This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Electron beam splitting at topological insulator surface states and a proposal for electronic Goos-Hänchen shift measurement

Hassan Ghadiri Department of Physics, North Tehran Branch, Islamic Azad University, 16511-53311, Tehran, Iran    Alireza Saffarzadeh [email protected] Department of Physics, Payame Noor University, P.O.Box 19395-3697 Tehran, Iran Department of Physics, Simon Fraser University, Burnaby, British Columbia, Canada V5A 1S6
Abstract

The hexagonal warping effect on transport properties and Goos-Hänchen (GH) lateral shift of electrons on the surface of a topological insulator with a potential barrier is investigated theoretically. Due to the warped Fermi surface for incident electron beams, we can expect two propagating transmitted beams corresponding to the occurrence of double refraction. The transmitted beams have spin orientations locked to their momenta so that one of the spin directions rotates compared to the incident spin direction. Based on a low-energy Hamiltonian near the Dirac point and considering Gaussian beams, we derive expressions for calculating lateral shifts in the presence of warping effect. We study the dependence of transmission probabilities and GH shifts of transmitted beams on system parameters in detail by giving an explanation for the appearance of large peaks in the lateral shifts corresponding to their transmission peaks. It is shown that the separation between two transmitted beams through their different GH shifts can be as large as a few micrometers which is large enough to be observed experimentally. Finally, we propose a method to measure the GH shift of electron beams based on the transverse magnetic focusing technique in which by tuning an applied magnetic field a detectable resonant path for electrons can be induced.

I Introduction

Topological insulators (TIs) are nonmagnetic insulators with conducting surface states as a consequence of the nontrivial topology of their bulk band structure, which in turn results from strong spin-orbit interaction. The surface states contain an odd number of spin-helical Dirac cones and are protected against any disturbance that maintains time-reversal symmetry Hasan ; Qi ; Hasan1 . In the vicinity of the Dirac point, the electron states can be well-described by massless Dirac equation, whereas at energies far enough away from the Dirac point, a distortion induced by surface spin-orbit coupling deforms the Fermi surface into a hexagonal snowflake shape Fu-2009 ; Chen-2009 ; Kuroda . This effect, called hexagonal warping, has been confirmed by angle-resolved photoemission spectroscopy Chen-2009 . Since the surface states close to the Fermi level play a decisive role in the electronic properties of two-dimensional (2D) materials, the hexagonal warping can affect transport properties on the surface of TIs. Therefore, the warping effect and also topologically protection of surface states may lead to a variety of interesting properties which are important from viewpoint of fundamental physics as well as novel device applications C. M. Wang ; J. Wang ; Pal ; An-2012 ; Rakyta ; Siu ; Roy ; Hassler ; Repin ; Yu ; Akzyanov ; Akzyanov1 .

It is well known that the totally reflected light beam from a dielectric interface undergoes a lateral displacement from the position predicted by the geometrical optics. The study of this phenomenon which is known as Goos-Hänchen (GH) shift Goos has been developed to partial reflections and also transmitting configuration Tamir ; Hsue ; Li ; Broe . When an electron beam is incident on a boundary separating two regions of different densities, the reflected/transmitted beam undergoes a GH shift similar to a light beam crossing a boundary between materials with different optical indices. Accordingly, the GH shift of electrons in condensed matter systems Wilson ; X. Chen ; X. Chen1 ; Gruszecki ; Yu2 especially in Dirac materials Beenakker ; Wu ; Zhai ; Ghosh ; Chen2 ; Song ; Chen3 ; Sun ; Ghadiri ; Ghadiri1 ; Qiu ; Azarova ; Kuai ; Jiang ; Zheng ; Liu ; Liu1 ; Chattopadhyay has been extensively studied.

The GH shift and transverse displacement, called Imbert-Fedorov (IF) shift, of a light beam on the surface of some Dirac materials, such as graphene and Weyl semimetals, have also been investigated Wu2017 ; Ye2019 . The results showed that the optical beam shifts provide a possible scheme for direct measurement of the parameters in these materials Wu2017 ; Ye2019 . Moreover, it was shown that the electronic IF shift can be utilized to characterize the Weyl semimetals Jiang . On the other hand, the results of electronic beam shifts (EBS) suggest a new generation of nanoelectronic devices based on transition metal dichalcogenides Sun ; Ghadiri ; Ghadiri1 and Weyl semimetals Jiang ; Zheng . Therefore, the study of EBS on topological insulators along with the ability of measuring EBS can potentially provide applications in characterizing the parameters of TIs as well as the fabrication of new TI-based nanodevices.

In this paper, we investigate the propagation of electrons through a square potential barrier on the surface of a TI by considering the hexagonal warping effect. We show that due to the warped Fermi surface, an electron wave impinging onto the barrier can have two transmitted waves, propagating with different momenta and hence in different directions, much like the double refraction of light in anisotropic crystals, demonstrating another optics-like property of electrons. We derive a formula for calculation of GH shifts of two transmitted beams in the presence of hexagonal warping. We show that the beams can be separated spatially due to their different GH shifts, while at the same time they have different spin orientations due to their different momenta. However, due to the difficulty in producing well collimated electron beam, the GH shift in electronic systems has not been measured so far Chen4 .

We present a proposal for experimentally measuring the GH shift based on the transverse magnetic focusing (TMF) technique in which by applying a transverse magnetic field one can focus the motion of electrons/holes in the ballistic regime Taychatanapat ; Milovanovic ; Tsoi . The TMF has been used to study the shape of Fermi surfaces Tsoi , Andreev reflection Tsoi ; Bhandari4 , spin-orbit interaction Rokhinson , the angle-resolved transmission probability in graphene S. Chen , imaging electron trajectories Bhandari1 ; Bhandari2 ; Bhandari3 ; Bhandari4 , as well as proposing a method for measuring warping strength in TIs Yu .

In this proposal, by applying a transverse magnetic field on incident region, the impact point and also incident angle of electrons at the first interface are controlled, similar to the experiment of Ref. [S. Chen, ] and also the proposal in Ref. [Yu, ]. The variation of transverse magnetic field applied on the transmission region induces a resonant conduction path (measured as a voltage peak) by which the entry point of electrons at the second interface is determined, and hence, the GH shift can be measured.

The paper is organized as follows. We introduce our model and formalism for obtaining transmission probability in Sec. II, where the scattering wave functions in the presence of warping effect and the transmission properties of incident electron waves are discussed in details. In Sec. III, we calculate the GH shift of electron beams and the spatial beam separation is investigated. In Sec. IV, we present our proposal for measuring the electronic GH shift based on the TMF phenomenon. We conclude our findings in Sec. V.

II Theoretical model and double refraction

Surface states of topological insulators with a single Dirac cone are generally described by the Hamiltonian (=1\hbar=1)

H=υF𝐤(z^×𝝈)+λ2(k+3+k3)σz,H=\upsilon_{F}{\bf k}\cdot(\hat{z}\times\boldsymbol{\sigma})+\frac{\lambda}{2}(k_{+}^{3}+k_{-}^{3})\sigma_{z}\ , (1)

where 𝐤=(kx,ky){\bf k}=(k_{x},k_{y}) is the wave vector of electron, 𝝈\boldsymbol{\sigma} is the Pauli matrix vector, k±=kx±ikyk_{\pm}=k_{x}\pm ik_{y} and z^\hat{z} is a unit vector normal to the surface. υF\upsilon_{F} and λ\lambda are the Fermi velocity and warping parameter, respectively. The linear term in 𝐤{\bf k} being similar to that of graphene, with exception that 𝝈\boldsymbol{\sigma} represents the real spin of electron and cubic terms in the Hamiltonian are responsible for hexagonal warping effect. This Hamiltonian which neglects multi-orbital structure of surface states is considered as a minimal model preserving given C3vC_{3v} symmetries Fu-2009 ; Akzyanov1 . Among different topological insulators, Bi2Te3 is found to have strong warping effect with Fermi velocity υF=2.55\upsilon_{F}=2.55 eV\cdotÅ and the hexagonal warping parameter λ=250\lambda=250 eV\cdotÅ3 that we consider here in the calculations Fu-2009 ; An-2012 . The eigenvalues of Eq. (1) give the upper and lower bands in 𝐤{\bf k} space as

E±(𝐤)=±(υFk)2+w2(𝐤),E_{\pm}({\bf k})=\pm\sqrt{(\upsilon_{F}k)^{2}+w^{2}({\bf k})}\ , (2)

where w(𝐤)=λkx(kx23ky2)w({\bf k})=\lambda k_{x}(k_{x}^{2}-3k_{y}^{2}). We have depicted the Dirac cone of fermions (Eq. (2)) and several constant energy contours (CECs) in Figs. 1(a) and 1(b), respectively. The energy of CEC is expressed in terms of E0=υF3/λE_{0}=\sqrt{\upsilon_{F}^{3}/\lambda} which is the characteristic energy introduced by hexagonal warping. At EE0E\ll E_{0} the warping effect is negligible and the CEC exhibits a circular shape. As the energy exceeds the critical value Ec=7634E00.69E0E_{c}=\frac{\sqrt{7}}{6^{\frac{3}{4}}}E_{0}\approx 0.69E_{0} the CEC deforms into a hexagonal shape, with inflection points satisfying the relation (kxky)E=(2kx2ky)E=0(\frac{\partial k_{x}}{\partial k_{y}})_{E}=(\frac{\partial^{2}k_{x}}{\partial^{2}k_{y}})_{E}=0. With further increasing EE, the rounded tips of the hexagon become sharper and the CEC exhibits a snowflake shape. The eigenspinors of Hamiltonian (1) can be written as

χ(𝐤,E±)ei𝐤𝐫=1N±(𝐤)(±(E±(𝐤)+w(𝐤))υF(ikxky))ei𝐤𝐫,\displaystyle\chi({\bf k},E_{\pm})\mathrm{e}^{i{\bf k}\cdot{\bf r}}=\frac{1}{N_{\pm}({\bf k})}\left(\begin{array}[]{c}\pm(E_{\pm}({\bf k})+w({\bf k}))\\ \upsilon_{F}(ik_{x}-k_{y})\\ \end{array}\right)\mathrm{e}^{i{\bf k}\cdot{\bf r}}, (5)

where N±(𝐤)N_{\pm}({\bf k}) are the normalization coefficients and the subscript +()+(-) corresponds to the upper (lower) band in Eq. (2). Note that the interaction between surface states and bulk states can be ignored since the surface Dirac point on TIs such as Bi2Te3 is closer to the bulk valence band than to the bulk conduction band Chen-2009 .

