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

Low threshold optical bistability based on topological edge state in photonic crystal heterostructure with Dirac semimetal

Open Access Open Access

Abstract

The special band structure of three-dimensional Dirac semimetal (3D DSM) makes it show strong nonlinear optical characteristics in the terahertz region, which provides a new way to develop terahertz nonlinear devices with low threshold. In this paper, we theoretically study the optical bistability (OB) of transmitted light in a multilayer structure with 3D DSM embedded in two one-dimensional photonic crystals (1D PhC). The topological edge state (TES) excited by the 1D PhC heterostructure significantly enhances the local electric field near the nonlinear 3D DSM, which provides a positive condition for the realization of low threshold OB. Through parameter optimization, we obtain a threshold electric field with an incident electric field of 106 V/m levels. Furthermore, the influences of the Fermi energy and thickness of 3D DSM and the angle of the incident light on the hysteretic behavior as well as the threshold of OB are clarified. 3D DSM-based optical devices with intrinsic OB provide a building block for future integrated optical and all-optical networks.

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

1. Introduction

Optical bistability (OB) refers to an optical phenomenon that produces two different stable state outputs for a given input. The two output states can interconvert with each other, showing an obvious hysteresis loop relationship [1]. As a typical optical nonlinear phenomenon, OB has potential applications in optical devices such as all-optical switches [2,3], optical transistors [4], optical storage [5], optical logical gates [6,7], etc., which has attracted extensive attention. However, the traditional optical bistable devices are limited by the small nonlinear coefficient of nonlinear materials, which makes the size larger, and the obvious OB phenomenon can be realized under the condition of strong incident light, which hinders the practicability of optical bistable devices. In recent years, with the development of micro-nano technology, researchers began to pay attention to OB phenomenon in micro-nano structures, such as Fabry-Perot (F-P) cavities [8], photonic crystals [9], metamaterials [10], optical ring cavities [11], etc. We know that the nonlinear refractive index of materials has an important influence on the threshold of OB. The application of materials with higher nonlinear refractive index can effectively reduce the threshold of OB. However, because the nonlinear coefficients of most materials are relatively low, it is difficult to achieve a lower threshold. In recent years, due to the excellent nonlinear characteristics of two-dimensional material graphene, the low threshold tunable OB based on graphene has been widely studied. For example, OB in photonic crystal/graphene/photonic crystal hybrid structure [12] and OB in graphene/waveguide hybrid structure [13] have been reported. The research of nonlinear graphene in the fields of OB and all-optical switch has made rapid development, but its two-dimensional properties have caused a bottleneck in large-area and high-quality preparation, which slows down the research progress of all-optical devices based on graphene [14]. Recently, the three-dimensional Dirac semimetal (3D DSM), a new strange topological quantum material, has attracted increasing attention. Its body electrons form a Dirac cone structure in three-dimensional space, so it is also called “3D graphene” [15,16]. Specifically, 3D DSM has high nonlinear refractive index, which is very similar to that of graphene. However, compared with graphene, 3D DSM has higher carrier mobility [17], higher Fermi velocity [18], easier preparation and stable performance [19]. Meanwhile, 3D DSM can also adjust the dielectric constant by changing the Fermi energy [20]. Therefore, the research of micro-nano optoelectronic devices based on 3D DSM has a wide application prospect. At present, many complexes have been proved to have 3D DSM properties, such as $\textrm{C}{\textrm{d}_\textrm{3}}\textrm{A}{\textrm{s}_\textrm{2}}$ [21], $\textrm{N}{\textrm{a}_\textrm{3}}\textrm{Bi}$ [22], $\textrm{ZrT}{\textrm{e}_\textrm{5}}$ [23], $\textrm{AlCuFe}$ [24], etc. The high nonlinear refractive index of 3D DSM is expected to play a crucial part in providing strong optical nonlinearity to realize low threshold OB.

In recent years, the topological edge state mode (TES) in photonic crystals has been widely concerned in micro-nano optoelectronic devices because of its unique topological protection and local field enhancement characteristics [25,26]. At present, researchers have proposed various structures and schemes theoretically or experimentally to construct photonic TES, such as plasmonic nanoparticles [27], optical waveguides [28], chiral hyperbolic metamaterials [29], optical ring resonators [30], two-dimensional photonic crystals (2D PhC) and three-dimensional photonic crystals (3D PhC). Although the TES of 2D and 3D PhC can enhance the local electric field of the structure, the application of TES of multi-dimensional topological PhC in integrated photonic devices is limited due to its complex design and manufacturing process. Compared with 2D and 3D PhC, one-dimensional photonic crystals (1D PhC) are easier to manufacture, and the expected local field enhancement can also be found at the interface of 1D topological PhC heterostructure [31]. This means that the emergence of TES can effectively reduce the threshold of OB. Therefore, we can consider such a question: Under the condition of TES excitation, whether OB with low threshold can be generated in 1D topological PhC based on 3D DSM, and whether the hysteresis loop of OB can be dynamically regulated by adjusting the parameters of 3D DSM.

In this paper, we study low threshold OB of transmitted light in a multilayer structure with 3D DSM embedded in two 1D PhCs. The results show that TES will be generated at the interface connecting two 1D PhCs, which significantly enhances the local electric field. Furthermore, inserting 3D DSM into the overall structure provides nonlinear conditions for the generation of OB. At the same time, the threshold of OB can be dynamically adjusted by changing the Fermi energy and relaxation time of 3D DSM. A low threshold OB scheme based on 3D DSM is proposed for the first time, which has the characteristics of simple structure and easy preparation. We believe that the low threshold OB scheme is expected to play an important role in the preparation of all-optical nonlinear photonic devices.

2. Theoretical model and method

We design a multilayered hybrid structure consisting of two 1D PhCs and a layer of 3D DSM. In this composite structure, 3D DSM is embedded between PhC 1 and PhC 2, as shown in Fig. 1. The two PhCs are composed of alternating medium A (TPX) and medium B (TiO2), and the period of PhC is defined as $\textrm{N} = \textrm{4}$. The refractive index of medium A is set to 1.46, that is ${n_{A1}} = {n_{A2}} = {n_A} = 1.46$; the refractive index of medium B is set to 2.82, that is ${n_{B1}} = {n_{B2}} = {n_B} = 2.82$ [12,32]. For PhC 1, the thicknesses of medium A and medium B meet ${d_{A1}} = 270\;{\mathrm{\mu} \mathrm{m}}$ and ${d_{B\textrm{1}}} = 121\;{\mathrm{\mu} \mathrm{m}}$ respectively, and for PhC 2, the thicknesses of medium A and medium B meet ${d_{A2}} = 275\;{\mathrm{\mu} \mathrm{m}}$ and ${d_{B2}} = 170\;{\mathrm{\mu} \mathrm{m}}$ respectively. The thickness of 3D DSM is set to ${d_{Dir}} = 20\;\textrm{nm}$. Based on current mature preparation technology of micro-nano multilayer dielectric structure, the layered structure defined by the above structural parameters is not difficult to construct. In addition, 3D DSM as a typical kerr nonlinear medium, here we use the third-order nonlinear susceptibility and linear refractive index to represent it. In order to obtain a larger nonlinear coefficient, we only consider the case of the terahertz region. According to Ref. [33], the linear intraband optical conductivity of 3D DSM can be resolved as:

