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

Bistable reflection and beam shifts with excitation of surface plasmons in a saturable absorbing medium

Open Access Open Access

Abstract

We investigate the nonlinear reflection of a light beam from a Kretschmann configuration with saturable absorbing medium. The absorption of medium has direct influence on the intrinsic loss of the system, thus affecting the reflectivity and the phase variation when the surface plasmons are resonantly excited. As the incident power changes, the reflectivity can be switched between high and low values and exhibits absorptive optical bistability as a result of the inherent positive feedback by the intensity-dependent saturation effect. The Goos-Hänchen and the Imbert-Fedorov shifts of the reflected beam have the same bistable behavior as the reflectance. The effects of the thickness of metal film and the linear absorption coefficient on the hysteresis loop are analyzed in detail by considering the system losses and the saturated absorption. The bistable reflection and beam shifts may have applications in all-optical devices, such as optical switching.

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

1. Introduction

When a bounded light impinges on the boundary of two media with different refractive indices, the reflected or refracted light beam undergoes displacements both in and perpendicular to the incident plane. The lateral shift was first observed by Goos and Hänchen in the experiment of totally internal reflection (TIR) [1] and then theoretically explained by Artmann [2]. The different plane wave components of a finite-sized light beam acquire different phase changes and the coherent superposition of these components then forms a reflected beam with a lateral displacement. The transverse Imbert-Fedorov shift [3,4], also known as the spin Hall effect of light (SHEL), results from the conservation of angular momentum and the spin-orbit interaction [57], and thus the photons with different spin (the left/right-hand circularly polarized) components accumulate on the two sides of the incident plane. Both Goos-Hänchen (GH) shift and Imbert-Fedorov (IF) shift are sensitive to the optical properties of materials and the structures of devices. The precise measurement of beam displacements has potential application in devising new types of optical devices. For example, the GH shift can be utilized for optical switching [8,9] and optical sensors [1012]; the IF shift is applied to identify graphene layers [13], edge detection [14] and other optical sensors [1517]. In general, the beam shifts in reflection or refraction from single interface are too small to be conveniently used in practice. Therefore, the enhancement and manipulation of beam shifts have attracted much attention in recent years. Different materials and various structures were employed to achieve large tunable GH and IF shifts, such as dielectric slab [1820], graphene [2123], surface plasmon resonance (SPR) systems [2430], multi-layer nanostructures [3134], coherent media [3540].

Optical bistability (OB) in reflection from the layered structure containing nonlinear medium is of great interest because there is no need of an optical cavity [4146]. Kaplan first explored the bistable reflection of light incidents on the interface of a nonlinear medium at grazing-angle [41]. The hysteretic response can result from the combined effect of the intensity-dependent nonlinear refractivity and the positive feedback from the change in the coupling efficiency of SPR [42,43] or guiding mode resonance [44,45]. Recently, multi-layer structures with graphene were proposed to realize OB [4648]. In addition to the bistable reflection, a hysteretic behavior between the GH/IF shifts and the incident intensity of light beam can also be produced owing to the excitation of surface plasmons (SPs) [25,49]. There are two types of OB, absorptive type and dispersive type. In most of the above investigations, Kerr media with intensity-dependent refractive index were utilized and then the resulting OB is dispersive-type. Up till now, few studies focused on the absorptive OB with SPs. To the very recent, the interplay between nonlinear gain media and SPs was investigated to achieve bistable enhanced reflectance [50,51]. Nevertheless, to the best of our knowledge, the influence of saturated absorption effect on the nonlinear response of SPR system has not been reported before.

In this paper, we study the bistable reflection and beam shifts in a Kretschmann configuration containing a saturable absorbing medium. When the light incidents upon the prism-metal interface with a specific angle, SPs can be resonantly excited in the medium below the metal, resulting a dip in the reflectance spectrum, in which the phase of reflected light changes rapidly. The reflectivity and phase variation in the SPR dip depend strongly on the absorption coefficient of medium. Owing to feedback from the intensity-dependent saturable absorption effect, the reflectivity as well as the GH/IF shifts can be switched from one stable state to the other stable state, thus exhibiting hysteretic feature. We also discuss and analyze the system parameters, such as the thickness of metal film, the linear absorbing coefficient, on the hysteresis curves. The results may be helpful for devising new type of bistable devices and SPR-based devices.

2. Model and equations

We consider a three-layer Kretschmann configuration supporting SPR as shown in Fig. 1. A thin metal film with thickness ${d_{1}}$ is sandwiched between a high-index prism and a low-index dielectric medium which has the characteristic of saturable absorption. BK7 glass with permittivity ${\varepsilon _{1}} = 2.28$ and silver with permittivity ${\varepsilon _{2}} ={-} 29.384 + 0.365i$ at incident wavelength $\lambda = 780{\textrm{nm}}$ are taken as the prism and the metal layer. The saturated absorbing medium is a homogeneously broadened atomic vapor of which the dielectric constant can be described as

