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

Thickness-dependent in-plane shift of photonic spin Hall effect in an anisotropic medium

Open Access Open Access

Abstract

As the in-plane spin splitting (IPSS) has a broad application for the precision measurement and sensing, it is extremely important to explore its enhancement mechanism via the photonic spin Hall effect (PSHE). However, for a multilayer structure, the thickness in most of previous works is generally set as a fixed value, lacking the deeply exploration of the influence of thickness on the IPSS. By contrast, here we demonstrate the comprehensive understanding of thickness-dependent IPSS in a three layered anisotropic structure. As thickness increases, near the Brewster angle, the enhanced in-plane shift exhibits a thickness-dependently periodical modulation, besides with much wider incident angle than that in an isotropic medium. While near the critical angle, it becomes thickness-dependently periodical or linear modulation under different dielectric tensors of the anisotropic medium, no longer keeps almost constant in an isotropic medium. In addition, as exploring the asymmetric in-plane shift with arbitrary linear polarization incidence, the anisotropic medium could bring more obvious and wider range of thickness-dependently periodical asymmetric splitting. Our results deepen the understanding of enhanced IPSS, which is expected to promise a pathway in an anisotropic medium for the spin control and integrated device based on PSHE.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

The physical mechanism of in-plane spin splitting (IPSS) originates from the photonic spin Hall effect (PSHE), which arises from the coupling of spin angular momentum and orbital angular momentum [13]. Similar to the Imbert-Fedorov (IF) shift along y-axis [46], IPSS induces the in-plane shift along x-axis, which fully considers the transverse wave vector component, has both spatial and angular characteristics in the geometrical space [7], enabling a distinctive method of optical spin control. The in-plane shift is equivalent to Goos–Hänchen (GH) shift when the shift of left- and right-circularly polarized components coincide, which has the advantage of direct observation compared with the transverse shift. It has been widely used in optical switches [8,9], temperature sensors [10,11], polarization beam splitters [12], and wavelength division multiplexer [13]. Naturally, a larger shift value is essential for improving the sensitivity of practical applications. Besides the conventional method of the Brewster effect [1416], researchers have explored various optical medium for enhancing the in-plane shift, such as metal surface plasmon resonators [17,18], metamaterials [1921], and two-dimensional materials [2224].

More recently, the anisotropic medium has shown significant potential in exploring and aggravating the physical mechanism of spin-orbit interaction due to the extra manipulating degree of freedom in the refractive index provided by the strong anisotropy [2527]. For example, tilted anisotropy plate [2830], hyperbolic metamaterials [3133], epsilon-near-zero materials [34] and two-dimensional anisotropic monolayer graphene surfaces [35] have been proposed to enable the realization of high-efficiency spin devices, optical switchers and optical sensors. Otherwise, as the study of single anisotropic interface is extended to the multilayer structure, such as layered nanostructure [3639], photonic crystal [40,41] and bilayer graphene [42], research has still adapted the fixed anisotropic medium without considering the role parameter of its thickness. In the previous work of anisotropic multilayer structure, the thickness affects the spin shift of PSHE [43,44], especially we found that the interference effect caused by the multiple scattering could strongly enhance the transverse shift of PSHE [45].Therefore, for enhancing and manipulating IPSS, the influence of thickness on the IPSS in multilayer structure needs to be developed in-depth, which is beneficial to the comprehensive understanding of IPSS and broadens the precision measurement method of thickness.

In this work, we focus on the thickness-dependent in-plane shift in a three layered anisotropic structure near the Brewster angle and critical angle. Firstly, it is proved that when considering an isotropic medium as the second layer, the thickness can periodically modulate both in-plane spatial and angular shifts near the Brewster angle, while the enhancement near the critical angle is almost thickness-independent. Then, when the anisotropic medium is substituted for the isotropic medium, the thickness-dependent in-plane shift near the Brewster angle can be realized with a wider incident angle range due to the strong anisotropy. By adjusting the dielectric tensor of the anisotropic medium, the enhanced in-plane shift near the critical angle could be modulated periodically or linearly by thickness. Through discussing and analyzing this thickness-dependent characteristic, we infer that (ζ-ϕp) and |∂rp/∂θi| of the Fresnel reflection coefficient rp are the direct factors leading to the evolution process of in-plane shift near the above two incident angles respectively. Finally, as the incident polarization is extended to arbitrary linear polarization, we investigate the asymmetric in-plane shift, and find that the anisotropic medium can effectively improve the performance of the asymmetric degree, which further verify the advantages of the anisotropic medium in IPSS.

2. Theoretical model and numerical method

Firstly, we establish a general three layered [ɛ1, ɛ2, ɛ3] structure model to reveal the PSHE, as illustrated in Fig. 1(a). The first layer is made of glass with a permittivity of ɛ1 = 1.5152, while the third layer is air with a permittivity of ɛ3 = 1. Importantly, the glass is set as a hemisphere to ensure that the light beam can directly incident on the interface between the first and second layers at any incident angle [46]. Here, ɛ1 > ɛ2, the partial reflection occurs for the critical angle. Only a part of incident light beam reflects from the interface while the remainder continues to transmit. When a linearly polarized light beam reflects at the interface, the gravity center of left- and right-handed circularly polarized light beam will not only show the in-plane spatial shift $\mathrm{\delta }_\mathrm{\ \pm }^{{x}}$, but also the in-plane angular shift $\mathrm{\Delta }_\mathrm{\ \pm }^{{x}}$. To explore the light propagation clearly, the three layered structure in Fig. 1 (a) can be simplified and equivalent to the schematic in Fig. 1(b). When the light incidents into medium 2 from the glass layer, the in-plane spatial and angular shifts will occur at the interface between the glass and medium 2. Here d denotes the thickness of the medium 2.

 figure: Fig. 1.

Fig. 1. (a)Schematic diagram of the PSHE of Gaussian light beam at the three layered structure. (b) Schematics of light transmission and reflection of the three layered structure. Here, “E” and “H” represent the electric field and magnetic field, which respectively parallel and perpendicular to incident plane.

Download Full Size | PDF

Then, the propagation process of the light beam in the three layered structure must be deduced before calculating the in-plane shift. The incident light is a general Gaussian light beam with a wavelength of 632.8 nm. Based on the angular spectrum theory, it can be written as

$$\widetilde {{{E}_{i}}}{=}\frac{{w}}{{\sqrt {{2\pi }} }}\textrm{exp[ - }\frac{{{{w}^{2}}{(k}_{{ix}}^{2}{ + k}_{{iy}}^{2}{)}}}{{4}}{],}$$
where w = 27um (about 42λ) represents the beam waist. kix and kiy are the wave vector components along x and y axes, respectively.

At any given plane z = const, the in-plane shift of the field centroid is given as [47]:

$$\left\langle {{{\text{X}}_{r{{ \pm }}}}} \right\rangle {\text{ = }}\frac{{\left\langle {{{\tilde{E}}_{r \pm }}{\text{|}}i\partial {k_{rx}}{\text{|}}{{\tilde{E}}_{r \pm }}} \right\rangle }}{{\left\langle {{{\tilde{E}}_{r \pm }}{\text{|}}{{\tilde{E}}_{r \pm }}} \right\rangle }}{\text{ + }}\frac{z}{{nk}}\frac{{\left\langle {{{\tilde{E}}_{r \pm }}{\text{|}}{k_{r{\text{x}}}}{\text{|}}{{\tilde{E}}_{r \pm }}} \right\rangle }}{{\left\langle {{{\tilde{E}}_{r \pm }}{\text{|}}{{\tilde{E}}_{r \pm }}} \right\rangle }}{\text{,}}$$

Here z = 250 mm is the transmission distance.