Refer to caption
Figure 1: (Color online) (a) The Dirac cone of fermions with Fermi energy EF=1.6E0E_{F}=1.6E_{0} on the surface of Bi2Te3 by including the hexagonal warping effect. (b) The constant-energy contours of Dirac cone at different energies measured from the Dirac point.

Now, we consider the propagation of electrons on the surface of TI scattered by a potential barrier V(y)=V0V(y)=V_{0} for 0yd0\leqslant y\leqslant d and V(y)=0V(y)=0 elsewhere (see Fig. 2). Such a barrier can be produced by a gate electrode deposited on top of the TI surface. We note that the electron transport in yy direction is coherent. Also, due to the translational invariance in xx direction, kxk_{x} is a good quantum number, and hence, the Fermi energy EFE_{F} of electron is conserved in the scattering process. In zero potential regions for given EFE_{F} and kxk_{x}, equation E+(𝐤)=EFE_{+}({\bf k})=E_{F} is quartic in terms of kyk_{y} and its roots determine the yy components of electron Fermi momentum. It gives two real symmetric roots and two imaginary symmetric roots in the case of EF<EcE_{F}<E_{c}, indicating that an incoming electron wave with an arbitrary incident angle θ\theta has one propagating reflected wave and one propagating transmitted wave (see Fig. 2a) as it is a normal case for most conventional materials. For EF>EcE_{F}>E_{c}, the CEC has concave segments and consequently, as shown in Fig. 2(b), there exists a critical incident angle θc\theta_{c} beyond which the equation E+(𝐤)=EFE_{+}({\bf k})=E_{F} has four symmetrical real roots. This means that for an incoming electron wave there exist two propagating reflected waves (double reflection) and simultaneously two propagating transmitted waves (double refraction) (see Fig. 2c). When θ<θc\theta<\theta_{c}, similar to the case of EF<EcE_{F}<E_{c}, single refraction happens as can be seen in Fig. 2(b). We should note that at θ>θc\theta>\theta_{c}, the Fermi momenta with bigger absolute values along yy axis are parallel to their corresponding group velocities defined by υy=(Eky)kx\upsilon_{y}=(\frac{\partial E}{\partial k_{y}})_{k_{x}}, therefore the corresponding states are electron-like. In contrast the Fermi momenta with smaller absolute values are antiparallel to their corresponding group velocities indicating hole-like propagating states. In the barrier region, the roots of kyk_{y} are obtained from the equation E+(𝐤)+V0=EFE_{+}({\bf k})+V_{0}=E_{F} depending on the amounts of EF,V0E_{F},V_{0} and incident angle θ\theta. They can be real, complex, or two real roots and two imaginary roots.

In order to obtain transmission probability and for the future purposes, we write down the generic scattering states for given EF,V0E_{F},V_{0} and θ=arctan(|kxky,1|)\theta=\arctan(|\frac{k_{x}}{k_{y,{1}}}|) as

ψ(y)={χ(ky,1,EF)eiky,1y+r1χ(ky,1,EF)eiky,1y+r2χ(ky,2,EF)eiky,2y,y0,n=14anχ(ky,n,EFV0)eiky,ny,0yd,t1χ(ky,1,EF)eiky,1(yd)+t2χ(ky,2,EF)eiky,2(yd),yd,\psi(y)=\left\{\begin{array}[]{ll}\chi(k_{y,1},E_{F})e^{ik_{y,1}y}+r_{1}\chi(-k_{y,1},E_{F})e^{-ik_{y,1}y}\\ +r_{2}\chi(k_{y,2},E_{F})e^{ik_{y,2}y},&\hbox{$y\leq 0$,}\\ \\ \sum_{n=1}^{4}a_{n}\chi(k_{y,n}^{{}^{\prime}},E_{F}-V_{0})e^{ik_{y,n}^{\prime}y},&\hbox{$0\leq y\leq d$,}\\ \\ t_{1}\chi(k_{y,1},E_{F})e^{ik_{y,1}(y-d)}\\ +t_{2}\chi(-k_{y,2},E_{F})e^{-ik_{y,2}(y-d)},&\hbox{$y\geq d$,}\end{array}\right. (6)

where χ(ky,1,EF)\chi(k_{y,1},E_{F}) is the incident state with Fermi momentum ky,1k_{y,1}, r1(r2)r_{1}(r_{2}) is the reflection amplitude corresponding to the Fermi momentum ky,1(ky,2)-k_{y,1}(k_{y,2}), and t1(t2)t_{1}(t_{2}) is the transmission amplitude corresponding to the Fermi momentum ky,1(ky,2)k_{y,1}(-k_{y,2}), while ana_{n} is the scattering amplitude corresponding to the momentum ky,nk_{y,n}^{{}^{\prime}} in the barrier region. As can be seen in Fig. 2, at EF>EcE_{F}>E_{c} and θc<θ<π3\theta_{c}<\theta<\frac{\pi}{3} (θ>π3)(\theta>\frac{\pi}{3}), ky,1k_{y,1} is a positive electron-like (negative hole-like) momentum and ky,2k_{y,2} is a positive hole-like (negative electron-like) momentum, while for θ<θc\theta<\theta_{c}, ky,1k_{y,1} is a positive real root and ky,2k_{y,2} is a negative imaginary root. In the case of EF<EcE_{F}<E_{c}, however, for every incident angle θ\theta, ky,1k_{y,1} is a positive real root and ky,2k_{y,2} is a negative imaginary root.

Refer to caption
Figure 2: (Color online) Scattering processes when an incident electron is reflected and transmitted from the barrier with width dd and height V0V_{0} as shown on top of (a) in the cases of (a,b) single and (c) double refractions. The insets show the CEC at (a) EF<EcE_{F}<E_{c}, (b) EF>EcE_{F}>E_{c}, θ<θc\theta<\theta_{c} and (c) EF>EcE_{F}>E_{c}, θ>θc\theta>\theta_{c}. The green (red) circles on CECs indicate propagating electron (hole)-like states. Also, the GH shift Δtr1\Delta_{tr_{1}} of beam 1 with the same momentum as that of the incident beam and GH shift Δtr2\Delta_{tr_{2}} of beam 2 with different momentum compared to the incident beam are shown.

The eigenvalue equation (H+V(y))ψ(y)=Eψ(y)(H+V(y))\psi(y)=E\psi(y) corresponding to the Hamiltonian (1) is a second-order partial differential equation with respect to yy, due to the warping effect. Therefore, by applying boundary conditions of continuity of ψ(y)\psi(y) and its first derivative with respect to yy at the two interfaces y=0y=0 and y=dy=d, the reflection and transmission amplitudes, and also, the scattering amplitude ana_{n} can be determined. The transmission probability which is defined as the ratio of yy-component of the probability current density of the transmitted waves and that of the incident wave can be expressed in terms of the transmission amplitudes and the yy-component of the corresponding group velocities as T=T1+T2=|t1|2+υy(ky,2,EF)υy(ky,1,EF)|t2|2T=T_{1}+T_{2}=|t_{1}|^{2}+\frac{\upsilon_{y}(-k_{y,2},E_{F})}{\upsilon_{y}(k_{y,1},E_{F})}|t_{2}|^{2} in the case of double refraction and T=T1=|t1|2T=T_{1}=|t_{1}|^{2} in the case of single refraction M. Wu ; Arabikhah . In these relations TT denotes the total transmission probability, while T1T_{1} (T2T_{2}) represents the transmission probability of transmitted wave with the same (different) momentum as (from) the incident wave momentum.

