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

Controlled unidirectional reflection in cold atoms via the spatial Kramers-Kronig relation

Open Access Open Access

Abstract

It is known that the Kramers-Kronig (KK) relation between real and imaginary parts of the optical susceptibility in the frequency domain can also be realized in the space domain, as first proposed in [Nat. Photonics 9(7), 436 (2015) [CrossRef]  ]. We here study a mechanism to implement spatial KK relations in a cold atomic sample and use it to control unidirectional reflectionless for probe light incident from either the left or right side of the sample at will. In our model, the complex frequency dependent atomic susceptibility is mapped into a spatially dependent one, employing a far-detuned driving field of intensity linearly varied in space. The reflection of an incident light from one side of the sample can then be set to vanish over a specific frequency band directly by changing the driving field parameters, such as its intensity and frequency. Also, by incorporating the Bragg scattering into the spatial KK relation, the reflectivity from the opposite side of the sample, though typically small for realistic atomic densities, can be made to increase to improve the reflectivity contrast. The present scheme bears potentials for all-optical network applications that require controllable unidirectional light propagation.

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

1. Introduction

Asymmetric reflection control of the flow of a light beam is a key technique to perform photonic and quantum communication manipulations. Unidirectional reflectionless, in particular, has attracted much research efforts because of its perspective applications in developing novel photonic devices [19]. Control over reflection of a light beam incident from opposite sides of a device is usually reciprocal and static, i.e., takes place with identical reflectivity values and cannot be changed. This can be achieved, e.g., using photonic band-gap materials exhibiting a given periodic structure of the real refractive index [10,11]. Tunable photonic band gap structures have also been proven to be viable, where control of the periodic structure of the complex susceptibility can be done, e.g., through electromagnetically induced transparency (EIT) [1214], with homogeneous atomic clouds dressed by standing-wave coupling fields [1522] or periodic atomic lattices dressed by traveling-wave coupling fields [2333]. Generally speaking, it is hard to achieve asymmetric light transport through standard linear optical processes [3436], though significant progresses have been made recently in coherently driven moving atomic lattices [1,2,32] and in materials exhibiting parity-time (PT) symmetry or asymmetry [3,4,3741]. Realistic implementations of these schemes are challenging owing to complex atom-light coupling configurations, precise light field arrangement in space and peculiar balanced gain and loss over a single period.

Alternatively, asymmetric and unidirectional reflection can be realized in an inhomogeneous continuous medium as the usual Kramers-Kronig (KK) relation, satisfied by its complex optical response function, is mapped from the frequency domain to the space domain [58,4249]. In the pioneering work [5], Horsley et al. found that a non-Hermitian medium would not reflect light from one side for all incident angles if its complex permittivity satisfies the spatial KK relation, which promises also the realization of omnidirectional perfect absorber [7] and media with no transmission [50]. When extended into discrete lattices, complex potentials exhibiting the spatial KK relation may further become invisible to support a bidirectional reflectionless behavior [48,49]. Notice, in addition, that the original idea has been already experimentally demonstrated via a suitable design of different inhomogeneous media [7,43,51,52]. All above findings not only deepen our understanding of light propagation, but also provide new platforms toward multi-functional optical elements, especially those requiring perfect antireflection properties. Yet, schemes so far studied rely on fixed complex refractive index, susceptibility, or permittivity structures and hence lack of dynamic tunability.

Here we propose an efficient scheme for mapping the KK relation of a probe susceptibility from the frequency domain into the space domain in a cold atomic sample. The scheme hinges on devising a position-dependent ground level shift with a far detuned external control field whose intensity can be linearly varied in space (see Sect. 2). Other spatially dependent EIT systems have also been used to manipulate structured light beams, albeit via phase-dependent dark-states [5355]. It is worth noting that in our scheme, depending on the probe frequency, spatial KK relations may be completely or partially fulfilled or even destroyed; in the following these are termed respectively as "unbroken", "transitional", and "broken" regimes. The unbroken regime is of particular interest here because it enables us to attain reflectionless manipulation of a probe beam incident on one sample end over a tunable range of frequencies. The transitional regime may instead be exploited to attain reflectionless manipulation from both sample side ends, albeit at different frequencies. Furthermore, changing the "sign" of the ground level shift enables us to directly swap the direction of the vanishing reflectivity whereas increasing the level shift "magnitude" leads to an enhancement of the nonzero reflectivity while maintaining (on the other side) a vanishing reflectivity. Bragg scattering may also be incorporated into the spatial KK relation to further increase the forward-backward reflectivity contrast.

2. Model and equations

We consider in Fig. 1(a) a cold atomic sample extending from $x=0$ to $x=L$, driven by a weak probe field of amplitude (frequency) $\mathbf {E}_{p}$ ($\omega _{p}$) and a strong control field of amplitude (frequency) $\mathbf {E}_{c}$ ($\omega _c$). The control field is assumed to illuminate the sample along the $-y$ direction while the probe field can travel through the sample along either $x$ or $-x$ direction. All atoms are driven into the three-level $V$ configuration, as shown in Fig. 1(b), characterized by Rabi frequencies (detunings) $\Omega _{p}=\mathbf {E}_{p}\cdot \mathbf {d}_{ge}/2\hbar$ on transition $\vert g\rangle \leftrightarrow \vert e\rangle$ ($\Delta _{p}=\omega _{p}-\omega _{eg}$) and $\Omega _{c}=\mathbf {E}_{c}\cdot \mathbf {d}_{ga}/2\hbar$ on transition $\vert g\rangle \leftrightarrow \vert a\rangle$ ($\Delta _{c}=\omega _{c}-\omega _{ag}$), being $\mathbf {d}_{\mu \nu }$ and $\omega _{\mu \nu }$ dipole moments and resonant frequencies of relevant transitions. We have also used $\Gamma _{ag}$ and $\Gamma _{eg}$ to describe the population decay rates from levels $|a\rangle$ and $|e\rangle$ to level $|g\rangle$, respectively. To be more concrete, levels $|a\rangle$, $|e\rangle$, and $|g\rangle$ may refer to states $\vert 5P_{3/2},F=3,m_{F}=3\rangle$, $\vert 5P_{1/2},F=1,m_{F}=1\rangle$, and $\vert 5S_{1/2},F=2,m_{F}=2\rangle$ of $^{87}$Rb atoms, respectively. This choice ensures that ($i$) dipole moments $|\mathbf {d}_{ge}|$ and $|\mathbf {d}_{ga}|$ take the largest possible values, which could relax the requirement of a very dense atomic sample for achieving a high reflectivity contrast; ($ii$) the control field doesn’t couple level $|g\rangle$ to a fourth level even in the case of a large $|\Delta _{c}|$, because no others except level $|a\rangle$ has $m_{F}=3$ on the $D_{2}$ line of $^{87}$Rb atoms. Most importantly, we will assume that the control field is linearly varied in intensity along the $x$ direction, e.g., by a neutral density filter (NDF). In this case, $\vert \Omega _{c}\vert ^{2}$ should be replaced by $\vert \Omega _{c}(x)\vert ^{2}=x\vert \Omega _{c0}\vert ^{2}/L$ with $\Omega _{c0}$ denoting the maximal Rabi frequency at $x=L$.

