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

Spin Hall effect of Laguerre-Gaussian beams in PT symmetric metamaterials

Open Access Open Access

Abstract

Spin Hall effect (SHE) of Laguerre-Gaussian (LG) beams reflected and transmitted in parity-time (PT) symmetric metamaterials are investigated near the coherent-perfect-absorption (CPA)-laser point and exceptional points (EPs). The numerical results show that large transverse shifts occur at the CPA-laser point regardless of the incident direction. But at EPs, the SHE increases at one side and disappears at the other side, thus achieving the intense SHE of the reflected light beams at the specified side incidence. In addition, it is found that Bragg oscillation can be generated by increasing the period number of PT symmetric metamaterial layers, thus increasing the number of formants in transverse displacement. In particular, the transverse shift peaks of the transmitted beams merge into a positive peak when the incident angle is close to $90^\circ$ and does not change basically with the increasing of Im(ɛ), which can also be considered as a strong tolerance to the variation of Im(ɛ). This feature is expected to realize a new type of optoelectronic devices with anti-interference performance. These results provide a feasible path for the modulation of spin Hall effect of light (SHEL) and provide the possibility for the development of new nanophotonic devices.

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

1. Introduction

When a linearly polarized beam with finite width propagates in an inhomogeneous medium, the gradient change of refractive index makes left-circularly polarized (LCP) and right-circularly polarized (RCP) components separate equally with opposite directions [1,2]. This phenomenon is called spin Hall effect of light (SHEL), since it is similar to the spin Hall effect in an electronic system. It is also called the Imbert-Fedorov (IF) effect as it was initially predicted by Fedorov and then verified by Imbert in the experiment [3].

Onoda et al. predicted theoretically the spatial separation of the two optical spins in 2004, and Bliokh et al. proposed an accurate expression for the shift of SHEL in 2006 [4]. However, the displacement is usually on the subwavelength scale. It was not until 2008 that Hosten et al. observed this weak effect [5] for the first time by using quantum weak measurement technology, which proved the correctness of previous studies. Subsequently, some better ways to manipulate and enhance SHEL have been explored due to its promising applications in precision metrology. Such as in references [68], it was demonstrated that the transverse shift can be enhanced by launching a Gaussian beam near the Brewster angle. The excitation of the surface plasmon resonance (SPR) can significantly control the spin splitting of the reflected beams [9]. Taking advantage of $\varepsilon$-near-zero materials, the large spin dependent splitting of transmitted beams can be obtained [10], and reference [11] showed that loss enhances spin dependent displacement up to 24.676 micrometers.

The concept of parity-time (PT) symmetry [12] originates from quantum mechanics, it requires that the operator of every observable quantity must be a Hermitian operator in quantum mechanics, but the Hamiltonian under PT symmetry is allowed to have real eigenvalues even if it is non-Hermitian [13]. In order to satisfy PT symmetry, the optical system needs to satisfy $n(x)=n^*(-x)$, it means that the real part of the refractive index of the metamaterial is even symmetric in space, while the imaginary part is odd symmetric. In the field of optics, PT symmetry has been applied in coherent perfect absorbers, unidirectional Bloch oscillations, optical switches, asymmetric transmission and topological optics, which has injected new vitality into the research of traditional optical devices [1416].

In this paper, we systematically study the asymmetric SHEL of the first-order LG beams at the interface of PT symmetric metamaterials. This asymmetry is manifested in two aspects, one is the asymmetric splitting of LCP and RCP components caused by the extra displacement which is result from the intrinsic orbital angular momentum (OAM) carried by the vortex beams. The second asymmetry is the transverse shifts of the reflected beams generated by the incidence from the gain and loss side of the PT symmetric metamaterials, it stems from the dependence of EPs on direction. It can be greatly enhanced at the CPA-laser point whether reflected or transmitted. In addition, it is found that Bragg oscillation can be generated by increasing the period number of PT symmetric metamaterial layers, thus increasing the number of transverse displacement peaks. This means that our results provide a feasible path for the modulation of SHEL.

2. Theory and models

We consider the transmission and reflection of a first-order LG beam at the interface of an air-PT symmetric metamaterial with the incident angle $\theta _i$, and the metamaterial is placed in the plane $z=0$, as shown in Fig. 1. Its background medium is air, the relative dielectric constant in air, gain and loss medium are denoted as $\varepsilon _{0}$, $\varepsilon _{1}$ and $\varepsilon _{2}$, respectively. Since we consider nonmagnetic medium, the relative permeability of each medium layer meets $\mu _{r}=1$. The thicknesses of both loss layer and gain layer are $d$.

 figure: Fig. 1.

Fig. 1. Schematic of SHEL in the periodic PT-symmetric metamaterial.

