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

Localized gap modes of coherently trapped atoms in an optical lattice

Open Access Open Access

Abstract

We theoretically investigate one-dimensional localized gap modes in a coherent atomic gas where an optical lattice is formed by a pair of counterpropagating far-detuned Stark laser fields. The atomic ensembles under study emerge as Λ-type three-level configuration accompanying the effect of electromagnetically induced transparency (EIT). Based on Maxwell-Bloch equations and the multiple scales method, we derive a nonlinear equation governing the spatial-temporal evolution of the probe-field envelope. We then uncover the formation and properties of optical localized gap modes of two kinds, such as the fundamental gap solitons and dipole gap modes. Furthermore, we confirm the (in)stability regions of both localized gap modes in the respective band-gap spectrum with systematic numerical simulations relying on linear-stability analysis and direct perturbed propagation. The predicted results may enrich the nonlinear horizon to the realm of coherent atomic gases and open up a new door for optical communication and information processing.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

The techniques for cooling and trapping atoms were initially achieved by utilizing the state-of-the-art laser technologies in the 1980s [1]. In particular, the rapid development of laser technologies has not only helped generate ultracold atoms including Bose-Einstein condensates (BECs) [1] but also contributed to the control and manipulation of their dynamics in optical lattices formed by counterpropagating laser beams [2]. In past decades, one of the great achievements of BECs loaded onto optical lattices was the realization of matter-wave gap solitons under self-defocusing nonlinearity induced by atom-atom interactions [3]. In recent years, much attention has been paid to revealing the formation, property and stability of gap solitons in various physical systems including BEC and optics [411]. For the latter, linear periodic potentials such as photonic crystals and lattices [4,7], prefabricated or wrote optically, inside which the optical gap solitons are localized, had been implemented experimentally [1215]. Nonlinear lattices [6], the nonlinear counterparts of periodic potentials, can be induced, respectively in BECs and optics, via Feshbach resonance mediated by spatially periodic fields [16] and properly made nonlinear photonic crystals [4]. The nonlinear lattices [6,1719] and the combined linear-nonlinear lattices [20,21] have recently been used to stabilize different kinds of solitons.

Self-induced transparency solitons [22,23] and gap ones [2426] had been well studied respectively in uniform and periodic resonantly atomic systems decades ago. Coherent atomic systems in the form of BECs whose atoms are prepared under electromagnetically induced transparency (EIT) condition with all-optical ways have recently been a hot topic in soliton studies [2741]. EIT is a typical quantum destructive interference effect emerging in the resonant multilevel systems with some suitable laser fields, which possesses many striking scientific properties, i.e., the large elimination of optical absorption [42], the obvious reduction of group velocity [43,44], a remarkable enhancement of Kerr nonlinearity with very low power light fields [45,46]. The study of EIT has led to a surge of research on various frontier domains due to its rich physical properties and important practical applications [4751].

It is relevant to point out that the self-induced transparency gap solitons in a resonantly absorbing Bragg reflector with EIT turning on had also been investigated by Kurizki and his colleagues years ago [52]. In addition, optical solitons have been predicted in a resonant three-level atomic system with a standing-wave control field which forms an optical lattice and excites EIT [5355], and in optical waveguides controlled by EIT [56], while the literal report on gap solitons in the coherent atomic system via EIT have yet to be explored, i.e., the stability and instability regions of localized gap modes in the band-gap spectrum. Theoretical and experimental works have demonstrated the ways for manipulating light pulses via dynamically controlled EIT-induced photonic band gap in coherently prepared atomic gases [5760]. Various effects including solitons have been widely studied in multilevel atomic systems with electromagnetically induced lattices formed by EIT in recent years [31,34,6163], while gap solitons are still missing. It is thus the main objective of this paper to survey localized gap modes in such coherent atomic ensembles with the EIT on.

In this paper, an one-dimensional (1D) coherent atomic system consisting of a $\Lambda$-type three-level atomic gases that are excited under EIT condition and trapped by an optical lattice formed by a pair of counterpropagating far-detuned Stark laser fields is proposed as a new platform to generate localized gap modes. The physical equation describing such system is in the framework of nonlinear Schrödinger equation, formulated and derived from the Maxwell-Bloch equations and the method of multiple scales [29,31] under suitable conditions. The model supports two types of localized gap modes, fundamental gap solitons and dipole ones, in the first two finite band gaps of the underlying linear spectrum. Both localized gap modes can be constructed as on-site and off-site modes, with their central profiles placing respectively into the maximum and minimum values of the optical lattice. The property and (in)stability of the predicted gap modes are evaluated by linear-stability analysis and direct perturbed simulations. The proposed physical scheme and the predicted gap modes therein can enlarge the nonlinear spectrum of coherent atomic gases and open up a new avenue for implications including optical communication and information processing.

The rest of the article is arranged as follows. The physical model under study is presented in Sec. 2. The nonlinear Schrödinger equation of the probe field is derived from Maxwell-Bloch equations, and its numerical schemes, including the linear-stability analysis and direct perturbed simulations, are introduced in Sec. 3. The existence and stability properties of localized gap modes in forms of fundamental gap solitons and dipole ones are presented numerically and discussed with an analysis in Sec. 4. A summary of this article is finally made in Sec. 5.

2. Physical model

We consider a lifetime-broadened three-level atomic gas in a $\Lambda$-type configuration interacting resonantly with two laser fields, i.e., the pulsed probe field $\textbf {E}_p$ couples the transitions $|1\rangle \leftrightarrow |3\rangle$ and the strong continuous-wave control field $\textbf {E}_c$ drives the transition $|2\rangle \leftrightarrow |3\rangle$, as depicted in Fig. 1(a). $\Gamma _{13}$ and $\Gamma _{23}$ are the spontaneous emission decay rate from $|3\rangle$ to $|1\rangle$ and $|3\rangle$ to $|2\rangle$, respectively. $\Delta _2$ and $\Delta _3$ are the two- and one-photon detunings, respectively. The atoms are firstly prepared in the ground state $|1\rangle$, and cooled to an extreme low temperature to restrain the relevant center-of-mass motion. Under these conditions, the physical system of such an atomic ensemble interacting with two laser fields constitutes the above-mentioned EIT where the absorption of the probe field can be greatly suppressed due to the quantum interference effect induced by the control field [42].

 figure: Fig. 1.

Fig. 1. (a) Theoretical scheme of a $\Lambda$-type three-level atomic system for setting on EIT, whose excitation mechanism is described in text. $\Gamma _{13}$ ($\Gamma _{23}$) is the spontaneous emission decay rate from $|3\rangle$ to $|1\rangle$ ($|3\rangle$ to $|2\rangle$). $\Delta _j$ ($j=2,\,3$) is the detuning. (b) Possible experimental arrangement of laser beams geometry. The weak probe and strong control fields (with angular frequency $\omega _p$ and $\omega _c$, respectively) propagate along the $z$ direction. The (purple) thick arrows denote a pair of counter-propagating Stark fields (both with angular frequency $\omega _s$), which form standing waves called optical lattice.