 figure: Fig. 1.

Fig. 1. (a) A cold atomic sample illuminated by a control beam $\mathbf {E}_{c}(x)$ along the $-y$ direction exhibits a strongly asymmetric reflection for a probe beam $\mathbf {E}_{p}$ incident either along the $+x$ or the $-x$ direction. (b) A three-level atomic system driven by a probe field of Rabi frequency $\Omega _{p}$ (detuning $\Delta _{p}$) and a control field of Rabi frequency $\Omega _c(x)$ (detuning $\Delta _c$) into the $V$ configuration. (c) A two-level atomic system with a dynamic shift $\delta (x)$ of level $\vert g\rangle$ upon the adiabatic elimination of level $\vert a\rangle$.

Download Full Size | PDF

With the electric-dipole and rotating-wave approximations, working in the weak probe limit, we can solve density matrix equations for the three-level $V$ configuration to attain the steady-state probe susceptibility

$$\chi_{3}(\Delta_{p},x)=i\frac{N_{0}\vert \mathbf{d}_{ge}\vert ^{2}}{\varepsilon _{0}\hbar }\frac{[(\gamma _{ga}^{2}+\Delta _{c}^{2}) +\vert \Omega _{c}(x)\vert ^{2}] [\gamma_{ea}-i(\Delta _{c}-\Delta _{p})]-\vert \Omega _{c}(x)\vert ^{2}(\gamma _{ga}-i\Delta_{c})} {[(\gamma_{ga}^{2}+\Delta _{c}^{2}) +\vert\Omega _{c}(x)\vert ^{2}] \{[\gamma _{ea}-i(\Delta _{c}-\Delta _{p})](\gamma _{ge}-i\Delta _{p})+\vert \Omega_{c}(x)\vert ^{2}\}},$$
where $N_{0}$ denotes the homogeneous atomic density while $\gamma _{\mu \nu }$ is the coherence dephasing rate on transition $|\mu \rangle \leftrightarrow |\nu \rangle$ with $\gamma _{ge}=\Gamma _{eg}/2$, $\gamma _{ga}=\Gamma _{ag}/2$, and $\gamma _{ea}=(\Gamma _{ag}+\Gamma _{eg})/2$. Now we consider that the control field is far detuned from transition $|g\rangle \leftrightarrow |a\rangle$ by requiring $\Delta _{c}\gg \Omega _{c0}, \gamma _{ga}$. In this case, level $\vert a\rangle$ can be adiabatically eliminated from the three-level $V$ configuration to yield a two-level system [see Fig. 1(c)], in which level $\vert g\rangle$ suffers a position-dependent energy shift $\delta (x)=\vert \Omega _{c}(x)\vert ^{2}/\Delta _{c}=x\delta _{0}/L$ with $\delta _{0}=\vert \Omega _{c0}\vert ^{2}/\Delta _{c}$. Then the probe susceptibility can be cast into a more compact form as given by
$$\chi_{2}(\Delta_{p},x) =i\frac{N_{0}\vert \mathbf{d}_{ge}\vert ^{2}}{\varepsilon _{0}\hbar }\frac{1}{\gamma _{ge}-i[\Delta_{p}+\delta (x)]},$$
whose real ($\chi _{2}^{\prime }$) and imaginary ($\chi _{2}^{\prime \prime }$) parts govern the local dispersive and absorption properties around the probe resonance, respectively. The possibility to control the effective detuning $\Delta _{p}^{eff}(x)=\Delta _{p}+\delta (x)$ by changing the ground level shift $\delta (x)$ amounts to a direct mapping from the spatially linear variation of the control intensity to that of the probe detuning.

The real and imaginary parts of $\chi _{2}$ are well known to be related via the KK relation in the frequency domain based on the causality principle and Cauchy’s theorem if we set $\delta (x)\equiv 0$ [56]. Thus, for appropriate probe detunings and sufficiently long samples such that the spatial variation of the effective probe detuning induced by $\delta (x)$ is fully developed, as shown by the red solid lines in Figs. 2(a) and 2(b), the KK relation holds in the space domain and can be defined by the following Cauchy’s principal value integral

$$\chi_{2}^{\prime}(\Delta_{p},x)=\frac{1}{\pi}\textrm{P}\int_{0}^{L}\frac{\chi_{2}^{\prime\prime}(\Delta_{p},s)}{s-x}ds$$
over spatial coordinate $s$ along the $x$ direction, where P indicates the principal part of the integral after excluding the singular point $s=x$.

 figure: Fig. 2.

Fig. 2. (a) Real and (b) imaginary parts of susceptibility $\chi _{2}$ against position $x$ with $\delta _{0}/\gamma _{ge}$=100 and $\Delta _p/\gamma _{ge}=-50$ (red-solid); $\delta _{0}/\gamma _{ge}$=20 and $\Delta _p/\gamma _{ge}=-10$ (blue-dashed). (c) Real and (d) imaginary parts of susceptibility $\chi _{2}$ against position $x$ and detuning $\Delta _p$ with $\delta _{0}/\gamma _{ge}$=100. Other parameters used in calculations are $\gamma _{ge}=2 \pi \times 2.87$ MHz, $d_{ge}=1.79\times 10^{-29}$ C$\cdot$m, $N_{0}=2.0 \times 10^{13}$ cm$^{-3}$, $L=5.0$ $\mu$m, and $\lambda _{p}=795$ nm.