$${\sigma ^{(1)}} = {\sigma _0}\frac{4}{{3{\pi ^2}}}\frac{\tau }{{1 - i\omega \tau }}\frac{{{{({{k_B}T} )}^2}}}{{{\hbar ^2}{v_F}}}\left[ {2L{i_2}\left( { - {e^{ - \frac{{{E_F}}}{{{k_B}T}}}}} \right) + {{\left( {\frac{{{E_F}}}{{{k_B}T}}} \right)}^2} + \frac{{{\pi^2}}}{3}} \right],$$
where ${\sigma _\textrm{0}} \equiv {{{e^2}} / {4\hbar }}$, $\omega$ is the angular frequency of the incident beam and ${k_B}$ represents the Boltzmann constant; T represents the temperature of 3D DSM, $\hbar$ represents the reduced Planck’s constant, and ${v_F}$ represents the Fermi velocity of electrons;$\tau$ and ${E_F}$ stand for the relaxation time and the Fermi energy of the 3D DSM, respectively. And $L{i_s}(z )$ is the polylogarithm. The third-order nonlinear optical conductivities of 3D DSM can be expressed as [33]:
$${\sigma ^{(3 )}} = {\sigma _0}\frac{{8{e^2}{v_F}}}{{5{\pi ^2}{\hbar ^2}}}\frac{{{\tau ^3}}}{{({1 + {\omega^2}{\tau^2}} )({1 - 2i\omega \tau } )}}\frac{1}{{1 + \textrm{exp} ({ - {{{E_F}} / {{k_B}T}}} )}}.$$

 figure: Fig. 1.

Fig. 1. Schematic diagram of TES based on 3D DSM/PhC heterostructure: (a) visual view and (b) side view.

Download Full Size | PDF

Thus, we can obtain the third-order nonlinear susceptibility ${\chi ^{(3)}}$ and linear refractive index ${n_D}$ of 3D DSM:

$$\left\{ {\begin{array}{*{20}{c}} {{\chi^{(3 )}} = i{\sigma^{(3 )}}/{\varepsilon_0}\omega }\\ {{n_D} = n + ik = \sqrt {1 + i{\sigma^{(1)}}/{\varepsilon_0}\omega } } \end{array}} \right.,$$
where ${\varepsilon _0}$ denotes the vacuum permittivity. It can be seen from the above formula that the third-order nonlinear susceptibility ${\chi ^{(3)}}$ and linear refractive index ${n_D}$ of 3D DSM are greatly affected by its Fermi energy and relaxation time, which provides us with an effective method to regulate optical bistable devices.

The transfer matrix method is used to evaluate the incident light intensity and its relationships to the transmitted light intensity and reflected light intensity. In this paper, we calculate the transmittance and reflectivity of the structure under TE polarization (TM can be obtained similarly). We assume that the 3D DSM is parallel to the plane of the x-axis and y-axis, and the electromagnetic field propagates along the z-axis. For linear medium A and medium B, we only need to know the thickness and refractive index of each layer to obtain the transfer matrix representing each layer of material [34]. As for the transfer matrix of the Dirac layer, considering that it is a Kerr type nonlinear material, its transfer matrix can be expressed as [35,36]:

$${M_D} = \frac{{{k_0}}}{{{k_{z + }} + {k_{z - }}}}\left[ {\begin{array}{*{20}{c}} {\frac{{{k_{z - }}}}{{{k_0}}}\textrm{exp} \left( { - i{k_{z + }}{d_D}} \right) + \frac{{{k_{z + }}}}{{{k_0}}}\textrm{exp} \left( {i{k_{z - }}{d_D}} \right)}&{\textrm{exp} \left( { - i{k_{z + }}{d_D}} \right) - \textrm{exp} \left( {i{k_{z - }}{d_D}} \right)}\\ {\frac{{{k_{z - }}{k_{z + }}}}{{k_0^2}}\textrm{exp} \left( { - i{k_{z + }}{d_D}} \right) - \textrm{exp} \left( {i{k_{z - }}{d_D}} \right)}&{\frac{{{k_{z + }}}}{{{k_0}}}\textrm{exp} \left( { - i{k_{z + }}{d_D}} \right) + \frac{{{k_{z - }}}}{{{k_0}}}\textrm{exp} \left( {i{k_{z - }}{d_D}} \right)} \end{array}} \right],$$
where ${k_0}$ refers to the wave vector in vacuum. ${k_{z \pm }}$ are the components of the propagating wave vector in the z direction, which can be expressed as
$${k_{z \pm }} = \sqrt {{{({{k_0}{n_D}} )}^2} - {{({{k_y}} )}^2}} {({1 + {U_ \pm } + 2{U_ \mp }} )^{\frac{1}{2}}},$$
with ${U_ \pm } = \frac{{{k_0}^2{\chi ^{(3)}}}}{{n_D^2k_0^2 - k_y^2}}{|{{A_ \pm }} |^2}$. Here ${A_ \pm }$ are the amplitudes of the propagating waves, which can be given by ${A_ \pm } = \left[ {{E_D}\frac{{{k_{z \mp }}}}{{{k_0}}} \pm {H_D}} \right]\frac{{{k_0}}}{{{k_{z + }} + {k_{z - }}}}$. It can be seen that the value of the wave vector ${k_{z \pm }}$ depends on ${U_ \pm }$. Therefore, to obtain the nonlinear transfer matrix ${M_D}$, we need to obtain a set of coupled nonlinear equations about ${U_ \pm }$ through fixed point iteration [37]. In order to obtain a stable solution, we take ${U_ \pm } = 0$ as the initial value and iterate multiple times in the calculation process. On this basis, it is not difficult for us to obtain the transfer matrix of the overall structure as follows:
$$M = {({{M_{B1}} \times {M_{A1}}} )^4} \times {M_D} \times {({{M_{B2}} \times {M_{A2}}} )^4}.$$

Therefore the transmission coefficient and reflection coefficient can be expressed by

$$\left\{ {\begin{array}{*{20}{c}} {t = \frac{{2{p_f}}}{{[{{M_{11}} + {M_{12}}{p_f}} ]{p_f} + [{{M_{21}} + {M_{22}}{p_f}} ]}}}\\ {r = \frac{{[{{M_{11}} + {M_{12}}{p_f}} ]{p_f} - [{{M_{21}} + {M_{22}}{p_f}} ]}}{{[{{M_{11}} + {M_{12}}{p_f}} ]{p_f} + [{{M_{21}} + {M_{22}}{p_f}} ]}}} \end{array}} \right.,$$
where ${p_f} = {({k_0^2 - k_y^2} )^{\frac{1}{2}}}/{k_0}$ and ${M_{ij}}$ are the elements of matrix M. Finally, we can get the relationship between the incident electric field ${E_{in}}$ and the transmitted electric field ${E_{out}}$ as well as that between ${E_{in}}$ and the reflected electric field ${E_{re}}$.