Download Full Size | PDF

Both the probe and control laser fields are assumed to propagate along the $z$ direction [see Fig. 1(b)], thus the electric-field vector can be written as $\textbf {E}=\textbf {E}_p+\textbf {E}_c=\hat {\mathbf {e}}_{p}{\cal E}_{p} e^{i(k_{p} z-\omega _{p} t)}+\hat {\mathbf {e}}_{c}{\cal E}_{c} e^{i(k_{c} z-\omega _{c} t)}+\textrm {c.c.}$. Here $\hat {\mathbf {e}}_{p}$ ($\hat {\mathbf {e}}_{c}$) is the unit vector of the probe (control) field with the envelope ${\cal E}_p$ (${\cal E}_c$), $\omega _p$ ($\omega _c$) is the angular frequency of the probe (control) field, and $k_{p}=\omega _{p}/c$ ($k_c=\omega _c/c$) is the wavenumber of the probe (control) field before entering the atomic gas. Furthermore, we apply two counter-propagating far-detuned laser fields (Stark fields), which have the same angular frequency $\omega _s$, to the EIT medium, to generate an optical lattice, onto which the optical localized gap modes can be loaded; See the possible experimental arrangement in Fig. 1(b). The Stark fields follow the form

$$\textbf{E}_{\textrm{Stark}}(x,t) = \hat{\textbf{e}}_s\sqrt{2} E_s(x) \cos(\omega_st),$$
where $\hat {\textbf {e}}_s$ and $E_s$ are the unit polarization vector and field amplitude, respectively. The Stark fields can cause a small energy shift for the level $|j\rangle$ in the $x$ direction, i.e., $\Delta E_{j,\textrm {Stark}}=-\alpha _j\langle \textbf {E}_{\textrm {Stark}}^2\rangle _t/2=-\alpha _j|E_s(x)|^2/2$. Here $\alpha _j$ is the scalar polarizability of the level $|j\rangle$ and $\langle \cdots \rangle$ denotes the time average in an oscillating cycle.

Under the electric-dipole and rotating-wave approximations, the Hamiltonian of the system in the interaction picture is $\hat {\cal H}=-\sum _{j=1}^{3}\hbar \Delta _j^\prime |j\rangle \langle j|-\hbar \,\left [\Omega _c|3\rangle \langle 2|+\Omega _{p}|3\rangle \langle 1|+\textrm {H.c.}\right ],$ with $\Delta _j^\prime =\Delta _j+\frac {1}{2}\alpha _{j1}|E_s(x)|^2$ and $\alpha _{jl}=(\alpha _{j}-\alpha _{l})/\hbar$. Here $\Delta _1=0$, $\Delta _2=\omega _{p}-\omega _{c}-\omega _{21}$, and $\Delta _3=\omega _{p}-\omega _{31}$ are the detunings, where $\omega _{jl}=(E_j-E_l)/\hbar$ with $E_j$ being the eigen energy of the state $|j\rangle$. $\Omega _c=(\mathbf {p}_{23}\cdot {\hat {\mathbf {e}}}_c){\cal E}_c/\hbar$ and $\Omega _{p}=(\mathbf {p}_{13}\cdot {\hat {\mathbf {e}}}_{p}){\cal E}_{p}/\hbar$ are half Rabi frequencies of the control and probe fields, where $\textbf {p}_{jl}$ is the electric dipole matrix element related to the transition from $|j\rangle$ to $|l\rangle$. Thus the equation of motion for density matrix $\sigma$ in the interaction picture is given by

$$\frac{\partial \sigma}{\partial t}=-\frac{i}{\hbar}[\hat{\mathcal{H}}_\textrm{int},\sigma]-\Gamma \sigma.$$
Note that $\sigma$ is a density matrix and $\Gamma$ represents the spontaneous emission rate and dephasing rate. The explicit expression for density-matrix elements $\sigma _{jl}$ is given in Appendix A..

Utilizing the slowly varying envelope approximation, the Maxwell equation for the probe-field Rabi frequency $\Omega _{p}$ is described as

$$i\left(\frac{\partial}{\partial z}+\frac{1}{c}\frac{\partial}{\partial t}\right)\Omega_{p}+\frac{c}{2\omega_{p}}\frac{\partial^2 \Omega_{p}}{\partial x^2} +\kappa_{13}\sigma_{31}=0,$$
where $\kappa _{13}={\cal N}_a\omega _{p}|\mathbf {p}_{13}\cdot \hat {\mathbf {e}}_{p}|^2/(2\hbar \varepsilon _0 c)$ with ${\cal N}_a$ being atomic density and $c$ the light speed in vacuum. Note that the transverse radius of the probe field in the $y$ direction is assumed to be large enough so that the diffraction in the $y$ direction can be neglected.

3. Nonlinear Schrödinger equation and numerical schemes

3.1 Derivation of a nonlinear Schrödinger equation

To investigate the EIT-based control of the nonlinear evolution properties of the probe field in the system, we have derived the relevant nonlinear envelope equation in the framework of Maxwell-Bloch equations by adopting the standard multiple scales method [29,31]. The asymptotic expansion has the following form: $\sigma _{jl}=\sum _{m=0}^{\infty }\epsilon ^m\sigma _{jl}^{(m)}$, with $\sigma _{jl}^{(0)}=\delta _{j1}\delta _{l1}$, $\Omega _{p}=\sum _{m=1}^{\infty }\epsilon ^m \Omega _{p}^{(m)}$, Here $\epsilon$ is the dimensionless small parameter characterizing the typical amplitude of the probe field. Note that all the quantities on the right-hand side of the expansion depend on the multi-scale variables $x_1=\epsilon x$, $z_{m}=\epsilon ^{m}z$, and $t_{m}=\epsilon ^{m}t$ ($m=0,\,2$). In addition, the space-dependent amplitude of Stark field [given by Eq. (1)] is assumed as $E_s=\epsilon E_s^{(1)}$. Thus $d_{jl}=d_{jl}^{(0)}+\epsilon ^{2}d_{jl}^{(2)}$, with $d_{jl}^{(0)}=\Delta _j-\Delta _l+i\gamma _{jl}$ and $d_{jl}^{(2)}= \frac {\alpha _{jl}}{2}|E_s^{(1)}|^2$. Substituting the expansion into Maxwell-Bloch equations and comparing the coefficients of $\epsilon ^m$ ($m=1,\,2,\,3,\,\ldots$), we obtain a set of linear but inhomogeneous equations for $\sigma _{jl}^{(m)}$ and $\Omega _{p}^{(m)}$, which can be solved order by order.

When carrying out the solution up to the third order, the Kerr nonlinearity plays an important role, and hence the propagation of the envelope $\mathcal {F}$ of the probe field is described by the nonlinear equation

$$i\left(\frac{\partial }{\partial z_2}+\frac{1}{V_{g}}\frac{\partial}{\partial t_2}\right)\mathcal{F}+\frac{c }{2\omega_{p}}\frac{\partial^2 \mathcal{F}}{\partial x_1^2}+W_{1}|\mathcal{F}|^2\mathcal{F} e^{-2\bar{a} z_2}+W_{2}|E_s^{(1)}|^2\mathcal{F}=0,$$
where $V_g=(\partial \mathcal {K}/\partial \omega )^{-1}$ is the complex group velocity of the envelope $\mathcal {F}$, and $\bar {a}=\epsilon ^{-2}a=\epsilon ^{-2}\textrm {Im}(\mathcal {K})$ with $\mathcal {K}$ being the linear dispersion relation. The explicit expressions of the self- and cross-phase modulation coefficients $W_{1}$ and $W_{2}$ and the solutions of the first and the second expansion are presented in Appendix B.

