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

Large photonic spin Hall effect in two dimensional semi-Dirac materials

Open Access Open Access

Abstract

A strong photonic spin Hall effect (PSHE) in the absence of external magnetic field is important to control the spin states of photons and design next-generation photonic devices based on spinotpics. Herein, we theoretically study the PSHE on the surface of semi-Dirac materials. We established a general model, by changing various incident conditions, to describe the spin-orbit interaction of light in semi-Dirac materials. When we made these changes, we found that a large PSHE arises from the intrinsic anisotropy in the dispersion of semi-Dirac materials. The in-plane (transverse) spin Hall shift is dozens times of λi, where λi is the wavelength of the incident photon. Both of them can be effectively tuned by adjusting the optical axis angle and the photon energy. The maximum of the total in-plane and transverse spin Hall shift is 83.91λi and 19.65λi, respectively, which is larger than those in conventional anisotropic two-dimensional materials and isotropic Dirac materials. Our results shed light on the spin-orbit coupling of light in semi-Dirac materials and pave the way for designing spin-optical devices.

© 2021 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Photonic spin Hall effect (PSHE) is a photonic analogy of electronic spin Hall effect (SHE), in which the photon spin plays the role of the electron spin, and the refractive index gradient plays the role of the external field [15]. PSHE manifests itself as spin-dependent splitting arising from the spin orbit coupling of light [35]. In particular, when a beam passes through an inhomogeneous medium with index gradient, photons with opposite spin angular momentum i.e., the left and right circular polarization, are spatially separated from each other, resulting in a spin dependent lateral shift, i.e., the PSHE [37]. PSHE provides a new way to manipulate photons and has many applications in nano-optics, quantum information, and semiconductor physics [5,8,9]. In 2004, Onoda et al. theoretically explained PSHE by considering the geometric Berry phase and the angular momentum conservation [3]. Subsequently, in 2008, Hosten and Kwiat firstly detect and verify this phenomenon for photons passing through an air–glass interface by using the weak measurement technique [10,11]. Since then, PSHE has attracted great scientific research interests and technological upgrading on spinoptics [5].

At first, the research on PSHE focuses on the surface of three-dimensional (3D) materials, especially the metasurfaces [1,2,1217]. With the rapid development of two-dimensional (2D) materials in the past decades, the research on PSHE has been extended to atomically thin 2D materials, such as the magnetically modulated PSHE in graphene [18,19], and the anisotropic optical response induced PSHE in black phosphorus [20,21]. Unfortunately, it is found that there is no in-plane spin Hall shift in nonmagnetic isotropic 2D materials due to the lack of cross conductivity arising from the restriction of time reversal symmetry [22,23]. Usually, for nonmagnetic 2D materials, we need to exert an external magnetic field to break the time reversal symmetry and obtain in-plane spin Hall shifts [18,19,21,24]. However, previous works [1217,20] show that there are finite in-plane and transversal spin Hall shifts in anisotropic materials such as hyperbolic metamaterials, tilted linear-dichroic plate, and 2D black phosphorus in the absence of magnetic fields due to their anisotropic optical response.

Recently, a new kind of highly anisotropic electron system called semi-Dirac material which have both the properties of the Dirac electron in graphene and Schördinger electron in convectional semiconductors attracts intensive attention [2530]. Semi-Dirac material hosts highly anisotropic two dimensional electron gas, of which the low energy dispersion is linear along one direction and parabolic along the perpendicular direction [31,32]. Semi-Dirac electron is firstly predicted in strained graphene [31] and recently observed in potassium-doped few-layer black phosphorus [33]. The unique dispersion results in highly anisotropic optical response to linearly polarized light [34]. The optical conductivity excited by linearly polarized light along the direction of linear dispersion is about two order larger than that of linearly polarized light along the direction of the parabolic dispersion [35]. From the previous work on the PSHE in monolayer black phosphorus [20], this highly anisotropic optical conductivity in semi-Dirac material surely indicates a strong PSHE with large in-plane spin Hall shift even in the absence of external magnetic fields. A large PSHE may allow us to observe the spin Hall shifts directly [1,3638], and avoid to use the quantum weak measurement which desires complex experimental setup. Based on the measured spin Hall shifts, one can extract the optical parameters of the system such as the thickness [39] and the optical conductivity [40]. However, up to date, the PSHE in semi-Dirac materials still remains elusive.

In this work, we study the PSHE in 2D semi-Dirac materials. In our model, the beam is irradiated on a semi-Dirac material, then the reflected beam will undergo in-plane and transverse spin splitting, in which the centroids of left and right circular polarized light are spatially separated. By changing various incident conditions, we find that there is a large PSHE on the surface of the semi-Dirac material in the absence of external magnetic field resulting from its anisotropic optical response. The in-plane and transverse spin Hall shift are dozens times of the wavelength of the incident photon, which can be effectively tuned by the optical axis angle and the energy of incident photon. The maximum of the total in-plane (transverse) spin Hall shift is 83.91 (19.65) times of the wavelength of the incident photon, which is larger than that on the surface on conventional anisotropic 2D materials [20] and isotropic Dirac materials [18,19,24]. Our results provide further understanding on the spin-orbit coupling of light in non-magnetic anisotropic materials and useful information in designing spin-optical devices based on them.