We have shown in Fig. 3 the contour plot of the transmission probability as a function of incident angle and the potential barrier height for two different Fermi energies EF<EcE_{F}<E_{c} and EF>EcE_{F}>E_{c}. At EF<EcE_{F}<E_{c}, there is only one propagating transmitted wave whose probability at the typical Fermi energy of 0.15 eV and the barrier width d=500d=500Å is shown in Fig. 3(a). As can be seen, there is a region with two boundaries inside which the total internal reflection (TIR) takes place. The boundaries represent the geometrical locations of a critical angle θTIR=θTIR(V0)\theta_{TIR}=\theta_{TIR}(V_{0}) such that when the incident angle reaches θTIR\theta_{TIR}, all four waves inside the barrier region become evanescent. Therefore, at sufficiently wide barrier, TIR begins. To obtain an expression for θTIR\theta_{TIR}, first we consider equation E+(𝐤)+V0=EFE_{+}({\bf k})+V_{0}=E_{F} which is quadratic in terms of ky2k_{y}^{2}. By solving the discriminant of this equation, kxk_{x} value corresponding to θTIR\theta_{TIR} can be obtained. Replacing kxk_{x} in equation E+(𝐤)=EFE_{+}{\bf(k)}=E_{F}, we obtain the corresponding kyk_{y} of θTIR\theta_{TIR}, resulting θTIR=arctan(kx/ky)\theta_{TIR}=\arctan{(k_{x}/k_{y})}. The obtained analytical expression for θTIR\theta_{TIR} is complicated. Therefore, we have not presented the resulting expression here. Instead, we explain below how θTIR\theta_{TIR} occurs and behaves as a function of V0V_{0}. At EF=0.15E_{F}=0.15 eV and the typical value V0=0.05V_{0}=0.05 eV, the CECs in the incident and barrier regions are shown in Fig. 4(a). The size of CEC in the barrier region (determined by |V0EF||V_{0}-E_{F}|) is smaller than the size of CEC in the incident region. The line kx=cte.kx=cte., corresponding to the incident angle θ\theta represents the conservation of kxk_{x} in the electron scattering process. As shown in Fig. 4(a), at a small incident angle θ\theta, the line kx=cte.k_{x}=cte. intersects the CEC of the barrier region at two points, representing two real wave numbers of propagating electrons inside the barrier. With increasing θ\theta, the corresponding line kx=cte.k_{x}=cte. moves upwards until it touches the CEC of the barrier region at a single point, indicating that the incident angle θ\theta reaches the critical angle θTIR\theta_{TIR}. When θ\theta exceeds θTIR\theta_{TIR}, the line of kx=cte.k_{x}=cte. can no longer intersects the CEC of the barrier region, meaning that all ky,nk^{\prime}_{y,n} are complex and consequently, TIR begins. With increasing V0V_{0} the size of CEC in the barrier region decreases, and hence, the critical angle becomes smaller. As V0V_{0} approaches EFE_{F}, the critical angle at which all ky,nk^{\prime}_{y,n} become complex approaches zero. However, due to the Klein tunneling effect (see discussion below) which prohibits backscattering near the normal incident angle, TIR cannot start at θ=0\theta=0^{\circ} (see Fig. 3(a)) Article3 . When V0V_{0} exceeds EFE_{F}, the size of CEC in the barrier region increases again as well as the critical angle, until |V0EF||V_{0}-E_{F}| reaches EFE_{F} . From now on, since the size of CEC in the barrier region becomes larger than that of the CEC in the incident region, TIR does not form at any incident angle.

As can be seen in Fig. 3(a), the barrier remains perfectly transparent at incident angles close to the normal incidence θ=0\theta=0^{\circ}, regardless of the amount of V0V_{0}. This process is known as Klein tunneling Klein and originates from spin conservation Katsnelson , since nonmagnetic barriers cannot change the spin direction of incident electrons in a scattering process. On the other hand, at almost normal incidence, the spin states of incident wave and propagating reflected wave are orthogonal (this can be easily deduced from Eq. (3)). Therefore, backscattering is forbidden and electrons can transmit perfectly. By increasing θ\theta, the spin states of incident and propagating reflected waves are no longer orthogonal, and hence, the electron reflection is allowed. Now, we consider the case of oblique incidence. When a wave impinges on the barrier at a given θ0\theta\neq 0, a part of wave transmits into the barrier and is multiply reflected at the two interfaces y=0y=0 and y=dy=d, therefore interference happens. As V0V_{0} increases from zero, the size of CEC in the barrier region changes, and hence, the acquired phase kydk’_{y}d of propagating waves inside the barrier region will change, causing oscillations in transmission probability. Whenever the waves interfere constructively, the Fabry-Pérot resonances with T(θ0)=1T(\theta\neq 0)=1 appears. As V0V_{0} reaches a value at which the line kx=cte.k_{x}=cte. becomes tangent to the CEC of barrier region (see Fig. 4(a)), the TIR begins. It continues until the line kx=cte.k_{x}=cte. again becomes tangent to the CEC, but this time at an amount of V0>EFV_{0}>E_{F}. With further increasing V0V_{0}, oscillations appear again in the transmission spectrum. Moreover, similar to the Schrödinger-type electrons, when the potential height is less than Fermi energy (corresponds to a n-n-n junction), the incident electrons generally pass through the potential barrier with larger transmission probabilities, compared to the case of V0>EFV_{0}>E_{F} (corresponds to a n-p-n junction). Also, we should mention that for θ>80\theta>80^{\circ} the group velocity of electrons vy=(Eky)kx=vx(kxky)E0v_{y}=(\frac{\partial E}{\partial k_{y}})_{k_{x}}=-v_{x}(\frac{\partial k_{x}}{\partial k_{y}})_{E}\simeq 0 (see CEC in incident region in Fig. 4(a)). Therefore, the electron transmission becomes very small at V0=0V_{0}=0. However, at some V0V_{0} values, TT can be considerably magnified, due to the interference effect. It is important to point out that for validity of minimal continuum model described by Hamiltonian (1), both EFE_{F} and |V0EF||V_{0}-E_{F}| must be considered less than 0.4 eV Chen-2009 ; An-2012 . However, in Fig. 3(a), we terminate V0V_{0} at 0.4 eV as at larger values no more features can be observed.

Refer to caption
Figure 3: (Color online) Calculated transmission probability T1T_{1} versus θ\theta and V0V_{0} for incident electrons on a potential barrier of width d=500d=500 Å at Fermi energy (a) EF=0.15E_{F}=0.15 eV<Ec<E_{c} and (b) EF=0.3E_{F}=0.3 eV>Ec>E_{c}. (Ec0.18E_{c}\simeq 0.18 eV)
Refer to caption
Figure 4: (Color online) (a) Solid curves represent CECs in the incident and barrier regions for EF=0.15E_{F}=0.15 eV and |V0EF|=0.1|V_{0}-E_{F}|=0.1 eV. The black (red) dashed line shows the conservation of transverse momentum kxk_{x} in the scattering process at incident angle θ(θTIR)\theta(\theta_{TIR}). (b) The CEC in the incident region corresponds to EF=0.3E_{F}=0.3 eV. The value of V0V_{0} is chosen such that the maximum of kxk_{x} on the CEC of the barrier region is bigger than the kxk_{x} value on CEC of the incident region at θ=π/2\theta=\pi/2, i.e. θTIR>θc\theta_{TIR}>\theta_{c}. Due to the warped CEC in the incident region, TIR is confined in the interval of θTIR<θ<θ\theta_{TIR}<\theta<\theta^{\prime}.
Refer to caption
Figure 5: (Color online) Calculated transmission probabilities (a) T1T_{1} and (b) T2T_{2} versus θ\theta and dd for incident electrons with Fermi energy EF=0.3E_{F}=0.3 eV on a potential barrier of V0=0.7V_{0}=0.7 eV.

The CEC of zero-potential regions is warped at EF>EcE_{F}>E_{c}, and hence, two transmitted waves can propagate. The transmission probability T1T_{1} at EF=0.3E_{F}=0.3 eV and with the same barrier width value as that in Fig. 3(a) is depicted in Fig. 3(b). First, we consider a given V0V_{0} at which the size of CEC in the barrier region is small enough compared to the CEC of the incident region. Similar to Fig. 4(a), as θ\theta increases from zero, the corresponding line kx=cte.k_{x}=cte. moves upwards and consequently, the acquired phase kydk^{\prime}_{y}d of propagating electron waves inside the barrier region will change, resulting oscillations in T1T_{1} and also Fabry-Pérot resonances in the constructive interference. When θ\theta reaches θTIR\theta_{TIR}, the line kx=cte.k_{x}=cte. touches the CEC of barrier region at a single point, that is the start of TIR. The TIR extends to π/2\pi/2 because for θ>θTIR\theta>\theta_{TIR} the constant kxk_{x} line can no longer intersect the CEC of the barrier region. Now, we consider an amount of V0V_{0} at which the size of CEC in the barrier region is close enough to the CEC of incident region, so that the maximum of kxk_{x} in the barrier region is larger than kxk_{x} at θ=π/2\theta=\pi/2 in the incident region (see Fig. 4(b)). In this case, when θ\theta exceeds θTIR\theta_{TIR}, the corresponding constant kxk_{x} line will start to intersect the CEC of the barrier region one more time at an angle θ>π/3\theta^{\prime}>\pi/3. Therefore, the TIR terminates at θ\theta^{\prime} and will not extend to π/2\pi/2. This happens at V0<0.03V_{0}<0.03 eV and 0.57 eV<V0<<V_{0}<0.6 eV in Fig. 3(b). At almost normal incident angle, the Klein tunneling happens, regardless of V0V_{0} value, similar to Fig. 3(a). However, at a given oblique incident angle θ0\theta\neq 0 when V0V_{0} varies outside the TIR region, the size of CEC in the barrier region changes, and hence, the acquired phase kydk’_{y}d of the propagating waves inside the barrier region may also change, causing an oscillatory behavior in T1T_{1}, similarly to the behavior of Fig. 3(a). Since (kx/ky)E(\partial k_{x}/\partial k_{y})_{E} approaches zero as θ\theta reaches 6262^{\circ} (see the CEC in the incident region in Fig. 4(b)), vyv_{y} and consequently T1T_{1} are very small, independent of V0V_{0} values. On the other hand, the transmission probability T2T_{2} (not shown here) is zero at θ<θc=54.7\theta<\theta_{c}=54.7^{\circ}, while it has the same features as T1T_{1} at θ>θc\theta>\theta_{c}, regardless of V0V_{0} values.