Download Full Size | PDF

When the beam propagates along the $z$ axis of the structure, taking TM wave as an example, based on the transfer matrix method (TMM), we can obtain [17]

$$\begin{pmatrix}H^{+}_{f}\\ H^{+}_{b} \end{pmatrix}=Q\begin{pmatrix} H^{-}_{f}\\ H^{-}_{b} \end{pmatrix}=\begin{pmatrix} 1 & 1 \\ \eta_{0} & -\eta_{0} \end{pmatrix}^{{-}1}M\begin{pmatrix} 1 & 1 \\ \eta_{0} & -\eta_{0} \end{pmatrix}\begin{pmatrix} H^{-}_{f}\\ H^{-}_{b} \end{pmatrix} ,$$
where $M=(m_{1}m_{2})^{N}$ is the total eigenmatrix of the structure in which plane waves pass through $N$ periods, and $m_{j}$ is the eigenmatrix at layer $j$ which can be expressed as
$$m_{j}=\begin{pmatrix} \cos\delta_{j} & -\frac{i\sin\delta_{j}}{\eta_{j}} \\ -i\eta_{j}\sin\delta_{j} & \cos\delta_{j} \end{pmatrix} ,$$
where $\delta _{j}=k_{zj}d$ denotes the phase retardance, $k_{zj}=k_0\sqrt {n_{j}^{2}-n_0^{2}{\sin }^{2}\theta _{i}}$ represents the $z$-direction component of the wavevector ${\vec {\boldsymbol {k}_{j}}}$ ($j=1,2$), $k_{0}=\frac {2\pi }{\lambda }$ is the wavenumber of vacuum. The refractive index of the $j$th layer medium is devoted as $n_{j}=\sqrt {\varepsilon _{j}\mu _{j}}$. The reflection coefficient and transmission coefficient, in the case of incidence from the gain side only, can be expressed as $r_{G}=H_{b}^{-}/H_{f}^{-}$, $t_{G}=H_{f}^{+}/H_{f}^{-}$. Inversely, it can be expressed as $r_{L}=H_{f}^{+}/H_{b}^{+}$, $t_{L}=H_{b}^{-}/H_{b}^{+}$ when it inputs from the loss side [18]. Thus, the reflection and transmission coefficients can be expressed as
$$t_{G}=t_{L}=\frac{1}{Q_{22}}=\frac{2}{M_{11}-\frac{1}{\eta_{0}}M_{21}-\eta_{0}M_{12}+M_{22}} ,$$
$$r_{G}={-}\frac{Q_{21}}{Q_{22}}=\frac{-M_{11}+\frac{1}{\eta_{0}}M_{21}-\eta_{0}M_{12}+M_{22}}{M_{11}-{\frac{1}{\eta_{0}}M_{21}}-\eta_{0}M_{12}+M_{22}} ,$$
$$r_{L}=\frac{Q_{12}}{Q_{22}}=\frac{M_{11}+\frac{1}{\eta_{0}}M_{21}-\eta_{0}M_{12}-M_{22}}{M_{11}-{\frac{1}{\eta_0}M_{21}}-\eta_{0}M_{12}+M_{22}} ,$$
similarly, the reflection and transmission coefficients of TE wave incidence can also be obtained. When the TM wave inputs, $\eta _{j}=\frac {k_{jz}}{\omega \varepsilon _{j}}$, and when the TE wave inputs, $\eta _{j}=-\frac {k_{jz}}{\omega \mu _{j}}$ ($j=0,1,2$), where, $\omega$ denotes the angular frequency of incident wave.

With LG beam incidence, the angular spectrum of p(H) polarization and s(V) polarization can be expressed as $\tilde {\boldsymbol {E}}_{i}^{H}=\tilde {\phi }_{l}\hat {\boldsymbol {e}}_{ix}$ and $\tilde {\boldsymbol {E}}_{i}^{V}=\tilde {\phi }_{l}\hat {\boldsymbol {e}}_{iy}$, where, $\hat {\boldsymbol {e}}_{ix}$, $\hat {\boldsymbol {e}}_{iy}$ are the basis vectors parallel and perpendicular to the incident plane. The scalar spectrum of LG beam is $\tilde {\phi }_{l}\propto [w_{0}(-ik_{x}+s_lk_{y})/\sqrt {2}]^{|l|}\textrm {exp}[-(k_{x}^{2}+k_{y}^{2})w_{0}^{2}/4]$, with $w_0$ being the beam waist, and $k_{x}$, $k_{y}$ being the wavevector components, $l$ and $s_{l}=\textrm {sign}(l)$ denote the OAM and the sign of $l$, respectively. The rectangular coordinate systems $(x,y,z)$, $(x_{i},y_{i},z_{i})$, and $(x_{r},y_{r},z_{r})$ are the laboratory coordinate systems, the coordinate systems of incidence and reflection, respectively. The angular spectrum of the reflected beams can be obtained through the transformation matrix between the incident and reflected angular spectrum [19]