Substituting the original variables to Eq. (4), one can obtain a dimensionless form

$$i\left(\frac{\partial}{\partial s}+\frac{1}{g}\frac{\partial}{\partial \tau}\right) u+\frac{1}{2}\frac{\partial^2 u}{\partial \xi^2}+g_{1}|u|^2u+g_{2}V(\xi) u=-i\mathcal{A}u,$$
with $u=\epsilon \mathcal {F}/u_0e^{-\bar {a} z_2}$, $s=z/L_\textrm {Diff}$, $\tau =t/\tau _0$, $g=V_{g}\tau _0/L_{\textrm {Diff}}$, $\xi =x/R$, $g_{1}=L_\textrm {Diff}/L_\textrm {Nonl}$, $g_{2}=L_\textrm {Diff}\textrm {Re}(W_{2})E_{0}^2$, $V(\xi )=[E_s(\xi )/E_0]^2$, and $\mathcal {A}=a L_\textrm {Diff}$. Here $u_0$, $\tau _0$, and $R$ are respectively the typical Rabi frequency, pulse duration, beam radius of probe field, and $E_{0}$ is the Stark field amplitude; $L_\textrm {Diff}=\omega _pR^2/c$ and $L_\textrm {Nonl}=1/|\textrm {Re}(W_{1})u_0^2|$ are the typical diffraction length and nonlinear length, respectively. We assume the wavepacket solution possesses the form $u(\tau ,s,\xi )=G(\tau ,s)v(\tau ,\xi )$ with $G(\tau ,s) = \sqrt [4]{2} \exp {[-(s-g\tau )^2/\rho _0^2]}$ and $\rho _0$ being a free real parameter [31,64]. After integrating over the variable $s$, Eq. (5) converts into
$$\left(\frac{i}{g}\frac{\partial}{\partial \tau}+\frac{1}{2}\frac{\partial^2 }{\partial \xi^2}\right) v+g_{1}|v|^2v+{g}_{2} V(\xi)v=-i\mathcal{A}v.$$
As we all know, the dimensionless coefficients in Eq. (6) are complex because of the system under study being a lifetime-broadened one, and hence it does not support stable localized nonlinear solutions. Fortunately, when the system works under the EIT circumstance, the imaginary parts of these coefficients can be made much smaller than their real parts and thus gap localized solutions are possible. The feasible model can be easily realized by selecting realistic physical systems. One of them is the ultracold $^{87}$Rb atoms tuned to D1-line transition with the energy levels $|1\rangle =|5^{2}\textrm {S}_{1/2}, F=1\rangle$, $|2\rangle =|5^{2}\textrm {S}_{1/2}, F=2\rangle$, and $|3\rangle =|5^{2}\textrm {P}_{1/2}, F=2\rangle$ [65]. And the corresponding system parameters are $\Gamma _2\simeq 1$ kHz, $\Gamma _3\simeq 5.75$ MHz, and $|\textbf {p}_{13}|=2.54\times 10^{-27}$ C cm.

Through choosing other parameters $\mathcal {N}_a\approx 3.69\times 10^{10}$ cm$^{-3}$, $\Omega _c=7.5\times 10^{7}$ Hz, $\Delta _2=1.1\times 10^7$ Hz, and $\Delta _3=2.5\times 10^8$ Hz, the complex coefficients in Eq. (4) are $\mathcal {K}=(3.83+0.043i)$ cm$^{-1}$, $W_{1}=-(7.34+0.096i)\times 10^{-15}$ cm$^{-1}$ s$^{2}$, and $W_{2}=(5.62+0.13i)\times 10^{-9}$ cm V$^{-2}$, where the imaginary parts of these quantities are indeed much smaller than their corresponding real parts. Furthermore, taking $u_0=8.68\times 10^{6}$ Hz, $\tau _0=1.26\,\mu$s, $R=48$ $\mu$m, and $E_0=9.92\times 10^{3}$ V cm$^{-1}$, we obtain ${g}=1-0.02i$, ${g}_{1}=-(1+0.01i)$, ${g}_{2}=1+0.02i$, and $\mathcal {A}=0.077$. In order to obtain nonlinear localized gap modes of the system, we assume the Stark fields have the form of sinusoidal function, i.e., $E_s(\xi )=E_{s0}\sin (\xi )$ ($E_{s0}$ is a real constant). Therefore, Eq. (6) can be reduced into

$$i\frac{\partial v}{\partial \tau}=-\frac{1}{2}\frac{\partial^2 v}{\partial \xi^2}+|v|^2v+V_{\textrm{OL}} (\xi) v.$$
Here $V_{\textrm {OL}} (\xi ) =-c_0\sin ^2 (\xi )$ with $\sqrt {c_0}=E_{s0}/E_{0}$ represents the 1D optical lattice induced by the Stark fields. A distinctive feature of this equation is the negative sign in front of the optical lattice, in contrast to the usual nonlinear Schrödinger equation specific to the BECs and optics.

3.2 Numerical schemes for stability analysis

We search stationary solution of field amplitude $v=U(\xi ) \exp (ib \tau )$ with $b$ being propagation constant, then Eq. (7) becomes

$$-b U=-\frac{1}{2}\frac{\partial^2 U}{\partial \xi^2}+|U|^2U+V_{\textrm{OL}}(\xi) U.$$
To inveatigate the localized gap modes and their properties in the model with optical lattice it is necessary to give the relevant band-gap structure. Linearizing the stationary Eq. (8), i.e., discarding Kerr nonlinear effect, results into an eigenvalue equation: $-b U=-\frac {1}{2}\frac {\partial ^2 U}{\partial \xi ^2}+V_{\textrm {OL}}(\xi ) U$. The linear eigenvalue equation (problem) can be gained as those in the conventional linear periodic system [10], the solutions of such equation comprises the Bloch bands through which the finite band gaps pass. Such a band-gap structure of linear spectrum $b(K)$ characterized by the momentum $K$ is displayed in Fig. 2(a) for a particular optical lattice strength $c_0=6$.

 figure: Fig. 2.

Fig. 2. (a) Band-gap structure of linear spectrum $b(K)$ characterized by the momentum $K$, which is induced by the EIT-based optical lattice $V_{\textrm {OL}}=-c_0 \sin ^2 (\xi )$ with strength $c_0=6$. The 1st BG and 2nd BG hereafter represent the first and second band gaps, respectively. (b) Power $P$ versus propagation constant $b$ for on-site gap solitons (red solid line) and off-site ones (blue dashed line). Panels (c) and (d) represent respectively the real parts of maximum eigenvalues, expressed by Re($\lambda$) as a function of $b$, for the corresponding off-site and on-site gap solitons. Profiles of gap solitons marked by dotted points (A1, A2, A3) and (B1, B2, B3) will be displayed below.

Download Full Size | PDF

The linear stability is a key to the question of whether the stationary solutions can be stable localized gap modes. And so, we take the perturbed amplitude as $v=[U+p(\xi )\exp {(\lambda \tau )}+q^{\ast }(\xi )\exp {(\lambda ^{\ast } \tau )}]\exp {(ib \tau )}$, here $U$ is the undisturbed field amplitude found from Eq. (8), and $p(\xi )$ and $q^{\ast }(\xi )$ are the small background perturbations at eigenvalue $\lambda$. Substituting such expression into Eq. (7) leads to the following linear eigenvalue problem:

$$\left( \begin{array}{cc} L & U^2\\ -U^{\ast 2} & -L \end{array} \right ) \left( \begin{array}{c} p\\ q \end{array} \right ) =i\lambda \left( \begin{array}{c} p\\ q \end{array} \right ),$$
where $L=b-\frac {1}{2}\frac {\partial ^2 }{\partial \xi ^2}+2|U|^2+V_{\textrm {OL}}(\xi )$. By numerically solving the eigenvalue Eqs. (9) with the Fourier collocation method [66], we can get the eigenvalues $\lambda$ and whose real parts, Re$(\lambda )$, determine the stability of the perturbed localized solutions. Specifically, the solutions are stable given that Re$(\lambda )=0$, and are unstable otherwise. The stability of these solutions is rechecked via direct numerical simulations of the perturbed evolution in Eq. (7) using the split-step Fourier method. As usual, the stationary solution is perturbed with a small random perturbation, i.e., $v=U(1+\varepsilon \psi )\exp (ib \tau )$, with $\varepsilon$ being the amplitude of the perturbation and $\psi$ being a random variable uniformly distributed in the interval $[0, 1]$. Before implementing such a numerical procedure, as stated before, the stationary solution (undisturbed field amplitude) $U$, which has the soliton power defined by $P=\int _{-\infty }^{+\infty }|U(\xi )|^2d\xi$, should be firstly found from Eq. (8), which we have solved through the modified squared-operator method [66].