$${\varepsilon _{3}} = 1 + i\alpha ({I_{3}})$$
with
$$\alpha ({I_3}) = \frac{{{\alpha _0}}}{{1 + \frac{{{I_3}}}{{{I_s}}}}},$$
where $\alpha$ and ${\alpha _0}$ are the nonlinear and linear absorption coefficients. ${I_3}$ and ${I_s}$ represent the intensity in the absorbing medium and saturable intensity of light, respectively. ${I_3}$ can be written as ${I_3} = {{{\mu _0}c{{|H |}^2}} / {({2{n_r}} )}}$, where ${n_r} = {\textrm{Re}} (\sqrt {{\varepsilon _3}} )$ is the real part of refractive index of the medium, H denotes the magnetic component of the electromagnetic field in the absorbing medium and ${\mu _0}$ is the permeability of vacuum. When the light beam incidents with an angle ${\theta _i}$ larger than the critical angle of total reflection, the intensity of evanescent wave ${I_3}(z)$ diminishes exponentially with the penetration depth, and thus the absorption coefficient $\alpha (z)$ in the medium increases gradually with z. The mean light intensity ${\tilde{I}_3}$ in the absorbing medium can be approximated as
$${\tilde{I}_3} \approx \frac{1}{2}{I_{3d}} = \frac{{c{\mu _0}}}{{4{n_r}}}{|{{H_d}} |^2},$$
where ${I_{3d}}$ and ${H_d}$ correspond to the intensity and magnetic component of electromagnetic filed at the metal-medium interface ($z = {d_1}$). Once the magnetic field ${H_d}$ is known, the incident light intensity can be obtained as
$$I = \frac{{c{\mu _0}}}{{2\sqrt {{\varepsilon _1}} }}{\left|{\frac{{{H_d}}}{t}} \right|^2},$$
with t being the transmission coefficient of the Kretschmann structure. For the p- and s-polarized incident light, the reflection coefficient ${r_p}$ and ${r_s}$, and the transmission coefficient ${t_p}$ and ${t_s}$, can be obtained by transfer matrix method and are given by
$${r_f} = \frac{{r_{12}^f + r_{23}^f{e^{2i{k_{2z}}d}}}}{{1 + r_{12}^fr_{23}^f{e^{2i{k_{2z}}d}}}},\textrm{ }f = p,s$$
$${t_f} = \frac{{t_{12}^ft_{23}^f{e^{i{k_{2z}}d}}}}{{1 + r_{12}^fr_{23}^f{e^{2i{k_{2z}}d}}}},\textrm{ }f = p,s$$
where $r_{ij}^f$ and $t_{ij}^f$ represent the Fresnel’s reflection and transmission coefficients at $i - j$ interface given by
$$r_{ij}^f = \left\{ \begin{array}{ll} \frac{{{{{k_{iz}}} / {{\varepsilon_i}}} - {{{k_{jz}}} / {{\varepsilon_j}}}}}{{{{{k_{iz}}} / {{\varepsilon_i}}} + {{{k_{jz}}} / {{\varepsilon_j}}}}}&\textrm{ for }f = p,\\ \frac{{{k_{iz}} - {k_{jz}}}}{{{k_{iz}} + {k_{jz}}}}&\textrm{ for }f = s, \end{array} \right.$$
$$t_{ij}^f = 1 + r_{ij}^f.$$

 figure: Fig. 1.

Fig. 1. Schematic diagram of Goos-Hänchen and Imbert-Fedorov effects in a three-layer Kretschmann configuration composed of a prism, a thin metal film backed by a saturable absorbing medium.

Download Full Size | PDF

Here ${k_{iz}} = {(k_0^2{\varepsilon _i} - k_x^2)^{1/2}}$ represents the normal wave vector in medium i, ${k_x} = \sqrt {{\varepsilon _1}} {k_0}\sin {\theta _i}$ denotes the wave vector along the x direction and ${k_0} = {{2\pi } / \lambda }$ is the wave vector in vacuum.

Because SPR can only be excited by light with p-polarization while the s-polarized light undergoes nearly total reflection, we focus on the reflectivity and lateral displacement of p-polarized light. According to the stationary phase theory [2], the GH shift, related to the reflected phase [${\varphi _p} = \arg ({r_p})$], is expressed as

$${S_{\textrm{GH}}} ={-} \frac{1}{{\sqrt {{\varepsilon _1}} {k_0}}}\frac{{\textrm{d}{\varphi _p}}}{{\textrm{d}{\theta _i}}}.$$

The transverse shift is spin-dependent. The transverse displacements of different spin components are the same in magnitude but opposite in direction. The IF shift of the left-hand circularly polarized (spin) component of the reflected light beam can be described as [52]:

$${S_{\textrm{IF}}} ={-} \frac{1}{{\sqrt {{\varepsilon _1}} {k_0}}}\left[ {1 + \left|{\frac{{{r_s}}}{{{r_p}}}} \right|\cos ({\varphi_s} - {\varphi_p})} \right]\cot {\theta _i}$$
with ${\varphi _s} = \arg ({r_s})$ being the reflected phase of s-polarized light. Owing to the nonlinear saturated absorption effect, the reflectivity and the GH/IF shifts are sensitive to the input power and exhibit hysteretic behavior. In the formation of OB, the intensity-dependent absorption is essential. For the nonlinear medium, we can derive the distribution of electric field from the differential equations given by
$$\frac{\textrm{d}}{{\textrm{d}z}}\left( {\begin{array}{{c}} H\\ {{{\tilde{E}}_x}} \end{array}} \right) = \textrm{i}k\left( {\begin{array}{{cc}} 0&{{\varepsilon_3}}\\ {1 - \frac{{{\varepsilon_1}}}{{{\varepsilon_3}}}{{\sin }^2}\theta }&0 \end{array}} \right)\left( {\begin{array}{{c}} H\\ {{{\tilde{E}}_x}} \end{array}} \right)$$
where ${\tilde{E}_x} = c{\varepsilon _0}{E_x}$ is the x component of electric intensity in the medium.