$$\begin{bmatrix} \tilde{\boldsymbol{E}}^{H}_{r}\\ \tilde{\boldsymbol{E}}^{V}_{r} \end{bmatrix}=M_{R}\begin{bmatrix} \tilde{\boldsymbol{E}}^{H}_{i}\\ \tilde{\boldsymbol{E}}^{V}_{i} \end{bmatrix}=\begin{bmatrix} r_{p} & {\frac{k_{ry}}{k_{0}}}\xi \\ -{\frac{k_{ry}}{k_{0}}}\xi & r_{s} \end{bmatrix}\begin{bmatrix} \tilde{\boldsymbol{E}}^{H}_{i}\\ \tilde{\boldsymbol{E}}^{V}_{i} \end{bmatrix} ,$$
with $\xi =(r_{p}+r_{s})\cot \theta _{i}$, $r_{p}(r_{s})$ denotes the reflection coefficient of p(s) polarization. According to the angular spectrum theory, one can obtain slightly different complex refraction or reflection coefficients since the plane wave in vortex beam propagates in slightly different directions. Therefore, under the premise of paraxial approximation, the Fresnel reflection coefficients $r_{p}$ and $r_{s}$ can be Taylor expanded at $k_{ix}=0$ and the first-order approximation can be taken as [20]
$$r_{p,s}(k_{ix}/k_{0})=r_{p,s}(0)+\frac{k_{ix}}{k_0}r'_{p,s}(0) ,$$
with $r'_{p,s}=\frac {\partial r_{p,s}}{\partial \theta _{i}}$. In the circular polarization basis, $\hat {\boldsymbol {e}}_{r\pm }=[\hat {\boldsymbol {e}}_{rx}\pm i\hat {\boldsymbol {e}}_{ry}]/2^{1/2}$, here, $\pm$ represent the LCP and RCP components, respectively. By taking the Fourier transformation of the angular spectrum, the reflected fields with p polarized inputting in the plane at $z_{r}=0$ can be obtained [21]
$$\begin{aligned}\boldsymbol{E}_{r\pm}^{H} &=\sqrt{\frac{2}{\pi w_{0}^{2}|l|!}}\textrm{exp}\left [-\frac{(x^{2}+y^{2})}{w_{0}}\right ] \\ & \left\{\left[r_{p}+\frac{ixr'_{p}\pm s_{l}y\xi}{k_{0}w_{0}^{2}/2}\right]\left[\frac{(x+is_{l}y)}{w_{0}}\right]^{|l|}\right . \left . +\,i|l|\frac{r'_{p}\pm s_{l}\xi}{k_{0}w_{0}}\left[\frac{(x+is_{l}y)}{w_{0}}\right]^{|l-1|}\right\} , \end{aligned}$$
The reflected light fields of LCP and RCP with s polarized inputting can be obtained by replacing $r_{p,s}$ with $r_{s,p}$. According to the definition of SHEL, the transverse shifts of the reflected light can be given as [19]
$$\delta_{{\pm}}^{H,V}=\frac{\iint y|\boldsymbol{E}_{{\pm}}^{H,V}|^{2}dxdy}{\iint |\boldsymbol{E}_{{\pm}}^{H,V}|^{2}dxdy} ,$$
After some straightforward calculations, we arrive
$${\delta_{r \pm}^{H,V}}=\frac{1}{k_{0}}\frac{\pm \textrm{Re}(r_{p,s}^{*}\xi)-l\textrm{Re}(r_{p,s}^{*}r'_{p,s})}{|r_{p,s}|^{2}+[|l|+1][|r'_{p,s}|^{2}+|\xi|^{2}]/k_{0}^2w_{0}^{2}} ,$$
Similarly, the spatial transverse shifts through the PT symmetric system can be obtain as
$${\delta_{t \pm}^{H,V}}=\frac{1}{k_{0}}\frac{\pm \textrm{Re}(t_{p,s}^{*}\zeta)-l\textrm{Re}(t_{p,s}^{*}t'_{p,s})}{|t_{p,s}|^2+[|l|+1][|t'_{p,s}|^{2}+|\zeta|^{2}]/k_{0}^{2}w_{0}^{2}} ,$$
with $\zeta =(t_{p}-t_{s})\textrm {cot}\theta _{i}$. As mentioned above, spin-orbit interaction is the intrinsic physical cause of SHEL. For vortex beams, since it also carries inherent OAM, the beams inevitably undergo internal and external OAM conversion. In this case, the orbit-orbit interaction is spin-independent, resulting in asymmetric splitting [22]. The first terms in Eqs. (10) and (11) are spin-dependent, which are derived from the Gaussian envelope and eventually lead to spin-splitting of equal magnitude and opposite directions of LCP and RCP, the second terms are the OAM($l$) related terms caused by the inherent angular momentum carried by the vortex beam, which causes the joint movement of LCP and RCP, resulting in asymmetric splitting.