Download Full Size | PDF

A medium described by $\chi _{2}$ in Eq. (2) is expected to exhibit asymmetric light transport features, which can be examined via the standard transfer-matrix method [5759] as sketched below. First, the atomic sample is partitioned along the $x$ direction into a large number ($J$) of slices labeled by $j\in [1,J]$, which exhibit slightly different susceptibilities $\bar {\chi }_{j}(\Delta _{p})=\chi _{2}(\Delta _{p},jl)$ but identical length $l=L/J=10$ nm. Second, a $2\times 2$ unimodular transfer matrix $M_{j}(\Delta _{p},l)$ characterized by $l$ and $\bar {\chi }_{j}(\Delta _{p})$ can be established to describe the propagation of an incident probe field of wavelength $\lambda _{p}$ through the $jth$ slice by

$$\left[ \begin{array}{c} E_{p}^{+}(\Delta_{p},jl) \\ E_{p}^{-}(\Delta_{p},jl) \end{array} \right] =M_{j}(\Delta_{p},l) \left[ \begin{array}{c} E_{p}^{+}(\Delta_{p},jl-l) \\ E_{p}^{-}(\Delta_{p},jl-l) \end{array} \right] ,$$
where $E_{p}^{+}$ and $E_{p}^{-}$ denote the forward and backward components of the probe field, respectively. Third, the total transfer matrix of the atomic sample turns out to be $M(\Delta _{p},L)=M_{J}(\Delta _{p},l)\cdot \cdot \cdot M_{j}(\Delta _{p},l)\cdot \cdot \cdot M_{1}(\Delta _{p},l)$ as a multiplication of the individual transfer matrices of all atomic slices. Finally, we can write the reflectivities ($R_{l}\ne R_{r}$) and the transmissivities ($T=T_{l}=T_{r}$) as
$$\begin{aligned} R_{l}(\Delta_{p},L) &= \left\vert r_{l}(\Delta_{p},L)\right\vert ^{2}=\left\vert \frac{M_{(21)}(\Delta_{p},L)}{M_{(22)}(\Delta_{p},L)}\right\vert ^{2},\\ R_{r}(\Delta_{p},L) &= \left\vert r_{r}(\Delta_{p},L)\right\vert ^{2}=\left\vert \frac{M_{(12)}(\Delta_{p},L)}{M_{(22)}(\Delta_{p},L)}\right\vert ^{2}, \\ T(\Delta_{p},L) &= \left\vert t(\Delta_{p},L)\right\vert ^{2}=\left\vert \frac{1}{M_{(22)}(\Delta_{p},L)}\right\vert^{2} \end{aligned}$$
in terms of the four matrix elements of $M(\Delta _{p},L)$. Here the subscripts ‘$l$’ and ‘$r$’ have been used to denote that the weak probe field is incident from the left and right sides along the $x$ and $-x$ directions, respectively.

3. Results and discussion

In this section, we show via numerical calculations how to implement the spatial KK relation in a narrow spectral range by tailoring the complex probe susceptibility, and how to implement the unidirectional reflection of a high reflectivity contrast by utilizing the spatial KK relation. All numerical calculations will be done with realistic parameters corresponding to the three states of $^{87}$Rb atoms chosen above, though our driving configuration can also be realized, e.g., in other alkali metal atoms.

First, we plot in Figs. 2(a) and 2(b), respectively, the real and imaginary parts of $\chi _{2}$ against position $x$ for two sets of parameters making the choice $\Delta _{p}=-\delta _{0}/2$, which allows the effective probe detuning $\Delta _{p}^{eff}(x)$ to be in the range of $\{-\delta _{0}/2,\delta _{0}/2\}$. It is clear that $\chi _{2}^{\prime }$ and $\chi _{2}^{\prime \prime }$ show an odd profile and an even profile, respectively, centered at $z=L/2$ and practically fully contained by the atomic sample, so they should satisfy the spatial KK relation described by Eq. (3). We also can see that smaller (larger) values of $\delta _{0}/\gamma _{12}$, a key dimensionless parameter in our reduced two-level system, will result in broader (narrower) spatial profiles of $\chi _{2}^{\prime }$ and $\chi _{2}^{\prime \prime }$ yet without changing their peak amplitudes. To reveal the frequency-dependent feature, we plot in Figs. 2(c) and 2(d), respectively, the real and imaginary parts of $\chi _{2}$ against both position $x$ and detuning $\Delta _{p}$ instead. The profiles of $\chi _{2}^{\prime }$ and $\chi _{2}^{\prime \prime }$ are found to move simultaneously toward the left (right) sample end with the increasing (decreasing) of $\Delta _{p}$. Accordingly, the spatial KK relation will be gradually destroyed because Eq. (3) becomes less and less satisfied.

In order to assess the extent to which the spatial KK relation is satisfied for different probe detunings in a finite atomic sample, here we propose the following integral

$$D_{kk}(\Delta_{p})=\frac{\int_{0}^{L}\left\{\chi_{2}^{\prime}(\Delta_{p},x) -\frac{1}{\pi}\textrm{P}\int_{0}^{L}\frac{\chi_{2}^{\prime\prime}(\Delta_{p},s)} {s-x}ds\right\} dx}{\left\vert \int_{0}^{L}\chi_{2}^{\prime}(\Delta_{p},x)dx\right\vert }$$
as a figure of merit for the spatial KK relation. In this definition, $D_{kk}=0$ denotes the unbroken regime where the spatial KK relation is fully satisfied while $D_{kk}=\pm 1$ denote the broken regime where the spatial KK relation is fully destroyed. According to Eq. (2) and Fig. 2, the spatial profiles of $\chi _{2}^{\prime }$ and $\chi _{2}^{\prime \prime }$ are well contained within our finite atomic sample only for a small range of $\Delta _{p}$, so the validity of the spatial KK relation is expected to increasingly deteriorate ($D_{kk}=0\to D_{kk}=\pm 1$) as $\Delta _{p}$ is gradually modulated out of this range. This is different from all previous works on spatial KK relation [58,4249], where the susceptibility or permittivity has been assumed to be fixed by design, i.e. not tunable.