The spatial and angular shift with H polarized incidence (transverse-magnetic, or p-polarized) can be derived based on the plane wave angular spectrum theory combining with enforcing boundary condition [48]:

$$\mathrm{\delta }_{x}^{H}\textrm{=- }\frac{{{k}{{w}^\textrm{2}}{p\chi \;\sin(\zeta -\ }{\phi _{p}}{)}}}{{{{k}^{2}}{{w}^{2}}{{p}^{2}}{ + }{{\chi }^{2}}{ + [}{{p}^{2}}{ + }{{s}^{2}}{ + 2ps\; \cos(}{\phi _{s}}{ - }{\phi _{p}}{)]}{{\cot}^{2}}{{\theta }_{i}}}}{,}$$
$$\Delta _x^H = - \frac{{2zp\chi \,\cos \left( {\zeta - {\phi _P}} \right)}}{{{{\sqrt \varepsilon}_1}\left\{ {{k^2}{w^2}{p^2} + {\chi ^2} + \left[ {{p^2} + {s^2} + 2ps\,\cos \left( {{\phi _s} - {\phi _p}} \right)} \right]{{\cot }^2}{\theta _i}} \right\}}}$$

Here ${{k}_{1}}{ = }\sqrt {{\mathrm{\varepsilon }_{1}}} \; {{k}_{0}}$ and ${{k}_{0}}\mathrm{\ =\ 2\pi /\lambda}$ denotes the wave vector. Only the shift of the right-handed circularly polarized component is considered while that of left-handed is the same with opposite sign. When the transmission distance is fixed, the in-plane angular shift here can be converted into distance, which can be normalized with the in-plane spatial shift about the wavelength. Then, the Fresnel reflection coefficient should be solved by first-order Taylor series expansion to distinguish the difference of plane wave components in different wave packets ${{r}_{{p, s }}}{ = }{{r}_{{p,s}}}{(}{{\theta }_{i}}{) + }\frac{{\partial {{r}_{{p, s}}}}}{{\partial {{\theta }_{i}}}}\frac{{{{k}_{{ix}}}}}{{{{k}_{0}}}}$ [49,50] with complex function ${{r}_{p}}{(}{{\theta }_{i}}{) = p \exp(i\; }{\phi _{p}}{)}$, ${{r}_{s}}{(}{{\theta }_{i}}{) = s\; \exp(i\; }{\phi _{s}}{)}$, $\partial {{r}_{p}}{/}\partial {{\theta }_{i}}{\ =\ \chi \;\exp(i\,\zeta ),}$ $\partial {{r}_{s}}{/}\partial {{\theta }_{i}}{\ =\ \eta \;\exp(i\,}{\epsilon }{)}$. In addition, when the polarization state of incident light beam is arbitrary, it should be represented by Jones matrix ${\mathrm{[cos\beta ,\,sin\beta ]}^\textrm{T}}$ (β here is the polarization angle). Combine with above, the in-plane spatial and angular shifts can be derived as:

$$\mathrm{\delta }_{\pm }^{x}{ = }\frac{{{k}{{w}^{2}}\mathrm{(A\ \pm B\ +\ C)}}}{{\mathrm{D\pm E\ +\ F}}}{,}$$
$$\Delta _ \pm ^x = \frac{{2z}}{{{{\sqrt \varepsilon}_1}}}\,\frac{{\textrm{G} \pm \textrm{H} + \textrm{J}}}{{\textrm{D} \pm \textrm{E} + \textrm{F}}},$$
where A=${p\chi \;\sin(}{\phi _{p}}{\ -\ \zeta )}{{\cos}^{2}}\mathrm{\beta }$, B = [${p\eta \;\cos(}\mathrm{\epsilon }{ - }{\phi _{p}}{)\ -\ s\chi \;\cos(}\mathrm{\epsilon }{ - }{\phi _{s}}{)}$] sin $\mathrm{\beta }$ cos $\mathrm{\beta }$, C=$s\eta$$\sin(\phi _{s}-{\epsilon}){{\sin}^{2}}\mathrm{\beta }$, D=${\{ }{{k}^{2}}{{w}^{2}}{{p}^{2}}{ + }{\mathrm{\chi }^{2}}{ + [}{{p}^{2}}{+ }{{s}^{2}}{ + 2ps\; \cos(}{\phi _{s}}{ - }{\phi _{p}}{)]}{{\cot}^{2}}{\mathrm{\theta }_{i}}{\}}{{\cos}^{2}}\mathrm{\beta }$, E = $[\mathrm{\eta \chi\; sin(}\mathrm{\epsilon }{\ -\ \zeta )\ +\ }$${{k}^{2}}{{w}^{2}}{ps \sin(}{\phi _{s}}{ - }{\phi _{p}}\mathrm{)]sin2\beta }$, F=${\{ }{{k}^{2}}{{w}^{2}}{{s}^{2}}{ + }{\mathrm{\eta }^{2}}{ + [}{{p}^{2}}{+ }{{s}^{2}}{ + 2ps\; \cos(}{\phi _{s}}{ - }{\phi _{p}}{)]}{{\cot}^{2}}{\mathrm{\theta }_{i}}{\}}{{\sin}^{2}}\mathrm{\beta }$}, G = $p\chi$$\cos(\phi _{p}-\zeta ){\textrm{cos}^{2}}\mathrm{\beta }$, H=${[p\eta \;\sin(}\mathrm{\epsilon }{ - }{\phi _{p}}{)+s\chi \;\sin(}\mathrm{\epsilon }{ - }{\phi _{s}}\mathrm{)]sin\mathrm \beta cos\mathrm \beta }$, J=${s\eta \;\cos(}{\phi _{s}}{ - }\mathrm{\epsilon }{)}{\textrm{sin}^{2}}\mathrm{\beta }$.

Here, focusing on the PSHE in multilayer structure, the interference effect of p- and s- polarized waves should be considered to correctly calculate the radiative characteristic. Medium 3 (air) is regarded as a semi-infinite substrate. The overall Fresnel reflection coefficient of p- and s-polarized waves rp,s based on the three layered structure can be rewritten as [40]:

$${r}_{{p,s}}^{{123}}{=r}_{{p,s}}^{{12}}{ + }\frac{{{t}_{{p,s}}^{{12}}{r}_{{p,s}}^{{23}}{t}_{{p,s}}^{{21}}{\exp(2i}{{k}_{{2z}}}{d)}}}{{{1 - r}_{{p,s}}^{{21}}{r}_{{p,s}}^{{23}}{\exp(2i}{{k}_{{2z}}}{d)}}}{,}$$
where r and t represent the reflection and transmission coefficients, respectively. The superscripts indicate the corresponding parameters of medium 1, 2, 3. The overall Fresnel reflection coefficients rp, s in the three layered structure [51] are obtained using the transfer matrix method, and are closely related to the thickness of the second layer medium.