The peculiarity of PT symmetric system is that, as long as the system parameters are properly adjusted, the system can undergo phase transformation, and EPs can obviously affect the dynamic characteristics of the system. In order to better reveal the underlying phenomena of PT symmetry, the scattering matrix is used as $S=\begin {pmatrix} t & r_{L} \\ r_{G} & t \end {pmatrix}$, with $S_{\pm }=t\pm \sqrt {r_{L}r_{G}}$ being the eigenvalues. The eigenvalues of the S-matrix are either reciprocal moduli pair ($|S_{+}|=1/|S_{-}|\neq 1$) or unimodular pair ($|S_{+}|=1/|S_{-}|=1$). For the EPs, they are exactly at the transition points between reciprocal moduli pair and unimodular pair [23,24].

3. Results and discussion

First, we chose the PT symmetric metamaterial with a single period ($N=1$) to study the SHE of first-order LG beams, which is shown in Fig. 2. In order to facilitate research, the scattering matrix eigenvalue, reflection and transmission coefficients are the logarithm of the module, and the displacements are multiples of wavelength, these will not be repeated in the following text. The PT symmetric system is optimized as $\varepsilon _{1}=0.1-0.1i, \varepsilon _{2}=0.1+0.1i$, $d=0.785\lambda$, and $w_{0}=15\lambda$. We find that the transverse shifts are zero at the EPs (the transition point between non-zero and zero values), but can be dramatically enhanced in their vicinity. The underlying physics is associated with the near-zero value and abrupt phase jump of the reflection coefficients at EPs. Because of the spontaneous PT symmetry breaking across the EPs, the giant transverse shifts have opposite signs at the two sides of EPs. Therefore, it is reasonable to argue that the dependence of sign change of the giant transverse shift is related to the phase transition in PT symmetric systems. Furthermore, numerical results show that near the EPs, the reflection spectrum increases only at one side, and at the other side disappears. Since $\delta _{r\pm }$ with p(s) polarization inputting is closely related $|r_{s}|/|r_{p}|$($|r_{p}|/|r_{s}|$) [25], a significant value of $|r_{s}|/|r_{p}|$($|r_{p}|/|r_{s}|$) near EPs would directly leads to the appearance of the giant SHEL for the p(s)-polarized incident beam. Thus, we can conclude that the sign switch of transverse shift accross the EPs is dependent on the propagation direction of the incident beam. This is the reason why EPs are often used to implement unidirectional reflectionlessness [16,26]. While at the CPA-laser point, large transverse shifts can be obtained regardless of the incident direction. In addition, although transverse shift of the LG beam is caused by both the spin-orbit angular momentum interaction and the orbital-orbit interaction, however, the transverse displacement generated by the intrinsic OAM is dominant. As a result, LCP and RCP will move in the same direction, and even almost coincide when it strikes at certain angles. Therefore, when the LG beam propagates in the PT symmetric metamaterial, we can not only obtain the asymmetric displacements of LCP and RCP, but also achieve EPs and CPA-laser points by adjusting the incident angle, so as to realize the intense SHE of the reflected beam when it inputs from the specified side.

 figure: Fig. 2.

Fig. 2. Dependences of the eigenvalue of the scattering matrix (the first row), the reflectivity (the second row), and the transverse displacement of reflected light (the third row) on $\theta$ for the cases of H polarization (the first column), and V polarization (the second column), respectively. The red (blue) line represents incident from the gain (loss) layer, and the solid (dashed) line represents reflectivity and the transverse displacement of LCP (RCP), respectively. In our calculations, $N=1$, $\varepsilon _{1}=0.1-0.1i, \varepsilon _{2}=0.1+0.1i$, $\mu _{r}=1$, $d=0.785\lambda$, and $w_{0}=15\lambda$.

Download Full Size | PDF

Next, we investigate the transverse displacement of transmitted light in the PT symmetric metamaterial with a single period ($N=1$), as shown in Fig. 3. Unlike the reflection case, the transverse shifts are the same regardless of the incidence from the loss or gain layer, since the transmission coefficients are the same for both sides. Similar to the reflected beam, The transverse shift of the transmitted beam is greatly enhanced at the CPA-laser point, the sign change of the enhanced transverse displacement of transmitted light across the laser mode is observed, which is caused by the phase change of $\pi$ at the laser mode [25]. But the difference is that it is not obvious near the EPs. Moreover, surprisingly, although the corresponding transmission coefficient is almost zero when the incident angle approaches $90^\circ$, the transverse displacement still reaches almost 6$\lambda$. This is a large lateral displacement caused by the intrinsic OAM carried by the vortex beam.

 figure: Fig. 3.