3. Results and discussions

We know that when the bandgaps of two PhCs share a common frequency range and the topological properties of the bandgaps are opposite, TES can be excited at the interface of PhC heterostructure [38]. In this part, we first discuss the conditions for the excitation of TES based on the proposed structure. Figure 2(a) and Fig. 2(b) show the relationship between PhC transmittance and frequency and that between PhC reflectivity and frequency, respectively, in which the red solid line and black dotted line represent the transmittance/reflectivity of PhC 1 and PhC 2 respectively. The results show that in the 0.97 ∼ 1.07 THz frequency range, both PhC 1 and PhC 2 show an obvious broad bandgap with nearly zero transmittance, and the bandgap of PhC 2 is slightly wider than that of PhC 1. Next, we continue to discuss the band structure of PhC. The dispersion relation of 1D PhC can be expressed by the following formula [39]:

$$\cos ({q\Lambda } )= \cos {k_a}{d_a}\cos {k_b}{d_b} - \frac{1}{2}\left( {\frac{{{z_a}}}{{{z_b}}} + \frac{{{z_b}}}{{{z_a}}}} \right)\sin {k_a}{d_a}\sin {k_b}{d_b},$$
where ${k_i} = {{\omega {n_i}} / c}$, ${z_i} = \sqrt {{{{\mu _i}} / {{\varepsilon _i}}}}$, ${\varepsilon _i} = {{n_i^2} / {{\mu _i}}}$, and $i = a\;\textrm{or}\;b$;$q$ and $\Lambda \textrm{ = }{d_a} + {d_b}$ are the bloch wave vector and the unit cell, respectively. We use COMSOL multiphysics to draw the band diagrams of PhC 1 and PhC 2, and use the method proposed in Ref. [38] to calculate the Zak phase ($\textrm{0}\;\textrm{or}\;\pi$) of each isolated band, which is marked in green in Fig. 2(e) and Fig. 2(f). In addition, we use the Roman numerals $\textrm{{I} }\textrm{ - }\textrm{{V} }$ to number the bandgaps; the topological phase symbol of the nth bandgap is determined by the sum of Zak phases of all isolated bands below the photonic bandgap, which can be simply expressed as [39]:
$${\mathop{\rm sgn}} [{{\varsigma^{(n )}}} ]= {( - 1)^n}{( - 1)^l}\textrm{exp} \left( {i\sum\limits_{m = 0}^{n - 1} {\theta_m^{Zak}} } \right),$$
where $\textrm{n = }\textrm{{I} }\textrm{ - }\textrm{{V} }$, l is the number of band intersections under the nth bandgap, and $l \equiv 0$ in the proposed structure. Figure 2(e) and Fig. 2(f) show the band structures of two PhCs respectively. The black solid line stands for isolated band, and the rectangular box stands for bandgaps of different topological properties. The red and blue rectangles represent the positive and negative topological phase of the bandgap, respectively. The figures show five bandgaps of two PhCs at $\textrm{0}\textrm{.4 - 1}\textrm{.3}$THz. We can see that the fourth bandgaps of PhC 1 and PhC 2 overlap each other near 1 THz, which coincide with the broad bandgaps at $\textrm{0}\textrm{.97 - 1}\textrm{.07}$ THz in Fig. 2(a) and Fig. 2(b). Meanwhile, the topological phase signs of the two PhCs bandgaps are opposite, satisfying the condition that the PhC heterostructure excites TES.

 figure: Fig. 2.

Fig. 2. (a) Transmission spectrum and (b) reflection spectrum of PhC 1 (red solid line) and PhC 2 (black dotted line); (c) transmission spectrum and (d) reflection spectrum of PhC 1 + 3D DSM + PhC 2 heterostructure (red solid line) and PhC 1 + PhC 2 heterostructure (black dotted line); the band structures of (e) PhC 1 and (f) PhC 2.

Download Full Size | PDF

Further, we draw the electric field diagram of the PhC heterostructure, as shown in Fig. 3. It can be seen that there is an obvious electric field localization near the interface between PhC 1 and PhC 2, which confirms that our proposed structure can indeed excite TES. In addition, the characteristics of TES make abnormal transmission phenomenon appear on the interface of the PhC heterostructure near 1 THz. As shown in Fig. 2(c) and Fig. 2(d), there is a sharp transmission peak and reflection valley in the transmission spectrum and reflection spectrum respectively, which are represented by black dotted lines in the figure, and the central frequencies of the transmission peak and reflection valley are just located at the overlapping part of the fourth bandgap in PhC 1 and PhC 2. Since the frequency of TES in topological PhC heterostructure will shift with the incident angle, we consider the case where incident angle $\theta = {0^\circ }$. And we will discuss the influence of angle on OB in detail later. Considering the local field enhancement characteristics of TES, we introduce 3D DSM at the interface of heterostructure to facilitate the realization of low threshold OB. The red solid lines in Fig. 2(c) and Fig. 2(d) show the transmission and reflection spectra of the overall structure. It can be seen that the addition of 3D DSM would not affect the excitation of TES, and its nonlinearity and tunability can provide a basis for the emergence and regulation of OB. It is worth noting that although the defect modes of PhC can also produce similar the transmission peak, they are different in essence. When there are defects in the PhC, a new defect mode conduction band will appear in the bandgaps. The propagation of frequency occurs in the frequency range where transmission was completely prohibited before, and it shows a sharp transmission peak in the macro view.

 figure: Fig. 3.

Fig. 3. Electric field distribution of PhC heterostructure.

Download Full Size | PDF

Subsequently, we study low threshold OB based on the proposed structure above. Since the third-order nonlinear optical conductivities of Dirac material decrease with the increase of incident light frequency, we study the influence of 3D DSM on OB based on 1 THz frequency range. At this time, the third-order nonlinear coefficient of 3D DSM is strong. Generally speaking, as a typical Kerr medium, 3D DSM's complex refractive index includes linear and nonlinear parts. The relationship between complex refractive index and electric field can be expressed as $n = {n_D} + \Delta n = {n_D} + {n_2}{|E |^2}$, where ${n_2}\textrm{ = }{{{\chi ^{( \textrm{3}) }}} / {\textrm{2}{n_D}}}$ is the nonlinear refractive index. Therefore, the local electric field enhancement caused by the excitation of TES plays a positive role in the nonlinear refractive index part of 3D DSM. It is not difficult to find that we can excite low threshold tunable OB phenomenon by embedding 3D DSM in 1D PhC heterostructure under the condition of exciting the TES.