3. Results and discussions

Before embarking on a detailed analysis of the nonlinear response of the SPR system, we first consider the linear case that the incident light is weak ($I \approx 0$) and then the nonlinearity of the absorbing medium can be ignorable ($\alpha \approx {\alpha _0}$). The excitation of SPR is essential to the formation of OB of reflectance and beam shifts. When the tangential component of incident wave vector in the prism ${k_x} = \sqrt {{\varepsilon _1}} {k_0}\sin {\theta _i}$ matches the wave vector of SPs, SPR can be excited by a p-polarized light of which the energy penetrates into the lower medium by evanescent wave, thereby resulting in attenuated total reflection (ATR) at the resonance angle ${\theta _R} = {42.38^ \circ }$ around which a large change occurs in the reflected phase spectrum as shown by the solid lines in Figs. 2(a) and 2(b). The resonance angle can be expressed as

$${\theta _R} = {\sin ^{ - 1}}\left[ {\frac{1}{{\sqrt {{\varepsilon_1}} }}{\textrm{Re}} \left( {\sqrt {\frac{{{\varepsilon_2}{\varepsilon_3}}}{{{\varepsilon_2} + {\varepsilon_3}}}} } \right)} \right].$$

 figure: Fig. 2.

Fig. 2. (a) Reflectivity and (b) reflected phase shift as a function of the incident angle. The thickness of metal film is ${d_1} = 60\textrm{ nm}$.

Download Full Size | PDF

In the case of small absorption, i.e. ${\alpha _0} < < 1$, ${\theta _R}$ is approximately independent of ${\alpha _0}$. Nevertheless, the absorption coefficient has a significant influence on the minimum reflectivity ${R_{\textrm{min}}}$ at ${\theta _R}$. As ${\alpha _0}$ increases, ${R_{\textrm{min}}}$ is increased and the ATR dip is broadened, thereby reducing the slope of phase variation as shown in Fig. 2(b). Consequently, the GH and IF shifts are closed related to the absorption of the absorbing medium according to Eqs. (9) and (10).

For typical Kretschmann configuration, the reflection properties can be described approximately by the wave vector of the structure, which consists two parts: the eigen wave vector $\beta$ of the SPs and the offset $\Delta \beta$ of eigen wave vector caused by the metal layer with finite thickness ${d_1}$, which are described as [53]

$$\beta = {k_0}\sqrt {\frac{{{\varepsilon _2}{\varepsilon _3}}}{{{\varepsilon _2} + {\varepsilon _3}}}} ,$$
$$\Delta \beta = {k_0}[r_{12}^p(\theta = {\theta _R})]\left( {\frac{2}{{{\varepsilon_3} - {\varepsilon_2}}}} \right){\left( {\frac{{{\varepsilon_2}{\varepsilon_3}}}{{{\varepsilon_2} + {\varepsilon_3}}}} \right)^{{3 / 2}}}\exp \left[ {i\frac{{4\pi {d_1}}}{\lambda }\frac{{{\varepsilon_2}}}{{{{({\varepsilon_2} + {\varepsilon_3})}^{{1 / 2}}}}}} \right].$$
$r_{12}^p(\theta = {\theta _R})$ is the reflective coefficient for light incident at resonance angle which can be calculated by Eq. (7). The system loss is composed of two parts: ${\Gamma _i} = {\mathop{\rm Im}\nolimits} (\beta )$, the intrinsic loss due to the absorption of metal and substrate; ${\Gamma _r} = {\mathop{\rm Im}\nolimits} (\Delta \beta )$, the radiation loss owing to the leakage from the thin metal film to the prism. By using these parameters, the reflectivity can be written as [54]:
$$R({\theta _i}) = 1 - \frac{{4{\Gamma _i}{\Gamma _r}}}{{{\Delta ^2} + {{({\Gamma _i} + {\Gamma _r})}^2}}},$$
where $\Delta $ is a difference between ${k_x}$ and ${\textrm{Re}} (\beta + \Delta \beta )$, the propagation constant of the structure. At the resonant angle ${\theta _R}$, the reflectivity drops to minimum, given by
$${R_{\textrm{min}}} = {\left( {\frac{{{\Gamma _i} - {\Gamma _r}}}{{{\Gamma _i} + {\Gamma _r}}}} \right)^2}.$$

When ${\Gamma _i} = {\Gamma _r}$, the minimum reflectivity goes to 0. In this case, we can obtain the optimal thickness of metal as ${d_{\textrm{opt}}}\textrm{ = }60.03\textrm{ nm}$. According to Eq. (15), the reflection spectrum is approximate to a Lorentzian profile of which the width is determined by the total loss ${\Gamma _i} + {\Gamma _r}$. In the case that $\alpha \ll 1$, the intrinsic loss can be derived from Eq. (13) by Taylor expansion of $\alpha$ (neglecting the high order items), given by

$${\Gamma _i} = {k_0}{\mathop{\rm Im}\nolimits} \left[ {\sqrt {\frac{{{\varepsilon_2}}}{{{\varepsilon_2} + 1}}} \left( {1 + i\frac{\alpha }{2}\frac{{{\varepsilon_2}}}{{{\varepsilon_2} + 1}}} \right)} \right].$$

With the increase of absorption coefficient, ${\Gamma _i}$ increases gradually, thus resulting in the broadening of the ATR dip and the increase of reflectance, which are coincident with the results shown in Fig. 2.

In the following, we consider the nonlinear response of the reflection to the incident light intensity. As shown in Fig. 3, the reflectivity and the GH/IF shifts exhibit hysteretic feature. When the input power increases from zero, the absorption of the substrate as well as the intrinsic loss of SPs is large, thus leading to a high reflectance. The wide ATR dip then results in a small lateral displacement. The transverse displacement is also small due to the low ratio of $|{{{{r_s}} / {{r_p}}}} |$. Near the point B, the saturated absorption effect emerges and is becoming serious. As the input intensity increases to the high threshold ${I_h}$, an abrupt switch from B down to point C takes place. Due to the saturation, a narrow ATR dip with low reflectivity occurs at point C. Accompanying with the switching process of reflectance, the GH and IF shifts jump to the state with large values as result of the steep variation of phase and the enhanced ratio of $|{{{{r_s}} / {{r_p}}}} |$. If the input power is gradually decreased, the absorbing medium remains saturated. The states of low reflectance and large beam shifts can be maintained until a low threshold ${I_l}$ is reached. As the input decreases further, the incident intensity is not large enough to make the absorbing medium saturated and the intrinsic loss increases suddenly. As a result, the reflectance is then switched from point D up to point A, the state with high reflectivity. Simultaneously, the GH and IF shifts jump to the state with small values. Over the range of incident intensity ${I_l} < {I_0} < {I_h}$, the reflectivity as well as the beam shifts has two stable outputs which are switchable by the input power.

 figure: Fig. 3.