Then we plot in Fig. 3(a) the reflection and transmission spectra for the parameters used in Figs. 2(a) and 2(b) based on Eq. (5). It is easy to see that these spectra can be divided into three regions: ($I$) where we have $R_{l}=R_{r}\to 0$ and $T\to 1$; ($II$) where $R_{l}\ne R_{r}$ and $T$ are sensitive to $\Delta _{p}$; ($III$) where $T\simeq 0.05$, $R_{l}\to 0$, but $R_{r}$ oscillates around $0.2$. The generation of three different regions can be understood by examining in Fig. 3(b) the figure of merit $D_{kk}$ against probe detuning $\Delta _{p}$, which clearly shows, as compared to Fig. 3(a), that $D_{kk}$ governs the relation between $R_{l}$, $R_{r}$, and $T$. The symmetric ($I$), asymmetric ($II$), and unidirectional ($III$) reflection regions correspond, respectively, to the broken ($D_{kk}=\pm 1$), transitional ($0<|D_{kk}|<1$), and unbroken ($D_{kk}=0$) regimes, and to the cases when the absorption ($\chi _{2}^{\prime \prime }$) and dispersion ($\chi _{2}^{\prime }$) profiles move out of the sample, lie at the sample boundaries, and are well contained by the sample. It is worth noting that in region ($III$) a left (right) incident probe beam is reflectionless (partially reflected) because it first sees the negative (positive) peak of $\chi _{2}^{\prime }$ [5], and the resonant absorption ($\chi _{2}^{\prime \prime }$) is already strong enough to yield $T \to 0.0$ for forward photons while the dispersion profile ($\chi _{2}^{\prime }$) is not too sharp to yield $R_{r}\to 1.0$ for backward photons. One way for further reducing $T$ and simultaneously increasing $R_{r}$ is to produce enhanced absorption profiles and sharper dispersion profiles in denser atomic samples. The restricted range of densities of cold atoms available in experiment, however, places a constraint on this approach.

 figure: Fig. 3.

Fig. 3. (a) Reflectivity $R_{l}$, reflectivity $R_{r}$, and half transmissivity $T$; (b) figure of merit $D_{kk}$ against detuning $\Delta _{p}$. Other parameters used in calculations are the same as in Fig. 2(a,b).

Download Full Size | PDF

Figure 4(a) further shows the different regimes on a diagram with $D_{kk}$ plotted against $\delta _{0}$ and $\Delta _{p}$, in which the green region ($D_{kk}=-1$) and the red region ($D_{kk}=1$) refer to the broken regime ($I$); the four narrow blue regions ($0<|D_{kk}|<1$) refer to the transitional regime ($II$); the two triangular yellow regions ($D_{kk}=0$) refer to the unbroken regime ($III$). It is also clear that the widths of two yellow regions depend critically on the magnitude of $\delta _{0}$; a broken regime may be converted into an unbroken regime and vice versa for a given $\Delta _{p}$ by changing the sign of $\delta _{0}$. Accordingly, it is viable to enlarge or reduce the reflectionless frequency band by varying the magnitude of $\delta _{0}$ and convert the sample from left reflectionless to right reflectionless or vice versa by changing the sign of $\delta _{0}$. This potentially dynamic controllability, a chief feature of our proposal, is well demonstrated in Figs. 4(b) and 4(c) in terms of reflectivities $R_{l}$ and $R_{r}$.

 figure: Fig. 4.

Fig. 4. (a) Figure of merit $D_{kk}$, (b) reflectivity $R_{l}$, and (c) reflectivity $R_{r}$ (c) against shift $\delta _{0}$ and detuning $\Delta _{p}$. Other parameters used in calculations are the same as in Figs. 2(a) and (b).

Download Full Size | PDF

It is also interesting to examine what could happen for reflectivities $R_{l}$ and $R_{r}$ when sample length $L$ is multiplied while atomic density $N_{0}$ remains invariant. In this case, we can see from $\delta (x)=x\delta _{0}/L$ that the linear variation occurs in a much larger range while its magnitude $\delta _{0}$ is unchanged. Then, as shown in Fig. 5, the spatially wider/smoother dispersion ($\chi _{2}^{\prime }$) and absorption ($\chi _{2}^{\prime \prime }$), of $L$-independent maxima and minima, together result in a notable reduction of $R_{r}$ while $R_{l}$ remains vanishing in the unbroken regime. The peak of $R_{l}$ ($R_{r}$) accompanied by $R_{r}\to 0$ ($R_{l}\to 0$) in the transitional regime has a $L$-independent position because it only appears as the main profiles of $\chi _{2}^{\prime }(x)$ and $\chi _{2}^{\prime \prime }(x)$ approach and even partially leave the left (right) sample end. It is clearly not a result of the spatial KK relation and allows two probe beams of different frequencies to be simultaneously reflected or not when they are incident upon the opposite sample ends. The damped oscillations of $R_{r}$ against $\Delta _{p}$ in the unbroken regime can be understood as a multiple interference effect due to the discontinuities of the probe susceptibility at the right sample end and at the resonant position inside the sample. It is clear that stronger (weaker) oscillations occur at larger (smaller) values of $|\Delta _{p}|$ because the resonant position of $\chi _{2}^{\prime }(x)$ and $\chi _{2}^{\prime \prime }(x)$ is close to (far from) the right sample end, yielding thus stronger (weaker) discontinuities. The oscillation period can be roughly estimated as $d\Delta _{p}\simeq \delta _{0}/L\cdot \lambda _{p}/2$ by considering that the interval $d\Delta _{p}$ of two adjacent maxima corresponds to a $2\pi$ phase shift ($\lambda _{p}/2$ spatial shift) gained by the reflected photons ($\chi _{2}^{\prime }(x)$ and $\chi _{2}^{\prime \prime }(x)$).

 figure: Fig. 5.