Then we discuss the influence of 3D DSM on OB phenomenon. We first explore the regulatory effect of Fermi energy of 3D DSM on OB. In Fig. 4(a), we show the relationship between the transmitted electric field and incident electric field under different Fermi energy of 3D DSM. We can see that when the Fermi energy gradually increases, the upper and lower thresholds of OB will increase, and the width of the hysteresis loop will also increase. In order to conveniently and qualitatively analyze the above OB phenomenon, we approximately equivalent 3D DSM to 2D film, similar to graphene. Then the bulk conductivity of 3D DSM would be equivalent to the surface conductivity, i.e., $\sigma \; = \;{d_{Dir}}({{\sigma^{(1 )}} + {\sigma^{(3 )}}{{|E |}^2}} )\;\textrm{ = }\;{\sigma _\textrm{1}}\textrm{ + }{\sigma _\textrm{3}}{|E |^2}$. According to Ref. [40], the expression of the transfer matrix of the equivalent 3D DSM is:

$${M_d} = \left[ {\begin{array}{*{20}{c}} {1 + \zeta }&\zeta \\ { - \zeta }&{1 - \zeta } \end{array}} \right].$$

Meanwhile, we can regard the two PhCs as two Bragg mirrors, so their transfer matrix can be expressed as [41]:

$${M_i} \approx \frac{1}{{{t_i}}}\left[ {\begin{array}{*{20}{c}} { - 1}&{ - |{{r_i}} |}\\ {|{{r_i}} |}&1 \end{array}} \right],$$
where $i = \{{1, 2 \}} $, ${t_i}$ and ${r_i}$ are the transmittance and reflectivity of PhC respectively, so the expression of the transfer matrix of the overall structure is: $M = {M_1}{M_d}{M_2}$. According to the definition of transmission coefficient, we write the relationship between transmission electric field ${E_t}$ and incident electric field ${E_i}$ as follows:
$$\frac{{{E_t}}}{{{E_i}}} = \frac{\textrm{1}}{{{M_{11}}}} = \frac{{{t_1}{t_2}}}{{\zeta [{1 - (|{{r_1}} |+ |{{r_2}} |)\textrm{ + }|{{r_1}} ||{{r_2}} |} ]+ 1 - |{{r_1}} ||{{r_2}} |}},$$
where $\zeta = {\mu _\textrm{0}}c({\sigma _\textrm{1}} + {\sigma _3}{|{{E_t}} |^2})/2$. Assuming that the transmitted electric field ${E_t}$ is a pure real number, let $Y = {|{{E_i}} |^2}$ and $X = {|{{E_t}} |^2}$, then
$$Y = {\left[ {\frac{{\zeta [{1 - (|{{r_1}} |+ |{{r_2}} |)\textrm{ + }|{{r_1}} ||{{r_2}} |} ]+ 1\textrm{ - }|{{r_1}} ||{{r_2}} |}}{{{t_1}{t_2}}}} \right]^2}X.$$

 figure: Fig. 4.

Fig. 4. Dependence of (a) the transmitted electric field and (b) transmittance on the incident electric field for different ${E_F}$ of 3D DSM. Here, $\lambda \mathrm{\ =\ 300\ \mu m}$, $\tau \textrm{ = 0}\textrm{.8 ps}$ and $\theta \textrm{ = 0 }^\circ$.

Download Full Size | PDF

At this time, let $k = [{1 - (|{{r_1}} |+ |{{r_2}} |)\textrm{ + }|{{r_1}} ||{{r_2}} |} ]$ and $Z = {\mu _0}c/2$. After calculation and simplification, the above formula can be written as:

$$\begin{array}{l} Y = \frac{{{{[{Zk({{\sigma_\textrm{1}} + {\sigma_3}X} )+ 1 - |{{r_1}} ||{{r_2}} |} ]}^2}}}{{{{({{t_1}{t_2}} )}^2}}}X\\ \;\;\; = \frac{{{{(Zk{\sigma _3})}^2}{X^2} + 2{\sigma _\textrm{1}}{\sigma _3}{{({Zk} )}^2}X + 2{\sigma _3}Zk({1 - |{{r_1}} ||{{r_2}} |} )X + {{(Zk{\sigma _\textrm{1}})}^2} + {{({1 - |{{r_1}} ||{{r_2}} |} )}^2} + 2Zk{\sigma _\textrm{1}}({1 - |{{r_1}} ||{{r_2}} |} )}}{{{{({{t_1}{t_2}} )}^2}}}X. \end{array}$$

It can be seen from Eq. (14) that Y is the cubic function of X, that is, an incident electric field may have three corresponding transmission electric fields. In order to satisfy the above situation, the discriminant $\Delta $ of the derivative of Eq. (14) must be greater than 0. The derivative function $Y^{\prime}$ of Eq. (14) can be expressed as:

$$\begin{array}{l} y = Y^{\prime}\\ \;\;\; = \frac{{3{{(Zk{\sigma _3})}^2}{X^2} + 4{\sigma _\textrm{1}}{\sigma _3}{{({Zk} )}^2}X + 4{\sigma _3}Zk({1 - |{{r_1}} ||{{r_2}} |} )X + {{(Zk{\sigma _\textrm{1}})}^2} + {{({1 - |{{r_1}} ||{{r_2}} |} )}^2} + 2Zk{\sigma _\textrm{1}}({1 - |{{r_1}} ||{{r_2}} |} )}}{{{{({{t_1}{t_2}} )}^2}}}. \end{array}$$

The discriminant $\Delta $ of derivative function $Y^{\prime}$ can be simplified as:

$$\Delta \textrm{ = }\frac{{{{[{\textrm{2}{\sigma_3}Zk({1 - |{{r_1}} ||{{r_2}} |} )\textrm{ + 2}{\sigma_\textrm{1}}{\sigma_3}{{({Zk} )}^2}} ]}^\textrm{2}}}}{{{{({{t_1}{t_2}} )}^4}}}.$$

From the above formula, we can clearly see that $\Delta $ is always greater than 0, which meets the condition for generating OB. Further, we can obtain the maximum and minimum points of function Y, that is, the two zeros of derivative function $y$:

$$\left\{ {\begin{array}{*{20}{c}} {{x_1} = \frac{{ - ({1 - |{{r_1}} ||{{r_2}} |} )- {\sigma_\textrm{1}}Zk}}{{3{\sigma_3}Zk}}}\\ {{x_2} = \frac{{ - ({1 - |{{r_1}} ||{{r_2}} |} )- {\sigma_\textrm{1}}Zk}}{{{\sigma_3}Zk}}} \end{array}} \right..$$

We study OB in the band near 1 THz where the individual PhC 1 and PhC 2 almost have total reflection, i.e., ${r_1} \approx \textrm{1}$ and ${r_2} \approx \textrm{1}$. In order to simplify the calculation, we set $1 - |{{r_1}} ||{{r_2}} |\approx \textrm{0}$. Equation (17), then, can be simplified to: ${x_1} \approx {{ - {\sigma _1}} / {3{\sigma _3}}}$ and ${x_2} \approx {{ - {\sigma _1}} / {{\sigma _3}}}$. Next, we bring the simplified maximum point and minimum point into Eq. (14) respectively to obtain the upper and lower thresholds of OB. Further, the width of hysteresis loop of OB can be expressed as:

$$\Delta Y\; = \;|{{Y_1} - {Y_2}} |\;\textrm{ = }\;\frac{{4\sigma _\textrm{1}^3{Z^2}{k^2}}}{{27{\sigma _3}t_1^2t_2^2}}.$$