2. Model and method

In our work, we establish a general model to describe the PSHE on the surface of the semi-Dirac materials as shown in Fig. 1(a). We consider a Gaussian wave packet with monochromatic frequency irradiate from the air to the surface of semi-Dirac material with incident angle $\theta _{i}$. We choose laboratory Cartesian frame $(x,y,z)$ to describe this question. The coordinates $({x_{i}},{y_{i}},{z_{i}})$ and $({x_{r}},{y_{r}},{z_{r}})$ represent the centers of the incident and reflected wave packets, respectively. The low-energy Hamiltonian of semi-Dirac material can be described by a two band model, which is given by [34]

$$H=\left( \begin{array}{cc} 0 & \frac{\hbar ^{2}k_{x}^{2}}{2m^{\ast}}-i\hbar v_{F}k_{y} \\ \frac{\hbar ^{2}k_{x}^{2}}{2m^{\ast}}+i\hbar v_{F}k_{y} & 0 \end{array} \right) ,$$
where $m^{\ast }$ is the effective mass, $v_{F}$ the Fermi velocity, and $\mathbf {k}=(k_{x},k_{y})$ the wave vector. The dispersion of Hamiltonian (1) is
$$E_{{\pm} }={\pm} \sqrt{\left( \hbar ^{2}k_{x}^{2}/2m^{{\ast} }\right) ^{2}+\hbar ^{2}v_{F}^{2}k_{y}^{2}},$$
where $+/-$ stands for the conduction/valence band. As plotted in Fig. 1(b), the dispersion is parabolic in $k_{x}$-direction and linear in $k_{y}$-direction. This peculiar dispersion has been recently observed in potassium-doped few-layer black phosphorus [33]. In this system, the two parameters are $v_F=3\times 10^5$ m/s, and $m^*=1.49$ $m_e$, where $m_e$ is the mass of free electron. Within the framework of linear response theory, the conductivity tensor given by the Kubo formula [41] is
$$\sigma _{\mu \nu }=\frac{e^{2}\hbar }{iS}\sum_{\mathbf{k,k}^{\prime }}\frac{ f\left( E_{\mathbf{k}}\right) -f\left( E_{\mathbf{k}^{\prime }}\right) }{ \left( E_{\mathbf{k}}-E_{\mathbf{k}^{\prime }}\right) }\frac{\left\langle \mathbf{k}\left\vert v_{\mu }\right\vert \mathbf{k}^{\prime }\right\rangle \left\langle \mathbf{k}^{\prime }\left\vert v_{\nu }\right\vert \mathbf{k} \right\rangle }{\left( E_{\mathbf{k}}-E_{\mathbf{k}^{\prime }}+\hbar \omega +i\Gamma \right) },$$
where $\mu,\nu \in \{x,y\}$, $S$ is the area of the sample, $E_{\mathbf {k} }\left ( \left \vert \mathbf {k}\right \rangle \right )$ the eigenvalue (eigenvector) of the system, $f\left ( E\right ) =\left [ e^{\left ( E-E_{F}\right ) /k_{B}T}+1\right ] ^{-1}$ the Fermi-Dirac distribution function with Boltzman constant $k_{B}$ and temperature $T$, $v_{\mu /\nu }=\partial H/\partial p_{\mu /\nu }$ the velocity operator, $\hbar \omega$ the photon energy, $\Gamma$ the level broadening induced by impurities.

 figure: Fig. 1.

Fig. 1. (a) Schematic illustration on the beam reflection on the surface of semi-Dirac materials in a Cartesian coordinate system with the incident angle $\theta _{i}$. The optical axis angle $\phi$ is defined as the angle between the $xoz$ plane and the incident plane which is determined by the incident light and the $z$-axis. The PSHE occurs on the reflecting surface and exhibits in-plane and transverse Hall shifts. (b) Schematic plot of the low energy dispersion of semi-Dirac electrons.

Download Full Size | PDF

Then, the conductivity tensor is given by

$$\sigma (\omega )=\left( \begin{array}{cc} \sigma _{xx} & \sigma _{xy} \\ \sigma _{yx} & \sigma _{yy} \end{array} \right) .$$

The off diagonal elements $\sigma _{xy}$ and $\sigma _{yx}$ are zero because of the time reversal symmetry [34,35]. The conductivity matrix is expressed as [20]

$$\sigma _{M}=M_{1}\sigma M_{2}=\left( \begin{array}{cc} \sigma _{pp} & \sigma _{ps} \\ \sigma _{sp} & \sigma _{ss} \end{array} \right) ,$$
with $M_{1}=\left ( \begin{array}{cc} \cos \phi & \sin (-\phi ) \\ \sin \phi & \cos \phi \end {array} \right )$, $M_{2}=\left ( \begin {array}{cc} \cos \phi & \sin \phi \\ \sin (-\phi ) & \cos \phi \end {array} \right )$, where $\phi$ is the optical axis angle as shown in Fig. 1. Then, we have
$$\sigma _{pp} =\sigma _{xx}\cos ^{2}\phi +\sigma _{yy}\sin ^{2}\phi ,$$
$$\sigma _{ss} =\sigma _{yy}\cos ^{2}\phi +\sigma _{xx}\sin ^{2}\phi ,$$
$$\sigma _{ps} =\sigma _{sp}=\left( \sigma _{xx}-\sigma _{yy}\right) \cos \phi \sin \phi ,$$
where $M_{1}$ and $M_{2}$ are coordinate transformation matrices, $\sigma _{pp}$, $\sigma _{ss}$, and $\sigma _{ps}$ ($\sigma _{sp}$) denote the transverse, longitudinal, and crossed conductivity, respectively. Notably, the crossed conductivity is none zero due to the highly anisotropic optical conductivities, i.e., $\sigma _{x x}\neq \sigma _{yy}$, which will result in a finite in-plane spin Hall shift as shall be discussed later.