Fig. 3. Hysteretic responses of (a) Reflectivity, (b) GH shift and (c) IF shift as a function of the input light intensity. The incident angle is ${\theta _i} = 42.38^\circ$ which corresponds to ${\theta _R}$, the thickness of metal film is ${d_1} = 60\textrm{ nm}$ and the linear absorption coefficient is ${\alpha _0} = 0.015$.

Download Full Size | PDF

In the scheme of OB with the excitation of SPs in Kerr media [25,42,43,46], the system is switched between ATR and TIR due to the movement of the resonance angle. However, in our scheme, the resonance angle remains unchanged [see Fig. 2(a)] and the two stable outputs is owing to the intensity-dependent absorption coefficient. In Fig. 4, we plot the internal electric field profile at the points marked with A, B, C and D in Fig. 3(a). In the process A-B, the internal electric field is weak due to the large absorptive loss of the absorbing medium. As the input power approaches to ${I_h}$, the saturation effect becomes remarkable. Meanwhile, the nonlinear absorption coefficient as well as the intrinsic loss of the medium is reduced, thus enhancing the electric field, which further strengthen the saturation effect. As a result, a positive feedback is established that causes an abrupt enhancement of the electric field from point B to point C. The hysteresis curve jumps from one stable state to the other stable state at the same time. In the process C-D, although the input power is decreasing, the enhanced electric field maintains the absorbing medium be deeply saturated. While the incident intensity is decreased to ${I_l}$, the saturation effect becomes weak and the intrinsic loss increases, then the electric field is reduced gradually from point C to point D. Due to the feedback mechanism, the state of low reflectance can no longer be sustained, thereby leading to the transition from point D to point A where the electric field is small and the absorption is large. The cooperative effect of the intensity-dependent absorption and absorption-dependent electric field gives rise to the inherent feedback and therefore a hysteretic response.

 figure: Fig. 4.

Fig. 4. Distribution of the internal electric field profile in the Kretschmann configuration. ${E_0}$ is the amplitude of the electric field of incident light. The four curves respectively correspond to the points A-D labeled in Fig. 3(a).

Download Full Size | PDF

The thickness of the metal film plays a crucial role in the excitation of SPs. In Fig. 5, we investigate the influence of the thickness ${d_1}$ on the bistable curves of reflectivity, GH shift and IF shift. With the increase of ${d_1}$, both ${I_l}$ and ${I_h}$ of the hysteretic response are increased, but ${I_h}$ increases faster than ${I_l}$. As a result, the region exhibiting two stable states becomes large, i.e. the width of hysteresis loop is broadened. In the Kretschmann configuration, the radiation loss ${\Gamma _r}$, due to the energy leakage from the metal film with finite thickness, decreases exponentially with the increase of ${d_1}$ as indicated by Eq. (14). According to Eq. (16), the minimum reflectivity ${R_{\min }}$ is increased and then the lower branch of hysteresis goes up as illustrated in Fig. 5(a). Meanwhile, the slope of reflected phase shift and the reflectivity ratio $|{{{{r_s}} / {{r_p}}}} |$ are reduced. Therefore, the upper branches of the hysteresis loops for GH/IF shifts are decreased correspondingly [see Figs. 5(b) and 5(c)]. Furthermore, if the thickness is large enough, the phenomenon of OB will disappear and we cannot obtain two stable outputs for any input intensity.

 figure: Fig. 5.

Fig. 5. (a) Reflectivity, (b) GH shift and (c) IF shift versus incident intensity with different thickness of metal film ${d_1}$. The incident angle and the linear absorption coefficient are the same as those in Fig. 3.

Download Full Size | PDF

As increasing the metal film thickness from ${d_1} = 60\textrm{ nm}$, the Kretschmann structure transit from the critical-coupled state to the under-coupled state, so a larger incident intensity is required to excite OB in the structure with a thick metal film. In Fig. 6, we plot the distributions of the internal electric field in the three-layer structure at the points C1, C2 and C3 shown in Fig. 5(a). The inputs at these points correspond to the high thresholds of the hysteresis loops. When ${d_1}$ is close to the optimal thickness ${d_{\textrm{opt}}}\textrm{ = }60.03\textrm{ nm}$, the electric field can be greatly enhanced in the saturable absorbing medium, and therefore the thresholds for the bistable response are small. As ${d_1}$ increases, the electric field strength decreases gradually, and therefore, a higher input power is necessary to make the absorbing medium saturated, resulting in the increase of the threshold ${I_h}$. Similarly, the minimum input power required to maintain deep saturation is smaller for a thin metal film, thus the threshold ${I_l}$ decreases with ${d_1}$.

 figure: Fig. 6.

Fig. 6. Distribution of the internal electric field in the Kretschmann configuration at the three points C1, C2 and C3 shown in Fig. 5(a).