In combination with the expressions of the linear optical conductivity and the third-order nonlinear optical conductivities of 3D DSM given by Eqs. (1) and (2), we can obtain the variation of OB hysteresis loop width with Fermi energy ${E_F}$ and relaxation time $\tau$:

$$\Delta Y\; = \;K\frac{{({1 + {w^2}{\tau^2}} )({1 - 2iw\tau } )}}{{{{({1 - iw\tau } )}^\textrm{3}}}}{\left[ {{{\left( {\frac{{{E_F}}}{{{k_B}T}}} \right)}^2} + \frac{{{\pi^2}}}{3}} \right]^3}\left( {1 + \textrm{exp} ( - \frac{{{E_F}}}{{{k_B}T}})} \right),$$
where $K\;\textrm{ = }\;\frac{{\textrm{160}{{({{d_{Dir}}{\sigma_0}Zk} )}^2}{{({{k_B}T} )}^6}}}{{729{{({\pi \hbar {v_F}} )}^4}{{({e{t_1}{t_2}} )}^2}}}$. It can be seen from the above formula that when the Fermi energy increases, the hysteresis loop width $\Delta Y$ will also increase, which is consistent with the law shown in Fig. 4. At the same time, Fig. 4(b) shows the law of the transmittance changing with the incident electric field. We can see that the transmittance of the incident wave tends to decrease with the increase of Fermi energy. This means that as the Fermi energy of 3D DSM increases, there would be a certain amount of loss in transmission. Therefore, when designing low threshold optical bistable devices, it is necessary to reasonably adjust the relevant parameters to reduce the loss caused by materials. In addition, from the variation law of Fig. 4, it is not difficult to find that when the Fermi energy increases, the transmittance decreases continuously, but the transmitted electric field shows an upward trend due to the local field enhancement effect caused by TES excitation. It is worth noting that the dynamically adjustable Fermi energy can be controlled by applying an external voltage. The common practice is to apply an electrode between the 3D DSM and the background material to manipulate the conductivity characteristics of 3D DSM. The above results show that the excitation of TES and the addition of 3D DSM provide the feasibility of realizing low threshold tunable OB phenomenon.

Combining Eq. (17) with the above derivation and analysis, we find that the OB phenomenon is not only affected by Fermi energy, but also closely related to the relaxation time $\tau$ of 3D DSM. Figure 5 describes the functional relationship between the transmitted electric field and the incident electric field when ${E_F} = 0.07\;\textrm{eV}$.The width of hysteresis loop and the upper and lower thresholds of OB change with the relaxation time, and this changing trend is consistent with our derivation results. It is obvious from Fig. 5 that as relaxation time $\tau$ increases from $\textrm{0}\textrm{.5}\;\textrm{ps}$ to $\textrm{0}\textrm{.8}\;\textrm{ps}$, the width of hysteresis loop of OB is also increasing, while the threshold of OB is gradually decreasing. Specifically, when $\tau \;\textrm{ = }\;\textrm{0}\textrm{.5 ps}$, the width of hysteresis loop $\Delta Y\; \approx \;\textrm{1}\textrm{.20} \times \textrm{1}{\textrm{0}^\textrm{6}}\;\textrm{V/m}$ and the lower threshold ${|{{E_i}} |_{down}}\; \approx \;\textrm{9}\textrm{.72} \times \textrm{1}{\textrm{0}^\textrm{5}}\;\textrm{V/m}$; when $\tau \;\textrm{ = }\;\textrm{0}\textrm{.8 ps}$, the width of hysteresis loop $\Delta Y\; \approx \;\textrm{1}\textrm{.31} \times \textrm{1}{\textrm{0}^\textrm{6}}\;\textrm{V/m}$ and the lower threshold ${|{{E_i}} |_{down}}\; \approx \;\textrm{7}\textrm{.40} \times \textrm{1}{\textrm{0}^\textrm{5}}\;\textrm{V/m}$. Overall, the width of hysteresis loop increases by nearly $\textrm{1}\textrm{.1} \times \textrm{1}{\textrm{0}^\textrm{5}}\;\textrm{V/m}$, and the threshold of OB decreases by nearly $\textrm{2}\textrm{.32} \times \textrm{1}{\textrm{0}^\textrm{5}}\;\textrm{V/m}$. The above results show that the threshold of OB can be effectively reduced by reasonably adjusting the relaxation time of 3D DSM, and the controllability of the hysteresis loop can be achieved simultaneously, which provides a new scheme for designing tunable bistable devices. However, it should be pointed out that even though relaxation time plays an obvious role in regulating OB threshold, it still has limitations in flexibility. After all, once the structure and 3D DSM are prepared, relaxation time has no regulatory characteristics. Nevertheless, the influence of relaxation time on OB can still provide reference for designing bistable devices.

 figure: Fig. 5.

Fig. 5. Dependence of (a) the transmitted electric field and (b) transmittance on the incident electric field for different $\tau$ of 3D DSM. Here, $\lambda \mathrm{\ =\ 300\ \mu m}$, ${E_F}\textrm{ = 0}\textrm{.07}\;\textrm{eV}$ and $\theta \textrm{ = 0 }^\circ$.

Download Full Size | PDF

We have previously discussed the influence of the parameters of 3D DSM on OB. It has been determined that the upper and lower thresholds of OB can be controlled by adjusting the Fermi energy or relaxation time of 3D DSM separately. Further, we found that OB is also affected by the thickness of 3D DSM and the angle of incident light. As can be seen from Fig. 6(a), the change of thickness of 3D DSM has little effect on the overall threshold, which perfectly meets our requirements for the design of bistable devices. As we all know, it is very difficult to accurately control the thickness of the device during the manufacturing process. In other words, it is beneficial to control the overall performance of the device if the slight variation in thickness does not have a decisive effect on the threshold and width of threshold. In addition, the change of incident light angle also has a certain impact on the width of threshold. As shown in Fig. 6(b), we draw the image of OB changing with incident light angle when ${E_F}\;\textrm{ = }\;\textrm{0}\textrm{.07 eV}$ and $\tau \;\textrm{ = }\;\textrm{0}\textrm{.8 ps}$. It is not difficult to find that as the incident angle $\theta$ increases from ${\textrm{0}^\circ }$ to ${\textrm{3}^\circ }$, the lower threshold of OB changes little, while the upper threshold increases and so does the width of threshold. Therefore, when designing a bistable device, the influence of the incident light angle on the device performance should also be considered.

 figure: Fig. 6.

Fig. 6. Influence of (a) the thickness of 3D DSM ${d_{Dir}}$ and (b) the angle of incident light $\theta$ on OB. Here, $\lambda \mathrm{\ =\ 300\ \mu m}$, ${E_F}\textrm{ = 0}\textrm{.07}\;\textrm{eV}$ and $\tau \textrm{ = 0}\textrm{.8 ps}$.