4. Numerical results and discussions

Having obtained the band-gap structure of the considered physical model above, we now turn to study the existence of optical localized gap modes within the given finite band gaps, with an emphasis on their property and stability in the first two band gaps. Here the localized gap modes of two types, the fundamental gap solitons and dipole ones, are discussed. As usual, the optical gap solitons can be classified into off-site and on-site solitons, with their central parts populating, respectively, in the minimum and maximum values of the optical lattice.

4.1 Off-site and on-site optical gap solitons

As shown in Fig. 2(b), where the dependence between the power $P$ and propagation constant $b$ for the on-site solitons (red solid line) and off-site (blue dashed line) fundamental optical gap solitons pinned in the relevant first and the second finite band gaps has been accumulated. One can obtain that the power $P$ increases with the decrease of propagation constant $b$, showing the obeying of the well-known inverted Vakhitov-Kolokolov stability criterion $dP/db<0$, which is a necessary condition for the stability of gap solitons supported by self-repulsive (self-defocusing) nonlinearity [20,67]. Insight scrutinization into the stability property of both gap solitons employing the linear stability analysis based on eigenvalue Eq. (9), we have obtained the relation of the maximal real part of eigenvalues Re$(\lambda )$ as a function of $b$ in Figs. 2(c) and 2(d) for the off-site and on-site localized gap modes, suggesting that both modes are robustly stable physical objects in the first finite band gap [recall that Re$(\lambda )=0$ as mentioned above], and in the second gap, however, the stability region is very limit for the former and the latter only has instability region because of Re$(\lambda )\neq 0$.

Typical examples of the 1D off-site and on-site optical gap solitons, with different values of propagation constant $b$ in the $U(\xi )$ profile, are respectively depicted in Figs. 3(a, b, c) and Figs. 3(d, e, f). For all the gap solitons lying in the first finite band gap [Figs. 3(a, d)], they exhibit a cusplike shape, i.e., they are accompanied by a weak modulation and such a modulation (of the gap solitons) increases near the gap edge [Figs. 3(b, e)] and in the second gap [Figs. 3(c, f)], resembling the same feature of gap solitons in other periodic physical systems [911,21]. In addition to the linear stability analysis, the stability properties of the fundamental gap solitons have also been examined in direct perturbed simulation using Eq. (7), and an agreement has been obtained between both methods, exemplified by the perturbed evolutions in Fig. 4 of these six gap solitons, with their eigenvalues being marked by points (A1, A2, A3) and (B1, B2, B3) in Figs. 2(c) and 2(d). The stable evolutions of perturbed gap solitons of both off-site and on-site types, which keep their profile over long time, are displayed in Figs. 4(a, c, d); their unstable counterparts, however, subject to decay and finally immerse into the background noise as radiating waves during evolution.

 figure: Fig. 3.

Fig. 3. Characteristic examples of off-site gap solitons (left column) and on-site ones (right column) with different propagation constant $b$. (a, d): $b=3.0$; (b, e): $b=1.78$; (c, f): $b=0.8$. The gap solitons in panels (c, f), marked by (A3, B3), lie at the second band gap, while the other four points (A1, A2, B1, B2) belong to the first band gap. The direct perturbed dynamics of these six gap modes will be shown in the following figure. The black dashed lines (in each panel), here and below, denote the shape of the optical lattice.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. Left column: Direct perturbed evolutions of stable off-site gap solitons (a, c) and unstable counterpart (b). Right column: Direct perturbed evolutions of unstable on-site gap solitons (e, f) and stable one (d). (a, d): $b=3.0$; (b, e): $b=1.78$; (c, f): $b=0.8$. $\xi \in [-20,\,20]$ for all panels.

Download Full Size | PDF

The cases we are talking about are so far limited to the strength of optical lattice $c_0=6$, revealing that there are no any stable on-site gap solitons in the second band gap. This may be explained by the fact that, a stronger Bragg scattering (in such gap), which impacts deeply on the double humps of on-site gap solitons in the evolution process, turn to destabilize rather them stabilize them. This can also be understood by the phenomenon that in the second band gap more optical waves are participating with the multiple constructive interference (Bragg scattering) so as to form gap solitons in the nonlinear regime, recalling that the power $P$ increases almost linearly through the first band gap to the second gap. We thus conjecture that an enhancement of the optical lattice strength $c_0$ may help to create stable on-site gap solitons in second gap. Given that $c_0=12$, whose band-gap structure of the underlying linear spectrum is shown in Fig. 5(a), suggesting wider gaps in both the first and second band gap as compared with the case of $c_0=6$ in Fig. 2(a). Our systematic numerical study, as those done in Fig. 2, has produced the dependence $P(b)$ for both the off-site and on-site gap solitons in Fig. 5(b), and their relevant eigenvalues via linear-stability analysis in Fig. 5(c) and in Fig. 5(d) respectively, demonstrating clearly that, in the second band gap, the on-site gap solitons are indeed stable and the off-site ones have a larger stability region. Direct simulations of the perturbed evolution of both gap solitons, stable and unstable cases, show that they share the similarities of those in Fig. 4, and it is unnecessary to show them here accordingly.

 figure: Fig. 5.

Fig. 5. (a) Band-gap structure of the EIT-based optical lattice $V_{\textrm {OL}}=-c_0 \sin ^2(\xi )$ with a larger strength $c_0=12$. (b) Power $P$ as a function of propagation constant $b$ for on-site gap solitons (red solid line) and off-site ones (blue dashed line). Panels (c) and (d) represent respectively the real parts of maximum eigenvalues, expressed by Re($\lambda$) versus $b$, for the corresponding off-site and on-site gap solitons.

Download Full Size | PDF

4.2 Dipole gap modes

The physical model may also upholds dipole gap modes, i.e., localized modes of two fundamental gap solitons arranged in a spatial distance $\bigtriangleup d$ and with opposite signs. Provided that such a distance $\bigtriangleup d$ doubling the period of the optical lattice, that is $\bigtriangleup d=2\pi$, the dipole gap modes can be stable objects, according to our systematic numerical calculations. Illustrated in Figs. 6(a) and 6(b) are, respectively, the power $P$ of off-site dipole gap localized modes versus propagation constant $b$ for the above-mentioned two different depths of the optical lattice depth $c_0=12$ and $c_0=6$. It is to observe once again that the dependence $P(b)$ follows the inverted Vakhitov-Kolokolov stability criterion $dP/db<0$ ($P$ increases when decreasing $b$) from the first band gap to the second gap, recall that which is a necessary but not sufficient condition for stable gap modes [20,67]. The corresponding relation between maximum real eigenvalues Re$(\lambda )$ and propagation constant $b$, obtained with linear-stability analysis in eigenvalue problem Eq. (9), are displayed in Figs. 6(c) and 6(d). The latter shows that the dipole gap solitons can not be stabilized in the second band gap at $c_0=6$, whereas their stable fundamental gap solitons are confined to a narrow region [see Fig. 2(c)]. Examples of such dipole gap solitons populating in the first and second band gaps for both cases [$c_0=12$ and $6$] are displayed in Figs. 6(e) and 6(f).

 figure: Fig. 6.