Download Full Size | PDF

In the case of ${d_1} < {d_{\textrm{opt}}}$, the bistable behavior can also be observed as shown in Fig. 7. As ${d_1}$ decreases, both the high and low thresholds are decreased. The width of the hysteresis loop is getting smaller and the OB response will disappear if ${d_1}$ is further decreased. At the high threshold ${I_h}$, the absorption transits from a large value to a small value where ${\Gamma _i}$ is still larger than ${\Gamma _r}$, then the GH shift at ${I_h}$ is negative [see Fig. 7(b)]. As the input intensity increases gradually from ${I_h}$, the absorption coefficient is decreased owing to the saturation effect. At the same time, ${\Gamma _i}$ becomes smaller and gets closed to ${\Gamma _r}$. Consequently, the GH shift is increasing with the input power. However, for a large enough input, ${\Gamma _i}$ can be smaller than ${\Gamma _r}$ and a positive lateral shift is then obtained [see the tails of the dashed and dotted lines in Fig. 7(b)]. The bistable response of IF shift exhibits similar tendency and the transverse displacement turns from positive to negative for a large input as shown in Fig. 7(c). Therefore, the phenomena of bistable beam shifts can be found in both the cases ${d_1} > {d_{\textrm{opt}}}$ and ${d_1} < {d_{\textrm{opt}}}$. Nevertheless, the hysteretic behaviors are more complex for the case ${d_1} < {d_{\textrm{opt}}}$.

 figure: Fig. 7.

Fig. 7. (a) Reflectivity, (b) GH shift and (c) IF shift versus incident intensity with different thickness of metal film ${d_1}$. The incident angle and the linear absorption coefficient are the same as those in Fig. 3.

Download Full Size | PDF

The effect of linear absorption coefficient ${\alpha _0}$ on the bistable reflectance, GH shift and IF shift is shown in Fig. 8. According to the previous analysis, the minimum reflectivity increases with ${\alpha _0}$ as a result of the increasing intrinsic loss ${\Gamma _r}$. Therefore, the initial reflectivity in the case of small signal ($I \approx 0$) is lower for small ${\alpha _0}$ [see Fig. 8(a)]. As ${\alpha _0}$ increases, the thresholds and the width of hysteretic loop increase accordingly. At the points C1, C2 and C3, the reflectance is smaller for large ${\alpha _0}$, which corresponds to a narrower ATR spectrum. With an input intensity a little beyond the switch-up threshold ${I_h}$, both the lateral and the transverse displacements are increased with ${\alpha _0}$ owing to the increasing angular variation of phase ${{\textrm{d}{\varphi _p}} / {\textrm{d}{\theta _i}}}$ and reflection ratio $|{{{{r_s}} / {{r_p}}}} |$. The change of the bistable curves is related to the electric field in the saturable absorbing medium. The distributions of the internal electric field at the points C1, C2 and C3 in Fig. 8(a) are plotted in Fig. 9. Although the electric field in the absorbing medium increases with ${\alpha _0}$, a large input intensity is still essential to achieve deep saturation, thus the hysteretic loop moves right, the direction of increasing incident intensity.

 figure: Fig. 8.

Fig. 8. (a) Reflectivity, (b) GH shift and (c) IF shift versus incident intensity with different linear absorption coefficient of absorbing medium ${\alpha _0}$. The incident angle and the thickness of metal film are the same as those in Fig. 3.

Download Full Size | PDF

 figure: Fig. 9.

Fig. 9. Distribution of the internal electric field in the Kretschmann configuration at the three points C1, C2 and C3 shown in Fig. 8(a).

Download Full Size | PDF

The thresholds of the hysteresis loop are related to the saturable intensity of the medium. The saturation effect can be explained by a two-level system, of which the saturable intensity is ${I_s} = {{{\hbar ^2}} / {({{\mu^2}{T_1}{T_2}} )}}$, where $\mu$ is the electric dipole moment of the transition, ${T_1} = {1 / {{\gamma _\parallel }}}$ represents the decay time of the population from the upper level to the lower level, and ${T_2} = {1 / {{\gamma _ \bot }}}$ denotes the decoherence time of the transition. ${\gamma _\parallel }$ and ${\gamma _ \bot }$ are the decay rate of population and the dephasing rate of the transition, respectively. ${\gamma _ \bot }$ is also the linewidth of the absorption spectrum. Therefore, the medium with a narrow absorption linewidth corresponds to a low saturable intensity and low thresholds of OB. For example, the saturable intensity is about $2.5\textrm{ mW/c}{\textrm{m}^2}$ for the D2 line of cold rubidium atoms of which the natural linewidth is ${\gamma _ \bot } = 2\pi \times 6\textrm{ MHz}$. However, for a homogeneously broadened media, the linewidth, mainly determined by the collision process between gas atoms or the lattice vibration of solid crystals, is of the order of GHz or even THz, then the saturable intensity is of the order of W/cm2 or kW/cm2. Moreover, the response time of OB with the excitation of SPs, a critical parameter for the operation of optical switching, is mainly determined by the nonlinear material and the structure [54]. For the saturable absorbing media, the response time is several times of ${T_2}$, which is of the order of $\mathrm{\mu s}$ for cold gas media, and $\textrm{ps - ns}$ for hot gas media and solid materials. If the response time of nonlinear absorption is short enough (sub-ps or fs), the switching time also depends on the structure of which the switching time can be sub-ps for OB with the excitation of SPs [55].

4. Conclusions

In summary, the nonlinear reflection of a light beam is explored in a Kretschmann configuration containing saturable absorbing medium. With the excitation of SPs, the reflectivity and the angular variation of reflected phase at resonance depend strongly on the intrinsic loss of the system, then the lateral and transverse displacements of the reflected beam are closely related to the absorption coefficient of medium. Owing to the intensity-dependent saturated absorption and the inherent feedback effect, the reflectance and the GH/IF shifts exhibit nonlinear hysteretic behavior. We also study the factors that affect the bistability of reflection. The thickness of metal film ${d_1}$ has influence on the radiation loss of the structure, which then leads to the change of switching thresholds as well as the width of hysteresis loop. The phenomenon of OB is easy to occur when ${d_1}$ is around the optimal thickness ${d_{\textrm{opt}}}$. The linear absorption coefficient ${\alpha _0}$ also affect the thresholds and the shape of bistable curve. Large ${\alpha _0}$ leads to the increase of thresholds, while the phenomenon of OB will disappear for a small ${\alpha _0}$.