Fig. 3. Dependences of the transmission coefficient (the first row), and the transverse displacement of transmitted light (the second row) on $\theta$ for the cases of H polarization (the first column), and V polarization (the second column), respectively. The solid (dashed) line represents the transverse displacement of LCP (RCP), respectively. Other parameters are the same as Fig. 2.

Download Full Size | PDF

The period number ($N$) of the metamaterial is an important parameter affecting the SHEL. In order to further study the influence of $N$ on the transverse displacement, we consider a periodic PT symmetric structure with $N=10$, and other parameters are the same as those in Fig. 2. The obtained results are shown in Fig. 4. With the increase of $N$ , we obtain more positive and negative transverse shift peaks than in the case of N=1, since the refractive index modulation along the propagation direction constitutes a PT Bragg grating structure. In this periodic PT symmetric structure, the incident light produces periodic transmission and reflection at the interface between the gain and loss media, the constant superpositions of these transmitted and reflected waves constitute the transmitted and reflected spectra of the Bragg grating, which lead to the oscillations of the positive and negative peaks of the transmitted and reflected transverse shifts. However, there is still only one EP in the case of s polarization, this is because the excitation of the surface mode at the interface between the gain and loss layers has a great effect on the p polarization, but has no effect on the s polarization. Even so, the transverse shifts on both sides still show more positive and negative peaks, this is because Bragg resonance causes positive and negative oscillations of reflection and transmission coefficients in the periodic medium. Therefore, we conclude that with the increase of $N$ we can obtain more severe Bragg oscillations, thus increase the number of formants in the transverse displacement.

 figure: Fig. 4.

Fig. 4. Dependences of the eigenvalue of the scattering matrix (first row), the reflectivity (second row), and the transverse displacement of reflected light (third row) on $\theta$ for the cases of H polarization (the first column), and V polarization (second column), respectively. The red (blue) line represents incident from the gain (loss) layer, and the solid (dashed) line represents reflectivity and displacement of LCP (RCP), respectively. Where, $N=10$, $\varepsilon _{1}=0.1-0.1i, \varepsilon _{2}=0.1+0.1i$, $\mu _{r}=1$, $d=0.785\lambda$, and $w_{0}=15\lambda$.

Download Full Size | PDF

Now let’s adjust the system parameters appropriately, change Im($\varepsilon$) and observe its effects on transverse displacement. For the sake of research, we first explore the changing trend of the material’s scattering properties with respect to the incident angle when the Im($\varepsilon$) changes. As shown in Fig. 5, for both p and s polarization incidence, an EP (the transition point between non-zero and zero values) appears in the region with $\theta _{i} < 10^{\circ }$ and Im$(\varepsilon ) < 0.20$. In addition, a CPA-laser point and two other EPs appear in the region with $\theta _{i} > 10^{\circ }$ and move towards a larger angle with the increase of Im($\varepsilon$) in the p polarized incidence.

 figure: Fig. 5.

Fig. 5. The eigenvalue of the scattering matrix contour for different Im($\varepsilon$) with H (a) and V (b) inputs, respectively. Here we choose $N=1$, $\Re (\varepsilon _{1})=\Re (\varepsilon _{2})=0.1$, $\mu _{r}=1$, $d=0.785\lambda$, and $w_{0}=15\lambda$.

Download Full Size | PDF

We have found that the incident angle corresponding to the CPA-laser point changes with the Im($\varepsilon$), taking LCP as an example, the peak of the transverse displacement also moves correspondingly. In the case of p polarization incidence, as shown in Fig. 6, the peak value of reflected light displacement appears near the CPA-laser point and EPs. At the segment with $\theta _{i}< 10^{\circ }$ and Im($\varepsilon$) $< 0.2$, it is worth noting that the displacement peak only occurs at the incidence from the loss side, this is because the propagation of light in the waveguide depends on the incident direction when the PT symmetry condition is introduced. Before the spontaneous PT symmetry breaking, the light intensity propagates periodically between the two waveguides in an oscillating manner, but when the gain and loss coefficients reach the spontaneous PT symmetry breaking, the light intensity always propagates locally in the gain waveguide, whether it inputs from the gain waveguide or the loss waveguide. At the same time, the energy in the waveguide increases exponentially and finally concentrates only in the gain waveguide. Jiang et al. have shown in [19] that there is a negative correlation between SHEL and energy, as a result, the huge difference in energy between the two sides results in a larger transverse shift on the loss side. However, when $\theta _{i} > 10^{\circ }$, the peak value of the displacement moves towards a larger angle as Im($\varepsilon$) increases, and roughly coincides with the CPA-laser point movement trend. This can more clearly proves that large transverse shifts occur at the CPA-laser point regardless of the incident direction. The same phenomenon occurs for s polarization incidence, and we can explain it by the same physical mechanism. Thus it can be concluded that SHEL can be enhanced by adjusting the parameters of Im($\varepsilon$) and incident angle simultaneously.

 figure: Fig. 6.