Fig. 6. Top: Power $P$ versus propagation constant $b$ of off-site dipole gap solitons under different optical lattice depth $c_0$: (a) $c_0=12$ and (b) $c_0=6$. Middle: The corresponding real part of maximum eigenvalue, Re($\lambda$), as a function of $b$ in (c) and (d). Below: Profiles of such dipole gap modes corresponding to the marked points (C1, D1) in the first band gap (e) and (C2, D2) in the second band gap (f). The relevant propagation constants are $b=8.0$ and $b=2.3$ for the marked points (C1, C2), and $b=3.5$ and $b=0.9$ for the marked points (D1, D2), respectively.

Download Full Size | PDF

We have found that the stable dipole gap modes keep their coherence during long time evolution, as depicted by Figs. 7(a) and 7(c). On the contrary, the unstable dipole gap solitons undergo weak oscillation initially and quickly transform themselves into radiating waves, see examples in Figs. 7(b) and 7(d). These systematic simulation results match up well with the relevant predictions from linear stability analysis in Figs. 6(c) and 6(d), demonstrating the efficiency of the numerical methods we adopted.

 figure: Fig. 7.

Fig. 7. Direct perturbed evolutions of stable off-site dipole gap solitons (a, c) and unstable ones (b, d). Depth of the optical lattices $c_0=12$ for (a, b) and $c_0=6$ for (c, d); propagation constant $b=8.0$, $b=2.3$, $b=3.5$ and $b=0.9$ in a sequential order for panels (a$\sim$d). $\xi \in [-20,\,20]$ for all.

Download Full Size | PDF

5. Conclusion

To summarize, we have presented a physical system consisting of a resonant atomic ensemble, which is in a $\Lambda$-type three-level configuration, interacting with two laser fields represented as control and probe fields to turn EIT on, and controlled by an optical lattice induced by a pair of counter-propagating far-detuned Stark laser fields. Using the typical Maxwell-Bloch equations and the multiple scales method, the governing model (equation) for the probe field envelope has been derived in the framework of nonlinear Schrödinger equation which, is found to support localized gap modes of two types, i.e., the fundamental gap solitons and dipole gap ones, in the first and second finite band gaps. Both gap modes can be constructed as on-site gap solitons and off-site ones, with their central humps residing on the maximum and minimum positions (values) of the optical lattice. Our systematic simulations basing on linear-stability analysis and direct perturbed simulations demonstrate the (in)stability regions of both localized gap modes in the respective linear band-gap spectrum. The theoretical results predicted here may enrich insights into the nonlinear properties of coherent atomic gases and open up a new door for implications ranging from optical communication to optical information processing.

An extension of this work is to consider soliton composites (alias multiple-peaked solitons) consisted of several or many gap solitons with equal spacing between each. It is also an interesting issue for the investigation of gap solitons and vortices in two-dimensional space.

Appendix

A. Equations of motion for density-matrix elements

The explicit equations of motion for density-matrix elements $\sigma _{jl}$ are given by

$$\begin{aligned} i\frac{\partial}{\partial t}\sigma_{11}-i\Gamma_{13}\sigma_{33}+\Omega_p^{\ast}\sigma_{31}-\Omega_p\sigma_{31}^{\ast}=0, \end{aligned}$$
$$\begin{aligned}i\frac{\partial}{\partial t}\sigma_{22}-i\Gamma_{23}\sigma_{33}+\Omega_{c}^{\ast}\sigma_{32}-\Omega_{c}\sigma_{32}^{\ast}=0, \end{aligned}$$
$$\begin{aligned}i\frac{\partial}{\partial t}\sigma_{33}+i(\Gamma_{13}+\Gamma_{23})\sigma_{33}-\Omega_p^{\ast}\sigma_{31}+\Omega_p\sigma_{31}^{\ast}-\Omega_c^{\ast}\sigma_{32}+\Omega_c\sigma_{32}^{\ast}=0, \end{aligned}$$
$$\begin{aligned}\left(i\frac{\partial}{\partial t}+d_{21}\right)\sigma_{21}-\Omega_p\sigma_{32}^{\ast}+\Omega_c^{\ast}\sigma_{31}=0, \end{aligned}$$
$$\begin{aligned}\left(i\frac{\partial}{\partial t}+d_{31}\right)\sigma_{31}-\Omega_p(\sigma_{33}-\sigma_{11})+\Omega_c\sigma_{21}=0, \end{aligned}$$
$$\begin{aligned}\left(i\frac{\partial}{\partial t}+d_{32}\right)\sigma_{32}-\Omega_c(\sigma_{33}-\sigma_{22})+\Omega_p\sigma_{21}^{\ast}=0, \end{aligned}$$
where $d_{jl}=\Delta _{j}^\prime -\Delta _{l}^\prime +i\gamma _{jl}$ and $\gamma _{jl}=(\Gamma _j+\Gamma _l)/2$ with $\Gamma _j=\sum _{E_i<E_j}\Gamma _{ij}$ being the spontaneous emission rate from the state $|j\rangle$ to all lower energy states $|i\rangle$.

B. Solutions of the first and the second asymptotic expansion

The first order ($m=1$) solution is given by $\Omega _{p}^{(1)}=\mathcal {F}e^{i\theta }$ and $\sigma _{j1}^{(1)}=\{[\delta _{j3}(\omega +d_{21}^{(0)})-\delta _{j2}\Omega _c^\ast ]/D\}\mathcal {F}e^{i\theta }$, where $D=|\Omega _c|^2-(\omega +d_{21}^{(0)})(\omega +d_{31}^{(0)})$ and $\theta =\mathcal {K}(\omega )z_0-\omega t_0$, with $\mathcal {K}(\omega )=\omega /c+\kappa _{13}(\omega +d_{21}^{(0)})/D$ being the linear dispersion relation. Note that the frequency and wave number of the probe field are given by $\omega _{p}+\omega$ and $k_{p}+\mathcal {K}(\omega )$, respectively. Thus $\omega =0$ corresponds to the center frequency of the probe field. In addition, $\mathcal {F}$ is a yet to be determined envelope function depending on the slow variables $x_1$, $z_2$, and $t_2$.

The second order ($m=2$) solution reads $\sigma _{21}^{(2)}=\sigma _{31}^{(2)}=0$, $\sigma _{jj}^{(2)}=a_{jj}^{(2)}|\mathcal {F}|^2e^{-2\bar {a} z_2}$ ($j=\,1,\,2$), and $\sigma _{32}^{(2)}=a_{32}^{(2)}|\mathcal {F}|^2e^{-2\bar {a} z_2}$, where

$$\begin{aligned}a_{11}^{(2)}=\frac{[i\Gamma_{23}-2|\Omega_c|^2\mathcal{P}]\mathcal{G}-i\Gamma_{13}|\Omega_c|^2\mathcal{Q}}{i\Gamma_{13}|\Omega_c|^2\mathcal{P}^\ast}, \end{aligned}$$
$$\begin{aligned}a_{22}^{(2)}=\frac{\mathcal{G}-i\Gamma_{13}a_{11}^{(2)}}{i\Gamma_{13}}, \end{aligned}$$
$$\begin{aligned}a_{32}^{(2)}=\frac{\Omega_c}{d_{32}^{(0)}}\left[\frac{1}{D^\ast}-(a_{11}^{(2)}+2a_{22}^{(2)})\right], \end{aligned}$$
and $\bar {a}=\epsilon ^{-2}\textrm {Im}[\mathcal {K}(\omega )]$, with $\mathcal {G}=(\omega +d_{21}^{(0)\ast })/D^\ast -(\omega +d_{21}^{(0)})/D$, $\mathcal {P}={1}/{d_{32}^{(0)}}-{1}/{d_{32}^{(0)\ast }}$, and $\mathcal {Q}={1}/({Dd_{32}^{(0)\ast }})-{1}/({D^\ast d_{32}^{(0)}})$.