Fig. 5. Reflectivities (a) $R_{l}$ and (b) $R_{r}$ against detuning $\Delta _{p}$ for $L=10$ $\mu$m (black-squares), $L=15$ $\mu$m (red-circles), and $L=20$ $\mu$m (blue-triangles). Other parameters used in calculations are the same as in Figs. 2(a) and (b). Black (red) curves are shown with a vertical offset $0.1$ ($0.05$) in both insets.

Download Full Size | PDF

Finally, we note that an experiment may typically have a lower atomic density and a larger sample length than in simulations presented so far. To overcome this difficulty, we need to find an alternative way to enhance the nonzero reflectivity in the unbroken regime. This can be done by loading cold atoms into a 1D optical lattice to create a spatially periodic density $N_{j}(x)$ as described in the caption of Fig. 6, yielding thus Bragg scattering incorporated into the spatial KK relation. As shown in Fig. 6(a) and Fig. 6(b), both dispersion $\chi _{2}^{\prime }(x)$ and absorption $\chi _{2}^{\prime \prime }(x)$ of one-order lower values now exhibit the comb-like spatial profiles while satisfying to a less extent the spatial KK relation. In this case, we can find from Fig. 6(c) that $R_{l}$ and $R_{r}$ are strongly asymmetric in a much smaller frequency range, e.g., with $R_{r}$ exhibiting a maximal value up to $0.54$ while $R_{l}\lesssim 0.01$ for $-25.5\lesssim \Delta _{p}/\gamma _{ge}\lesssim -11.0$. Figure 6(d) further shows that the reflectivity contrast $C=(R_{r}-R_{l})/(R_{r}+R_{l})$, an important figure of merit on the asymmetric reflection, could be up to $0.97$ and is over $0.90$ for $-27.0\lesssim \Delta _{p}/\gamma _{ge}\lesssim -4.5$. It is noticeable that the incorporation of Bragg scattering, typically yielding symmetric reflectivities, has negligible effects on the vanishing reflectivity but largely enhances the nonzero reflectivity and the reflectivity contrast. That means, replacing a constant density $N_{0}$ with a periodic density $N_{j}(x)$ does not hamper the implementation of spatial KK relation, which is essential for developing nonreciprocal optical devices requiring a high reflectivity contrast.

 figure: Fig. 6.

Fig. 6. Spatial profiles of (a) $\chi _{2}^{\prime }$ and (b) $\chi _{2}^{\prime \prime }$ as well as corresponding spectra of (c) reflectivities $R_{l,r}$ and (d) contrast $C$ for an atomic lattice of density $N_{j}(x)=N_{0}e^{-(x-x_{j})^{2}/\delta x^{2}}$ in the $jth$ trap of center $x_{j}=(j-1/2)\Lambda$, width $\delta x$, and period $\Lambda$. Parameters are the same as in Fig. 2(b) except $N_{0}=2.0 \times 10^{12}$ cm$^{-3}$, $L=60$ $\mu$m, $\delta _{0}/\gamma _{ge}=30$, $\Lambda =400$ nm, and $\delta x=\Lambda /6$.

Download Full Size | PDF

It has been shown that atoms could be trapped and guided using nanofabricated wires and surfaces to form atom chips [60]. These chips provide a versatile experimental platform with cold atoms and constitute the basis for wide and robust applications ranging from atom optics to quantum optics. They have been used, for instance, in diverse experiments involving quantum simulation, metrology, and information processing [6163]. We then believe that our proposal is well poised to atom-chip implementations in integrated optical devices.

4. Conclusions

In summary, we have investigated the spatial KK relation and relevant reflection features in a short and dense sample of cold $^{87}$Rb atoms. This nontrivial relation in regard of the probe susceptibility is enabled by generating a position-dependent ground level shift $\delta (x)$ with a far detuned control field of intensity linearly varied along the $x$ direction. We find, in particular, that the figure of merit $D_{kk}$ characterizing the spatial KK relation may switch from the unbroken regime of unidirectional reflection, via the transitional regime of asymmetric reflection, to the broken regime of symmetric reflection, or vice versa. This is attained by increasing the maximal level shift $\delta _{0}$ from a negative value to a positive value or considering an inverse process, depending on the sign of probe detuning $\Delta _{p}$. A swapping between the nonzero reflectivity and the vanishing reflectivity at opposite sample ends is also viable by changing the sign of maximal level shift $\delta _{0}$. It is of more interest that the nonzero reflectivity can be well enhanced to result in a high reflectivity contrast for lower densities and larger lengths in a cold atomic lattice, indicating that Bragg scattering does not hamper or spoil the main effects of spatial KK relation. Such a feasible scheme may inspire further applications in nonreciprocal optical devices and provide a promising platform for atom-chip implementations.

Funding

National Natural Science Foundation of China (11534002, 11674049, 11704064, 11861131001); Ministero degli Affari Esteri e della Cooperazione Internazionale (PGR00960); Program of State Key Laboratory of Quantum Optics and Quantum Optics Devices (KF201807); Fundamental Research Funds for the Central Universities (2412019FZ045).

Disclosures

The authors declare that there are no conflicts of interest related to this article.