Fig. 6. Transverse shift contour (integer multiples of wavelength) of LCP reflected light for different Im($\varepsilon$) with H (a) and (c) and V (b) and (d) inputs, respectively. Here, the first (second) row is incident from the loss(gain) side. Other parameters are the same as Fig. 5.

Download Full Size | PDF

Finally, we study the transverse displacement of transmitted light in a PT symmetric metamaterial with a single period ($N=1$). It can be seen from Fig. 7 that the peak of the lateral displacement will shift to a larger angle with the increase of Im($\varepsilon$) for both p and s polarization incidence, and there will be a sudden change from negative to positive (or from positive to negative) peak. In particular, as the incident angle continues to increase, we can observe that the transverse shift of the transmitted beam coalesces into a positive peak near $90^{\circ }$, and with the increase of Im($\varepsilon$), the peak does not change basically. This huge transverse shift is caused by the extra OAM carried by LG beams. It can also be considered as a strong tolerance to the variation of Im($\varepsilon$), which is expected to enable new optoelectronic devices with anti-interference performance.

 figure: Fig. 7.

Fig. 7. Transverse shift contour (integer multiples of wavelength) of LCP transmitted light for different Im($\varepsilon$) with H (a) and V (b) inputs, respectively. Other parameters are the same as Fig. 5.

Download Full Size | PDF

4. Conclusion

In summary, we have shown the asymmetric SHE of the first-order LG beams at the interface of PT symmetric metamaterials. By adjusting the parameters of the structure and the incident angle of the light, we can not only obtain the asymmetric displacement of LCP and RCP, but also realize the intense SHE of the reflected beam when it incident on the specified side. In addition, with the increase of the number of periodic layers, due to the Bragg resonance, the transmission and reflection coefficients achieve violent oscillations. Therefore, the transverse shift peaks are increased and enhanced effectively. Finally, we have found that the peak of the transverse shift can always appear near the CPA-laser point, even though it will move to a larger angle as Im($\varepsilon$) increases. In particular, we can observe that the transverse shift peak of the transmitted beam coalesces into a positive peak near $90^\circ$, this is caused by the extra OAM carried by the LG beams and is largely unaffected by the variation of Im($\varepsilon$). Our results provide a feasible path for the modulation of SHEL and provide the possibility for the development of new nanophotonic devices in some relevant fields.

Funding

Guangzhou Municipal Science and Technology Project (201707010403); Program for University Key Laboratory of Guangdong Province (2017B030311003); Natural Science Foundation of Guangdong Province (2016A030313439, 2018A030313480).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. K. Y. Bliokh and A. Aiello, “Goos-Hänchen and Imbert-Fedorov beam shifts: an overview,” J. Opt. 15(1), 014001 (2013). [CrossRef]  

2. K. Y. Bliokh and Y. P. Bliokh, “Conservation of angular momentum, transverse shift, and spin Hall effect in reflection and refraction of an electromagnetic wave packet,” Phys. Rev. Lett. 96(7), 073903 (2006). [CrossRef]  

3. S. M. Ornigotti and A. Szameit, “Goos-Hänchen and Imbert-Fedorov beam shifts for Gaussian beams impinging on graphene-coated surfaces,” Opt. Express 23(23), 30195–30203 (2015). [CrossRef]  

4. M. Onoda, S. Murakami, and N. Nagaosa, “Hall effect of light,” Phys. Rev. Lett. 93(8), 083901 (2004). [CrossRef]  

5. O. Hosten and P. Kwiat, “Observation of the spin hall effect of light via weak measurements,” Science 319(5864), 787–790 (2008). [CrossRef]  

6. H. Luo, X. Zhou, W. Shu, S. Wen, and D. Fan, “Enhanced and switchable spin Hall effect of light near the Brewster angle on reflection,” Phys. Rev. A 84(4), 043806 (2011). [CrossRef]  

7. X. Zhou, H. Luo, and S. Wen, “Weak measurements of a large spin angular splitting of light beam on reflection at the Brewster angle,” Opt. Express 20(14), 16003 (2012). [CrossRef]  