According to the electromagnetic boundary condition, the Fresnel reflection coefficients are obtained as [22]

$$r_{pp} =\frac{M_{+}^{1}M_{-}^{2}+N}{M_{+}^{1}M_{+}^{2}+N},$$
$$r_{ss} =-\frac{M_{-}^{1}M_{+}^{2}+N}{M_{+}^{1}M_{+}^{2}+N},$$
$$r_{ps} =-r_{sp}=\frac{R}{M_{+}^{1}M_{+}^{2}+N}.$$

Here $M_{\pm }^{1}=k_{tz}\pm k_{iz}+\omega \mu _{0}\sigma _{ss}$ and $M_{\pm }^{2}=\left ( k_{iz}\varepsilon \pm k_{tz}\varepsilon _{0}+k_{iz}k_{tz}\sigma _{pp}/\omega \right ) /\varepsilon _{0}$, $N=-\mu _{0}k_{iz}k_{tz}\sigma _{ps}^{2}/\varepsilon _{0}$, $R=2Z_{0}k_{iz}k_{tz}\sigma _{ps}$, $k_{iz}=k_{i}\cos \theta _{i}$ and $k_{tz}=k_{t}\cos \theta _{t}$, $k_{i}=k_{t}=\omega \left ( \varepsilon _{0}\mu _{0}\right ) ^{1/2}$, $Z_{0}= (\mu _{0}/\varepsilon _{0})^{1/2}$, and $\cos \theta _{t}=(1-\sin ^{2}\theta _{i}/4)^{1/2}$. While, $\theta _{i}\left ( \theta _{t}\right )$ is the incidence (refraction) angle, $\varepsilon _{0}\left ( \mu _{0}\right )$ the permittivity (permeability) in the vacuum, $\varepsilon _{1}$ the permittivity of refractive media, and $Z_{0}$ the impedance in the vacuum. The $r_{ps}$ and $r_{sp}$ sensitively depend on the crossed conductivity. A finite $r_{ps}$ and $r_{sp}$ require none zero crossed conductivity $\sigma _{sp}$.

Considering horizontal polarization $(H)$ and vertical polarization $(V)$ Gaussian incident beam reflected by the surface between the air and semi-Dirac materials, the angular spectrum of the incident Gaussian beam in momentum space is defined by

$$\widetilde{E}_{i}=\left(\begin{array}{c} \widetilde{E}_{i}^{H} \\ \widetilde{E}_{i}^{V} \end{array}\right)=\frac{w_{0}}{\sqrt{2 \pi}} \exp \left[-\frac{w_{0}^{2}\left(k_{i x}^{2}+k_{i y}^{2}\right)}{4}\right]\left(\begin{array}{c} \cos \alpha_{i} \\ \sin \alpha_{i} \end{array}\right),$$
where $w_{0}$ is the beam waist, $\alpha _{i}$ the polarization angle, and ${E}_{i}^{H}$ and ${E}_{i}^{V}$ the horizontal and vertical components of incident electric field, respectively. In our calculation, the beam waist is chosen as $w_{0}=20 \lambda _{i}$, where $\lambda _{i}$ is the wavelength of the incident photon. The polarization states of wave packets are determined by [42]
$$\left(\begin{array}{c} \widetilde{E}_{r}^{H} \\ \widetilde{E}_{r}^{V} \end{array}\right)=\left(\begin{array}{cc} r_{p p} & r_{p s}+\frac{k_{i y}\left(r_{r p}+r_{s s}\right) \cot \theta_{i}}{k_{0}} \\ r_{s p}-\frac{k_{i y}\left(r_{p p}+r_{s s}\right) \cot \theta_{i}}{k_{0}} & r_{s s} \end{array}\right)\left(\begin{array}{c} \widetilde{E}_{i}^{H} \\ \widetilde{E}_{i}^{V} \end{array}\right),$$
where $k_{0}=\omega /c$ is the wave vector in vacuum, $k_{ix}$ and $k_{iy}$ ( $k_{rx}$ and $k_{ry}$) represent the wave-vector components of incident (reflected) beam along $x_{i}$ and $y_{i}$ ($x_{r}$ and $y_{r}$) axes, respectively. The boundary conditions are $k_{rx}=-k_{ix}$ and $k_{ry}=k_{iy}$. To obtain accurate spin Hall shifts, we Taylor expand the Fresnel coefficients $r_{ab}$ to the first order, which are given by
$$r_{a b}=r_{a b}(\theta_{i})+\frac{\partial r_{a b}}{\partial \theta_{i}} \cdot \frac{k_{i x}}{k_{0}},$$
where $ab \in \{p p, s s, p s, s p\}$. In the spin basis set
$$\widetilde{E}_{r\pm}=\widetilde{E}_{r}^{H}\pm i \widetilde{E}_{r}^{V},$$
where $|+\rangle$ and $|-\rangle$ stand for the left- and right-circular polarization components, respectively. Then, the centroid of the beams for a given plane ($z$ = constant) is [43]
$$\left\langle\mathbf{r}_{{\perp}}\right\rangle=\frac{\left\langle\tilde{\mathbf{E}}\left|i \partial_{\mathbf{k}_{{\perp}}}\right| \tilde{\mathbf{E}}\right\rangle}{\langle\tilde{\mathbf{E}} \mid \tilde{\mathbf{E}}\rangle}+\frac{z}{n k} \frac{\left\langle\tilde{\mathbf{E}}\left|\mathbf{k_{0}}_{{\perp}}\right| \tilde{\mathbf{E}}\right\rangle}{\langle\tilde{\mathbf{E}} \mid \tilde{\mathbf{E}}\rangle},$$
where $n$ is the refractive index, $\mathbf {r}_{\perp }=x \hat {\mathbf {x}}+y \hat {\mathbf {y}}, \partial _{\mathbf {k}_{\perp }}=\frac {\partial }{\partial k_{rx}} \hat {\mathbf {x}}+\frac {\partial }{\partial k_{ry}} \hat {\mathbf {y}}$. The first and second terms in Eq. (16) are the spatial and angular shifts, respectively. In our work, we focus on the spatial shifts. Among the spatial shifts, the lateral one is usually associated with Goos-Hänchen shift regardless of the polarization of light, and the transverse one is known as Imbert-Fedorov shift, i.e., the spin Hall shifts considered here, indicating the separation of the left and right circular polarized light [12]. Moreover, the Imbert-Fedorov shift has two components, namely the in-plane and the transverse one, in the presence of cross Fresnel coefficients [20].