Further, when the second layer medium is anisotropy, ɛ2 should be written as a dielectric tensor: [ɛO, ɛO, ɛE], which represent the uniaxial relative permittivity components for ordinary and extraordinary waves respectively. In our work, the optical axis is set along z axis, then the Fresnel reflection coefficient should be re-deduced as [38]:

$$\scalebox{0.86}{$\displaystyle r_p^{12} = \frac{{{\mathrm{\varepsilon }_\textrm{O}}\,\cos {\theta _i} - \sqrt {{\mathrm{\varepsilon }_1}} \sqrt {\,{\mathrm{\varepsilon }_\textrm{O}} - \,{\mathrm{\varepsilon }_1}({{\mathrm{\varepsilon }_\textrm{O}}/{\mathrm{\varepsilon }_\textrm{E}}} )\,{{\sin }^2}{\theta _i}\,} }}{{{\mathrm{\varepsilon }_\textrm{O}}\,\cos {\theta _i} + \sqrt {{\mathrm{\varepsilon }_1}} \sqrt {\,{\mathrm{\varepsilon }_\textrm{O}} - \,{\mathrm{\varepsilon }_1}({{\mathrm{\varepsilon }_\textrm{O}}/{\mathrm{\varepsilon }_\textrm{E}}} )\,{{\sin }^2}{\theta _i}\,\,} }},\,r_p^{23} = \frac{{\sqrt {{\mathrm{\varepsilon }_3}} \sqrt {\,{\mathrm{\varepsilon }_\textrm{O}} - \,{\mathrm{\varepsilon }_1}({{\mathrm{\varepsilon }_\textrm{O}}/{\mathrm{\varepsilon }_\textrm{E}}} )\,{{\sin }^2}{\theta _i}\,} - {\mathrm{\varepsilon }_\textrm{O}}\sqrt {{\mathrm{\varepsilon }_3} - {\mathrm{\varepsilon }_1}\,{{\sin }^2}{\theta _i}} }}{{\sqrt {{\mathrm{\varepsilon }_3}} \sqrt {\,{\mathrm{\varepsilon }_\textrm{O}} - \,{\mathrm{\varepsilon }_1}({{\mathrm{\varepsilon }_\textrm{O}}/{\mathrm{\varepsilon }_\textrm{E}}} )\,{{\sin }^2}{\theta _i}\,} + {\mathrm{\varepsilon }_\textrm{O}}\sqrt {{\mathrm{\varepsilon }_3} - {\mathrm{\varepsilon }_1}\,{{\sin }^2}{\theta _i}} }}$}$$
$$r_\textrm{s}^{\textrm{12}}\textrm{ = }\frac{{\sqrt {{\mathrm{\varepsilon }_\textrm{1}}} \textrm{cos}{\mathrm{\theta }_{i}}\textrm{ - }\sqrt {{\mathrm{\varepsilon }_\textrm{O}}\textrm{ - }{\mathrm{\varepsilon }_\textrm{1}}\; \textrm{si}{\textrm{n}^\textrm{2}}{\mathrm{\theta }_{i}}} }}{{\sqrt {{\mathrm{\varepsilon }_\textrm{1}}} \textrm{cos}{\mathrm{\theta }_{i}}\textrm{ + }\sqrt {{\mathrm{\varepsilon }_\textrm{O}}\textrm{ - }{\mathrm{\varepsilon }_\textrm{1}}\; \textrm{si}{\textrm{n}^\textrm{2}}{\mathrm{\theta }_{i}}} }},{r}_\textrm{s}^{\textrm{23}}\textrm{ = }\frac{{\sqrt {{\mathrm{\varepsilon }_\textrm{O}}\textrm{ - }{\mathrm{\varepsilon }_\textrm{1}}\; \textrm{si}{\textrm{n}^\textrm{2}}{\mathrm{\theta }_{i}}} \textrm{ - }\sqrt {{\mathrm{\varepsilon }_\textrm{3}}\textrm{ - }{\mathrm{\varepsilon }_\textrm{1}}\; \textrm{si}{\textrm{n}^\textrm{2}}{\mathrm{\theta }_{i}}} }}{{\sqrt {{\mathrm{\varepsilon }_\textrm{O}}\textrm{ - }{\mathrm{\varepsilon }_\textrm{1}}\; \textrm{si}{\textrm{n}^\textrm{2}}{\mathrm{\theta }_{i}}} \textrm{ + }\sqrt {{\mathrm{\varepsilon }_\textrm{3}}\textrm{ - }{\mathrm{\varepsilon }_\textrm{1}}\; \textrm{si}{\textrm{n}^\textrm{2}}{\mathrm{\theta }_{i}}} }}\textrm{.}$$

3. Results and discussion

For comparison, we first demonstrate the thickness-dependent in-plane shift in an isotropic medium, where the ɛ2 is set as 1.4322 (GaF2). Here the Brewster angle and critical angle are approximately 32.76° and 41.304°, respectively. We plot the thickness-dependent in-plane shift according to Eqs. (3) and (4), as shown in Figs. 2(a) and (b). With the change of incident angle, both the in-plane spatial and angular shifts are obviously enhanced near the Brewster angle and critical angle. For the enhancement near the Brewster angle, as thickness increases, both in-plane spatial and angular shifts have a periodical intensity modulation with a period of approximately 0.4λ, which is the result of interference effect and satisfy the Fabry-Pérot resonance in the structure (also occurring when medium 2 is anisotropic) [52]. In Fig. 2(a), as the thickness increases within one period, the peak value of the in-plane spatial shift near the Brewster angle decreases from the maximum (approximately 20λ) firstly and then increases again to the maximum. Subsequently, after experiencing a zero gap, the peak value repeats the process with opposite direction. The incident angle with maximum shift also moves slightly and periodically near the Brewster angle with the change of thickness. In Fig. 2(b), as thickness increases, the periodical evolution process of the in-plane angular shift near Brewster angle is similar to that of in-plane spatial shift, however, the direction of in-plane angular shift is only altered with the motion of incident angle near Brewster angle, which is independent of the thickness. It should be noted that the in-plane spatial and angular shifts described in Eqs. (3) and (4) are directly related to the sin(ζ-ϕp) and cos(ζ-ϕp) (ζ=arg(∂rp/∂θi), ϕp represents the phase of rp), respectively. Therefore, we show the corresponding sin(ζ-ϕp) and cos(ζ-ϕp) in Figs. 2(c) and (d), which are consistent with the evolution of the in-plane spatial and angular shifts. Therefore, the (ζ-ϕp) is the direct factor leading to the evolution process of in-plane shift near the Brewster angle, and we can choose the corresponding thickness of the second layer medium to manipulate the shift value and direction of both the in-plane spatial and angular shifts.

 figure: Fig. 2.

Fig. 2. (a) In-plane spatial and (b) angular shift with different thickness and incident angles under [ɛ1, ɛ2, ɛ3] = [1.5152, 1.4322, 1]. (c) and (d) shows the phase of Fresnel reflection coefficient of rp.