The transmission probabilities T1T_{1} and T2T_{2} for the two transmitted waves 1 and 2 as functions of incident angle θ\theta and the barrier width dd, at EF=0.3E_{F}=0.3 eV and V0=0.7V_{0}=0.7 eV are shown in Figs. 5(a) and 5(b), respectively. Since |V0EF|>EF|V_{0}-E_{F}|>E_{F}, the TIR does not occur and for a given dd value at almost normal incident angle, the Klein tunneling with perfect transmission probability for T1T_{1} happens, as shown in Fig. 5(a). As θ\theta is increased from zero, the acquired phase kydk’_{y}d of the propagating waves inside the barrier region changes. This causes oscillations in T1T_{1}, emerging Fabry-Pérot resonances when constructive interference takes place. If dd varies at a fixed oblique incident angle, the acquired phase kydk’_{y}d will change so that T1T_{1} again exhibits oscillations and Fabry-Pérot resonances can emerge. Here, θc=54.7\theta_{c}=54.7^{\circ} is the same as that in Fig. 3(b) because θc\theta_{c} depends only on EFE_{F}. When θ\theta exceeds θc\theta_{c}, T2T_{2} in Fig. 5 (b) takes non-zero values and shows oscillations due to the change of acquired phase kydk’_{y}d, with varying θ\theta or dd. In the vicinity of θ=62\theta=62^{\circ}, both T1T_{1} and T2T_{2} are very small for the same reason explained in Fig. 3(b).

Refer to caption
Figure 6: (Color online) (a,b) Calculated transmission probabilities T1,2T_{1,2} (blue curves) and the Gh shifts Δtr1,2\Delta_{tr_{1,2}} (red curves) in transmission as a function of θ\theta. (c) The spatial separation δΔtr\delta\Delta_{tr} (blue curve) and the angle difference α\alpha (red curve) between spin orientation of transmitted beams. The parameters are EF=0.35E_{F}=0.35 eV, V0=0.72V_{0}=0.72 eV, and d=d=490 Å.

III Goos-Hänchen shift and beam splitting

The GH shift for a plane wave of electrons is not detectable due to its infinite spatial width. Therefore, to calculate the GH lateral shift, we consider a beam of electrons instead of a plane wave. We model an incident electron beam by using a Gaussian wave packet of surface states as

ψin(𝐫)=+𝑑kxf(kxkx0)χ(ky,1(kx),EF)ei(kxx+ky,1(kx)y),\psi_{in}({\bf r})=\int_{-\infty}^{+\infty}dk_{x}f(k_{x}-k_{x_{0}})\chi(k_{y_{,1}}(k_{x}),E_{F})e^{i(k_{x}x+k_{y_{,1}}(k_{x})y)}, (7)

where f(kxkx0)=(2πΔkx)1e(kxkx0)2/2Δkx2f(k_{x}-k_{x_{0}})=(\sqrt{2\pi}\Delta_{k_{x}})^{-1}e^{-(k_{x}-k_{x_{0}})^{2}/2\Delta_{k_{x}}^{2}} shows the Gaussian angular distribution of width Δkx\Delta_{k_{x}} around central incident angel θ0=arctan(|kx0ky,1|)\theta_{0}=\arctan(|\frac{k_{x_{0}}}{k_{y,{1}}}|). Analogously, the wave functions of transmitted electron beams can be written as

ψtr1(𝐫)\displaystyle\psi_{tr_{1}}({\bf r}) =\displaystyle= +𝑑kxf(kxkx0)t1(kx)\displaystyle\int_{-\infty}^{+\infty}dk_{x}f(k_{x}-k_{x_{0}})t_{1}(k_{x}) (8)
×χ(ky,1(kx),EF)ei(kxx+ky,1(kx)(yd)),\displaystyle\times\chi(k_{y_{,1}}(k_{x}),E_{F})e^{i(k_{x}x+k_{y_{,1}}(k_{x})(y-d))},

and

ψtr2(𝐫)\displaystyle\psi_{tr_{2}}({\bf r}) =\displaystyle= +𝑑kxf(kxkx0)t2(kx)\displaystyle\int_{-\infty}^{+\infty}dk_{x}f(k_{x}-k_{x_{0}})t_{2}(k_{x}) (9)
×χ(ky,2(kx),EF)ei(kxxky,2(kx)(yd)).\displaystyle\times\chi(-k_{y_{,2}}(k_{x}),E_{F})e^{i(k_{x}x-k_{y_{,2}}(k_{x})(y-d))}.

For the well collimated electron beams, f(kxkx0)f(k_{x}-k_{x_{0}}) is sharp around kx0k_{x_{0}} such that the spinor components χ±=χ±eiφ±\chi^{\pm}=\mid\chi^{\pm}\mid e^{i\varphi^{\pm}} can be converted into an exponential form and approximated by keeping the first two terms of the Taylor expansion of its exponent around kx0k_{x_{0}} as

χ±(ky,1(kx))\displaystyle\chi^{\pm}(k_{y_{,1}}(k_{x})) =\displaystyle= exp[lnχ±(ky,1(kx))]\displaystyle\exp[\ln\chi^{\pm}(k_{y_{,1}}(k_{x}))] (10)
\displaystyle\simeq χ±(ky,1(kx0))exp{[|χ˙±(ky,1(kx0))||χ±(ky,1(kx0))|\displaystyle\chi^{\pm}(k_{y_{,1}}(k_{x_{0}}))\exp\{[\frac{|\dot{\chi}^{\pm}(k_{y_{,1}}(k_{x_{0}}))|}{|\chi^{\pm}(k_{y_{,1}}(k_{x_{0}}))|}
+iφ˙±(ky,1(kx0))](kxkx0)},\displaystyle+i\dot{\varphi}^{\pm}(k_{y_{,1}}(k_{x_{0}}))](k_{x}-k_{x_{0}})\},

where φ˙±(ky,1(kx0))(|χ˙±(ky,1(kx0))|)\dot{\varphi}^{\pm}(k_{y_{,1}}(k_{x_{0}}))(|\dot{\chi}^{\pm}(k_{y_{,1}}(k_{x_{0}}))|) denotes derivative of φ±(ky,1(kx))(|χ±(ky,1(kx))|)\varphi^{\pm}(k_{y_{,1}}(k_{x}))(|\chi^{\pm}(k_{y_{,1}}(k_{x}))|) with respect to kxk_{x}, evaluated at kx=kx0k_{x}=k_{x_{0}}. Substituting Eq. (8) into Eq. (5), using the approximation ky,1(kx)ky,1(kx0)+k˙y,1(kx0)(kxkx0)k_{y_{,1}}(k_{x})\simeq k_{y_{,1}}(k_{x_{0}})+\dot{k}_{y_{,1}}(k_{x_{0}})(k_{x}-k_{x_{0}}) and then evaluating the integral we obtain the spatial form of the components of the incident beam as

ψin±(𝐫)\displaystyle\psi_{in}^{\pm}({\bf r}) =\displaystyle= χ±(ky,1(kx0))\displaystyle\chi^{\pm}(k_{y_{,1}}(k_{x_{0}})) (11)
×e[x+φ˙±(ky,1(kx0))+k˙y,1(kx0)y]2Δkx2/2\displaystyle\times e^{-[x+\dot{\varphi}^{\pm}(k_{y_{,1}}(k_{x_{0}}))+\dot{k}_{y_{,1}}(k_{x_{0}})y]^{2}\Delta_{k_{x}}^{2}/2}
×eγ±2/2Δkx2eiγ±φ˙±(ky,1(kx0))\displaystyle\times e^{\gamma^{\pm^{2}}/2\Delta_{k_{x}}^{2}}e^{i\gamma^{\pm}\dot{\varphi}^{\pm}(k_{y_{,1}}(k_{x_{0}}))}
×ei[(ky,1(kx0)+γ±k˙y,1(kx0))y+(kx0+γ±)x],\displaystyle\times e^{i[(k_{y_{,1}}(k_{x_{0}})+\gamma^{\pm}\dot{k}_{y_{,1}}(k_{x_{0}}))y+(k_{x_{0}}+\gamma^{\pm})x]},

where γ±=Δkx2|χ˙±(ky,1(kx0))||χ±(ky,1(kx0))|\gamma^{\pm}=\frac{\Delta_{k_{x}}^{2}|\dot{\chi}^{\pm}(k_{y_{,1}}(k_{x_{0}}))|}{|\chi^{\pm}(k_{y_{,1}}(k_{x_{0}}))|}. As can be seen from the second factor in Eq. (9), the incident beam has a Gaussian shape and the peak location of its upper and lower components at the interface y=0y=0 is given by xin±=φ˙±(ky,1(kx0))x_{in}^{\pm}=-\dot{\varphi}^{\pm}(k_{y_{,1}}(k_{x_{0}})). Therefore, the average location of incident beam at the interface y=0y=0 can be expressed as

x¯in\displaystyle\bar{x}_{in} =\displaystyle= φ˙+(ky,1(kx0))|χ+(ky,1(kx0))|2\displaystyle-\dot{\varphi}^{+}(k_{y_{,1}}(k_{x_{0}}))|\chi^{+}(k_{y_{,1}}(k_{x_{0}}))|^{2} (12)
φ˙(ky,1(kx0))|χ(ky,1(kx0))|2.\displaystyle-\dot{\varphi}^{-}(k_{y_{,1}}(k_{x_{0}}))|\chi^{-}(k_{y_{,1}}(k_{x_{0}}))|^{2}.