The explicit expressions of the self- and cross-phase modulation coefficients $W_{1}$ and $W_{2}$ are written respectively as

$$\begin{aligned} W_{1}=\kappa_{13}\frac{\Omega_ca_{32}^{(2)\ast}+(\omega+d_{21}^{(0)})(2a_{11}^{(2)}+a_{22}^{(2)})}{D}, \end{aligned}$$
$$\begin{aligned}W_{2}=\kappa_{13}\frac{(\omega+d_{21}^{(0)})^2\alpha_{31}+|\Omega_c|^2\alpha_{21}}{2D^2}. \end{aligned}$$

Funding

East China University of Technology (DHBK2016118); Natural Science Foundation of Jiangxi Province (20202BABL211013); National Natural Science Foundation of China (11704066, 12074063, 12074423).

Disclosures

The authors declare no conflicts of interest.

References

1. C. J. Pethick and H. Smith, Bose-Einstein Condensation in Dilute Gases (Cambridge University, 2008).

2. O. Morsch and M. Oberthaler, “Dynamics of Bose-Einstein condensates in optical lattices,” Rev. Mod. Phys. 78(1), 179–215 (2006). [CrossRef]  

3. B. Eiermann, Th. Anker, M. Albiez, M. Taglieber, P. Treutlein, K. P. Marzlin, and M. K. Oberthaler, “Bright Bose-Einstein gap solitons of atoms with repulsive interaction,” Phys. Rev. Lett. 92(23), 230401 (2004). [CrossRef]  

4. Y. S. Kivshar and G. P. Agrawal, Optical Solitons: From Fibers to Photonic Crystals (Academic, 2003).

5. B. A. Malomed, Soliton Management in Periodic Systems (Springer, 2006).

6. Y. V. Kartashov, B. A. Malomed, and L. Torner, “Solitons in nonlinear lattices,” Rev. Mod. Phys. 83(1), 247–305 (2011). [CrossRef]  

7. I. L. Garanovich, S. Longhi, A. A. Sukhorukov, and Y. S. Kivshar, “Light propagation and localization in modulated photonic lattices and waveguides,” Phys. Rep. 518(1-2), 1–79 (2012). [CrossRef]  

8. Y. V. Kartashov, G. E. Astrakharchik, B. A. Malomed, and L. Torner, “Frontiers in multidimensional self-trapping of nonlinear fields and matter,” Nat. Rev. Phys. 1(3), 185–197 (2019). [CrossRef]  

9. L. Zeng and J. Zeng, “Gap-type dark localized modes in a Bose-Einstein condensate with optical lattices,” Adv. Photonics 1(4), 046004 (2019). [CrossRef]  

10. J. Shi and J. Zeng, “Asymmetric localized states in periodic potentials with a domain-wall-like Kerr nonlinearity,” J. Phys. Commun. 3(3), 035003 (2019). [CrossRef]  

11. L. Zeng and J. Zeng, “Preventing critical collapse of higher-order solitons by tailoring unconventional optical diffraction and nonlinearities,” Commun. Phys. 3(1), 26 (2020). [CrossRef]  

12. B. J. Eggleton, R. E. Slusher, C. M. de Sterke, P. A. Krug, and J. E. Sipe, “Bragg grating solitons,” Phys. Rev. Lett. 76(10), 1627–1630 (1996). [CrossRef]  

13. E. A. Ostrovskaya, J. Abdullaev, M. D. Fraser, A. S. Desyatnikov, and Y. S. Kivshar, “Self-localization of polariton condensates in periodic potentials,” Phys. Rev. Lett. 110(17), 170407 (2013). [CrossRef]  

14. E. A. Cerda-Méndez, D. Sarkar, D. N. Krizhanovskii, S. S. Gavrilov, K. Biermann, M. S. Skolnick, and P. V. Santos, “Exciton-polariton gap solitons in two-dimensional lattices,” Phys. Rev. Lett. 111(14), 146401 (2013). [CrossRef]  

15. D. Tanese, H. Flayac, D. Solnyshkov, A. Amo, A. LemaÎtre, E. Galopin, R. Braive, P. Senellart, I. Sagnes, G. Malpuech, and J. Bloch, “Polariton condensation in solitonic gap states in a one-dimensional periodic potential,” Nat. Commun. 4(1), 1749 (2013). [CrossRef]  

16. C. Chin, R. Grimm, P. Julienne, and E. Tiesinga, “Feshbach resonances in ultracold gases,” Rev. Mod. Phys. 82(2), 1225–1286 (2010). [CrossRef]  

17. L. Zeng and J. Zeng, “One-dimensional solitons in fractional Schrödinger equation with a spatially periodical modulated nonlinearity: nonlinear lattice,” Opt. Lett. 44(11), 2661–2664 (2019). [CrossRef]  

18. J. Shi and J. Zeng, “1D solitons in saturable nonlinear media with space fractional derivatives,” Ann. Phys. (Berlin, Ger.) 532(1), 1900385 (2020). [CrossRef]  

19. J. Chen and J. Zeng, “One-dimensional localized modes of spin-orbit-coupled Bose-Einstein condensates with spatially periodic modulated atom-atom interactions: Nonlinear lattices,” Commun. Nonlinear Sci. Numer. Simulat. 85, 105217 (2020). [CrossRef]  

20. H. Sakaguchi and B. A. Malomed, “Solitons in combined linear and nonlinear lattice potentials,” Phys. Rev. A 81(1), 013624 (2010). [CrossRef]  

21. J. Shi and J. Zeng, “Self-trapped spatially localized states in combined linear-nonlinear periodic potentials,” Front. Phys. 15(1), 12602 (2020). [CrossRef]  

22. J. Zeng, J. Zhou, G. Kurizki, and T. Opatrny, “Backward self-induced transparency in metamaterials,” Phys. Rev. A 80(6), 061806 (2009). [CrossRef]  

23. H. Leblond and D. Mihalache, “Models of few optical cycle solitons beyond the slowly varying envelope approximation,” Phys. Rep. 523(2), 61–126 (2013). [CrossRef]  

24. A. Kozhekin and G. Kurizki, “Self-induced transparency in Bragg reflectors-gap solitons near absorption resonances,” Phys. Rev. Lett. 74(25), 5020–5023 (1995). [CrossRef]  

25. A. E. Kozhekin, G. Kurizki, and B. A. Malomed, “Standing and moving gap solitons in resonantly absorbing gratings,” Phys. Rev. Lett. 81(17), 3647–3650 (1998). [CrossRef]  

26. J. Zeng, J. Zhou, G. Kurizki, and T. Opatrny, “Generation of a self-pulsed picosecond solitary wave train from a periodically amplifying Bragg structure,” Phys. Rev. A 78(1), 011803 (2008). [CrossRef]  

27. Y. Wu and L. Deng, “Ultraslow optical solitons in a cold four-state medium,” Phys. Rev. Lett. 93(14), 143904 (2004). [CrossRef]  

28. I. Friedler, G. Kurizki, O. Cohen, and M. Segev, “Spatial Thirring-type solitons via electromagnetically induced transparency,” Opt. Lett. 30(24), 3374–3376 (2005). [CrossRef]  