Download Full Size | PDF

In addition, we also calculate the hysteresis loop relationship of reflective OB as shown in Fig. 7, where the relevant parameters are consistent with those in the calculation of transmissive OB. Figure 7(a) and Fig. 7(b) show the hysteresis loop relationships of the incident electric field to the reflected electric field and the reflectivity at $\tau \;\textrm{ = }\;\textrm{0}\textrm{.8 ps}$. When the Fermi energy increases from $\textrm{0}\textrm{.07}\;\textrm{eV}$ to $0.10\;\textrm{eV}$, the width of hysteresis loop of OB increases, and so do the upper and lower thresholds. This is the same as the variation law of transmissive OB shown in Fig. 4. Similarly, Fig. 7(c) and Fig. 7(d) are the hysteresis loop relationships of the incident electric field to the reflected electric field and the reflectivity when ${E_F}\;\textrm{ = }\;\textrm{0}\textrm{.07 eV}$. With the increase of relaxation time, the width of hysteresis loop of OB will increase, while the upper and lower thresholds will decrease, which is basically consistent with the law in Fig. 5.

 figure: Fig. 7.

Fig. 7. Dependence of (a) the reflected electric field and (b) reflectance on the incident electric field for different ${E_F}$ of 3D DSM; here, $\lambda \mathrm{\ =\ 300\ \mu m}$, $\tau \textrm{ = 0}\textrm{.8 ps}$ and $\theta \textrm{ = 0 }^\circ$. Dependence of (c) the reflected electric field and (d) reflectance on the incident electric field for different $\tau$ of 3D DSM; here, $\lambda \mathrm{\ =\ 300\ \mu m}$, ${E_F}\textrm{ = 0}\textrm{.07}\;\textrm{eV}$ and $\theta \textrm{ = 0 }^\circ$.

Download Full Size | PDF

4. Conclusion

In this paper, we are the first to achieve the dynamically tunable low threshold OB by exciting TES based on a structure composed of 3D DSM and two PhCs in the terahertz region. The results show that TES caused by PhC heterostructure can effectively enhance the local electric field, which has a positive effect on realizing low threshold OB. Meanwhile, the large nonlinear refractive index of 3D DSM provides nonlinear conditions for the emergence of OB phenomenon; the threshold and hysteresis loop width of OB can be adjusted by changing its relevant parameters. In addition, we further found that the threshold and width of threshold are regulated by the Fermi energy and relaxation time of 3D DSM, and are closely related to the angle of incident light, while the change of 3D DSM thickness has little effect on the overall threshold. The low threshold OB scheme proposed in this paper has the characteristics of simple structure and easy preparation. It is believed that this low threshold OB scheme has excellent potential for applications in related nonlinear optical devices.

Funding

National Natural Science Foundation of China (11704119, 61875133, 11874269); Scientific Research Fund of Hunan Provincial Education Department (21B0048); Natural Science Foundation of Hunan Province (2021JJ30135, 2021JJ30149); Science and Technology Project of Shenzhen (JCYJ20190808143801672, JCYJ20190808150803580, JCYJ20180508152903208); Changsha Natural Science Foundation (kq2202236); National College Students’ innovation and entrepreneurship training program (202110542014).

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Grant Nos. 11704119, 61875133 and 11874269), Scientific Research Fund of Hunan Provincial Education Department (Grant No. 21B0048), Natural Science Foundation of Hunan Province (2021JJ30135 and 2021JJ30149), and the Science and Technology Project of Shenzhen (Grant Nos. JCYJ20190808143801672, JCYJ20190808150803580, JCYJ20180508152903208), the Changsha Natural Science Foundation (Grant No. kq2202236) and National College Students’ innovation and entrepreneurship training program (Grant No. 202110542014).

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 corresponding author upon reasonable request.

References

1. H. M. Gibbs, Optical Bistability: Controlling Light with Light (Academic, 1985).

2. M. Notomi, A. Shinya, S. Mitsugi, G. Kira, and T. Tanabe, “Optical bistable switching action of Si high-Q photonic-crystal nanocavities,” Opt. Express 13(7), 2678–2687 (2005). [CrossRef]  

3. D. A. Mazurenko, R. Kerst, J. I. Dijkhuis, A. V. Akimov, V. G. Golubev, D. A. Kurdyukov, A. B. Pevtsov, and A. V. Sel’kin, “Ultrafast optical switching in three-dimensional photonic crystals,” Phys. Rev. Lett. 91(21), 213903 (2003). [CrossRef]  

4. M. F. Yanik, “All-optical transistor action with bistable switching in a photonic crystal cross-waveguide geometry,” Opt. Lett. 28(24), 2506–2508 (2003). [CrossRef]  

5. T. Tanabe, M. Notomi, S. Mitsugi, A. Shinya, and E. Kuramochi, “Fast bistable all-optical switch and memory on a silicon photonic crystal on-chip,” Opt. Lett. 30(19), 2575–2577 (2005). [CrossRef]  

6. P. Y. Wen, M. Sanchez, M. Gross, and S. Esener, “Vertical-cavity optical AND gate,” Opt. Commun. 219(1-6), 383–387 (2003). [CrossRef]  

7. W. L. Zhang, Y. Jiang, Y. Y. Zhu, F. Wang, and Y. J. Rao, “All-optical bistable logic control based on coupled Tamm plasmons,” Opt. Lett. 38(20), 4092–4095 (2013). [CrossRef]  

8. H. Yuan, X. Jiang, F. Huang, and X. Sun, “Ultralow threshold optical bistability in metal/randomly layered media structure,” Opt. Lett. 41(4), 661–664 (2016). [CrossRef]  

9. J. Shiri, J. Khalilzadeh, and S. H. Asadpour, “Optical bistability in reflection of the laser pulse in a 1D photonic crystal doped with four-level InGaN/GaN quantum dots,” Laser Phys. 31(3), 036202 (2021). [CrossRef]  

10. M. Kim, S. Kim, and S. Kim, “Optical bistability based on hyperbolic metamaterials,” Opt. Express 26(9), 11620–11632 (2018). [CrossRef]  

11. Z. P. Wang and B. L. Yu, “Optical bistability and multistability in polaritonic materials doped with nanoparticles,” Laser Phys. Lett. 11(11), 115903 (2014). [CrossRef]  

12. Y. X. Peng, J. Xu, S. P. Wang, H. Dong, Y. J. Xiang, X. Y. Dai, J. Guo, S. Y. Qian, and L. Y. Jiang, “Low-threshold and tunable optical bistability based on topological edge state in one-dimensional photonic crystal heterostructure with graphene,” IEEE Access 8, 196386–196393 (2020). [CrossRef]  

13. H. Q. Deng, C. P. Ji, X. Y. Zhang, P. Chen, L. C. Wu, J. Jiang, H. S. Tian, and L. Y. Jiang, “Low threshold optical bistability in graphene/waveguide hybrid structure at terahertz frequencies,” Opt. Commun. 499, 127282 (2021). [CrossRef]  