According to Eqs. (12)–(16) and taking into account the paraxial-approximation, we can evaluate the spin Hall shifts of the reflected beam. Generally, the spin Hall shifts are produced by horizontally and vertically polarized incidence [44]. The shift contributed by the vertically polarized incidence is usually small and can be neglected. Hence, we can only focus on the shifts of the horizontal polarization. The in-plane and transverse spin Hall shifts are

$$\left\langle x_{r \pm}\right\rangle={\mp} \frac{1}{k_{0}} \frac{\left(\operatorname{Re}\left[r_{p p} \frac{\partial r_{p s}^{*}}{\partial \theta}-r_{p s}^{*} \frac{\partial r_{p p}}{\partial \theta}\right] \pm \operatorname{Im}\left[r_{p p} \frac{\partial r_{p p}^{*}}{\partial \theta}+r_{p s} \frac{\partial r_{p s}^{*}}{\partial \theta}\right]\right)}{\left(\left|r_{p p}\right|^{2} \mp 2 \operatorname{Im}\left[r_{p s}^{*} r_{p p}\right]+\left|r_{p s}\right|^{2}\right)+\frac{\left|r_{p p}+r_{s s}\right|^{2} \cot ^{2} \theta}{w_{0}^{2} k_{0}^{2}}},$$
$$\left\langle y_{r \pm}\right\rangle={\pm} \frac{\cot \theta}{k_{0}} \frac{\left(\left|r_{p p}\right|^{2}+\operatorname{Re}\left[r_{p p}^{*} r_{s s}\right] \pm \operatorname{Im}\left[r_{p s}\left(r_{p p}^{*}+r_{s s}^{*}\right)\right]\right)}{\left(\left|r_{p p}\right|^{2} \mp 2 \operatorname{Im}\left[r_{p s}^{*} r_{p p}\right]+\left|r_{p s}\right|^{2}\right)+\frac{\left|r_{p p}+r_{s s}\right|^{2} \cot ^{2} \theta}{w_{0}^{2} k_{0}^{2}}},$$
where the subscripts $+/-$ indicates the left/right polarized beams. As expressed in Eqs. (17)–(18), the shifts of the left and right circular polarized light are not symmetric relative to the incident point. The separation between the left and right circular polarized light along the in-plane and transverse direction is
$$\left\langle\Delta X\right\rangle=\left\langle x_{r+}\right\rangle-\left\langle x_{r-}\right\rangle, \left\langle\Delta Y\right\rangle=\left\langle y_{r+}\right\rangle-\left\langle y_{r-}\right\rangle,$$
which is just the total shifts usually detected in real experiment.

3. Results and discussions

In this section, we present the numerical results for the conductivity spectra, Fresnel coefficients and photonic spin Hall shifts as a function of photon energy or optical axis angle. Hereafter, unless explicitly specified, the conductivities are all in units of $\sigma _0$ = 2$e^2/h$, the temperature is set as T = 4 K, the level broadening factor is $\Gamma$=1 meV, and the Fermi energy is $E_F$ =0 eV.

First, we consider how the conductivities vary with the photon energy. Based on our calculation, we find the intraband transition is very weak and it can be ignored for certain Fermi energy, which is consistent with previous work [34,35]. Hence, we mainly present the conductivity contributed by interband transitions. Figure 2 plots the real and imaginary part of the conductivity as a function of photon energy with different optical axis angle $\phi$. According to Eqs. (6)–(7), the conductivities $\sigma _{pp}$ and $\sigma _{ss}$ equal to $\sigma _{yy}$ and $\sigma _{xx}$, respectively, when the optical axis angle $\phi$ is $0^{\circ }$. As shown in Figs. 2(a)-(b), arising from the highly anisotropic dispersion expressed by Eq. (2), the optical conductivity ($\sigma _{yy}$) excited by linearly polarized light along the linear dispersion direction is about two order larger than ($\sigma _{xx}$) excited by linearly polarized light along the parabolic dispersion direction in low frequency regime. By contrast, they tend to be the same in high frequency regime. For $\phi =0^{\circ }$, originating from the highly anisotropic behavior of $\sigma _{xx}$ and $\sigma _{yy}$, the longitudinal conductivity $\sigma _{pp}$ is also about two order larger than the transverse conductivity $\sigma _{ss}$ in low frequency regime. However, the cross conductivity $\sigma _{ps}$ is zero in this case according to Eq. (8) , which means there is no in-plane spin Hall shifts [22]. For $\phi =60^{\circ }$ [see Figs. 2(c)-(d)], the longitudinal conductivity $\sigma _{pp}$ is dominated by the optical conductivity $\sigma _{yy}$. It is larger than the transverse conductivity $\sigma _{ss}$ in this case. More importantly, there is a finite cross conductivity $\sigma _{ps}$ which induces the coupling between p- and s-polarised waves, due to the anisotropic optical conductivities ($\sigma _{xx}$ and $\sigma _{yy}$) resulting from the highly anisotropic dispersion.

 figure: Fig. 2.

Fig. 2. The real (a) and imaginary (b) part of the conductivity as a function of photon energy with optical axis angle $\phi =0^{\circ }$ and chemical potential $E_{F}$=0 eV. (c)/(d) is the counterpart to (a)/(b) with optical axis angle $\phi =60^{\circ }$.