29. G. Huang, L. Deng, and M. G. Payne, “Dynamics of ultraslow optical solitons in a cold three-state atomic system,” Phys. Rev. E 72(1), 016617 (2005). [CrossRef]  

30. H. Michinel, M. J. Paz-Alonso, and V. M. Pérez-García, “Turning light into a liquid via atomic coherence,” Phys. Rev. Lett. 96(2), 023903 (2006). [CrossRef]  

31. C. Hang and G. Huang, “Stern-Gerlach effect of weak-light ultraslow vector solitons,” Phys. Rev. A 86(4), 043809 (2012). [CrossRef]  

32. Z. Chen and G. Huang, “Trapping of weak signal pulses by soliton and trajectory control in a coherent atomic gas,” Phys. Rev. A 89(3), 033817 (2014). [CrossRef]  

33. D. Xu, Z. Chen, and G. Huang, “Ultraslow weak-light solitons and their storage and retrieval in a kagome-structured hollow-core photonic crystal fiber,” Opt. Express 25(16), 19094–19111 (2017). [CrossRef]  

34. K. Zhang, Y. Liang, J. Lin, and H. Li, “Controlling the stability of nonlinear optical modes via electromagnetically induced transparency,” Phys. Rev. A 97(2), 023844 (2018). [CrossRef]  

35. Y. Qi, Y. Niu, F. Zhou, H. Sun, and S. Gong, “Thirring-type spatial optical solitons in asymmetric quantum wells,” J. Phys. B: At., Mol. Opt. Phys. 51(2), 025504 (2018). [CrossRef]  

36. Z. Bai, W. Li, and G. Huang, “Stable single light bullets and vortices and their active control in cold Rydberg gases,” Optica 6(3), 309–317 (2019). [CrossRef]  

37. Z. Chen, H. Xie, Q. Li, and G. Huang, “Stern-Gerlach deflection of optical Thirring solitons in a coherent atomic system,” Phys. Rev. A 100(1), 013827 (2019). [CrossRef]  

38. K. M. Devi, G. Kumar, and A. K. Sarma, “Surface polaritonic solitons and breathers in a planar plasmonic waveguide structure via electromagnetically induced transparency,” J. Opt. Soc. Am. B 36(8), 2160–2166 (2019). [CrossRef]  

39. J. Ru, Z. Wu, Y. Zhang, F. Wen, and Y. Gu, “Talbot effect in nonparaxial self-accelerating beams with electromagnetically induced transparency,” Front. Phys. 15(5), 52503 (2020). [CrossRef]  

40. H. Xu, C. Hang, and G. Huang, “Nonlocal nonlinear optical X waves and their active control in a Rydberg atomic gas,” Phys. Rev. A 101(5), 053832 (2020). [CrossRef]  

41. Z. Gu, Q. Liu, Y. Zhou, and C. Tan, “Symmetric and antisymmetric surface plasmon polariton solitons in a metal-dielectric-metal waveguide with incoherent pumping,” Eur. Phys. J. D 74(4), 78 (2020). [CrossRef]  

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

43. 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 (London) 409(6819), 490–493 (2001). [CrossRef]  

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

45. M. D. Lukin and A. Imamoğlu, “Nonlinear optics and quantum entanglement of ultraslow single photons,” Phys. Rev. Lett. 84(7), 1419–1422 (2000). [CrossRef]  

46. H. Kang and Y. Zhu, “Observation of large Kerr nonlinearity at low light intensities,” Phys. Rev. Lett. 91(9), 093601 (2003). [CrossRef]  

47. M. Afzelius, N. Gisin, and H. De Riedmatten, “Quantum memory for photons,” Phys. Today 68(12), 42–47 (2015). [CrossRef]  

48. C. Murray and T. Pohl, “Quantum and nonlinear optics in strongly interacting atomic ensembles,” in Advances in Atomic, Molecular, and Optical Physics65 (Academic, 2016), pp. 321–372.

49. Q. Wang, L. Ma, W. Cui, M. Chen, and S. Zou, “Ultra-narrow electromagnetically induced transparency in the visible and near-infrared regions,” Appl. Phys. Lett. 114(21), 213103 (2019). [CrossRef]  

50. C. Wang, X. Jiang, G. Zhao, M. Zhang, C. W. Hsu, B. Peng, A. D. Stone, L. Jiang, and L. Yang, “Electromagnetically induced transparency at a chiral exceptional point,” Nat. Phys. 16(3), 334–340 (2020). [CrossRef]  

51. Y. Hu, T. Jiang, H. Sun, M. Tong, J. You, X. Zheng, Z. Xu, and X. Cheng, “Ultrafast frequency shift of electromagnetically induced transparency in Terahertz metaphotonic devices,” Laser Photonics Rev. 14(3), 2070019 (2020). [CrossRef]  

52. G. Kurizki, D. Petrosyan, T. Opatrny, M. Blaauboer, and B. Malomed, “Self-induced transparency and giant nonlinearity in doped photonic crystals,” J. Opt. Soc. Am. B 19(9), 2066–2074 (2002). [CrossRef]  

53. J. Wang, C. Hang, and G. Huang, “Weak-light gap solitons in a resonant three-level system,” Phys. Lett. A 366(4-5), 528–533 (2007). [CrossRef]  

54. Y. Zhang, Z. Wang, H. Zheng, C. Yuan, C. Li, K. Lu, and M. Xiao, “Four-wave-mixing gap solitons,” Phys. Rev. A 82(5), 053837 (2010). [CrossRef]  

55. C. Hang, V. V. Konotop, and G. Huang, “Spatial solitons and instabilities of light beams in a three-level atomic medium with a standing-wave control field,” Phys. Rev. A 79(3), 033826 (2009). [CrossRef]  

56. Y. Li, B. A. Malomed, M. Feng, and J. Zhou, “Arrayed and checkerboard optical waveguides controlled by electromagnetically induced transparency,” Phys. Rev. A 82(6), 063813 (2010). [CrossRef]  

57. D. Petrosyan, “Tunable photonic band gaps with coherently driven atoms in optical lattices,” Phys. Rev. A 76(5), 053823 (2007). [CrossRef]  

58. J. Nunn, U. Dorner, P. Michelberger, K. F. Reim, K. C. Lee, N. K. Langford, I. A. Walmsley, and D. Jaksch, “Quantum memory in an optical lattice,” Phys. Rev. A 82(2), 022327 (2010). [CrossRef]  

59. Z. Zhang, Y. Zhang, J. Sheng, L. Yang, M. Miri, D. N. Christodoulides, B. He, Y. Zhang, and M. Xiao, “Observation of patity-time in optically induced atomic lattices,” Phys. Rev. Lett. 117(12), 123601 (2016). [CrossRef]  

60. H. Yang, T. Zhang, Y. Zhang, and J. Wu, “Dynamically tunable three-color reflections immune to disorder in optical lattices with trapped cold 87Rb atoms,” Phys. Rev. A 101(5), 053856 (2020). [CrossRef]  

61. L. Zhao, “All-optical spin-orbit coupling of light using electromagnetically induced transparency,” Phys. Rev. A 100(1), 013822 (2019). [CrossRef]  

62. Z. Zhang, F. Li, G. Malpuech, Y. Zhang, O. Bleu, S. Koniakhin, C. Li, Y. Zhang, M Xiao, and D. D. Solnyshkov, “Particlelike behavior of topological defects in linear wave packets in photonic graphene,” Phys. Rev. Lett. 122(23), 233905 (2019). [CrossRef]  

63. Z. Zhang, R. Wang, Y. Zhang, Y. V. Kartashov, F. Li, H. Zhong, H. Guan, K. Gao, F. Li, Y. Zhang, and M. Xiao, “Observation of edge solitons in photonic graphene,” Nat. Commun. 11(1), 1902 (2020). [CrossRef]  

64. Y. Guo, L. Zhou, L. M. Kuang, and C. P. Sun, “Magneto-optical Stern-Gerlach effect in an atomic ensemble,” Phys. Rev. A 78(1), 013833 (2008). [CrossRef]  

65. D. A. Steck, “Rubidium 87 D Line Data, available online at http://steck.us/alkalidata,” (2019).

66. J. Yang, Nonlinear Waves in Integrable and Nonintegrable Systems (Society for Industrial and Applied Mathematics, 2011).