Download Full Size | PDF

While for the enhancement near the critical angle, both the in-plane spatial and angular shifts show evident thickness-independent enhancement (fixed at about 20λ and 700λ respectively as the thickness increases). The sin(ζ-ϕp) and cos(ζ-ϕp) here keep constant, showing the thickness-independence. Even though the ideal in-plane spatial and angular shifts can always be obtained, the thickness insensitivity of the in-plane shift here still hinder its tunability.

Then, in order to obtain a more flexible thickness-dependent modulation and further explain the enhancement mechanism of PSHE near the two incident angles, the anisotropic medium is introduced into the three layered structure. Here, to exclude the influence of other factors and compare with the isotropic medium, we choose two conditions of [ɛO, ɛE] = [1, 1.4322] and [1.4322, 1]. Figure 3 shows the three-dimensional relationship of the in-plane shift with incident angle and thickness, in which the density similar to Fig. 2 is plotted at the bottom. In Figs. 3(a) and (c), the in-plane spatial shift under [ɛO, ɛE] = [1, 1.4322] and [1.4322, 1] both show the thickness-dependent enhancement near the Brewster angle. The similar behavior can also be found in Figs. 3(b) and (d) for the in-plane angular shift. However, unlike the nearly sine-shaped in Figs. 2(a) and (b), the enhancement of in-plane shift near the Brewster angle as thickness increases is fan-shaped. The enhancement range is no longer limited around the Brewster angle. It can even exceed the critical angle under [ɛO, ɛE] = [1, 1.4322] (Fig. 3(a)), while that in Fig. 3(c) is limited less than the critical angle, which is beneficial for the wide-angle enhanced PSHE. Besides, at some fixed thickness, as the incident angle increases, both in-plane spatial and angular shifts in Figs. 3(c) and (d) near the Brewster angle are enhanced more than once (twice for the spatial shift, even quartic for the angular shift), along with the change of direction.

 figure: Fig. 3.

Fig. 3. The relationship of in-plane shift with different thickness and incident angles. (a), (b) and (c), (d) respectively describe the in-plane shift under [ɛO, ɛE] = [1,1.4322] and [1.4322,1].

Download Full Size | PDF

While for the enhancement near the critical angle, both in-plane spatial and angular shifts exhibit obvious modulation as thickness increases, no longer keep almost constant as shown in Fig. 2. When [ɛO, ɛE] = [1, 1.4322] shown in Figs. 3(a) and (b), the in-plane spatial and angular shifts also unfold periodical intensity modulation as thickness increases, which is similar to the in-plane shift near the Brewster angle. Then under [ɛO, ɛE] = [1.4322, 1] in Figs. 3(c) and (d), the in-plane spatial and angular shifts both show a nearly linear variation trend versus thickness. As the thickness increases gradually to 2λ, the in-plane spatial and angular shifts decrease slowly from 17λ and 700λ to approximately 12λ and 600λ, respectively.

In short, these observations illustrate that the introduction of strong anisotropy broadens the incident angle range for the thickness-dependently periodical enhancement of in-plane shift near the Brewster angle, besides makes the enhanced in-plane shift near the critical angle become thickness-dependently periodical to ameliorate the thickness insensitivity in the isotropic medium.

Then, we focus on the enhancement mechanism in the anisotropic medium near the Brewster angle and critical angle. Due to the property of (ζ-ϕp) and the fact that the in-plane angular shift only includes an additional multiplication term related to the transmission distance z, the mechanism of in-plane spatial and angular shifts is similar. Therefore, we focus our analysis solely on the in-plane spatial shift in order to thoroughly investigate the enhancement mechanism. Figures 4(a) and (c) display the sin(ζ-ϕp) with different thickness and incident angles under [ɛO, ɛE] = [1, 1.4322] and [1.4322, 1], respectively. When near the Brewster angle, the sin(ζ- ϕp) directly corresponds to the in-plane spatial shift in Figs. 3(a) and (c). As the sin(ζ-ϕp) reaches the peak value of 1, the corresponding in-plane spatial shift in Figs. 3(a) and (c) also approaches the peak value. Simultaneously, the direction of sin(ζ-ϕp) also manipulates the direction of in- plane spatial shift. Therefore, these findings can further verify that (ζ-ϕp) is the main determinant of the in-plane shift near the Brewster angle. Certainly, compared with the evolution of the isotropic medium in Fig. 2, the influence of strong anisotropy on interference effect here changes the phase difference between these two reflected lights from the upper and lower interfaces, which ultimately leads to the wider incident range of thickness-dependent enhancement of the in-plane shift near the Brewster angle.

 figure: Fig. 4.

Fig. 4. [(a), (c)] sin(ζ-ϕp) and [(b), (d)] ϕp with different thickness and incident angles under [ɛO, ɛE] = [1, 1.4322] and [1.4322, 1].

Download Full Size | PDF

However, it should be noted that near the critical angle shown in Figs. 4(a) and (c), sin(ζ-ϕp)=-1, which is irrelevant to the in-plane shift obviously. Then, we plot the corresponding ϕp in Figs. 4(b) and (d), which can further demonstrate this fan-shaped enhancement effect in the anisotropic medium. Under [1, 1.4322] in Fig. 4(b), the instantly jumps of ϕp could still exists when the incident angle exceeds the critical angle, that is to say the fan-shaped enhancement still occurs. Attention, the ϕp here still cannot directly explain the thickness-dependent modulation near the critical angle under different dielectric tensors of the anisotropic medium.

Therefore, for explaining the origin of the thickness-dependent enhancement of the in-plane shift near the critical angle, we then pay attention to on the χ=|∂rp/∂θi| (Taylor’s first partial derivative of the Fresnel reflection coefficient rp). When the incident angle exceeds the critical angle, here p = 1 (absolute value of rp) and the χ can no longer be ignored. As a result, the in-plane spatial shift in Eq. (3) can be simplified as $\mathrm{\delta }_{x}^{H}{ = (k}{{w}^{2}}\mathrm{\chi )/\{\ }{{k}^{2}}{{w}^{2}}{ + }{\mathrm{\chi }^{2}}{ + 2[1 + \cos(}{\phi _{s}}{ - }{\phi _{p}}{)]}{{\cot}^{2}}{\mathrm{\theta }_{i}}{\} }{.}$ When ${[1 + \cos(}{\phi _{s}}{ - }{\phi _{p}}{)]}{\textrm{cot}^{2}}{\mathrm{\theta }_{i}}$ is small enough to be neglected, χ is the sole variable parameter to affect the in-plane spatial shift near the critical angle. Therefore, in the inset of Figs. 5(a) and (c), we show the relationship of χ versus thickness and incident angle, in which χ increases sharply near the critical angle. We then choose four different curves shown in Fig. 5 to describe the relationship between the in-plane spatial shift and χ near the critical angle. The black curve shows the conditions when the incident angle is slightly less than the critical angle (θi = 41.30°), the blue, green and red curves when exceeds the critical angle (θi = 41.31°, 41.32° and 41.33°, respectively).

 figure: Fig. 5.

Fig. 5. (a) and (c) show the χ=|∂rp/∂θi| with different thickness under [ɛO, ɛE] = [1, 1.4322] and [1.4322, 1]; inset show the three-dimensional image of χ with respect to thickness and incident angle. (b) and (d) illustrate the corresponding in-plane spatial shift with the change of thickness.