Download Full Size | PDF

Figures 3(a)-(b) plot the real and imaginary part of the conductivity as a function of Fermi energy with photon energy $\hbar \omega$=0.3 eV and optical axis angle $\phi =60^{\circ }$. As shown in the figures, two different regimes in the conductivity spectra can be identified. There is a prominent inflection point at the half of the photon energy ($\hbar \omega$/2=0.15 eV). When the Fermi energy is lower (higher) than half of the photon energy [see Eq. (3)], the conductivity is dominated by the interband (intraband) transitions [34]. The result shows that the intraband conductivity is much weaker than the interband one. In our work, we focus on the PSHE on the surface between air and the pristine semi-Dirac material ($E_{F}$=0 eV). In this case, the electromagnetic response of the system is dominated by the interband transitions. For certain photon energy and Fermi level, the conductivity can be modulated by the optical axis angle as shown in Figs. 3(c)-(d). The conductivities versus optical axis angle is a trigonometric function. Importantly, there is a finite crossed conductivity $\sigma _{ps}$($\sigma _{sp}$) except for $\phi =0^{\circ }$ and $180^{\circ }$ arising from the intrinsic anisotropy of the semi-Dirac materials, which means it can be effectively tuned by the optical axis angle. The crossed conductivity will result in the in-plane and transverse spin Hall shifts and play an important role in spin-orbit interaction of light.

 figure: Fig. 3.

Fig. 3. The real (a) and imaginary (b) part of the conductivity as a function of Fermi energy with photon energy $\hbar \omega$=0.3 eV and optical axis angle $\phi =60^{\circ }$. (c)/(d) counterpart to (a)/(b) as a function of Fermi energy with photon energy $\hbar \omega$=0.3 eV and Fermi energy $E_{F}$=0 eV.

Download Full Size | PDF

Substituting various conductivities into Eqs. (9)–(11), we obtain the Fresnel coefficients which are directly related to the PSH shifts. Figs. 4(a)-(b) plot the Fresnel coefficients and the ratio between the transverse and longitudinal one $|r_{ss}/r_{pp}|$ as a function of the incident angle with photon energy $\hbar \omega$=0.3eV and optical axis angle $\phi =60^{\circ }$. From Fig. 4(a), we find the Brewster angle is $\theta _i=63.4^{\circ }$ where the longitudinal Fresnel coefficient ($|r_{pp}|$) takes the minimum [8]. However, the transverse Fresnel coefficient $|r_{ss}|$ increases with the increase of incident angle. It equals to 1 for glancing incidence ($\theta _i=90^{\circ }$) according to Eqs. (10). Importantly, there is a finite cross Fresnel coefficient ($|r_{sp}|$) arising from the anisotropic optical response of semi-Dirac material. Although, $|r_{sp}|$ is very small compared with $|r_{pp}|$ or $|r_{ss}|$, it directly determines whether there is an in-plane spin Hall shift and plays important role in the PSHE according to Eqs. (17)–(18). Further, the first derivative of cross Fresnel coefficient is closely related to the magnitude of the in-plane spin Hall shift. The first derivative takes the maximum at the Brewster angle, corresponding to the inflection point of $|r_{sp}|$. Another quantity closely related to the magnitude of the spin Hall shift is the ratio between the longitudinal Fresnel coefficient and the transverse one $|r_{ss}/r_{pp}|$ as depicted in Fig. 4(b). The ratio $|r_{ss}/r_{pp}|$ has a peak around the Brewster angle because the transverse Fresnel coefficient $|r_{pp}|$ takes the minimum at Brewster angle [see Fig. 4(a)]. Since both the first derivative of $|r_{sp}|$ and the ratio $|r_{ss}/r_{pp}|$ take the maximum at Brewster angle, hereafter we only focus on the spin Hall shifts around the Brewster angle. Figures 4(c)-(d) plot the Fresnel coefficients $r_{pp}$ and the ratio $|r_{ss}/r_{pp}|$ as functions of photon energy and optical axis angle by setting the incident angle near the Brewster angle. As shown in the figure, $|r_{pp}|$ and $|r_{ss}/r_{pp}|$ sensitively depend on the optical axis angle and photon energy because the reflection coefficients are closely related to the conductivity matrix. The $|r_{pp}|$ has a similar variation tendency with the $\sigma _{pp}$ as a function of optical axis angle.

 figure: Fig. 4.

Fig. 4. (a) the Fresnel coefficients and (b) the ratio between the transverse and longitudinal one $|r_{ss}/r_{pp}|$ as a function of the incident angle with photon energy $\hbar \omega$=0.3eV and optical axis angle $\phi =60^{\circ }$. The Fresnel’s coefficients (c) $r_{pp}$ and (d) the ratio $|r_{ss}/r_{pp}|$ as functions of photon energy and optical axis angle with incident angle $\theta _i=63.4^{\circ }$.

Download Full Size | PDF

To verify whether the PSH shifts take the maximum at the Brewster angle, we consider the total PSH shifts as a function of incident angle for certain photon energy and optical axis angle as presented in Fig. 5. From the figure, apart from the usually observed transverse shift, we find that there is a large in-plane shift due to the anisotropic optical response of the semi-Dirac material, which is similar to that in hyperbolic metamaterials [13,14,16,17] and tilted linear-dichroic plates [12,15], and monolayer black phosphorus [20]. Both the in-plane and transverse spin Hall shifts take the maximum near the Brewster’s angle because the first derivative of $r_{sp}$ and the ratio $|r_{ss}/r_{pp}|$ take the maximum at the Brewster angle. The maximum of the total in-plane and transverse spin Hall shift is 83.91$\lambda _i$ and 19.65$\lambda _i$, respectively, where $\lambda _i$ is the wavelength of the incident photon. In particular, for photon energy $\hbar \omega$=0.03 eV, the wavelength $\lambda _i$=41 $\mu$m. Then, the maximum of the total in-plane shift ($\langle \Delta X\rangle$) is 3.4 mm, which may be observable directly. However, the maximum of ($\langle \Delta Y\rangle$) is smaller than the beam diameter (2$w_0$=40$\lambda _i$) and it is difficult to observe them directly. Fortunately, the quantum weak measurement technique can amplify the spin Hall shifts and detect them precisely [10,11]. The first derivative of transverse Fresnel coefficients $\partial r_{pp} / \partial \theta _{i}$ also takes the maximum at the Brewster angle, which can be inferred from Fig. 4(a). Moreover, according to Eq. (18), the magnitude of transverse shift is closely related to the value of the ratio $r_{ss}/r_{pp}$.

 figure: Fig. 5.