References

1. D.-W. Wang, H.-T. Zhou, M.-J. Guo, J. X. Zhang, J. Evers, and S.-Y. Zhu, “Optical diode made from a moving photonic crystal,” Phys. Rev. Lett. 110(9), 093901 (2013). [CrossRef]  

2. S. A. R. Horsley, J.-H. Wu, M. Artoni, and G. C. La Rocca, “Optical nonreciprocity of cold atom bragg mirrors in motion,” Phys. Rev. Lett. 110(22), 223602 (2013). [CrossRef]  

3. J.-H. Wu, M. Artoni, and G. C. La Rocca, “Non-Hermitian degeneracies and unidirectional reflectionless atomic lattices,” Phys. Rev. Lett. 113(12), 123004 (2014). [CrossRef]  

4. J.-H. Wu, M. Artoni, and G. C. La Rocca, “Parity-time-antisymmetric atomic lattices without gain,” Phys. Rev. A 91(3), 033811 (2015). [CrossRef]  

5. S. A. R. Horsley, M. Artoni, and G. C. La Rocca, “Spatial Kramers-Kronig relations and the reflection of waves,” Nat. Photonics 9(7), 436–439 (2015). [CrossRef]  

6. S. Longhi, “Bidirectional invisibility in Kramers-Kronig optical media,” Opt. Lett. 41(16), 3727–3730 (2016). [CrossRef]  

7. W. Jiang, Y. Ma, J. Yuan, G. Yin, W. Wu, and S. He, “Deformable broadband metamaterial absorbers engineered with an analytical spatial Kramers-Kronig permittivity profile,” Laser Photonics Rev. 11(1), 1600253 (2017). [CrossRef]  

8. S. A. R. Horsley and S. Longhi, “Spatiotemporal deformations of reflectionless potentials,” Phys. Rev. A 96(2), 023841 (2017). [CrossRef]  

9. A. Calabrese, F. Ramiro-Manzano, H. M. Price, S. Biasi, M. Bernard, M. Ghulinyan, I. Carusotto, and L. Pavesi, “Unidirectional reflection from an integrated taiji microresonator,” Photonics Res. 8(8), 1333–1341 (2020). [CrossRef]  

10. S. John, “Strong localization of photons in certain disordered dielectric superlattices,” Phys. Rev. Lett. 58(23), 2486–2489 (1987). [CrossRef]  

11. K. Sakoda, Optical properties of photonic crystals (Springer, 2001).

12. S. E. Harris, “Electromagnetically induced transparency,” Phys. Today 50(7), 36–42 (1997). [CrossRef]  

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

14. N. Kukharchyk, D. Sholokhov, O. Morozov, S. L. Korableva, A. A. Kalachev, and P. A. Bushev, “Electromagnetically induced transparency in a mono-isotopic 167Er:7LiYF4 crystal below 1 Kelvin: microwave photonics approach,” Opt. Express 28(20), 29166–29177 (2020). [CrossRef]  

15. A. André and M. D. Lukin, “Manipulating light pulses via dynamically controlled photonic band gap,” Phys. Rev. Lett. 89(14), 143602 (2002). [CrossRef]  

16. X.-M. Su and B. S. Ham, “Dynamic control of the photonic band gap using quantum coherence,” Phys. Rev. A 71(1), 013821 (2005). [CrossRef]  

17. M. Artoni and G. C. La Rocca, “Optically tunable photonic stop bands in homogeneous absorbing media,” Phys. Rev. Lett. 96(7), 073905 (2006). [CrossRef]  

18. J. Wu, Y. Liu, D.-S. Ding, Z.-Y. Zhou, B.-S. Shi, and G. C. Guo, “Light storage based on four-wave mixing and electromagnetically induced transparency in cold atoms,” Phys. Rev. A 87(1), 013845 (2013). [CrossRef]  

19. Y. Zhang, Y.-M. Liu, X.-D. Tian, T.-Y. Zheng, and J.-H. Wu, “Tunable high-order photonic band gaps of ultraviolet light in cold atoms,” Phys. Rev. A 91(1), 013826 (2015). [CrossRef]  

20. D.-W. Wang, H. Cai, L. Yuan, S.-Y. Zhu, and R.-B. Liu, “Topological phase transitions in superradiance lattices,” Optica 2(8), 712–715 (2015). [CrossRef]  

21. F. Wen, H. Ye, X. Zhang, W. Wang, S. Li, H. Wang, Y. Zhang, and C. Qiu, “Optically induced atomic lattice with tunable near-field and far-field diffraction patterns,” Photonics Res. 5(6), 676–683 (2017). [CrossRef]  

22. H. Yang, T. Zhang, Y. Zhang, and J.-H. 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]  

23. G. Birkl, M. Gatzke, I. H. Deutsch, S. L. Rolston, and W. D. Phillips, “Bragg scattering from atoms in optical lattices,” Phys. Rev. Lett. 75(15), 2823–2826 (1995). [CrossRef]  

24. I. H. Deutsch, R. J. C. Spreeuw, S. L. Rolston, and W. D. Phillips, “Photonic band gaps in optical lattices,” Phys. Rev. A 52(2), 1394–1410 (1995). [CrossRef]  

25. P. M. Visser and G. Nienhuis, “Band gaps and group velocity in optical lattices,” Opt. Commun. 136(5-6), 470–479 (1997). [CrossRef]  

26. S. Slama, C. von Cube, M. Kohler, C. Zimmermann, and P. W. Courteille, “Multiple reflections and diffuse scattering in Bragg scattering at optical lattices,” Phys. Rev. A 73(2), 023424 (2006). [CrossRef]  

27. M. Antezza and Y. Castin, “Fano-Hopfield model and photonic band gaps for an arbitrary atomic lattice,” Phys. Rev. A 80(1), 013816 (2009). [CrossRef]  

28. A. Schilke, C. Zimmermann, P. W. Courteille, and W. Guerin, “Photonic band gaps in one-dimensionally ordered cold atomic vapors,” Phys. Rev. Lett. 106(22), 223903 (2011). [CrossRef]  

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

30. D. Yu, “Photonic band structure of the three-dimensional 88Sr atomic lattice,” Phys. Rev. A 84(4), 043833 (2011). [CrossRef]  

31. A. Schilke, C. Zimmermann, and W. Guerin, “Photonic properties of one-dimensionally-ordered cold atomic vapors under conditions of electromagnetically induced transparency,” Phys. Rev. A 86(2), 023809 (2012). [CrossRef]  

32. L. Yang, Y. Zhang, X.-B. Yan, Y. Sheng, C.-L. Cui, and J.-H. Wu, “Dynamically induced two-color nonreciprocity in a tripod system of a moving atomic lattice,” Phys. Rev. A 92(5), 053859 (2015). [CrossRef]  

33. Y. Zhang, Y.-M. Liu, T.-Y. Zheng, and J.-H. Wu, “Light reflector, amplifier, and splitter based on gain-assisted photonic band gaps,” Phys. Rev. A 94(1), 013836 (2016). [CrossRef]  