Download Full Size | PDF

In Fig. 5(a), it can be seen that under [ɛO, ɛE] = [1, 1.4322], due to the interference effect, χ exhibits a periodical evolution with the increase of thickness, while the peak value moves up. Obviously, the curve evolution of the in-plane spatial shift in Fig. 5(b) coincides with the corresponding χ in Fig. 5(a), which can confirm that χ is the direct result of thickness-dependently periodical modulation near the critical angle. When the incident angle exceeds the critical angle, the curve of χ gradually decreases with the increase of the incident angle, resulting in the decrease of in-plane spatial shift in Fig. 5(b). It should be noted that in Fig. 5(b), although the thickness-dependently periodical modulation always exits near the critical angle, the more obvious and distinguishable modulation only occurs when the incident angle exceeds the critical angle (the blue, green and red curves). The maximum in-plane spatial shift can approach the upper limit: half of beam waist (the blue curve of θi = 41.31°).

Then, under [ɛO, ɛE] = [1.4322, 1], as thickness increases, the evolution of in-plane spatial shift and χ both keep flat near the critical angle, which absent the similar periodical effects in Figs. 5(a) and (b). As ɛE3 = 1, the Fresnel reflection coefficient ${r}_{p}^{{23}}$ in Eq. (8) is simplified to a pure real number. Here ${r}_{p}^{{23}}$ cannot introduce additional phase when calculating the overall Fresnel reflection coefficient rp in the three layered structure, eventually suppressing the interference effect near the critical angle. In addition, the in-plane spatial shift and χ show a good consistency when the incident angle exceeds the critical angle. They all gradually decrease with the increase of thickness, besides, their maximums gradually decrease with the increase of incident angle. The incident angle should be kept close to the critical angle (the blue curve of θi = 41.31° in Figs. 5(b) and (d)) to obtain the optimal in-plane spatial shift. Therefore, the above discoveries verify that χ=abs(∂rp/∂θi) is the fundamental reason for the enhancement of in-plane shift near the critical angle, and it can be controlled by adjusting the dielectric tensor of the anisotropic medium to realize the periodical or linear thickness-dependent modulation near the critical angle. It can be concluded that the anisotropic medium has more valid and unique modulation effect than the isotropic medium, which provides new insights into the in-plane shift in multilayer structure.

Additionally, when the incident light has an arbitrary linear polarization state, the in-plane shifts of left- and right-circularly polarized components are not the same any more, which is the asymmetric shift. The asymmetric degree is described by the difference shift values, denoted by “$\mathrm{|\delta }_{ + }^{x}\mathrm{|-\ |\delta }_{ - }^{x}{|}$” and “$\mathrm{|\Delta }_{ + }^{x}\mathrm{|-\ |\Delta }_{ - }^{x}{|}$” for the in-plane spatial and angular shift respectively [53]. It is worth noting that at a single glass-air interface, the asymmetric degree disappears near the Brewster angle, where the in-plane spatial shifts of two polarization components possess the same magnitude but opposite direction, while for the in-plane angular shifts, they are always the same. Even near the critical angle, the asymmetric degree of both in-plane spatial and angular shifts is too weak to distinguish. In short, we attempt to improve the performance of these asymmetric degrees in the three layered structure.

Firstly, we demonstrate the asymmetric degree of in-plane shifts in an isotropic medium as stating before (ɛ2 = 1.4322) with different polarization angles, shown in Figs. 6(a) and (b). The left and right sides of each figure show the results near the Brewster angle (θi = 33.42°) and critical angle (θi = 41.32°). As can be seen in Figs. 6(a) and (b), near the Brewster angle (left side), both the value and direction of asymmetry degree exhibit the thickness-dependent characteristic with a period of approximately 0.4λ, but the large asymmetric degrees of both in-plane spatial and angular shifts only occur near β=0 (H polarization). While near the critical angle (right side), as thickness increases, although the asymmetry degrees can almost occur in the polarization angle range of 0-90° besides with thickness-dependently periodical modulation, the overall performance of these asymmetry degrees are still relatively weak (the peak values in right side of Figs. 6(a) and (b) are 1λ and 50λ respectively), which hinder the modulation effect of the thickness.

 figure: Fig. 6.

Fig. 6. asymmetric degree of [(a), (c), (e)] in-plane spatial and [(b), (d), (f)] angular shift with different polarization angles and thickness in the three layered structure, respectively.

Download Full Size | PDF

To address these, we replace the second layer medium with the anisotropic medium under [ɛO, ɛE] = [1, 1.4322] and [1.4322, 1], as shown in Figs. 6(c)-(f). With the similar-evolution of thickness-dependently periodical modulation, near the Brewster angle (left side of Figs. 6(c)-(e)), the strong anisotropy can effectively expand the range of large asymmetric degree of in-plane shift, comparing with Figs. 6(a) and (b). It is worth noting that the range under [1.4322, 1] (Figs. 6(e) and (f)) is much more than that under [1, 1.4322] (Figs. 6(c) and (d)). When set 2λ as the minimum discernible shift value, the widest range under [1.4322, 1] could reach 70°, while 30° under [1, 1.4322]. The range in the isotropic medium (Figs. 6(a) and (b)) is only 20°. It can be concluded that this thickness-dependent asymmetric degree near the Brewster angle can be effectively controlled and expanded by further adjusting the dielectric tensor of the anisotropic medium.

Near the critical angle (right side of Figs. 6(c)-(f)), it is intuitive that the asymmetric degree of in-plane spatial and angular shifts is enhanced significantly (the peak values of asymmetric degree can reach 20λ and 600λ respectively), much more than that in the isotropic medium. Besides, in the right side of Figs. 6(e) and (f), the peak value of asymmetric degree under [1.4322, 1] gradually decrease with the increase of thickness, while the period number is larger than that under [1, 1.4322] in Figs. 6(c) and (d). Therefore, the anisotropic medium reveals an efficient and flexible method to control the asymmetric degree of in-plane shift, and the thickness in multilayer structure should be fully considered to obtain an ideal asymmetric degree.

4. Conclusions

In conclusion, we have investigated the thickness-dependent in-plane spatial and angular shift in a multilayer structure near the Brewster angle and critical angle. On the basis of expression of in-plane shift with H polarized incidence, the phase and Taylor’s first partial derivative of the Fresnel reflection coefficient rp can directly affect IPSS. For this thickness-dependent characteristic due to the interference effect, the enhancement of in-plane shift near the Brewster angle can be expanded with a wider range when introducing the anisotropic medium instead of the isotropic medium. By virtue of this strong anisotropy, the in-plane shift emerges a periodical or linear thickness-modulation near the critical angle, which ameliorates the almost thickness-independence in the isotropic medium. In addition, when considering the arbitrary linear polarization incidence, compared with the isotropic medium, the anisotropic medium not only expands the polarization angle range of large thickness-dependent asymmetric degree near the Brewster angle from 20° to 70°, but also enhances the asymmetric degree near the critical angle from 1λ and 50λ to 20λ and 600λ, respectively. These findings further reveal that the anisotropic medium-based platform can provide an efficient and flexible method for the thickness-dependent modulation, and may provide a perspective in probing the spin-orbit interaction.