Fig. 5. The total (a) in-plane and (b) transverse spin-dependent shifts in unit of incident wavelength $\lambda _i$ as a function of the incident angle with optical axis angle $\phi =5^{\circ }$, $0^{\circ }$ and photon energy $\hbar \omega$=0.03, 0.15 eV, respectively.

Download Full Size | PDF

Figure 6 shows the total (a) in-plane $\left \langle \Delta X\right \rangle$ and (b) transverse $\left \langle \Delta Y\right \rangle$ shifts on the surface of semi-Dirac materials as functions of optical axis angle and photon energy with incident angle setting near the Brewster angle and Fermi energy $E_F$=0 eV. As shown in the figure, the $\left \langle \Delta X\right \rangle$ is zero for $\phi =0$ or $\pi$ due to the absence of the cross conductivity, which can also be inferred from Eqs. (8), (11), (17) and (19). The $\left \langle \Delta X\right \rangle$ ($\left \langle \Delta Y\right \rangle$) is antisymmetrical (symmetrical) with respect to $\phi =\pi /2$ under certain photon energy. The total in-plane spin Hall shift is larger than the transverse one arising from the anisotropic optical response of semi-Dirac material. The spin Hall shifts depend more sensitively on the optical axis angle than the energy of the incident photon. The $\left \langle \Delta Y\right \rangle$ takes the maximum when the optical axis angle is set around $\phi ={0}^{\circ }$ or ${180}^{\circ }$ while the $\left \langle \Delta X\right \rangle$ takes the maximum when the optical axis angle is set around $\phi ={5}^{\circ }$ or ${175}^{\circ }$. The maximum of $\left \langle \Delta X\right \rangle$ and $\left \langle \Delta Y\right \rangle$ is 83.91$\lambda _i$ and 19.65$\lambda _i$, respectively, where $\lambda _i$ is the wavelength of the incident photon. The in-plane and transverse shift together contribute to a large PSH shift, when compared with those in traditional anisotropic two-dimensional materials [20] and isotropic Dirac materials [18,19,24]. This large photonic spin Hall shift is easier to be measured in real experiment. It is also useful in studying the spin-orbit coupling of light and designing spin-optical devices based on semi-Dirac materials.

 figure: Fig. 6.

Fig. 6. The total (a) in-plane and (b) transverse shifts on the surface between air and the semi-Dirac materials as functions of optical axis angle and photon energy with incident angle $\theta _i=63.4^{\circ }$ and Fermi energy $E_F$=0 eV.

Download Full Size | PDF

4. Conclusions

In conclusion, we established a general model to study the PSHE on the surface between air and semi-Dirac materials. We found there is a large PSHE originating from the intrinsic anisotropy of electronic states in semi-Dirac materials. Owing to the highly anisotropic optical response of semi-Dirac materials, there is a large in-plane spin Hall shift compared with that in the isotropic systems [22,23]. It was found that the in-plane and transverse spin Hall shift are dozens times of the incident wavelength, which can be effectively tuned by the optical axis angle and the energy of incident photon. The maximum of the total in-plane and transverse spin Hall shift is 83.91$\lambda _i$ and 19.65$\lambda _i$, respectively, resulting in a large PSH shift compared with those in conventional anisotropic two-dimensional materials [20] and isotropic Dirac materials [18,19,24]. This large PSH shift is more conveniently to be measured in real experiment. Our result is useful in understanding the spin-orbit coupling of light in semi-Dirac material and designing spin-optical devices based on it.

Funding

National Natural Science Foundation of China (Grant Nos. 11804092, and 11774085); China Postdoctoral Science Foundation (Grant Nos. BX20180097, and 2019M652777).

Acknowledgments

The authors are grateful to Dr. Xinxing Zhou for insight discussions.

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. X. B. Yin, Z. l. Ye, J. Rho, Y. Wang, and X. Zhang, “Photonic Spin Hall Effect at Metasurfaces,” Science 339(6126), 1405–1407 (2013). [CrossRef]  

2. Y. C. Liu, Y. G. Ke, H. L. Luo, and S. C. Wen, “Photonic spin Hall effect in metasurfaces: a brief review,” Nanophotonics 6(1), 51–70 (2017). [CrossRef]  

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

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

5. K. Y. Bliokh, F. J. Rodriguez-Fortuno, F. Nori, and A. V. Zayats, “Spin-orbit interactions of light,” Nat. Photonics 9(12), 796–808 (2015). [CrossRef]  

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

7. W. J. Long, J. T. Pan, X. Y. Guo, X. H. Liu, H. Lin, H. D. Zheng, J. H. Yu, H. Y. Guan, H. H. Lu, Y. C. Zhong, S. H. Fu, L. Zhang, W. G. Zhu, and Z. Chen, “Optimized weak measurement of orbital angular momentum-induced beam shifts in optical reflection,” Photonics Res. 7(11), 1273–1278 (2019). [CrossRef]  

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

9. X. X. Zhou, Z. C. Xiao, H. l. Luo, and S. C. Wen, “Experimental observation of the spin Hall effect of light on a nanometal film via weak measurements,” Phys. Rev. A 85(4), 043809 (2012). [CrossRef]  

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

11. Y. Aharonov, D. Z. Albert, and L. Vaidman, “How the Result of a Measurement of a Component of the Spin of a Spin-$\frac{1}{2}$ Particle Can Turn Out to be 100,” Phys. Rev. Lett. 60(14), 1351–1354 (1988). [CrossRef]  