It is worthy to make a comparison between our scheme and the previous schemes of OB with the excitation of SPs in Kerr nonlinear medium [25,42,43,46]. For the Kretschmann structure with Kerr medium, the system is switched between two modes, attenuated total reflection and totally internal reflection, by tuning the resonance angle via power-dependent refractive index of Kerr medium. Hence OB is dispersive-type in these investigations. However, the resonance angle in our scheme remains unchanged, and an absorptive OB is obtained. The two stable states of system results from the intensity-dependent intrinsic loss which is controlled by the saturation effect of absorbing medium. The results may provide an alternative avenue for controlling light by light.

Funding

National Natural Science Foundation of China (11204367); Natural Science Foundation of Shaanxi Province (2018JQ1051).

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. F. Goos and H. Hänchen, “Ein neuer und fundamentaler Versuch zur Totalreflexion,” Ann. Phys. 436(7-8), 333–346 (1947). [CrossRef]  

2. K. Artmann, “Berechnung der Seitenversetzung des totalreflektierten Strahles,” Ann. Phys. 437(1-2), 87–102 (1948). [CrossRef]  

3. F. I. Fedorov, “On the theory of total internal reflection,” Dokl. Akad. Nauk SSSR 105, 465–468 (1955).

4. C. Imbert, “Calculation and experimental proof of the transverse shift induced by total internal reflection of a circularly polarized light beam,” Phys. Rev. D 5(4), 787–796 (1972). [CrossRef]  

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

6. K. Y. Bliokh and Y. P. Bliokh, “Conservation of Angular Momentum, Transverse Shift, and Spin Hall Effect in Reflection and Refraction of Electromagnetic Wave Packet,” Phys. Rev. Lett. 96(7), 073903 (2006). [CrossRef]  

7. O. Hosten and P. Kwiat, “Observation of the Spin Hall Effect of Light via Weak Measurements,” Science 319(5864), 787–790 (2008). [CrossRef]  

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

9. Y. H. Wan, Z. Zheng, W. J. Kong, X. Zhao, and J. S. Liu, “Fiber-to-fiber optical switching based on gigantic Bloch-surface-wave-induced Goos-Hänchen shifts,” IEEE Photonics J. 5(1), 7200107 (2013). [CrossRef]  

10. Y. Wang, H. Li, Z. Cao, T. Yu, Q. Shen, and Y. He, “Oscillating wave sensor based on the Goos-Hänchen effect,” Appl. Phys. Lett. 92(6), 061117 (2008). [CrossRef]  

11. T. Yu, H. Li, Z. Cao, Y. Wang, Q. Shen, and Y. He, “Oscillating wave displacement sensor using the enhanced Goos-Hänchen effect in a symmetrical metal-cladding optical waveguide,” Opt. Lett. 33(9), 1001–1003 (2008). [CrossRef]  

12. T. Shui, W. X. Yang, Q. Y. Zhang, X. Liu, and L. Li, “Squeezing-induced giant Goos-Hänchen shift and hypersensitized displacement sensor in a two-level atomic system,” Phys. Rev. A 99(1), 013806 (2019). [CrossRef]  

13. X. X. Zhou, X. H. Ling, H. L. Luo, and S. C. Wen, “Identifying graphene layers via spin Hall effect of light,” Nat. Neurosci. 101(25), 251602 (2012). [CrossRef]  

14. T. Zhu, Y. Lou, Y. Zhou, J. Zhang, J. Huang, Y. Li, H. Luo, S. Wen, S. Zhu, Q. Gong, M. Qiu, and Z. Ruan, “Generalized Spatial Differentiation from the Spin Hall Effect of Light and Its Application in Image Processing of Edge Detection,” Phys. Rev. Appl. 11(3), 034043 (2019). [CrossRef]  

15. X. Jiang, Q. K. Wang, J. Guo, S. Q. Chen, X. Y. Dai, and Y. J. Xiang, “Enhanced Photonic Spin Hall Effect with a Bimetallic Film Surface Plasmon Resonance,” Plasmonics 13(4), 1467–1473 (2018). [CrossRef]  

16. X. X. Zhou, L. J. Sheng, and X. H. Ling, “Photonic spin Hall effect enabled refractive index sensor using weak measurements,” Sci. Rep. 8(1), 1221 (2018). [CrossRef]  

17. C. K. Liang, G. H. Wang, D. M. Deng, and T. T. Zhang, “Controllable refractive index sensing and multi-functional detecting based on the spin Hall effect of light,” Opt. Express 29(18), 29481–29491 (2021). [CrossRef]  

18. L. G. Wang, H. Chen, and S. Y. Zhu, “Large negative Goos-Hänchen shift from a weakly absorbing dielectric slab,” Opt. Lett. 30(21), 2936–2938 (2005). [CrossRef]  

19. W. G. Zhu and W. L. She, “Enhanced spin Hall effect of transmitted light through a thin epsilon-near-zero slab,” Opt. Lett. 40(13), 2961–2964 (2015). [CrossRef]  

20. M. J. Jiang, W. G. Zhu, H. Y. Guan, J. H. Yu, H. H. Lu, J. Y. 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), 3259–3262 (2017). [CrossRef]  

21. L. Y. Jiang, Q. K. Wang, Y. J. Xiang, X. Y. Dai, and S. C. Wen, “Electrically tunable Goos-Hänchen shift of light beam reflected from a graphene-on-dielectric surface,” IEEE Photonics J. 5(3), 6500108 (2013). [CrossRef]  

22. W. J. M. Kort-Kamp, N. A. Sinitsyn, and D. A. R. Dalvit, “Quantized beam shifts in graphene,” Phys. Rev. B 93(8), 081410 (2016). [CrossRef]  