Funding

National Natural Science Foundation of China (11604095); Shenzhen Government’s Plan of Science and Technology (JCYJ20180305124927623, JCYJ20190808150205481).

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. X. Zhou and X. Ling, “Unveiling the photonic spin Hall effect with asymmetric spin-dependent splitting,” Opt. Express 24(3), 3025–3036 (2016). [CrossRef]  

3. M. Merano, N. Hermosa, A. Aiello, and J. P. Woerdman, “Demonstration of a quasi-scalar angular Goos-Hänchen effect,” Opt. Lett. 35(21), 3562–3564 (2010). [CrossRef]  

4. J. Li, T. Tang, L. Luo, J. Shen, C. Li, J. Qin, and L. Bi, “Weak measurement of the magneto-optical spin Hall effect of light,” Photonics Res. 7(9), 1014–1018 (2019). [CrossRef]  

5. T. Li, Q. Wang, A. Taallah, S. Zhang, T. Yu, and Z. Zhang, “Measurement of the magnetic properties of thin films based on the spin Hall effect of light,” Opt. Express 28(20), 29086–29097 (2020). [CrossRef]  

6. Y. Wu, S. Liu, S. Chen, H. Luo, and S. Wen, “Examining the optical model of graphene via the photonic spin Hall effect,” Opt. Lett. 47(4), 846–849 (2022). [CrossRef]  

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

8. T. Sakata, T. Hiroyoshi, and F. Shimokawa, “Reflection-type 2× 2 optical waveguide switch using the Goos–Hänchen shift effect,” Appl. Phys. Lett. 76(20), 2841–2843 (2000). [CrossRef]  

9. X. Wang, C. Yin, J. Sun, J. Gao, M. Huang, and Z. Cao, “Reflection-type space-division optical switch based on the electrically tuned Goos–Hänchen effect,” J. Opt. (Bristol, U. K.) 15(1), 014007 (2013). [CrossRef]  

10. X. Wang, C. Yin, J. Sun, H. Li, Y. Wang, M. Ran, and Z. Cao, “High-sensitivity temperature sensor using the ultrahigh order mode-enhanced Goos-Hänchen effect,” Opt. Express 21(11), 13380–13385 (2013). [CrossRef]  

11. X. Zhou, P. Tang, C. Yang, S. Liu, and Z. Luo, “Temperature-dependent Goos-Hänchen shifts in a symmetrical graphene-cladding waveguide,” Results Phys. 24, 104100 (2021). [CrossRef]  

12. X. Chen, M. Shen, Z. Zhang, and C. Li, “Tunable lateral shift and polarization beam splitting of the transmitted light beam through electro-optic crystals,” J. Appl. Phys. (Melville, NY, U. S.) 104(12), 123101 (2008). [CrossRef]  

13. H. Sattari, S. Ebadollahi-Bakhtevar, and M. Sahrai, “Proposal for a 1× 3 Goos-Hänchen shift-assisted de/multiplexer based on a multilayer structure containing quantum dots,” J. Appl. Phys. (Melville, NY, U. S.) 120(13), 133102 (2016). [CrossRef]  

14. 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: At., Mol., Opt. Phys. 84(4), 043806 (2011). [CrossRef]  

15. X. Qiu, Z. Zhang, L. Xie, J. Qiu, F. Cao, and J. Du, “Incident-polarization-sensitive and large in-plane-photonic-spin-splitting at the Brewster angle,” Opt. Lett. 40(6), 1018–1021 (2015). [CrossRef]  

16. Y. Shu, Y. Song, Z. Wen, Y. Zhang, S. Liu, J. Liu, and Z. Luo, “Theory of quantized photonic spin Hall effect in strained graphene under a sub-Tesla external magnetic field,” Opt. Express 31(5), 8805–8819 (2023). [CrossRef]  

17. X. Tan and X. Zhu, “Enhancing photonic spin Hall effect via long-range surface plasmon resonance,” Opt. Lett. 41(11), 2478–2481 (2016). [CrossRef]  

18. Y. Xiang, X. Jiang, Q. You, and X. Dai, “Enhanced spin Hall effect of reflected light with guided-wave surface plasmon resonance,” Photonics Res. 5(5), 467–472 (2017). [CrossRef]  

19. D. Liu, C. Liang, D. Deng, G. Wang, and L. Zhang, “Asymmetric spin splitting of Laguerre-Gaussian beams in chiral PT-symmetric metamaterials,” Opt. Express 30(23), 41821–41831 (2022). [CrossRef]  

20. Y. Huang, W. Dong, L. Gao, and C. Qiu, “Large positive and negative lateral shifts near pseudo-Brewster dip on reflection from a chiral metamaterial slab,” Opt. Express 19(2), 1310–1323 (2011). [CrossRef]  

21. Q. Kong, H. Shi, J. Shi, and X. Chen, “Goos-Hänchen and Imbert-Fedorov shifts at gradient metasurfaces,” Opt. Express 27(9), 11902–11913 (2019). [CrossRef]  

22. H. Lin, B. Chen, S. Yang, W. Zhu, J. Yu, H. Guan, H. Lu, Y. Luo, and Z. Chen, “Photonic spin Hall effect of monolayer black phosphorus in the Terahertz region,” Nanophotonics 7(12), 1929–1937 (2018). [CrossRef]  

23. W. Zhang, W. Wu, S. Chen, J. Zhang, X. Ling, W. Shu, H. Luo, and S. Wen, “Photonic spin Hall effect on the surface of anisotropic two-dimensional atomic crystals,” Photonics Res. 6(6), 511–516 (2018). [CrossRef]  

24. A. Das and M. Pradhan, “Goos–Hänchen shift for Gaussian beams impinging on monolayer-MoS 2-coated surfaces,” J. Opt. Soc. Am. B 35(8), 1956–1962 (2018). [CrossRef]  

25. L. Xu, Y. Bai, G. Gong, F. Song, Z. Li, Y. Han, L. Ling, and Z. Tian, “Strong anisotropic Hall effect in single-crystalline CeMn2Ge2 with helical spin order,” Phys. Rev. B 105(7), 075108 (2022). [CrossRef]  

26. M. Mazanov, O. Yermakov, I. Deriy, O. Takayama, A. Bogdanov, and A. V. Lavrinenko, “Photonic spin Hall effect: contribution of polarization mixing caused by anisotropy,” Quantum Rep. 2(4), 489–500 (2020). [CrossRef]  

27. Y. Zhang, Y. Sun, H. Yang, J. Zelezny, S. Parkin, C. Felser, and B. Yan, “Strong anisotropic anomalous Hall effect and spin Hall effect in the chiral antiferromagnetic compounds Mn3X (X = Ge, Sn, Ga, Ir, Rh, and Pt),” Phys. Rev. B 95(7), 075128 (2017). [CrossRef]  

28. K. Y. Bliokh, C. Samlan, C. Prajapati, G. Puentes, N. K. Viswanathan, and F. Nori, “Spin-Hall effect and circular birefringence of a uniaxial crystal plate,” Optica 3(10), 1039–1047 (2016). [CrossRef]  

29. O. Takayama and G. Puentes, “Enhanced spin Hall effect of light by transmission in a polymer,” Opt. Lett. 43(6), 1343–1346 (2018). [CrossRef]  