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

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

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

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

16. M. Kim, D. Lee, T. H. Kim, Y. Yang, H. J. Park, and J. Rho, “Observation of Enhanced Optical Spin Hall Effect in a Vertical Hyperbolic Metamaterial,” ACS Photonics 6(10), 2530–2536 (2019). [CrossRef]  

17. M. Mazanov, O. Yermakov, I. Deriy, O. Takayama, A. Bogdanov, and A. V. Lavrinenko, “Photonic Spin Hall Effect: Contribution of Polarization Mixing Caused by Anisotropy,” Quantum Rep. 2(4), 489–500 (2020). [CrossRef]  

18. T. T. Tang, J. Li, L. Luo, P. Sun, and J. Q. Yao, “Magneto-Optical Modulation of Photonic Spin Hall Effect of Graphene in Terahertz Region,” Adv. Opt. Mater. 6(7), 1701212 (2018). [CrossRef]  

19. 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 graphene,” Phys. Rev. A 95(1), 013809 (2017). [CrossRef]  

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

21. G. Y. Jia, G. Li, Y. Zhou, X. L. Miao, and X. Y. Zhou, “Landau quantisation of photonic spin Hall effect in monolayer black phosphorus,” Nanophotonics 9(1), 225–233 (2019). [CrossRef]  

22. M. X. Liu, L. Cai, S. Z. Chen, Y. C. Liu, H. L. Luo, and S. C. Wen, “Strong spin-orbit interaction of light on the surface of atomically thin crystals,” Phys. Rev. A 95(6), 063827 (2017). [CrossRef]  

23. T. Low, A. Chaves, J. D. Caldwell, A. Kumar, N. X. Fang, P. Avouris, T. F. Heinz, F. Guinea, L. Martin-Moreno, and F. Koppens, “Polaritons in layered two-dimensional material,” Nat. Mater. 16(2), 182–194 (2017). [CrossRef]  

24. X. X. Zhou, J. Zhang, X. H. Ling, S. Z. Chen, H. L. Luo, and S. C. Wen, “Photonic spin Hall effect in topological insulators,” Phys. Rev. A 88(5), 053840 (2013). [CrossRef]  

25. S. Banerjee and W. E. Pickett, “Phenomenology of a semi-Dirac semi-Weyl semimetal,” Phys. Rev. B 86(7), 075124 (2012). [CrossRef]  

26. F. Zhai and J. Wang, “Shot noise in systems with semi-Dirac points,” J. Appl. Phys. 116(6), 063704 (2014). [CrossRef]  

27. Z. Li, H. Cao, and L. B. Fu, “Zitterbewegung for ultracold atoms in the merging of Dirac points,” Phys. Rev. A 91(2), 023623 (2015). [CrossRef]  

28. Q. H. Liu, X. W. Zhang, L. B. Abdalla, A. Fazzio, and A. Zunger, “Switching a Normal Insulator into a Topological Insulator via Electric Field with Application to Phosphorene,” Nano Lett. 15(2), 1222–1228 (2015). [CrossRef]  

29. B. Ghosh, B. Singh, R. Prasad, and A. Agarwal, “Electric-field tunable Dirac semimetal state in phosphorene thin films,” Phys. Rev. B 94(20), 205426 (2016). [CrossRef]  

30. H. J. Doh and H. J. Choi, “Dirac-semimetal phase diagram of two-dimensional black phosphorus,” 2D Mater. 4(2), 025071 (2017). [CrossRef]  

31. P. Dietl, F. Piechon, and G. Montambaux, “New Magnetic Field Dependence of Landau Levels in a Graphenelike Structure,” Phys. Rev. Lett. 100(23), 236405 (2008). [CrossRef]  

32. G. Montambaux, F. Piéchon, J. N. Fuchs, and M. O. Goerbig, “Merging of Dirac points in a two-dimensional crystal,” Phys. Rev. B 80(15), 153412 (2009). [CrossRef]  

33. J. Kim, S. S. Baik, S. H. Ryu, Y. Sohn, S. Park, B.-G. Park, J. Denlinger, Y. Yi, H. J. Choi, and K. S. Kim, “Observation of tunable band gap and anisotropic Dirac semimetal state in black phosphorus,” Science 349(6249), 723–726 (2015). [CrossRef]  

34. J. P. Carbotte, K. R. Bryenton, and E. J. Nicol, “Optical properties of a semi-Dirac material,” Phys. Rev. B 99(11), 115406 (2019). [CrossRef]  

35. J. H. Jang, S. Ahn, and H. Min, “Optical conductivity of black phosphorus with a tunable electronic structure,” 2D Mater. 6(2), 025029 (2019). [CrossRef]  

36. J. L. Ren, B. Wang, Y. F. Xiao, Q. H. Gong, and Y. Li, “Direct observation of a resolvable spin separation in the spin Hall effect of light at an air-glass interface,” Appl. Phys. Lett. 107(11), 111105 (2015). [CrossRef]  

37. X. X. Zhou, L. G Xie, S. J. Cheng, X. H. Ling, Z. Y. Zhang, H. L. Luo, and H. D. Sun, “Large in-plane asymmetric spin angular shifts of a light beam near the critical angle,” Opt. Lett. 44(2), 207–210 (2019). [CrossRef]  

38. H. L. Dai, L. Q. Yuan, C. Yin, Z. Q. Cao, and X. F. Chen, “Direct Visualizing the Spin Hall Effect of Light via Ultrahigh-Order Modes,” Phys. Rev. Lett. 124(5), 053902 (2020). [CrossRef]  

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

40. S. Z. Chen, X. H. Ling, W. X. Shu, H. L. Luo, and S. C. Wen, “Precision Measurement of the Optical Conductivity of Atomically Thin Crystals via the Photonic Spin Hall Effect,” Phys. Rev. A 13(1), 014057 (2020). [CrossRef]  