23. L. Cai, M. X. Liu, S. Z. Chen, Y. C. Liu, W. X. Shu, H. L. Luo, and S. C. Wen, “Quantized photonic spin Hall effect in grapheme,” Phys. Rev. A 95(1), 013809 (2017). [CrossRef]  

24. X. B. Yin, L. Hesselink, Z. W. Liu, N. Fang, and X. Zhang, “Large positive and negative lateral optical beam displacements due to surface plasmon resonance,” Appl. Phys. Lett. 85(3), 372–374 (2004). [CrossRef]  

25. H. C. Zhou, X. Chen, P. Hou, and C. F. Li, “Giant bistable lateral shift owing to surface-plasmon excitation in Kretschmann configuration with a Kerr nonlinear dielectric,” Opt. Lett. 33(11), 1249–1251 (2008). [CrossRef]  

26. R. G. Wan and M. S. Zubairy, “Tunable and enhanced Goos-Hänchen shift via surface plasmon resonance assisted by a coherent medium,” Opt. Express 28(5), 6036–6047 (2020). [CrossRef]  

27. L. Salasnich, “Enhancement of four reflection shifts by a three-layer surface-plasmon resonance,” Phys. Rev. A 86(5), 055801 (2012). [CrossRef]  

28. X. X. Zhou and X. H. Ling, “Enhanced photonic spin Hall effect due to surface plasmon resonance,” IEEE Photonics J. 8(1), 1–8 (2016). [CrossRef]  

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

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

31. L. Chen, Z. Q. Cao, F. Ou, H. G. Li, Q. H. Shen, and H. C. Qiao, “Observation of large positive and negative lateral shifts of a reflected beam from symmetrical metal-cladding waveguides,” Opt. Lett. 32(11), 1432–1434 (2007). [CrossRef]  

32. H. L. Luo, S. C. Wen, W. X. Shu, and D. Y. Fan, “Spin Hall effect of light in photon tunneling,” Biomed. Opt. Express 82(4), 043825 (2010). [CrossRef]  

33. B. Wang, Y. Li, M. M. Pan, J. Li Ren, Y. F. Xiao, H. Yang, and Q. H. Gong, “Spin displacements of a Gaussian beam at an air-multilayer-film interface,” Phys. Rev. A 88(4), 043842 (2013). [CrossRef]  

34. T. T. Tang, J. Li, Y. F. Zhang, C. Y. Li, and L. Luo, “Spin Hall effect of transmitted light in a three-layer waveguide with lossy epsilon-near-zero metamaterial,” Opt. Express 24(24), 28113–28121 (2016). [CrossRef]  

35. L. G. Wang, M. Ikram, and M. S. Zubairy, “Control of the Goos-Hänchen shift of a light beam via a coherent driving field,” Phys. Rev. A 77(2), 023811 (2008). [CrossRef]  

36. Ziauddin, S. Qamar, and M. S. Zubairy, “Coherent control of the Goos-Hänchen shift,” Acta Neuropathol. 81(2), 023821 (2010). [CrossRef]  

37. W. X. Yang, S. P. Liu, Z. H. Zhu, R. K. Ziauddin, and Lee, “Tunneling-induced giant Goos-Hänchen shift in quantum wells,” Opt. Lett. 40(13), 3133–3136 (2015). [CrossRef]  

38. S. Asiri, J. Xu, M. Al-Amri, and M. S. Zubairy, “Controlling the Goos-Hänchen and Imbert-Fedorov shifts via pump and driving fields,” Phys. Rev. A 93(1), 013821 (2016). [CrossRef]  

39. X. J. Zhang, H. H. Wang, C. Z. Liu, G. W. Zhang, L. Wang, and J. H. Wu, “Controlling transverse shift of the reflected light via high refractive index with zero absorption,” Opt. Express 25(9), 10335–10344 (2017). [CrossRef]  

40. R. G. Wan and M. S. Zubairy, “Controlling photonic spin Hall effect based on tunable surface plasmon resonance with an N type coherent medium,” PLoS Biol. 101(3), 033837 (2020). [CrossRef]  

41. A. E. Kaplan, “Theory of hysteresis reflection and refraction of light by a boundary of a nonlinear medium,” Photoacoustics 45, 896–905 (1977).

42. G. Wysin, J. Simon, and R. Deck, “Optical bistability with surface plasmons,” Opt. Lett. 6(1), 30–32 (1981). [CrossRef]  

43. R. Hickernell and D. Sarid, “Optical bistability using prism-coupled, long-range surface plasmons,” J. Opt. Soc. Am. B 3(8), 1059–1068 (1986). [CrossRef]  

44. V. Montemayor and R. Deck, “Optical bistability with the waveguide mode,” J. Opt. Soc. Am. B 2(6), 1010–1013 (1985). [CrossRef]  

45. P. Dannberg and E. Broese, “Observation of optical bistability by prism excitation of a nonlinear film-guided wave,” Appl. Opt. 27(8), 1612–1614 (1988). [CrossRef]  

46. Y. J. Xiang, X. Y. Dai, J. Guo, S. C. Wen, and D. Y. Tang, “Tunable optical bistability at the graphene-covered nonlinear interface,” Appl. Phys. Lett. 104(5), 051108 (2014). [CrossRef]  

47. X. Y. Dai, L. Y. Jiang, and Y. J. Xiang, “Low threshold optical bistability at terahertz frequencies with graphene surface plasmons,” Sci. Rep. 5(1), 12271 (2015). [CrossRef]  

48. K. J. Ahn and F. Rotermund, “Terahertz optical bistability of graphene in thin layers of dielectrics,” Opt. Express 25(8), 8484–8490 (2017). [CrossRef]  