30. K. Y. Bliokh, C. Prajapati, C. T. Samlan, N. K. Viswanathan, and F. Nori, “Spin-Hall effect of light at a tilted polarizer,” Opt. Lett. 44(19), 4781–4784 (2019). [CrossRef]  

31. O. Takayama, J. Sukham, R. Malureanu, A. V. Lavrinenko, and G. Puentes, “Photonic spin Hall effect in hyperbolic metamaterials at visible wavelengths,” Opt. Lett. 43(19), 4602–4605 (2018). [CrossRef]  

32. M. Kim, D. Lee, Y. Wang, Y. Kim, and J. Rho, “Reaching the highest efficiency of spin Hall effect of light in the near-infrared using all-dielectric metasurfaces,” Nat. Commun. 13(1), 2036 (2022). [CrossRef]  

33. M. Kim, D. Lee, T. H. Kim, Y. Yang, H. J. Park, and J. Rho, “Observation of enhanced optical spin Hall effect in a vertical hyperbolic metamaterial,” ACS Photon. 6(10), 2530–2536 (2019). [CrossRef]  

34. H. Chen, D. Guan, W. Zhu, H. Zheng, J. Yu, Y. Zhong, and Z. Chen, “High-performance photonic spin Hall effect in anisotropic epsilon-near-zero metamaterials,” Opt. Lett. 46(17), 4092–4095 (2021). [CrossRef]  

35. Z. Chen, H. Zhang, X. Zhang, H. Li, W. Zhang, and L. Xi, “Cross-coupling effect induced beam shifts for polarized vortex beam at two-dimensional anisotropic monolayer graphene surface,” Opt. Express 28(6), 8308–8323 (2020). [CrossRef]  

36. 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]  

37. X. Tan, “Longitudinal spin splitting of vortex beam in surface plasmon resonance,” Plasmonics 14(6), 1411–1417 (2019). [CrossRef]  

38. S. Shahnawaz, X. Lin, L. Shen, M. Renuka, B. Zhang, and H. Chen, “Interferenceless polarization splitting through nanoscale van der Waals heterostructures,” Phys. Rev. Applied 10(3), 034025 (2018). [CrossRef]  

39. W. Zhu, M. Jiang, H. Guan, J. Yu, H. Lu, and Z. Chen, “Tunable spin splitting of Laguerre–Gaussian beams in graphene metamaterials,” Photonics Res. 5(6), 684–688 (2017). [CrossRef]  

40. J. Li, T. Tang, L. Luo, and J. Yao, “Enhancement and modulation of photonic spin Hall effect by defect modes in photonic crystals with graphene,” Carbon 134, 293–300 (2018). [CrossRef]  

41. S. Chen, X. Ling, W. Shu, H. Luo, and S. Wen, “Precision measurement of the optical conductivity of atomically thin crystals via the photonic spin Hall effect,” Phys. Rev. Applied 13(1), 014057 (2020). [CrossRef]  

42. M. Cheng, P. Fu, and S. Chen, “Enhanced and tunable photonic spin Hall effect in metasurface bilayers,” J. Opt. Soc. Am. B 39(1), 316–323 (2022). [CrossRef]  

43. T. Tang, C. Li, and L. Luo, “Enhanced spin Hall effect of tunneling light in hyperbolic metamaterial waveguide,” Sci. Rep. 6(1), 30762–8 (2016). [CrossRef]  

44. T. Tang, Y. Zhang, J. Li, and L. Luo, “Spin Hall effect enhancement of transmitted light through an anisotropic metamaterial slab,” IEEE Photonics J. 9(4), 1–10 (2017). [CrossRef]  

45. Z. Chen, S. Lin, J. Hong, L. Sheng, Y. Chen, and X. Zhou, “Enhanced photonic spin Hall effect via singularity induced by destructive interference,” Opt. Lett. 46(19), 4883–4886 (2021). [CrossRef]  

46. Y. Wu, L. Sheng, L. Xie, S. Li, P. Nie, Y. Chen, X. Zhou, and X. Ling, “Actively manipulating asymmetric photonic spin Hall effect with graphene,” Carbon 166, 396–404 (2020). [CrossRef]  

47. A. Andrea, “Goos–Hänchen and Imbert–Fedorov shifts: a novel perspective,” New J. Phys. 14(1), 013058 (2012). [CrossRef]  

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

49. J. M. Ménard, A. E. Mattacchione, M. Betz, and H. M. van Driel, “Imaging the spin Hall effect of light inside semiconductors via absorption,” Opt. Lett. 34(15), 2312–2314 (2009). [CrossRef]  

50. Y. Qin, Y. Li, X. Feng, Y. Xiao, H. Yang, and Q. Gong, “Observation of the in-plane spin separation of light,” Opt. Express 19(10), 9636–9645 (2011). [CrossRef]  

51. J. A. Kong, “Electromagnetic Wave Theory,” EMW Publishing, Cambridge, MA (2008).

52. T. Tang, J. Li, L. Luo, P. Sun, and J. Yao, “Magneto-optical modulation of photonic spin Hall effect of graphene in terahertz region,” Adv. Opt. Mater. 6(7), 1701212 (2018). [CrossRef]  

53. X. Zhou, L. Xie, X. Ling, S. Cheng, Z. Zhang, H. Luo, and H. Sun, “Large in-plane asymmetric spin angular shifts of a light beam near the critical angle,” Opt. Lett. 44(2), 207–210 (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 (6)

Fig. 1.
Fig. 1. (a)Schematic diagram of the PSHE of Gaussian light beam at the three layered structure. (b) Schematics of light transmission and reflection of the three layered structure. Here, “E” and “H” represent the electric field and magnetic field, which respectively parallel and perpendicular to incident plane.
Fig. 2.
Fig. 2. (a) In-plane spatial and (b) angular shift with different thickness and incident angles under [ɛ1, ɛ2, ɛ3] = [1.5152, 1.4322, 1]. (c) and (d) shows the phase of Fresnel reflection coefficient of rp.
Fig. 3.
Fig. 3. The relationship of in-plane shift with different thickness and incident angles. (a), (b) and (c), (d) respectively describe the in-plane shift under [ɛO, ɛE] = [1,1.4322] and [1.4322,1].
Fig. 4.
Fig. 4. [(a), (c)] sin(ζ-ϕp) and [(b), (d)] ϕp with different thickness and incident angles under [ɛO, ɛE] = [1, 1.4322] and [1.4322, 1].
Fig. 5.
Fig. 5. (a) and (c) show the χ=|∂rp/∂θi| with different thickness under [ɛO, ɛE] = [1, 1.4322] and [1.4322, 1]; inset show the three-dimensional image of χ with respect to thickness and incident angle. (b) and (d) illustrate the corresponding in-plane spatial shift with the change of thickness.
Fig. 6.
Fig. 6. asymmetric degree of [(a), (c), (e)] in-plane spatial and [(b), (d), (f)] angular shift with different polarization angles and thickness in the three layered structure, respectively.