14. Z. H. Xu, D. Wu, Y. M. Liu, C. Liu, Z. Y. Yu, L. Yu, and H. Ye, “Design of a tunable ultra-broadband terahertz absorber based on multiple layers of graphene ribbons,” Nanoscale Res. Lett. 13(1), 143–151 (2018). [CrossRef]  

15. S. M. Young, S. Zaheer, J. C. Y. Teo, C. L. Kane, E. J. Mele, and A. M. Rappe, “Dirac semimetal in three dimensions,” Phys. Rev. Lett. 108(14), 140405 (2012). [CrossRef]  

16. B. J. Yang and N. Nagaosa, “Classification of stable three-dimensional Dirac semimetal with nontrivial topology,” Nat. Commun. 5(1), 3045 (2014). [CrossRef]  

17. T. Liang, Q. Gibson, M. N. Ali, M. Liu, R. J. Cava, and N. P. Ong, “Ultrahigh mobility and giant magnetoresistance in the Dirac semimetal Cd3As2,” Nat. Mater. 14(3), 280–284 (2015). [CrossRef]  

18. M. Veldhorst, M. Snelder, M. Hoek, T. Gang, V. K. Guduru, X. L. Wang, U. Zeitler, W. G. van der Wiel, A. A. Golubov, H. Hilgenkamp, and A. Brinkman, “Josephson supercurrent through a topological insulator surface state,” Nat. Mater. 11(5), 417–421 (2012). [CrossRef]  

19. M. Neupane, S. Y. Xu, R. Sankar, N. Alidoust, G. Bian, C. Liu, I. Belopolski, T. R. Chang, H. T. Jeng, H. Lin, A. Bansil, F. C. Chou, and M. Z. Hasan, “Observation of a three-dimensional topological Dirac semimetal phase in high-mobility Cd3As2,” Nat. Commun. 5(1), 3786 (2014). [CrossRef]  

20. T. L. Wang, M. Y. Cao, Y. P. Zhang, and H. Y. Zhang, “Tunable polarization-nonsensitive electromagnetically induced transparency in Dirac semimetal metamaterial at terahertz frequencies,” Opt. Mater. Express 9(4), 1562–1576 (2019). [CrossRef]  

21. L. P. He, X. C. Hong, J. K. Dong, J. Pan, Z. Zhang, J. Zhang, and S. Y. Li, “Quantum transport evidence for the three-dimensional Dirac semimetal phase in Cd3As2,” Phys. Rev. Lett. 113(24), 246402 (2014). [CrossRef]  

22. Z. K. Liu, B. Zhou, Y. Zhang, Z. J. Wang, H. M. Weng, D. Prabhakaran, S. K. Mo, Z. X. Shen, Z. Fang, X. Dai, Z. Hussain, and Y. L. Chen, “Discovery of a three-dimensional topological Dirac semimetal, Na3Bi,” Science 343(6173), 864–867 (2014). [CrossRef]  

23. R. Y. Chen, G. D. Gu, S. J. Zhang, J. A. Schneeloch, C. Zhang, Q. Li, and N. L. Wang, “Optical spectroscopy study of three dimensional Dirac semimetal ZrTe5,” Phys. Rev. B 92(7), 075107 (2015). [CrossRef]  

24. T. Timusk, J. P. Carbotte, C. C. Homes, D. N. Basov, and S. G. Sharapov, “Three-dimensional Dirac fermions in quasicrystals as seen via optical conductivity,” Phys. Rev. B 87(23), 235121 (2013). [CrossRef]  

25. C. Li, X. Hu, H. Yang, and Q. Gong, “Unidirectional transmission in 1D nonlinear photonic crystal based on topological phase reversal by optical nonlinearity,” AIP Adv. 7(2), 025203 (2017). [CrossRef]  

26. A. V. Poshakinskiy, A. N. Poddubny, L. Pilozzi, and E. L. Ivchenko, “Radiative topological states in resonant photonic crystals,” Phys. Rev. Lett. 112(10), 107403 (2014). [CrossRef]  

27. L. Lu, C. Fang, L. Fu, S. G. Johnson, J. D. Joannopoulos, and M. Soljacic, “Symmetry-protected topological photonic crystal in three dimensions,” Nat. Phys. 12(4), 337–340 (2016). [CrossRef]  

28. M. Verbin, O. Zilberberg, Y. E. Kraus, Y. Lahini, and Y. Silberberg, “Observation of topological phase transitions in photonic quasicrystals,” Phys. Rev. Lett. 110(7), 076403 (2013). [CrossRef]  

29. W. Gao, M. Lawrence, B. Yang, F. Liu, F. Fang, B. Béri, J. Li, and S. Zhang, “Topological photonic phase in chiral hyperbolic Metamaterials,” Phys. Rev. Lett. 114(3), 037402 (2015). [CrossRef]  

30. M. Hafezi, S. Mittal, J. Fan, A. Migdall, and J. M. Taylor, “Imaging topological edge states in silicon photonics,” Nat. Photonics 7(12), 1001–1005 (2013). [CrossRef]  

31. W. Gao, X. Y. Hu, C. Li, J. H. Yang, Z. Chai, J. Y. Xie, and Q. H. Gong, “Fano-resonance in one-dimensional topological photonic crystal heterostructure,” Opt. Express 26(7), 8634–8644 (2018). [CrossRef]  

32. D. R. Lide, Handbook of Chemistry and Physics (CRC Press, 1995).

33. K. J. A. Ooi, Y. S. Ang, Q. Zhai, D. T. H. Tan, L. K. Ang, and C. K. Ong, “Nonlinear plasmonics of three-dimensional Dirac semimetal,” APL Photonics 4(3), 034402 (2019). [CrossRef]  

34. N. H. Liu, S. Y. Zhu, H. Chen, and X. Wu, “Superluminal pulse propagation through one-dimensional photonic crystals with a dispersive defect,” Phys. Rev. E 65(4), 046607 (2002). [CrossRef]  

35. G. S. Agarwal and S. D. Gupta, “Effect of nonlinear boundary conditions on nonlinear phenomena in optical resonators,” Opt. Lett. 12(10), 829–831 (1987). [CrossRef]  

36. S. D. Gupta, “Optical multistability and solitonlike intensity distribution in a nonlinear superlattice for oblique incidence,” J. Opt. Soc. Am. B 6(10), 1927–1931 (1989). [CrossRef]  

37. S. D. Gupta and D. S. Ray, “Optical multistability in a nonlinear Fibonacci multilayer,” Phys. Rev. B 38(5), 3628–3631 (1988). [CrossRef]  

38. V. D. Lago, M. Atala, and L. E. F. F. Torres, “Floquet topological transitions in a driven one-dimensional topological insulator,” Phys. Rev. A 92(2), 023624 (2015). [CrossRef]  

39. M. Xiao, Z. Q. Zhang, and C. T. Chan, “Surface impedance and bulk band geometric phases in one-dimensional systems,” Phys. Rev. X 4(2), 021017 (2014). [CrossRef]  