49. A. Kar, N. Goswami, and A. Saha, “Tunable low-threshold bistable Goos-Hänchen shift and Imbert-Fedorov shift using long-range graphene surface plasmons within the terahertz region,” Appl. Opt. 58(34), 9376–9383 (2019). [CrossRef]  

50. S. Kim and K. Kim, “Giant enhancement of reflectance due to the interplay between surface confined wave modes and nonlinear gain in dielectric media,” Opt. Express 25 318161 (2017). [CrossRef]  

51. H. C. Zhou, J. Guo, K. Xu, Z. Li, J. Q. Tang, and S. Q. Man, “Bistable enhanced total reflection in Kretschmann configuration containing a saturable gain medium,” Opt. Express 26(5), 5253–5264 (2018). [CrossRef]  

52. Y. Qin, Y. Li, H. He, and Q. Gong, “Measurement of spin hall effect of reflected light,” Opt. Lett. 34(17), 2551–2553 (2009). [CrossRef]  

53. K. Kurihara and K. Suzuki, “Theoretical understanding of an absorption-based surface plasmon resonance sensor based on Kretchmann’s theory,” Anal. Chem. 74(3), 696–701 (2002). [CrossRef]  

54. S. Herminghaus, M. Klopfleisch, and H. J. Schmidt, “Attenuated total reflectance as a quantum interference phenomenon,” Opt. Lett. 19(4), 293–295 (1994). [CrossRef]  

55. C. Min, P. Wang, C. Chen, Y. Deng, Y. Lu, H. Ming, T. Ning, Y. Zhou, and G. Yang, “All-optical switching in subwavelength metallic grating structure containing nonlinear optical materials,” Opt. Lett. 33(8), 869–871 (2008). [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 (9)

Fig. 1.
Fig. 1. Schematic diagram of Goos-Hänchen and Imbert-Fedorov effects in a three-layer Kretschmann configuration composed of a prism, a thin metal film backed by a saturable absorbing medium.
Fig. 2.
Fig. 2. (a) Reflectivity and (b) reflected phase shift as a function of the incident angle. The thickness of metal film is ${d_1} = 60\textrm{ nm}$.
Fig. 3.
Fig. 3. Hysteretic responses of (a) Reflectivity, (b) GH shift and (c) IF shift as a function of the input light intensity. The incident angle is ${\theta _i} = 42.38^\circ$ which corresponds to ${\theta _R}$, the thickness of metal film is ${d_1} = 60\textrm{ nm}$ and the linear absorption coefficient is ${\alpha _0} = 0.015$.
Fig. 4.
Fig. 4. Distribution of the internal electric field profile in the Kretschmann configuration. ${E_0}$ is the amplitude of the electric field of incident light. The four curves respectively correspond to the points A-D labeled in Fig. 3(a).
Fig. 5.
Fig. 5. (a) Reflectivity, (b) GH shift and (c) IF shift versus incident intensity with different thickness of metal film ${d_1}$. The incident angle and the linear absorption coefficient are the same as those in Fig. 3.
Fig. 6.
Fig. 6. Distribution of the internal electric field in the Kretschmann configuration at the three points C1, C2 and C3 shown in Fig. 5(a).
Fig. 7.
Fig. 7. (a) Reflectivity, (b) GH shift and (c) IF shift versus incident intensity with different thickness of metal film ${d_1}$. The incident angle and the linear absorption coefficient are the same as those in Fig. 3.
Fig. 8.
Fig. 8. (a) Reflectivity, (b) GH shift and (c) IF shift versus incident intensity with different linear absorption coefficient of absorbing medium ${\alpha _0}$. The incident angle and the thickness of metal film are the same as those in Fig. 3.
Fig. 9.
Fig. 9. Distribution of the internal electric field in the Kretschmann configuration at the three points C1, C2 and C3 shown in Fig. 8(a).

Equations (17)

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

ε 3 = 1 + i α ( I 3 )
α ( I 3 ) = α 0 1 + I 3 I s ,
I ~ 3 1 2 I 3 d = c μ 0 4 n r | H d | 2 ,
I = c μ 0 2 ε 1 | H d t | 2 ,
r f = r 12 f + r 23 f e 2 i k 2 z d 1 + r 12 f r 23 f e 2 i k 2 z d ,   f = p , s
t f = t 12 f t 23 f e i k 2 z d 1 + r 12 f r 23 f e 2 i k 2 z d ,   f = p , s
r i j f = { k i z / ε i k j z / ε j k i z / ε i + k j z / ε j  for  f = p , k i z k j z k i z + k j z  for  f = s ,
t i j f = 1 + r i j f .
S GH = 1 ε 1 k 0 d φ p d θ i .
S IF = 1 ε 1 k 0 [ 1 + | r s r p | cos ( φ s φ p ) ] cot θ i
d d z ( H E ~ x ) = i k ( 0 ε 3 1 ε 1 ε 3 sin 2 θ 0 ) ( H E ~ x )
θ R = sin 1 [ 1 ε 1 Re ( ε 2 ε 3 ε 2 + ε 3 ) ] .
β = k 0 ε 2 ε 3 ε 2 + ε 3 ,
Δ β = k 0 [ r 12 p ( θ = θ R ) ] ( 2 ε 3 ε 2 ) ( ε 2 ε 3 ε 2 + ε 3 ) 3 / 2 exp [ i 4 π d 1 λ ε 2 ( ε 2 + ε 3 ) 1 / 2 ] .
R ( θ i ) = 1 4 Γ i Γ r Δ 2 + ( Γ i + Γ r ) 2 ,
R min = ( Γ i Γ r Γ i + Γ r ) 2 .
Γ i = k 0 Im [ ε 2 ε 2 + 1 ( 1 + i α 2 ε 2 ε 2 + 1 ) ] .
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.