Equations (9)

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

$$\widetilde {{{E}_{i}}}{=}\frac{{w}}{{\sqrt {{2\pi }} }}\textrm{exp[ - }\frac{{{{w}^{2}}{(k}_{{ix}}^{2}{ + k}_{{iy}}^{2}{)}}}{{4}}{],}$$
$$\left\langle {{{\text{X}}_{r{{ \pm }}}}} \right\rangle {\text{ = }}\frac{{\left\langle {{{\tilde{E}}_{r \pm }}{\text{|}}i\partial {k_{rx}}{\text{|}}{{\tilde{E}}_{r \pm }}} \right\rangle }}{{\left\langle {{{\tilde{E}}_{r \pm }}{\text{|}}{{\tilde{E}}_{r \pm }}} \right\rangle }}{\text{ + }}\frac{z}{{nk}}\frac{{\left\langle {{{\tilde{E}}_{r \pm }}{\text{|}}{k_{r{\text{x}}}}{\text{|}}{{\tilde{E}}_{r \pm }}} \right\rangle }}{{\left\langle {{{\tilde{E}}_{r \pm }}{\text{|}}{{\tilde{E}}_{r \pm }}} \right\rangle }}{\text{,}}$$
$$\mathrm{\delta }_{x}^{H}\textrm{=- }\frac{{{k}{{w}^\textrm{2}}{p\chi \;\sin(\zeta -\ }{\phi _{p}}{)}}}{{{{k}^{2}}{{w}^{2}}{{p}^{2}}{ + }{{\chi }^{2}}{ + [}{{p}^{2}}{ + }{{s}^{2}}{ + 2ps\; \cos(}{\phi _{s}}{ - }{\phi _{p}}{)]}{{\cot}^{2}}{{\theta }_{i}}}}{,}$$
$$\Delta _x^H = - \frac{{2zp\chi \,\cos \left( {\zeta - {\phi _P}} \right)}}{{{{\sqrt \varepsilon}_1}\left\{ {{k^2}{w^2}{p^2} + {\chi ^2} + \left[ {{p^2} + {s^2} + 2ps\,\cos \left( {{\phi _s} - {\phi _p}} \right)} \right]{{\cot }^2}{\theta _i}} \right\}}}$$
$$\mathrm{\delta }_{\pm }^{x}{ = }\frac{{{k}{{w}^{2}}\mathrm{(A\ \pm B\ +\ C)}}}{{\mathrm{D\pm E\ +\ F}}}{,}$$
$$\Delta _ \pm ^x = \frac{{2z}}{{{{\sqrt \varepsilon}_1}}}\,\frac{{\textrm{G} \pm \textrm{H} + \textrm{J}}}{{\textrm{D} \pm \textrm{E} + \textrm{F}}},$$
$${r}_{{p,s}}^{{123}}{=r}_{{p,s}}^{{12}}{ + }\frac{{{t}_{{p,s}}^{{12}}{r}_{{p,s}}^{{23}}{t}_{{p,s}}^{{21}}{\exp(2i}{{k}_{{2z}}}{d)}}}{{{1 - r}_{{p,s}}^{{21}}{r}_{{p,s}}^{{23}}{\exp(2i}{{k}_{{2z}}}{d)}}}{,}$$
$$\scalebox{0.86}{$\displaystyle r_p^{12} = \frac{{{\mathrm{\varepsilon }_\textrm{O}}\,\cos {\theta _i} - \sqrt {{\mathrm{\varepsilon }_1}} \sqrt {\,{\mathrm{\varepsilon }_\textrm{O}} - \,{\mathrm{\varepsilon }_1}({{\mathrm{\varepsilon }_\textrm{O}}/{\mathrm{\varepsilon }_\textrm{E}}} )\,{{\sin }^2}{\theta _i}\,} }}{{{\mathrm{\varepsilon }_\textrm{O}}\,\cos {\theta _i} + \sqrt {{\mathrm{\varepsilon }_1}} \sqrt {\,{\mathrm{\varepsilon }_\textrm{O}} - \,{\mathrm{\varepsilon }_1}({{\mathrm{\varepsilon }_\textrm{O}}/{\mathrm{\varepsilon }_\textrm{E}}} )\,{{\sin }^2}{\theta _i}\,\,} }},\,r_p^{23} = \frac{{\sqrt {{\mathrm{\varepsilon }_3}} \sqrt {\,{\mathrm{\varepsilon }_\textrm{O}} - \,{\mathrm{\varepsilon }_1}({{\mathrm{\varepsilon }_\textrm{O}}/{\mathrm{\varepsilon }_\textrm{E}}} )\,{{\sin }^2}{\theta _i}\,} - {\mathrm{\varepsilon }_\textrm{O}}\sqrt {{\mathrm{\varepsilon }_3} - {\mathrm{\varepsilon }_1}\,{{\sin }^2}{\theta _i}} }}{{\sqrt {{\mathrm{\varepsilon }_3}} \sqrt {\,{\mathrm{\varepsilon }_\textrm{O}} - \,{\mathrm{\varepsilon }_1}({{\mathrm{\varepsilon }_\textrm{O}}/{\mathrm{\varepsilon }_\textrm{E}}} )\,{{\sin }^2}{\theta _i}\,} + {\mathrm{\varepsilon }_\textrm{O}}\sqrt {{\mathrm{\varepsilon }_3} - {\mathrm{\varepsilon }_1}\,{{\sin }^2}{\theta _i}} }}$}$$
$$r_\textrm{s}^{\textrm{12}}\textrm{ = }\frac{{\sqrt {{\mathrm{\varepsilon }_\textrm{1}}} \textrm{cos}{\mathrm{\theta }_{i}}\textrm{ - }\sqrt {{\mathrm{\varepsilon }_\textrm{O}}\textrm{ - }{\mathrm{\varepsilon }_\textrm{1}}\; \textrm{si}{\textrm{n}^\textrm{2}}{\mathrm{\theta }_{i}}} }}{{\sqrt {{\mathrm{\varepsilon }_\textrm{1}}} \textrm{cos}{\mathrm{\theta }_{i}}\textrm{ + }\sqrt {{\mathrm{\varepsilon }_\textrm{O}}\textrm{ - }{\mathrm{\varepsilon }_\textrm{1}}\; \textrm{si}{\textrm{n}^\textrm{2}}{\mathrm{\theta }_{i}}} }},{r}_\textrm{s}^{\textrm{23}}\textrm{ = }\frac{{\sqrt {{\mathrm{\varepsilon }_\textrm{O}}\textrm{ - }{\mathrm{\varepsilon }_\textrm{1}}\; \textrm{si}{\textrm{n}^\textrm{2}}{\mathrm{\theta }_{i}}} \textrm{ - }\sqrt {{\mathrm{\varepsilon }_\textrm{3}}\textrm{ - }{\mathrm{\varepsilon }_\textrm{1}}\; \textrm{si}{\textrm{n}^\textrm{2}}{\mathrm{\theta }_{i}}} }}{{\sqrt {{\mathrm{\varepsilon }_\textrm{O}}\textrm{ - }{\mathrm{\varepsilon }_\textrm{1}}\; \textrm{si}{\textrm{n}^\textrm{2}}{\mathrm{\theta }_{i}}} \textrm{ + }\sqrt {{\mathrm{\varepsilon }_\textrm{3}}\textrm{ - }{\mathrm{\varepsilon }_\textrm{1}}\; \textrm{si}{\textrm{n}^\textrm{2}}{\mathrm{\theta }_{i}}} }}\textrm{.}$$
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.