34. P. Yeh, Optical Waves in Layered Media (Wiley-Interscience, 2005).

35. L. Chang, X. Jiang, S. Hua, C. Yang, J. Wen, L. Jiang, G. Li, G. Wang, and X. Min, “Parity-time symmetry and variable optical isolation in active-passive-coupled microresonators,” Nat. Photonics 8(7), 524–529 (2014). [CrossRef]  

36. B. Peng, S. K. Özdemir, F. Lei, F. Monifi, M. Gianfreda, G. L. Long, S. Fan, F. Nori, C. M. Bender, and L. Yang, “Parity-time-symmetric whispering-gallery microcavities,” Nat. Phys. 10(5), 394–398 (2014). [CrossRef]  

37. Z. Lin, H. Ramezani, T. Eichelkraut, T. Kottos, H. Cao, and D. N. Christodoulides, “Unidirectional invisibility induced by PT-symmetric periodic structures,” Phys. Rev. Lett. 106(21), 213901 (2011). [CrossRef]  

38. L. Feng, Y.-L. Xu, W. S. Fegadolli, M.-H. Lu, J. E. B. Oliveira, V. R. Almeida, Y.-F. Chen, and A. Scherer, “Experimental demonstration of a unidirectional reflectionless parity-time metamaterial at optical frequencies,” Nat. Mater. 12(2), 108–113 (2013). [CrossRef]  

39. X. Wang and J. -H. Wu, “Optical PT-symmetry and PT-antisymmetry in coherently driven atomic lattices,” Opt. Express 24(4), 4289–4298 (2016). [CrossRef]  

40. S. K. Gupta, Y. Zou, X.-Y. Zhu, M.-H. Lu, L.-J. Zhang, X.-P. Liu, and Y.-F. Chen, “Parity-Time Symmetry in Non-Hermitian Complex Optical Media,” Adv. Mater. 32, 1903639 (2019). [CrossRef]  

41. Yuan Lijun and Yan Lu Ya, “Unidirectional reflectionless transmission for two-dimensional PT-symmetric periodic structures,” Phys. Rev. A 100(5), 053805 (2019). [CrossRef]  

42. A. Akyurtlu and A.-G. Kussow, “Relationship between the Kramers-Kronig relations and negative index of refraction,” Phys. Rev. A 82(5), 055802 (2010). [CrossRef]  

43. D. Ye, C. Cao, T. Zhou, and J. Huangfu, “G. Zheng and L. Ran, Observation of reflectionless absorption due to spatial Kramers-Kronig profile,” Nat. Commun. 8(1), 51 (2017). [CrossRef]  

44. S. Longhi, “Wave reflection in dielectric media obeying spatial Kramers-Kronig relations,” EPL 112(6), 64001 (2015). [CrossRef]  

45. T. G. Philbin, “All-frequency reflectionlessness,” J. Opt. 18(1), 01LT01 (2016). [CrossRef]  

46. S. A. R. Horsley, M. Artoni, and G. C. La Rocca, “Reflection of waves from slowly decaying complex permittivity profiles,” Phys. Rev. A 94(6), 063810 (2016). [CrossRef]  

47. C. G. King, S. A. R. Horsley, and T. G. Philbin, “Perfect transmission through disordered media,” Phys. Rev. Lett. 118(16), 163201 (2017). [CrossRef]  

48. S. Longhi, “Kramers-Kronig potentials for the discrete Schrödinger equation,” Phys. Rev. A 96(4), 042106 (2017). [CrossRef]  

49. S. Longhi, “Reflectionless and invisible potentials in photonic lattices,” Opt. Lett. 42(16), 3229–3232 (2017). [CrossRef]  

50. C. G. King, S. A. R. Horsley, and T. G. Philbin, “Zero reflection and transmission in graded index media,” J. Opt. 19(8), 085603 (2017). [CrossRef]  

51. K. G. Makris, I. Krešić, A. Brandstötter, and S. Rotter, “Scattering-free channels of invisibility across non-Hermitian media,” Optica 7(6), 619–623 (2020). [CrossRef]  

52. D. Liu, Y. Huang, H. Hu, L. Liu, D. Gao, L. Ran, D. Ye, and Y. Luo, “Designing spatial Kramers–Kronig media using transformation optics,” IEEE Trans. Antennas Propag. 68(4), 2945–2949 (2020). [CrossRef]  

53. N. Radwell, T. W. Clark, B. Piccirillo, S. M. Barnett, and S. Franke-Arnold, “Spatially dependent electromagnetically induced transparency,” Phys. Rev. Lett. 114(12), 123603 (2015). [CrossRef]  

54. S. Sharma and T. N. Dey, “Phase-induced transparency-mediated structured-beam generation in a closed-loop tripod configuration,” Phys. Rev. A 96(3), 033811 (2017). [CrossRef]  

55. H. R. Hamedi, V. Kudriašov, J. Ruseckas, and G. Juzel unas, “Azimuthal modulation of electromagnetically induced transparency using structured light,” Opt. Express 26(22), 28249–28262 (2018). [CrossRef]  

56. L. D. Landau and E. M. Lifshitz, Electrodynamics of continuous media (Butterworth-Heinemann, 2004).

57. M. Born and E. Wolf, Principles of optics (Cambridge University Press, 1980), 6th ed.

58. M. Artoni, G. La Rocca, and F. Bassani, “Resonantly absorbing one-dimensional photonic crystals,” Phys. Rev. E 72(4), 046604 (2005). [CrossRef]  

59. Y. Zhang, Y. Xue, G. Wang, C.-L. Cui, R. Wang, and J.-H. Wu, “Steady optical spectra and light propagation dynamics in cold atomic samples with homogeneous or inhomogeneous densities,” Opt. Express 19(3), 2111–2119 (2011). [CrossRef]  

60. R. Folman, P. Kruger, D. Cassettari, B. Hessmo, T. Maier, and J. Schmiedmayer, “Controlling cold atoms using nanofabricated surfaces: atom chips,” Phys. Rev. Lett. 84(20), 4749–4752 (2000). [CrossRef]  

