Re ( K ω ) Im ( K ω ) ,
T ( ω ) = | d k 2 π G ~ ( k , ω ) e i ( k l ω t ) | 2 / | d k 2 π G ~ ( 0 ) ( k , ω ) e i ( k l ω t ) | 2 = | G ~ ( K ω , ω ) e i ( K ω l ω t ) | 2 / | G ~ ( 0 ) ( ω c , ω ) e i ( ω c l ω t ) | 2 = ω 2 | K ω | 2 exp ( 2 Im ( K ω ) l ) ,
J ( t ) = J c [ 1 + 2 γ sin ( Ω t ) ] e i ω c t = J c ( t ) + J + ( t ) + J ( t ) , with J c ( t ) = J c e i ω c t   and   J ± = γ J c e i ( ω c ± Ω ) t ,
A i , j ( ω ) = g 4 2 A ~ ( ω i , J i ) A ~ ( ω j , J j ) d ω 2 π Δ ~ ( 2 ) , 3 ( + ) ( ω + ω ω i + ω j ; J c , ω c ) Δ ~ ( 2 ) , 2 ( + ) ( ω + ω ω i ; J c , ω c ) . Δ ~ ( 2 ) , 3 ( + ) ( ω + ω ; J c , ω c ) Δ ~ ( 2 ) , 1 ( ) ( ω ; J c , ω c ) ,
G ~ ( K ω , ω ) e i ( K ω x ω t ) + i , j { c , ± } A i , j ( ω ) e i [ ω c x ( ω ω i + ω j ) t ] .
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.