It is worth mentioning that the last factor in Eq. (9) shows that the propagation direction of incident-beam components deviates from the central angle θ0\theta_{0} by the amount of δ±tanδ±γ±ky,1(kx0)\delta^{\pm}\approx\tan\delta^{\pm}\approx\frac{\gamma^{\pm}}{k_{y_{,1}}(k_{x_{0}})}. This deflection is due to the warping effect as in the absence of warping |χ±||\chi^{\pm}| is constant and δ±=0\delta^{\pm}=0. Moreover, the third factor in Eq. (9) reveals that the magnitude of incident beam is adjusted by warping as well.

By comparing Eqs. (6) and (7) with Eq. (5) we can write an expression for the transmitted beam components, similar to Eq. (9), by the substitutions χ±χ±t1(2)\chi^{\pm}\mapsto\chi^{\pm}t_{1(2)}, φ±φ±+φt1(2)\varphi^{\pm}\mapsto\varphi^{\pm}+\varphi_{t_{1(2)}}and |χ±||χ±||t1(2)||\chi^{\pm}|\mapsto|\chi^{\pm}||t_{1(2)}| in Eq. (9) where φt1(2)\varphi_{t_{1(2)}} represent the phase of transmission amplitude t1(2)t_{1(2)}. Therefore, the transmitted beams find also Gaussian shapes just like incident beam. The average locations of the transmitted beams at the interface y=dy=d read as

x¯tr1\displaystyle\bar{x}_{tr_{1}} =\displaystyle= φ˙+(ky,1(kx0))|χ+(ky,1(kx0))|2φt1˙(kx0)\displaystyle-\dot{\varphi}^{+}(k_{y_{,1}}(k_{x_{0}}))|\chi^{+}(k_{y_{,1}}(k_{x_{0}}))|^{2}-\dot{\varphi_{t_{1}}}(k_{x_{0}}) (13)
φ˙(ky,1(kx0))|χ(ky,1(kx0))|2,\displaystyle-\dot{\varphi}^{-}(k_{y_{,1}}(k_{x_{0}}))|\chi^{-}(k_{y_{,1}}(k_{x_{0}}))|^{2},

and

x¯tr2\displaystyle\bar{x}_{tr_{2}} =\displaystyle= φ˙+(ky,2(kx0))|χ+(ky,2(kx0))|2φt2˙(kx0)\displaystyle-\dot{\varphi}^{+}(-k_{y_{,2}}(k_{x_{0}}))|\chi^{+}(-k_{y_{,2}}(k_{x_{0}}))|^{2}-\dot{\varphi_{t_{2}}}(k_{x_{0}}) (14)
φ˙(ky,2(kx0))|χ(ky,2(kx0))|2.\displaystyle-\dot{\varphi}^{-}(-k_{y_{,2}}(k_{x_{0}}))|\chi^{-}(-k_{y_{,2}}(k_{x_{0}}))|^{2}.
Refer to caption
Figure 7: (Color online) (a,b) Calculated transmission probabilities T1,2T_{1,2} (blue curves) and the GH shifts Δtr1,2\Delta_{tr_{1,2}} (red curves) in transmission as a function of barrier width dd. (c) The spatial separation δΔtr\delta\Delta_{tr} between transmitted beams. The parameters are EF=0.3E_{F}=0.3 eV, V0=0.65V_{0}=0.65 eV, and θ=59\theta=59^{\circ}.

The GH lateral shift Δtr\Delta_{tr} is defined as the displacement of the peak of transmitted beam at the interface y=dy=d relative to the peak of incident beam at the interface y=0y=0 Broe (see Fig. 2) which is different from classical prediction of electron optics, i.e., Snell’s shift dtanθd\tan\theta^{\prime}, where θ\theta^{\prime} is the refraction angle. Therefore, the GH shift of the transmitted beam 1(2) with the same (different) momentum as (from) that of the incident beam can be obtained from Eqs. (10)-(12) as

Δtr1=φt1˙(kx0),\Delta_{tr_{1}}=-\dot{\varphi_{t_{1}}}(k_{x_{0}}), (15)

and

Δtr2\displaystyle\Delta_{tr_{2}} =\displaystyle= φt2˙(kx0)φ˙(ky,2(kx0))|χ(ky,2(kx0))|2\displaystyle-\dot{\varphi_{t_{2}}}(k_{x_{0}})-\dot{\varphi}^{-}(-k_{y_{,2}}(k_{x_{0}}))|\chi^{-}(-k_{y_{,2}}(k_{x_{0}}))|^{2} (16)
+φ˙(ky,1(kx0))|χ(ky,1(kx0))|2.\displaystyle+\dot{\varphi}^{-}(k_{y_{,1}}(k_{x_{0}}))|\chi^{-}(k_{y_{,1}}(k_{x_{0}}))|^{2}.

In deriving Eq. (14) we have used φ˙+(ky,1(kx0))=φ˙+(ky,2(kx0))=0\dot{\varphi}^{+}(k_{y_{,1}}(k_{x_{0}}))=\dot{\varphi}^{+}(-k_{y_{,2}}(k_{x_{0}}))=0 because in the case of double refraction upper spinor components in zero potential regions are real. The spatial splitting between the two beams will occur when they have different GH shifts. In this case, the spatial separation between the two beams is given by δΔtr=Δtr1Δtr2\delta\Delta_{tr}=\Delta_{tr_{1}}-\Delta_{tr_{2}}.

The electron spin orientation can be obtained from Eq. (3) as 𝐬=<𝝈>=E+1(υFky,υFkx,w(𝐤))\mathbf{s}=<\boldsymbol{\sigma}>=E_{+}^{-1}(-\upsilon_{F}k_{y},\upsilon_{F}k_{x},w({\bf k})) indicating the spin-momentum locking of surface electrons in TIs, due to the spin-orbit coupling. Consequently, the spin direction of the transmitted beam 2 is rotated relative to the spin direction of both transmitted beam 1 and incident beam by the amount of α=arccos(𝐬1𝐬2)\alpha=\arccos(\mathbf{s}_{1}\cdot\mathbf{s}_{2}) where 𝐬1\mathbf{s}_{1} and 𝐬2\mathbf{s}_{2} are spin orientations of transmitted beams 1 and 2, respectively.

Due to the warping term, Δtr1(2)\Delta_{tr_{1(2)}} cannot be derived in compact analytical expressions. Therefore, these quantities are calculated numerically using Eqs. (13) and (14). Typical results for GH shifts and the corresponding transmission probabilities are shown in Figs. (6) and (7). The parameters are chosen to avoid TIR and that two transmitted beams propagate. Figs. 6(a) and (b) show the transmission probabilities and the corresponding GH values of the two transmitted beams 1 and 2 in terms of incident angle. Due to the interference effect, T1,2T_{1,2} show an oscillatory behavior and some sharp maxima and minima appear for both transmitted beams. In fact, by changing the incident angle the acquired phase (kydk_{y}^{\prime}d) of every propagating wave along the barrier region varies, which leads to the oscillation of transmission probabilities. The peak positions of the two beams are almost the same. The corresponding GH shifts (red lines) of the two beams exhibit some strong peaks beside usual ones with positive and negative values. In order to explain qualitatively the behavior of lateral shifts and the occurrence of their peaks, we rewrite the formula (13) and (14) Article as

Δtr=ddkxtanφt1+tan2φt,\Delta_{tr}=\frac{\frac{d}{dk_{x}}\tan\varphi_{t}}{1+\tan^{2}\varphi_{t}}\ , (17)

where tanφt=Im[t(kx)]/Re[t(kx)]\tan\varphi_{t}=\mathrm{Im}[t(k_{x})]/\mathrm{Re}[t(k_{x})]. We approximate Re[t(kx)]\mathrm{Re}[t(k_{x})] and Im[t(kx)]\mathrm{Im}[t(k_{x})] around a given point kx0k_{x_{0}} by retaining the first and second terms of their Taylor expansion as Re[t(kx)]aR+bR(kxkx0)\mathrm{Re}[t(k_{x})]\simeq a_{R}+b_{R}(k_{x}-k_{x_{0}}) and Im[t(kx)]aI+bI(kxkx0)\mathrm{Im}[t(k_{x})]\simeq a_{I}+b_{I}(k_{x}-k_{x_{0}}), where aR,bR,aIa_{R},b_{R},a_{I} and bIb_{I} are coefficients of the expansions. By inserting these approximations in Eq. (15) we obtain