67. M. Vakhitov and A. Kolokolov, “Stationary solutions of the wave equation in a medium with nonlinearity saturation,” Radiophys. Quantum Electron. 16(7), 783–789 (1973). [CrossRef]  

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 (7)

Fig. 1.
Fig. 1. (a) Theoretical scheme of a $\Lambda$ -type three-level atomic system for setting on EIT, whose excitation mechanism is described in text. $\Gamma _{13}$ ( $\Gamma _{23}$ ) is the spontaneous emission decay rate from $|3\rangle$ to $|1\rangle$ ( $|3\rangle$ to $|2\rangle$ ). $\Delta _j$ ( $j=2,\,3$ ) is the detuning. (b) Possible experimental arrangement of laser beams geometry. The weak probe and strong control fields (with angular frequency $\omega _p$ and $\omega _c$ , respectively) propagate along the $z$ direction. The (purple) thick arrows denote a pair of counter-propagating Stark fields (both with angular frequency $\omega _s$ ), which form standing waves called optical lattice.
Fig. 2.
Fig. 2. (a) Band-gap structure of linear spectrum $b(K)$ characterized by the momentum $K$ , which is induced by the EIT-based optical lattice $V_{\textrm {OL}}=-c_0 \sin ^2 (\xi )$ with strength $c_0=6$ . The 1st BG and 2nd BG hereafter represent the first and second band gaps, respectively. (b) Power $P$ versus propagation constant $b$ for on-site gap solitons (red solid line) and off-site ones (blue dashed line). Panels (c) and (d) represent respectively the real parts of maximum eigenvalues, expressed by Re( $\lambda$ ) as a function of $b$ , for the corresponding off-site and on-site gap solitons. Profiles of gap solitons marked by dotted points (A1, A2, A3) and (B1, B2, B3) will be displayed below.
Fig. 3.
Fig. 3. Characteristic examples of off-site gap solitons (left column) and on-site ones (right column) with different propagation constant $b$ . (a, d): $b=3.0$ ; (b, e): $b=1.78$ ; (c, f): $b=0.8$ . The gap solitons in panels (c, f), marked by (A3, B3), lie at the second band gap, while the other four points (A1, A2, B1, B2) belong to the first band gap. The direct perturbed dynamics of these six gap modes will be shown in the following figure. The black dashed lines (in each panel), here and below, denote the shape of the optical lattice.
Fig. 4.
Fig. 4. Left column: Direct perturbed evolutions of stable off-site gap solitons (a, c) and unstable counterpart (b). Right column: Direct perturbed evolutions of unstable on-site gap solitons (e, f) and stable one (d). (a, d): $b=3.0$ ; (b, e): $b=1.78$ ; (c, f): $b=0.8$ . $\xi \in [-20,\,20]$ for all panels.
Fig. 5.
Fig. 5. (a) Band-gap structure of the EIT-based optical lattice $V_{\textrm {OL}}=-c_0 \sin ^2(\xi )$ with a larger strength $c_0=12$ . (b) Power $P$ as a function of propagation constant $b$ for on-site gap solitons (red solid line) and off-site ones (blue dashed line). Panels (c) and (d) represent respectively the real parts of maximum eigenvalues, expressed by Re( $\lambda$ ) versus $b$ , for the corresponding off-site and on-site gap solitons.
Fig. 6.
Fig. 6. Top: Power $P$ versus propagation constant $b$ of off-site dipole gap solitons under different optical lattice depth $c_0$ : (a) $c_0=12$ and (b) $c_0=6$ . Middle: The corresponding real part of maximum eigenvalue, Re( $\lambda$ ), as a function of $b$ in (c) and (d). Below: Profiles of such dipole gap modes corresponding to the marked points (C1, D1) in the first band gap (e) and (C2, D2) in the second band gap (f). The relevant propagation constants are $b=8.0$ and $b=2.3$ for the marked points (C1, C2), and $b=3.5$ and $b=0.9$ for the marked points (D1, D2), respectively.
Fig. 7.
Fig. 7. Direct perturbed evolutions of stable off-site dipole gap solitons (a, c) and unstable ones (b, d). Depth of the optical lattices $c_0=12$ for (a, b) and $c_0=6$ for (c, d); propagation constant $b=8.0$ , $b=2.3$ , $b=3.5$ and $b=0.9$ in a sequential order for panels (a $\sim$ d). $\xi \in [-20,\,20]$ for all.

Equations (20)

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

E Stark ( x , t ) = e ^ s 2 E s ( x ) cos ( ω s t ) ,
σ t = i [ H ^ int , σ ] Γ σ .
i ( z + 1 c t ) Ω p + c 2 ω p 2 Ω p x 2 + κ 13 σ 31 = 0 ,
i ( z 2 + 1 V g t 2 ) F + c 2 ω p 2 F x 1 2 + W 1 | F | 2 F e 2 a ¯ z 2 + W 2 | E s ( 1 ) | 2 F = 0 ,
i ( s + 1 g τ ) u + 1 2 2 u ξ 2 + g 1 | u | 2 u + g 2 V ( ξ ) u = i A u ,
( i g τ + 1 2 2 ξ 2 ) v + g 1 | v | 2 v + g 2 V ( ξ ) v = i A v .
i v τ = 1 2 2 v ξ 2 + | v | 2 v + V OL ( ξ ) v .
b U = 1 2 2 U ξ 2 + | U | 2 U + V OL ( ξ ) U .
( L U 2 U 2 L ) ( p q ) = i λ ( p q ) ,
i t σ 11 i Γ 13 σ 33 + Ω p σ 31 Ω p σ 31 = 0 ,
i t σ 22 i Γ 23 σ 33 + Ω c σ 32 Ω c σ 32 = 0 ,
i t σ 33 + i ( Γ 13 + Γ 23 ) σ 33 Ω p σ 31 + Ω p σ 31 Ω c σ 32 + Ω c σ 32 = 0 ,
( i t + d 21 ) σ 21 Ω p σ 32 + Ω c σ 31 = 0 ,
( i t + d 31 ) σ 31 Ω p ( σ 33 σ 11 ) + Ω c σ 21 = 0 ,
( i t + d 32 ) σ 32 Ω c ( σ 33 σ 22 ) + Ω p σ 21 = 0 ,
a 11 ( 2 ) = [ i Γ 23 2 | Ω c | 2 P ] G i Γ 13 | Ω c | 2 Q i Γ 13 | Ω c | 2 P ,
a 22 ( 2 ) = G i Γ 13 a 11 ( 2 ) i Γ 13 ,
a 32 ( 2 ) = Ω c d 32 ( 0 ) [ 1 D ( a 11 ( 2 ) + 2 a 22 ( 2 ) ) ] ,
W 1 = κ 13 Ω c a 32 ( 2 ) + ( ω + d 21 ( 0 ) ) ( 2 a 11 ( 2 ) + a 22 ( 2 ) ) D ,
W 2 = κ 13 ( ω + d 21 ( 0 ) ) 2 α 31 + | Ω c | 2 α 21 2 D 2 .
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.