8. L. Xie, X. Zhou, X. Qiu, L. Luo, X. Liu, Z Li, Y. He, J. Du, Z. Zhang, and D. Wang, “Unveiling the spin Hall effect of light in Imbert-Fedorov shift at the Brewster angle with weak measurements,” Opt. Express 26(18), 22934 (2018). [CrossRef]  

9. Y. Xiang, X. Jiang, Q. You, J. Guo, and X. Dai, “Enhanced spin Hall effect of reflected light with guided-wave surface plasmon resonance,” Photonics Res. 5(5), 467 (2017). [CrossRef]  

10. W. Zhu and W. She, “Enhanced spin Hall effect of transmitted light through a thin epsilon-near-zero slab,” Opt. Lett. 40(13), 2961 (2015). [CrossRef]  

11. T. Tang, L. J L. Luo, P. Sun, and Y. Zhang, “Loss enhanced spin Hall effect of transmitted light through anisotropic epsilon-and mu-zero metamaterial slab,” Opt. Express 25(3), 2347 (2017). [CrossRef]  

12. Y. D. Chong, L. Ge, and A. D. Stone, “PT-symmetry breaking and laser-absorber modes in optical scattering systems,” Phys. Rev. Lett. 106(9), 093902 (2011). [CrossRef]  

13. L. Feng, R. El-Ganainy, and L. Ge, “Non-Hermitian photonics based on parity-time symmetry,” Nat. Photonics 11(12), 752–762 (2017). [CrossRef]  

14. S. Longhi, “PT-symmetric laser absorber,” Phys. Rev. A 82(3), 031801 (2010). [CrossRef]  

15. M. Sakhdari, M. Hajizadegan, Q. Zhong, D. N. Christodoulides, R. El-Ganainy, and P. -Y. Chen, “Experimental Observation of PT Symmetry Breaking near Divergent Exceptional Points,” Phys. Rev. Lett. 123(19), 193901 (2019). [CrossRef]  

16. Z. Lin, H. Ramezani, T. Eichelkraut, T. Kottos, H. Cao, and D. N. Christodoulides, “Unidirectional Invisibility Induced by PT-Symmetric Periodic Structres,” Phys. Rev. Lett. 106(21), 213901 (2011). [CrossRef]  

17. P. Ma and L. Gao, “Large and tunable lateral shifts in one-dimensional PT-symmetric layered structures,” Opt. Express 25(9), 9676 (2017). [CrossRef]  

18. A. Mostafazadeh, “Spectral Singularities of Complex Scattering Potentials and Infinite Reflection and Transmission Coefficients at Real Energies,” Phys. Rev. Lett. 102(22), 220402 (2009). [CrossRef]  

19. H. Luo, X. Ling, X. Zhou, W. Shu, S. Wen, and D. Fan, “Enhancing or suppressing the spin Hall effect of light in layered nanostructures,” Phys. Rev. A 84(3), 033801 (2011). [CrossRef]  

20. M. Jiang, H. Lin, L. Zhuo, W. Zhu, H. Guan, J. Yu, H. Lu, J. Tan, and Z. Chen, “Chirality induced asymmetric spin splitting of light beams reflected from an air-chiral interface,” Opt. Express 26(6), 6593 (2018). [CrossRef]  

21. W. Zhu, L. Zhuo, M. Jiang, H. Guan, J. Yu, H. Lu, Y. Luo, J. Zhang, and Z. Chen, “Controllable symmetric and asymmetric spin splitting of Laguerre-Gaussian beams assisted by surface plasmon resonance,” Opt. Lett. 42(23), 4869 (2017). [CrossRef]  

22. M. Jiang, W. Zhu, H. Guan, J. Yu, H. Lu, J. Tan, J. Zhang, and Z. Chen, “Giant spin splitting induced by orbital angular momentum in an epsilon-near-zero metamaterial slab,” Opt. Lett. 42(17), 3259 (2017). [CrossRef]  

23. Y. Cao, Y. Fu, Q. Zhou, Y. Xu, L. Gao, and H. Chen, “Giant Goos-Hänchen shift induced by bounded states in optical PT-symmetric bilayer structures,” Opt. Express 27(6), 7857 (2019). [CrossRef]  

24. O. V. Shramkova and G. P. Tsironis, “Scattering properties of PT-symmertric layered periodic structrures,” J. Opt. 18(10), 105101 (2016). [CrossRef]  

25. Y. Fu, Y. Fei, D. Dong, and Y. Liu, “Photonic spin Hall effect in PT symmetric metamaterials,” Front. Phys. 14(6), 62601 (2019). [CrossRef]  