61. M. F. Riedel, P. Bohi, Y. Li, T. W. Hansch, A. Sinatra, and P. Treutlein, “Atom-chip-based generation of entanglement for quantum metrology,” Nature 464(7292), 1170–1173 (2010). [CrossRef]  

62. J. D. Carter and J. D. D. Martin, “Coherent manipulation of cold Rydberg atoms near the surface of an atom chip,” Phys. Rev. A 88(4), 043429 (2013). [CrossRef]  

63. H. Hattermann, D. Bothner, L. Y. Ley, B. Ferdinand, D. Wiedmaier, L. Sarkany, R. Kleiner, D. Koelle, and J. Fortagh, “Coupling ultracold atoms to a superconducting coplanar waveguide resonator,” Nat. Commun. 8(1), 2254 (2017). [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 (6)

Fig. 1.
Fig. 1. (a) A cold atomic sample illuminated by a control beam $\mathbf {E}_{c}(x)$ along the $-y$ direction exhibits a strongly asymmetric reflection for a probe beam $\mathbf {E}_{p}$ incident either along the $+x$ or the $-x$ direction. (b) A three-level atomic system driven by a probe field of Rabi frequency $\Omega _{p}$ (detuning $\Delta _{p}$ ) and a control field of Rabi frequency $\Omega _c(x)$ (detuning $\Delta _c$ ) into the $V$ configuration. (c) A two-level atomic system with a dynamic shift $\delta (x)$ of level $\vert g\rangle$ upon the adiabatic elimination of level $\vert a\rangle$ .
Fig. 2.
Fig. 2. (a) Real and (b) imaginary parts of susceptibility $\chi _{2}$ against position $x$ with $\delta _{0}/\gamma _{ge}$ =100 and $\Delta _p/\gamma _{ge}=-50$ (red-solid); $\delta _{0}/\gamma _{ge}$ =20 and $\Delta _p/\gamma _{ge}=-10$ (blue-dashed). (c) Real and (d) imaginary parts of susceptibility $\chi _{2}$ against position $x$ and detuning $\Delta _p$ with $\delta _{0}/\gamma _{ge}$ =100. Other parameters used in calculations are $\gamma _{ge}=2 \pi \times 2.87$ MHz, $d_{ge}=1.79\times 10^{-29}$ C $\cdot$ m, $N_{0}=2.0 \times 10^{13}$ cm $^{-3}$ , $L=5.0$ $\mu$ m, and $\lambda _{p}=795$ nm.
Fig. 3.
Fig. 3. (a) Reflectivity $R_{l}$ , reflectivity $R_{r}$ , and half transmissivity $T$ ; (b) figure of merit $D_{kk}$ against detuning $\Delta _{p}$ . Other parameters used in calculations are the same as in Fig. 2(a,b).
Fig. 4.
Fig. 4. (a) Figure of merit $D_{kk}$ , (b) reflectivity $R_{l}$ , and (c) reflectivity $R_{r}$ (c) against shift $\delta _{0}$ and detuning $\Delta _{p}$ . Other parameters used in calculations are the same as in Figs. 2(a) and (b).
Fig. 5.
Fig. 5. Reflectivities (a) $R_{l}$ and (b) $R_{r}$ against detuning $\Delta _{p}$ for $L=10$ $\mu$ m (black-squares), $L=15$ $\mu$ m (red-circles), and $L=20$ $\mu$ m (blue-triangles). Other parameters used in calculations are the same as in Figs. 2(a) and (b). Black (red) curves are shown with a vertical offset $0.1$ ( $0.05$ ) in both insets.
Fig. 6.
Fig. 6. Spatial profiles of (a) $\chi _{2}^{\prime }$ and (b) $\chi _{2}^{\prime \prime }$ as well as corresponding spectra of (c) reflectivities $R_{l,r}$ and (d) contrast $C$ for an atomic lattice of density $N_{j}(x)=N_{0}e^{-(x-x_{j})^{2}/\delta x^{2}}$ in the $jth$ trap of center $x_{j}=(j-1/2)\Lambda$ , width $\delta x$ , and period $\Lambda$ . Parameters are the same as in Fig. 2(b) except $N_{0}=2.0 \times 10^{12}$ cm $^{-3}$ , $L=60$ $\mu$ m, $\delta _{0}/\gamma _{ge}=30$ , $\Lambda =400$ nm, and $\delta x=\Lambda /6$ .

Equations (6)

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

χ 3 ( Δ p , x ) = i N 0 | d g e | 2 ε 0 [ ( γ g a 2 + Δ c 2 ) + | Ω c ( x ) | 2 ] [ γ e a i ( Δ c Δ p ) ] | Ω c ( x ) | 2 ( γ g a i Δ c ) [ ( γ g a 2 + Δ c 2 ) + | Ω c ( x ) | 2 ] { [ γ e a i ( Δ c Δ p ) ] ( γ g e i Δ p ) + | Ω c ( x ) | 2 } ,
χ 2 ( Δ p , x ) = i N 0 | d g e | 2 ε 0 1 γ g e i [ Δ p + δ ( x ) ] ,
χ 2 ( Δ p , x ) = 1 π P 0 L χ 2 ( Δ p , s ) s x d s
[ E p + ( Δ p , j l ) E p ( Δ p , j l ) ] = M j ( Δ p , l ) [ E p + ( Δ p , j l l ) E p ( Δ p , j l l ) ] ,
R l ( Δ p , L ) = | r l ( Δ p , L ) | 2 = | M ( 21 ) ( Δ p , L ) M ( 22 ) ( Δ p , L ) | 2 , R r ( Δ p , L ) = | r r ( Δ p , L ) | 2 = | M ( 12 ) ( Δ p , L ) M ( 22 ) ( Δ p , L ) | 2 , T ( Δ p , L ) = | t ( Δ p , L ) | 2 = | 1 M ( 22 ) ( Δ p , L ) | 2
D k k ( Δ p ) = 0 L { χ 2 ( Δ p , x ) 1 π P 0 L χ 2 ( Δ p , s ) s x d s } d x | 0 L χ 2 ( Δ p , x ) d x |
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.