40. L. Y. Jiang, J. Guo, L. M. Wu, X. Y. Dai, and Y. J. Xiang, “Manipulating the optical bistability at terahertz frequency in the Fabry-Perot cavity with graphene,” Opt. Express 23(24), 31181–31191 (2015). [CrossRef]  

41. A. Ferreira, N. M. R. Peres, R. M. Ribeiro, and T. Stauber, “Graphene-based photodetector with two cavities,” Phys. Rev. B 85(11), 115438 (2012). [CrossRef]  

Data availability

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

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1.
Fig. 1. Schematic diagram of TES based on 3D DSM/PhC heterostructure: (a) visual view and (b) side view.
Fig. 2.
Fig. 2. (a) Transmission spectrum and (b) reflection spectrum of PhC 1 (red solid line) and PhC 2 (black dotted line); (c) transmission spectrum and (d) reflection spectrum of PhC 1 + 3D DSM + PhC 2 heterostructure (red solid line) and PhC 1 + PhC 2 heterostructure (black dotted line); the band structures of (e) PhC 1 and (f) PhC 2.
Fig. 3.
Fig. 3. Electric field distribution of PhC heterostructure.
Fig. 4.
Fig. 4. Dependence of (a) the transmitted electric field and (b) transmittance on the incident electric field for different ${E_F}$ of 3D DSM. Here, $\lambda \mathrm{\ =\ 300\ \mu m}$, $\tau \textrm{ = 0}\textrm{.8 ps}$ and $\theta \textrm{ = 0 }^\circ$.
Fig. 5.
Fig. 5. Dependence of (a) the transmitted electric field and (b) transmittance on the incident electric field for different $\tau$ of 3D DSM. Here, $\lambda \mathrm{\ =\ 300\ \mu m}$, ${E_F}\textrm{ = 0}\textrm{.07}\;\textrm{eV}$ and $\theta \textrm{ = 0 }^\circ$.
Fig. 6.
Fig. 6. Influence of (a) the thickness of 3D DSM ${d_{Dir}}$ and (b) the angle of incident light $\theta$ on OB. Here, $\lambda \mathrm{\ =\ 300\ \mu m}$, ${E_F}\textrm{ = 0}\textrm{.07}\;\textrm{eV}$ and $\tau \textrm{ = 0}\textrm{.8 ps}$.
Fig. 7.
Fig. 7. Dependence of (a) the reflected electric field and (b) reflectance on the incident electric field for different ${E_F}$ of 3D DSM; here, $\lambda \mathrm{\ =\ 300\ \mu m}$, $\tau \textrm{ = 0}\textrm{.8 ps}$ and $\theta \textrm{ = 0 }^\circ$. Dependence of (c) the reflected electric field and (d) reflectance on the incident electric field for different $\tau$ of 3D DSM; here, $\lambda \mathrm{\ =\ 300\ \mu m}$, ${E_F}\textrm{ = 0}\textrm{.07}\;\textrm{eV}$ and $\theta \textrm{ = 0 }^\circ$.

Equations (19)

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

σ ( 1 ) = σ 0 4 3 π 2 τ 1 i ω τ ( k B T ) 2 2 v F [ 2 L i 2 ( e E F k B T ) + ( E F k B T ) 2 + π 2 3 ] ,
σ ( 3 ) = σ 0 8 e 2 v F 5 π 2 2 τ 3 ( 1 + ω 2 τ 2 ) ( 1 2 i ω τ ) 1 1 + exp ( E F / k B T ) .
{ χ ( 3 ) = i σ ( 3 ) / ε 0 ω n D = n + i k = 1 + i σ ( 1 ) / ε 0 ω ,
M D = k 0 k z + + k z [ k z k 0 exp ( i k z + d D ) + k z + k 0 exp ( i k z d D ) exp ( i k z + d D ) exp ( i k z d D ) k z k z + k 0 2 exp ( i k z + d D ) exp ( i k z d D ) k z + k 0 exp ( i k z + d D ) + k z k 0 exp ( i k z d D ) ] ,
k z ± = ( k 0 n D ) 2 ( k y ) 2 ( 1 + U ± + 2 U ) 1 2 ,
M = ( M B 1 × M A 1 ) 4 × M D × ( M B 2 × M A 2 ) 4 .
{ t = 2 p f [ M 11 + M 12 p f ] p f + [ M 21 + M 22 p f ] r = [ M 11 + M 12 p f ] p f [ M 21 + M 22 p f ] [ M 11 + M 12 p f ] p f + [ M 21 + M 22 p f ] ,
cos ( q Λ ) = cos k a d a cos k b d b 1 2 ( z a z b + z b z a ) sin k a d a sin k b d b ,
sgn [ ς ( n ) ] = ( 1 ) n ( 1 ) l exp ( i m = 0 n 1 θ m Z a k ) ,
M d = [ 1 + ζ ζ ζ 1 ζ ] .
M i 1 t i [ 1 | r i | | r i | 1 ] ,
E t E i = 1 M 11 = t 1 t 2 ζ [ 1 ( | r 1 | + | r 2 | )  +  | r 1 | | r 2 | ] + 1 | r 1 | | r 2 | ,
Y = [ ζ [ 1 ( | r 1 | + | r 2 | )  +  | r 1 | | r 2 | ] + 1  -  | r 1 | | r 2 | t 1 t 2 ] 2 X .
Y = [ Z k ( σ 1 + σ 3 X ) + 1 | r 1 | | r 2 | ] 2 ( t 1 t 2 ) 2 X = ( Z k σ 3 ) 2 X 2 + 2 σ 1 σ 3 ( Z k ) 2 X + 2 σ 3 Z k ( 1 | r 1 | | r 2 | ) X + ( Z k σ 1 ) 2 + ( 1 | r 1 | | r 2 | ) 2 + 2 Z k σ 1 ( 1 | r 1 | | r 2 | ) ( t 1 t 2 ) 2 X .
y = Y = 3 ( Z k σ 3 ) 2 X 2 + 4 σ 1 σ 3 ( Z k ) 2 X + 4 σ 3 Z k ( 1 | r 1 | | r 2 | ) X + ( Z k σ 1 ) 2 + ( 1 | r 1 | | r 2 | ) 2 + 2 Z k σ 1 ( 1 | r 1 | | r 2 | ) ( t 1 t 2 ) 2 .
Δ  =  [ 2 σ 3 Z k ( 1 | r 1 | | r 2 | )  + 2 σ 1 σ 3 ( Z k ) 2 ] 2 ( t 1 t 2 ) 4 .
{ x 1 = ( 1 | r 1 | | r 2 | ) σ 1 Z k 3 σ 3 Z k x 2 = ( 1 | r 1 | | r 2 | ) σ 1 Z k σ 3 Z k .
Δ Y = | Y 1 Y 2 |  =  4 σ 1 3 Z 2 k 2 27 σ 3 t 1 2 t 2 2 .
Δ Y = K ( 1 + w 2 τ 2 ) ( 1 2 i w τ ) ( 1 i w τ ) 3 [ ( E F k B T ) 2 + π 2 3 ] 3 ( 1 + exp ( E F k B T ) ) ,
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.