26. X. Zhou, X. Lin, Z. Xiao, T. Low, A. Alù, B. Zhang, and A. H. Sun, “Controlling photonic spin Hall effect via exceptional points,” Phys. Rev. B 100(11), 115429 (2019). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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. Schematic of SHEL in the periodic PT-symmetric metamaterial.
Fig. 2.
Fig. 2. Dependences of the eigenvalue of the scattering matrix (the first row), the reflectivity (the second row), and the transverse displacement of reflected light (the third row) on $\theta$ for the cases of H polarization (the first column), and V polarization (the second column), respectively. The red (blue) line represents incident from the gain (loss) layer, and the solid (dashed) line represents reflectivity and the transverse displacement of LCP (RCP), respectively. In our calculations, $N=1$, $\varepsilon _{1}=0.1-0.1i, \varepsilon _{2}=0.1+0.1i$, $\mu _{r}=1$, $d=0.785\lambda$, and $w_{0}=15\lambda$.
Fig. 3.
Fig. 3. Dependences of the transmission coefficient (the first row), and the transverse displacement of transmitted light (the second row) on $\theta$ for the cases of H polarization (the first column), and V polarization (the second column), respectively. The solid (dashed) line represents the transverse displacement of LCP (RCP), respectively. Other parameters are the same as Fig. 2.
Fig. 4.
Fig. 4. Dependences of the eigenvalue of the scattering matrix (first row), the reflectivity (second row), and the transverse displacement of reflected light (third row) on $\theta$ for the cases of H polarization (the first column), and V polarization (second column), respectively. The red (blue) line represents incident from the gain (loss) layer, and the solid (dashed) line represents reflectivity and displacement of LCP (RCP), respectively. Where, $N=10$, $\varepsilon _{1}=0.1-0.1i, \varepsilon _{2}=0.1+0.1i$, $\mu _{r}=1$, $d=0.785\lambda$, and $w_{0}=15\lambda$.
Fig. 5.
Fig. 5. The eigenvalue of the scattering matrix contour for different Im($\varepsilon$) with H (a) and V (b) inputs, respectively. Here we choose $N=1$, $\Re (\varepsilon _{1})=\Re (\varepsilon _{2})=0.1$, $\mu _{r}=1$, $d=0.785\lambda$, and $w_{0}=15\lambda$.
Fig. 6.
Fig. 6. Transverse shift contour (integer multiples of wavelength) of LCP reflected light for different Im($\varepsilon$) with H (a) and (c) and V (b) and (d) inputs, respectively. Here, the first (second) row is incident from the loss(gain) side. Other parameters are the same as Fig. 5.
Fig. 7.
Fig. 7. Transverse shift contour (integer multiples of wavelength) of LCP transmitted light for different Im($\varepsilon$) with H (a) and V (b) inputs, respectively. Other parameters are the same as Fig. 5.

Equations (11)

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

( H f + H b + ) = Q ( H f H b ) = ( 1 1 η 0 η 0 ) 1 M ( 1 1 η 0 η 0 ) ( H f H b ) ,
m j = ( cos δ j i sin δ j η j i η j sin δ j cos δ j ) ,
t G = t L = 1 Q 22 = 2 M 11 1 η 0 M 21 η 0 M 12 + M 22 ,
r G = Q 21 Q 22 = M 11 + 1 η 0 M 21 η 0 M 12 + M 22 M 11 1 η 0 M 21 η 0 M 12 + M 22 ,
r L = Q 12 Q 22 = M 11 + 1 η 0 M 21 η 0 M 12 M 22 M 11 1 η 0 M 21 η 0 M 12 + M 22 ,
[ E ~ r H E ~ r V ] = M R [ E ~ i H E ~ i V ] = [ r p k r y k 0 ξ k r y k 0 ξ r s ] [ E ~ i H E ~ i V ] ,
r p , s ( k i x / k 0 ) = r p , s ( 0 ) + k i x k 0 r p , s ( 0 ) ,
E r ± H = 2 π w 0 2 | l | ! exp [ ( x 2 + y 2 ) w 0 ] { [ r p + i x r p ± s l y ξ k 0 w 0 2 / 2 ] [ ( x + i s l y ) w 0 ] | l | + i | l | r p ± s l ξ k 0 w 0 [ ( x + i s l y ) w 0 ] | l 1 | } ,
δ ± H , V = y | E ± H , V | 2 d x d y | E ± H , V | 2 d x d y ,
δ r ± H , V = 1 k 0 ± Re ( r p , s ξ ) l Re ( r p , s r p , s ) | r p , s | 2 + [ | l | + 1 ] [ | r p , s | 2 + | ξ | 2 ] / k 0 2 w 0 2 ,
δ t ± H , V = 1 k 0 ± Re ( t p , s ζ ) l Re ( t p , s t p , s ) | t p , s | 2 + [ | l | + 1 ] [ | t p , s | 2 + | ζ | 2 ] / k 0 2 w 0 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.