41. S. Barati and S. H. Abedinpour, “Optical conductivity of three and two dimensional topological nodal-line semimetals,” Phys. Rev. B 96(15), 155150 (2017). [CrossRef]  

42. S. Z. Chen, C. Q. Mi, W. J. Wu, W. S. Zhang, W. X. Shu, H. L. Luo, and S. C. Wen, “Weak-value amplification for Weyl-point separation in momentum space,” New J. Phys. 20(10), 103050 (2018). [CrossRef]  

43. Z. C. Xiao, H. L. Luo, and S. C. Wen, “Goos-Hänchen and Imbert-Fedorov shifts of vortex beams at air-left-handed-material interfaces,” Phys. Rev. A 85(5), 053822 (2012). [CrossRef]  

44. Z. R. Qin, Q. G. Liu, C. Yue, and Y. P. Lang, “Modified model of photonic spin Hall effect of Gaussian beam reflected from a dielectric interface,” Appl. Phys. Express 12(6), 062008 (2019). [CrossRef]  

Data availability

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

Cited By

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

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. (a) Schematic illustration on the beam reflection on the surface of semi-Dirac materials in a Cartesian coordinate system with the incident angle $\theta _{i}$. The optical axis angle $\phi$ is defined as the angle between the $xoz$ plane and the incident plane which is determined by the incident light and the $z$-axis. The PSHE occurs on the reflecting surface and exhibits in-plane and transverse Hall shifts. (b) Schematic plot of the low energy dispersion of semi-Dirac electrons.
Fig. 2.
Fig. 2. The real (a) and imaginary (b) part of the conductivity as a function of photon energy with optical axis angle $\phi =0^{\circ }$ and chemical potential $E_{F}$=0 eV. (c)/(d) is the counterpart to (a)/(b) with optical axis angle $\phi =60^{\circ }$.
Fig. 3.
Fig. 3. The real (a) and imaginary (b) part of the conductivity as a function of Fermi energy with photon energy $\hbar \omega$=0.3 eV and optical axis angle $\phi =60^{\circ }$. (c)/(d) counterpart to (a)/(b) as a function of Fermi energy with photon energy $\hbar \omega$=0.3 eV and Fermi energy $E_{F}$=0 eV.
Fig. 4.
Fig. 4. (a) the Fresnel coefficients and (b) the ratio between the transverse and longitudinal one $|r_{ss}/r_{pp}|$ as a function of the incident angle with photon energy $\hbar \omega$=0.3eV and optical axis angle $\phi =60^{\circ }$. The Fresnel’s coefficients (c) $r_{pp}$ and (d) the ratio $|r_{ss}/r_{pp}|$ as functions of photon energy and optical axis angle with incident angle $\theta _i=63.4^{\circ }$.
Fig. 5.
Fig. 5. The total (a) in-plane and (b) transverse spin-dependent shifts in unit of incident wavelength $\lambda _i$ as a function of the incident angle with optical axis angle $\phi =5^{\circ }$, $0^{\circ }$ and photon energy $\hbar \omega$=0.03, 0.15 eV, respectively.
Fig. 6.
Fig. 6. The total (a) in-plane and (b) transverse shifts on the surface between air and the semi-Dirac materials as functions of optical axis angle and photon energy with incident angle $\theta _i=63.4^{\circ }$ and Fermi energy $E_F$=0 eV.

Equations (19)

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

H = ( 0 2 k x 2 2 m i v F k y 2 k x 2 2 m + i v F k y 0 ) ,
E ± = ± ( 2 k x 2 / 2 m ) 2 + 2 v F 2 k y 2 ,
σ μ ν = e 2 i S k , k f ( E k ) f ( E k ) ( E k E k ) k | v μ | k k | v ν | k ( E k E k + ω + i Γ ) ,
σ ( ω ) = ( σ x x σ x y σ y x σ y y ) .
σ M = M 1 σ M 2 = ( σ p p σ p s σ s p σ s s ) ,
σ p p = σ x x cos 2 ϕ + σ y y sin 2 ϕ ,
σ s s = σ y y cos 2 ϕ + σ x x sin 2 ϕ ,
σ p s = σ s p = ( σ x x σ y y ) cos ϕ sin ϕ ,
r p p = M + 1 M 2 + N M + 1 M + 2 + N ,
r s s = M 1 M + 2 + N M + 1 M + 2 + N ,
r p s = r s p = R M + 1 M + 2 + N .
E ~ i = ( E ~ i H E ~ i V ) = w 0 2 π exp [ w 0 2 ( k i x 2 + k i y 2 ) 4 ] ( cos α i sin α i ) ,
( E ~ r H E ~ r V ) = ( r p p r p s + k i y ( r r p + r s s ) cot θ i k 0 r s p k i y ( r p p + r s s ) cot θ i k 0 r s s ) ( E ~ i H E ~ i V ) ,
r a b = r a b ( θ i ) + r a b θ i k i x k 0 ,
E ~ r ± = E ~ r H ± i E ~ r V ,
r = E ~ | i k | E ~ E ~ E ~ + z n k E ~ | k 0 | E ~ E ~ E ~ ,
x r ± = 1 k 0 ( Re [ r p p r p s θ r p s r p p θ ] ± Im [ r p p r p p θ + r p s r p s θ ] ) ( | r p p | 2 2 Im [ r p s r p p ] + | r p s | 2 ) + | r p p + r s s | 2 cot 2 θ w 0 2 k 0 2 ,
y r ± = ± cot θ k 0 ( | r p p | 2 + Re [ r p p r s s ] ± Im [ r p s ( r p p + r s s ) ] ) ( | r p p | 2 2 Im [ r p s r p p ] + | r p s | 2 ) + | r p p + r s s | 2 cot 2 θ w 0 2 k 0 2 ,
Δ X = x r + x r , Δ Y = y r + y r ,
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.