Δtr=aRbIaIbR|t(kx)|2.\Delta_{tr}=\frac{a_{R}b_{I}-a_{I}b_{R}}{|t(k_{x})|^{2}}. (18)

Moreover, by approximating ddkx|t(kx)|aRbR+aIbI|t(kx)|\frac{d}{dk_{x}}|t(k_{x})|\simeq\frac{a_{R}b_{R}+a_{I}b_{I}}{|t(k_{x})|}, Eq. (16) can be written as

Δtr=aRbIaIbR(aRbR+aIbI)2(ddkx|t(kx)|)2.\Delta_{tr}=\frac{a_{R}b_{I}-a_{I}b_{R}}{(a_{R}b_{R}+a_{I}b_{I})^{2}}(\frac{d}{dk_{x}}|t(k_{x})|)^{2}. (19)

From Eqs. (16) and (17), the local properties of Δtr\Delta_{tr} in the vicinity of a given point kx0k_{x_{0}}, and hence, θ0\theta_{0} can be studied. According to Eq. (17), the absolute value of Δtr\Delta_{tr} at any point depends on the absolute value of the slope of transmission probability at that point. By approaching the sharp maxima and minima points in the blue curves, the slope of the T1,2T_{1,2} rapidly finds very large values. Therefore, the absolute values of the corresponding GH shifts near these points in the red curves suddenly increase, creating sharp maxima and minima. The sign of GH shift is determined by the sign of the numerator in Eq. (17). Some deep minima (not exactly zero) for T1T_{1} and T2T_{2} appear in Figs. 6(a) and (b), specially for T2T_{2}, i.e. |t(kx)|0|t(k_{x})|\approx 0. According to Eq. (16), the absolute value of the corresponding lateral shifts at these points can become large and therefore, local maxima appear at these points, as seen in red curves. Spatial separation between the two beams and the angle between their spin orientations as a function of incident angle are shown in Fig. 6(c) with blue and red curves, respectively. One can see that at the given angle window, δΔtr\delta\Delta_{tr} exhibits several pronounced positive peaks, which make the observation of well-separated beams practically more feasible. Note that although at these points the transmission probabilities are far from the perfect splitter case with T1=T2=0.5T_{1}=T_{2}=0.5, these values are practically considerable. As an example, for the incident angle θ=76.3\theta=76.3^{\circ} at which the obtained transmission probabilities for the two beams are 0.20.2 and 0.110.11 (see Figs. 6(a) and (b)), the spatial separation is about 6.5μm6.5\mu m which is large enough to measure experimentally. At this point, the angle between spin orientations of the two beams is 77.877.8^{\circ}.

Refer to caption
Figure 8: (Color online) Schematic view showing the procedure of GH shift measurement. The cyclotron radius rir_{i} in incident region, fixed by both transverse magnetic field 𝐁𝐢\mathbf{B_{i}} and the Fermi energy, determines the incident angle θ\theta and also the impact point of electron trajectory at the interface y=0y=0. A varying transverse magnetic field 𝐁𝐭\mathbf{B_{t}} induces a resonant path with a specific radius rtr_{t} which determines the entry point of electron at the interface y=dy=d.

The transmission probabilities of the two transmitted beams and the corresponding GH shifts in terms of barrier width dd are depicted in Figs. 7(a) and (b). By varying the width of the barrier, the acquired phase, kydk_{y}^{\prime}d, changes, and hence, T1T_{1} and T2T_{2} oscillate by revealing several sharp maxima and deep minima. We note that kyk_{y}^{\prime} does not change at a fixed incident angle. The behavior of GH shifts in Figs. 7(a) and (b), compared to their corresponding transmission probabilities, are similar to the behavior of GH shifts in Fig. 6(a) and (b), compared to their corresponding transmissions. That means, near the sharp maxima of transmission probabilities where their slope rapidly increases with dd, the absolute of the corresponding GH shifts increases as well. Also, near the deep minima of transmission probabilities where |t(d)|0|t(d)|\approx 0, the absolute of the corresponding lateral shifts can be large. Such a similarity is explained as follows: The dependence of transmission coefficients, and hence, their phases on dd is through the exponential function eiky,nde^{ik_{y,n}^{\prime}d}, when the boundary conditions of continuity of the wave function in Eq. (4) and its derivative at the interface y=dy=d are applied. Consequently, the dependence of GH shifts on dd should be through exponential functions eiky,nde^{ik_{y,n}^{\prime}d} as well. On the other hand, the dependence of transmission coefficients as well as GH shifts on kx(θ)k_{x}(\theta) comes from eiky,nde^{ik_{y,n}^{\prime}d} and also from other terms which vary slowly compared to the exponential functions. Therefore, when dd changes, the dependence of GH shift on the transmission probability will be similar to the dependence of GH shift on the transmission probability, when kxk_{x} or θ\theta changes.

Spatial separation between the two electron beams as a function of dd is shown in Fig. 7(c). One can see that at d775Ad\sim 775A^{\circ}, the separation between the two beams is almost 10μm10\mu m and the transmission probabilities of T1T_{1} and T2T_{2} are 0.5\sim 0.5 and 0.20.2, respectively (see Figs. 7(a) and (b)). Also, the absolute value of δΔtr\delta\Delta_{tr} peaks between 720 Å and 800 Å exhibits a considerable width, similar to T1T_{1} and T2T_{2}. Moreover, the angle between spin orientations of two beams is 59.159.1^{\circ} which is independent of dd and can be obtained by the values of EFE_{F} and θ\theta. Therefore, our findings reveal that TIs with hexagonal warping effects can be utilized to design an electron beam splitter with the ability of spatial separation as large as a few micrometers with high chance of observation of the well-separated beams.

In this paper we focused on a barrier with interfaces along xx direction (Γ\GammaK direction in 𝐤{\bf k}-space) which resulted in double refraction and double GH shifts of electron beams. If the barrier extends along yy direction (Γ\GammaM direction), due to the highly anisotropic nature of hexagonally warped Fermi surface, triple refraction An-2012 ; M. Wu and consequently triple GH shifts can emerge. The occurrence of double and triple GH shifts in different directions can be a signature of hexagonally warped Fermi surface, while they do not occur in other cases such as trigonally or tetragonally warped Fermi surfaces. Also, observing gaps in the GH shift measurements in terms of electron energy can indicate the existence of energy gap in the band structure of matterials Ghadiri1 . Nevertheless, determining whether the shape of the Fermi surface can be identified with GH shift measurements, requires more research. It is worth mentioning that in Weyl semimetals it has been shown that the GH and IF shifts of a reflected beam from a gapped medium can provide a probe of the topological Fermi arc at the reflecting surface Chattopadhyay .

IV A proposal for GH shift measurement

To the best of our knowledge the GH shift in electronic systems has not been experimentally measured yet, due to the smallness of GH shift values and the difficulty in producing a well collimated electron beam Chen4 ; Yu3 . Although the magnitude of GH shift in total reflection from a single-interface (step potential) is about Fermi wavelength of electron which impedes its direct measurement, it can be enlarged by considering a system acting as a waveguide which causes accumulation of shifts in multiple reflection of electron beam from the waveguide boundaries Beenakker ; Sun ; Ghadiri1 ; Yu3 . Also, in the process of transmitting electrons through potential barrier/well, transmission resonances can occur which enhance the GH shift value considerably Chen2 ; Song ; Ghadiri ; Ghadiri1 ; Zheng ; Article1 . Note that similar and other mechanisms for amplifying optical GH shifts are considered in literatures (see Ref. [Li2004, ] and [ChenAPL2017, ]). On the other hand, to directly measure GH shift values, we need a collimator to generate collimated electron beam and then detect the transmitted/reflected beam from the interface by local gates. Although various proposals for electron collimation in 2D materials Park ; Cheianov ; Moghaddam ; M. Liu ; Betancur and surface states of TIs Hassler ; M. Liu are suggested, a decent production of narrow and well collimated electron beam in such materials has not been attained yet S. Chen ; Libisch ; Zhou ; LaGasse .

Despite the lack of efficient collimation, Chen et al. S. Chen achieved a direct measurement of angle-dependent transmission probability based on TMF measurement scheme. They applied a transverse magnetic field on electrostatically defined nn-nn^{\prime} (p-n) junction on graphene and measured the transresistance proportional to the transmission of electrons between an injection electrode (at nn side) and a collection electrode (at nn^{\prime} side), while sweeping magnetic field and gate voltage of nn^{\prime} side. In this way, they reached a map in which the first and higher-order resonant peaks appeared. Moreover, using a semiclassical Billiard model they performed a simulation of electron trajectories whose result was well-matched with that of experimental data. As a result, they reverse-engineered the first-order resonant transport by considering a trajectory for electrons similar to the one that we consider in Fig. 8 which clearly gives the peak positions observed in the experiment as well as in the simulation.

Here, by applying a similar TMF measurement scheme we propose a procedure for electrons’ GH shift measurement on the surface of a TI junction, as schematically shown in Fig. 8. We consider a positive GH shift which mostly occur in n-n-n case. Before explaining the procedure, we give a brief discussion about survivability of surface states in the presence of a transverse magnetic field. In TMF phenomenon, it is assumed that the motion of electrons is ballistic, following the classical trajectories Taychatanapat ; Milovanovic ; Tsoi . This is justified when the electron mean free path lel_{e} is larger than the width of the device in xx direction as well as the separation between the electrodes and interfaces (li(l_{i} and lt)l_{t}). The length of lel_{e} is estimated 120 nm for surface electrons in Bi2Te3 Xie-2017 . When the surface classical electrons are subjected to a transverse magnetic field BB, they follow circular cyclotron orbits with radius r=EF/(evFB)r=E_{F}/(ev_{F}B) given by Lorentz force, where ee is the charge of electron. If the system is treated quantum mechanically, these orbitals get quantized into Landau levels giving rise to chiral edge states. However when the magnetic field is not too high, the Landau levels undergo a collapse transition and the edge states can be avoided Akzyanov ; Li-2015 . To match our procedure with the above-mentioned experiment S. Chen , we consider electrons with low Fermi energy in the incident and transmission regions with a circular shape of energy contour similar to that of graphene, which makes an accurate control of electron trajectories in the presence of magnetic field Yu . In the barrier (nn^{\prime}) region, there is no magnetic field and the warping effect can be remarkable for large enough V0V_{0} values. Under a transverse magnetic field BiB_{i}, applied on incident electrons injected from a narrow injection electrode, located at the bottom of this region, the electrons undergo a cyclotron motion with radius ri=EFeυFBir_{i}=\frac{E_{F}}{e\upsilon_{F}B_{i}}. With some simple calculations, the impact point of electrons and their incident angle at the interface y=0y=0 can be determined as xi=li(2rili)x_{i}=\sqrt{l_{i}(2r_{i}-l_{i})} and θ=arctan(rilixi)\theta=\arctan(\frac{r_{i}-l_{i}}{x_{i}}), respectively, where lil_{i} is the distance between the injection electrode and the interface (see Fig. 8).

Now we apply an independent transverse magnetic field BtB_{t} on the transmission region. BtB_{t} bends the transmitted electrons’ path downward the device into a cyclotron orbit. If the electrons enter the collection electrode placed at the bottom of this region, a peak in the transresistance between the injection and collection electrodes (or corresponding voltage) will occur. Therefore, by tuning BtB_{t} in the experiment, it is possible to obtain the amount of BtB_{t} and the corresponding radius rt=EFeυFBtr_{t}=\frac{E_{F}}{e\upsilon_{F}B_{t}} for which a resonance in magnetic focusing takes place. Having rtr_{t} and knowing the angle of incoming electrons (equal to the incident angle) into this region and the distance ltl_{t} between the collection electrode and the interface y=dy=d, after some straightforward algebraic computations, one can determine the position of entry point of electrons at the interface y=dy=d as xt=rtcosθ+rt2(ltrtsinθ)2x_{t}=r_{t}\cos\theta+\sqrt{r_{t}^{2}-(l_{t}-r_{t}\sin\theta)^{2}}. Finally the GH shift can be obtained by Δtr=xtxi\Delta_{tr}=x_{t}-x_{i}.

It is important to note that there is a correspondence between the variables used in Chen et al. experiment S. Chen and the variables in our proposal. In their experiment, the same magnetic field BB is applied to both incident and transmission regions, and the magnetic field BB as well as the gate voltage of transmission region are varied. Also, the angle of entry of electrons into the transmission region (θ)(\theta^{\prime}) is different from the incident angle θ\theta. Moreover, θ\theta^{\prime} is a function of θ\theta according to the Snell’s law sinθ=((EFVi)/(EFVt))sinθ\sin\theta^{\prime}=((E_{F}-V_{i})/(E_{F}-V_{t}))\sin\theta, where ViV_{i} and VtV_{t} are the gate voltage of incident and transmission regions, respectively. In our proposal, different magnetic fields BiB_{i} and BtB_{t} are applied to the incident and transmission regions, respectively. The gate voltage of the barrier region is fixed, while BiB_{i} and BtB_{t} are variable during the experiment. On the other hand, the transmitted electrons enter the transmission region with the same angle as the incident angle θ\theta but with a lateral shift Δtr\Delta_{tr} which is a function of incident angle. Because of such correspondences, we expect that the electrons contributing in the transresistance peak to be those electrons that leave the injection electrode vertically, just like the Chen et al. experiment S. Chen .

The above discussion can be presented in a general form as follows. Consider the electrons that leave the injection electrode with arbitrary injection angle β\beta (with respect to yy axis) at a given BiB_{i}. The incident angle of these electrons can be calculated as θ=arcsin(sinβli/ri)\theta=\arcsin(\sin\beta-l_{i}/r_{i}). At a fixed BtB_{t}, due to the dependence of impact point of electrons on the interface at y=0y=0 and also Δtr\Delta_{tr} to θ\theta the electrons reach the edge of the device (yy axis) in the transmission region at a position that depends on their injection angle β\beta. Calculating the derivative of θ\theta with respect to β\beta, we obtain dθ/dβ=cosβ/cos2β+2(li/ri)sinβli2/ri2d\theta/d\beta=\cos\beta/\sqrt{\cos^{2}\beta+2(l_{i}/r_{i})\sin\beta-l_{i}^{2}/r_{i}^{2}} whose magnitude is zero at β=π/2\beta=\pi/2. This means that the electrons which leave the injection electrode in small vicinity of angle β=π/2\beta=\pi/2, have the same incident angle θ\theta, and hence, the same lateral shift Δtr\Delta_{tr}, resulting the largest density of electrons at a point on the edge of the device in the transmission region. By sweeping the magnetic field BtB_{t}, rtr_{t} is varied, so that at a fixed collection electrode position, a peak in transresistance belonging to the electrons that leave the injection electrode vertically, appears. By tuning BtB_{t} on larger amounts, the cyclotron radius rtr_{t} is reduced, so that the electrons can reach the collection electrode after one or more specular reflection from the interface and/or the edge of the device, leading to the formation of next peaks Tsoi .

Although the present proposal of GH shift measurement was applied to the surface state of Bi2Te3 consisting of a single nondegenerate Dirac cone, this approach can also be utilized in 2D conventional systems such as graphene and other single-layer hexagonal crystals. Nevertheless, in materials consisting of multivalleys, multirefraction can appear, making the observation of GH shift more complicated than the present study. Moreover, surface states of TIs are topologically protected against non-magnetic perturbations compared to the conventional surface states which are sensitively dependent on the geometry of surface structure.

Since in most Dirac materials the GH shift is spin and/or valley dependent which originates from spin-orbit coupling Sun ; Ghadiri ; Ghadiri1 ; Azarova ; Article1 ; Article2 , the measurement of GH shift can provide the possibility of fabrication of spin/valley devices based on electronic beam shifts.

V Conclusion

In summary, we studied theoretically the influence of hexagonal warping effect on the transport properties and lateral shifts of electrons at the surface of a TI n-n-n (n-p-n) junction. It is shown that double refractions occur when the Fermi energy and incident angle of electron beams exceed their critical values. We establish an expression for calculating GH shift values and show that a deflection of propagation direction of beams from their central propagation directions appears due to the hexagonal warping effect. The dependence of lateral shifts and the corresponding transmissions on system parameters such as incident angle, height and the width of potential barrier are carefully examined. We show that the system can produce two spatially separated beams with different spin orientations as a result of GH effect. Therefore our findings provide an alternative way to construct an electron beam splitter on the basis of TI junctions. Using the physics of TMF phenomenon, we also introduce a procedure for experimentally measuring the GH shift of electron beams in 2D electronic systems which may pave a new route in spin/valley-tronics.

VI ACKNOWLEDGMENTS

This work was supported by Iran National Science Foundation: INSF (Grant No. 96017337).

References

  • (1) M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010).
  • (2) X.-L. Qi and S.-C. Zhang, Rev. Mod. Phys. 83, 1057 (2011).
  • (3) M. Z. Hasan and J. E. Moore, Annu. Rev. Condens. Matter Phys. 2, 55 (2011).
  • (4) L. Fu, Phys. Rev. Lett. 103, 266801 (2009).
  • (5) Y. L. Chen, J. G. Analytis, J. H. Chu, Z. K. Liu, S. K.Mo, X. L. Qi, H. J. Zhang, D. H. Lu, X. Dai, Z. Fang, S. C. Zhang, I. R. Fisher, Z. Hussain, and Z.-X. Shen, Science 325, 178 (2009).
  • (6) K. Kuroda, M. Arita, K. Miyamoto, M. Ye, J. Jiang, A. Kimura, E. E. Krasovskii, E. V. Chulkov, H. Iwasawa, T. Okuda, K. Shimada, Y. Ueda, H. Namatame, and M. Taniguchi, Phys. Rev. Lett. 105, 076802 (2010).
  • (7) C. M. Wang and F. J. Yu, Phys. Rev. B 84, 155440 (2011).
  • (8) J. Wang, W. Li, P. Cheng, C. Song, T. Zhang, P. Deng, X. Chen, X. Ma, K. He, J. F. Jia, Q.-K. Xue, and B. F. Zhu, Phys. Rev. B 84, 235447 (2011).
  • (9) H. K. Pal, V. I. Yudson, and D. L. Maslov, Phys. Rev. B 85, 085439 (2012).
  • (10) J. An and C. S. Ting, Phys. Rev. B 86, 165313 (2012).
  • (11) P. Rakyta, A. Palyi, and J. Cserti, Phys. Rev. B 86, 085456 (2012).
  • (12) Z. B. Siu, M. B. A. Jalil, and S. G. Tan, Sci. Rep. 4, 5062 (2014).
  • (13) S. Roy, A. Soori, and S. Das, Phys. Rev. B 91, 041109(R) (2015).
  • (14) F. Hassler, A. R. Akhmerov, and C.W. J. Beenakker, Phys. Rev. B 82, 125423 (2010).
  • (15) E. V. Repin and I. S. Burmistrov, J. Exp. Theor. Phys. 121, 509 (2015).
  • (16) Z.-M. Yu, D.-S. Ma, H. Pan, Y. Yao, Phys. Rev. B 96, 125152 (2017).
  • (17) R. S. Akzyanov and A. L. Rakhmanov, Phys. Rev. B 97, 075421 (2018).
  • (18) R. S. Akzyanov and A. L. Rakhmanov, Phys. Rev. B 99, 045436 (2019).
  • (19) F. Goos and H. Hänchen, Ann. Phys. (Leipzig) 436, 333 (1947).
  • (20) T. Tamir, J. Opt. Soc. Am. A 3, 558(1986).
  • (21) C. W. Hsue and T. Tamir, J. Opt. Soc. Am. A 2, 1809 (1985).
  • (22) C.-F. Li, Phys. Rev. Lett. 91, 133903 (2003).
  • (23) J. Broe and O. Keller, J. Opt. Soc. Am. A 19, 1212 (2002).
  • (24) D. W. Wilson, E. N. Glytsis, and T. K. Gaylord, IEEE J. Quantum Electron. 29, 1364 (1993).
  • (25) X. Chen, C.-F. Li, and Y. Ban, Phys. Rev. B 77, 073307 (2008).
  • (26) X. Chen, X.-J. Lu, Y. Wang, and C.-F. Li, Phys. Rev. B 83, 195409 (2011).
  • (27) P. Gruszecki, M. Mailyan, O. Gorobets, and M. Krawczyk, Phys. Rev. B 95, 014421 (2017).
  • (28) Z. M. Yu, Y. Liu, Y. Yao, and S. A. Yang, Phys. Rev. Lett. 121, 176602 (2018)
  • (29) C. W. J. Beenakker, R. A. Sepkhanov, A. R. Akhmerov and J. Tworzydlo, Phys. Rev. Lett. 102, 146804 (2009).
  • (30) Z. Wu, F. Zhai, F. M. Peeters, H. Q. Xu and K. Chang, Phys. Rev. Lett. 106, 176802 (2011).
  • (31) F. Zhai, Y. Ma, and K. Chang, New J. Phys. 13, 083029 (2011).
  • (32) S. Ghosh and M. Sharma, J. Phys.: Condens. Matter 21, 292204 (2009).
  • (33) X. Chen, J.-W. Tao, and Y. Ban, Eur. Phys. J. B 79, 203 (2011).
  • (34) Y. Song, H.-C. Wu, and Y. Guo, Appl. Phys. Lett. 100, 253116 (2012).
  • (35) X. Chen, P. L. Zhao, X. J. Lu, Eur. Phys. J. B 86, 223 (2013).
  • (36) J. F. Sun and F. Cheng, J. Appl. Phys. 115, 133703 (2014).
  • (37) H. Ghadiri and A. Saffarzadeh, J. Phys.: Condens. Matter 29, 115303 (2017).
  • (38) H. Ghadiri and A. Saffarzadeh, J. Appl. Phys. 123, 104301 (2018).
  • (39) X. J. Qiu , Q. Lv and Z. Z. Cao, J. Phys.: Condens. Matter 31, 225303 (2019).
  • (40) E. S. Azarova and G. M. Maksimova ,J. Phys. Chem. Solids 100, 143 (2016).
  • (41) J. Kuai and H. X. Da, J. Magn. Magn. Mater. 354, 355 (2014).
  • (42) Q. D. Jiang, H. Jiang, H. Liu, Q. F. Sun, and X. C. Xie, Phys. Rev. Lett. 115, 156602 (2015).
  • (43) R.-F. Zheng, L. Zhou, and W. Zhang, Phys. Lett. A 381, 3798 (2017).
  • (44) Y. Liu, Z. M. Yu, H. Jiang, and S. A. Yang, Phys. Rev. B 98, 075151 (2018)
  • (45) Y. Liu, Z. M. Yu, J. Liu, H. Jiang, and S. A. Yang, Phys. Rev. B 98, 195141 (2018)
  • (46) U. Chattopadhyay, L. K. Shi, B. Zhang, J. C. W. Song, and Y. D. Chong, Phys. Rev. Lett. 122, 066602 (2019).
  • (47) W. Wu, S. Chen, C. Mi, W. Zhang, H. Luo, and S. Wen, Phys. Rev. A 96, 043814 (2017).
  • (48) G. Ye, W. Zhang, W. Wu, S. Chen, W. Shu, H. Luo, and S. Wen, Phys. Rev. A 99, 023807 (2019).
  • (49) X. Chen, X. J. Lu, Y. Ban, and C. F. Li, J. Opt. 15, 033001 (2013)
  • (50) T. Taychatanapat, K. Watanabe, T. Taniguchi, P. Jarillo-Herrero, Nat. Phys. 9, 225 (2013).
  • (51) S. P. Milovanovic, M. Ramezani Masir, F. M. Peeters, J. Appl. Phys. 115, 043719 (2014).
  • (52) V. S. Tsoi, J. Bass, and P. Wyder, Rev. Mod. Phys. 71, 1641 (1999).
  • (53) S. Bhandari, G.-H. Lee, K. Watanabe, T. Taniguchi, P. Kim, and R. M. Westervelt, Nano Lett. 20, 4890 (2020).
  • (54) L. P. Rokhinson, V. Larkina, Y. B. Lyanda-Geller, L. N. Pfeiffer, and K. W. West, Phys. Rev. Lett. 93, 146601 (2004).
  • (55) S. Chen, Z. Han, M. M. Elahi, K. M. M. Habib, L. Wang, B. Wen, Y. Gao, T. Taniguchi, K.Watanabe, J. Hone, A.W. Ghosh, and C. R. Dean, Science 353, 1522 (2016).
  • (56) S. Bhandari, G. H. Lee, A. Klales, K. Watanabe, T. Taniguchi, E. Heller, P. Kim, and R. M. Westervelt, Nano Lett. 16, 1690 (2016).
  • (57) S. Bhandari, G. H. Lee, P. Kim, R. M. Westervelt, J. Electron. Mater. 46, 3837 (2017).
  • (58) S. Bhandari, G. H. Lee, K. Watanabe, T. Taniguchi. P. Kim, and R. M. Westervelt, 2D Mater. 5, 021003 (2018).
  • (59) M.-M. Wu, Z.-M. Yu and H. Pan, Phys. Lett. A 383, 237 (2019).
  • (60) M. Arabikhah and A. Saffarzadeh, J. Phys.: Condens. Matter 31, 445001 (2019).
  • (61) At V0=EFV_{0}=E_{F}, the angle in which TIR begins is dependent on dd value. By increasing dd, although the angle decreases, it never reaches zero due to the Klein tunneling effect.
  • (62) O. Klein, Z. Phys. 53, 157 (1929).
  • (63) M. I. Katsnelson, K. S. Novoselov, and A. K. Geim, Nature Phys. 2, 620 (2006).
  • (64) In Eq. (14), the second and the third terms can be ignored in the vicinity of the peak positions, compared to the first term.
  • (65) Z.-M. Yu, Y. Liu, and S. A. Yang, Front. Phys. 14, 33402 (2019).
  • (66) See the result of this paper in Sec. III .
  • (67) C.-F. Li and Q. Wang, Phys. Rev. E 69, 055601(R) (2004).
  • (68) S. Chen, C. Mi, L. Cai, M. Liu, H. Luo, and S. Wen, Appl. Phys. Lett. 110, 031105 (2017).
  • (69) C.-H. Park, Y.-W. Son, L. Yang, M. L. Cohen, and S. G. Louie, Nano Lett. 8, 2920 (2008).
  • (70) V. V. Cheianov, V. Falíko, and B. L. Altshuler, Science 315, 1252 (2007).
  • (71) A. G. Moghaddam, and M. Zareyan, Phys. Rev. Lett. 105, 146803 (2010).
  • (72) M.-H. Liu, C. Gorini, and K. Richter, Phys. Rev. Lett. 118, 066801 (2017).
  • (73) Y. Betancur-Ocampo, F. Leyvraz, and T. Stegmann, Nano Lett. 19, 7760 (2019).
  • (74) F. Libisch, T. Hisch, R. Glattauer, L. A. Chizhova, and J. Burgdorfer, J. Phys.: Condens. Matter 29, 114002 (2017).
  • (75) X. Zhou, A. Kerelsky, M. M. Elahi, D. Wang, K. M. M. Habib, R. N. Sajjad, P. Agnihotri, J. U. Lee, A. W. Ghosh, F. M. Ross, and A. N. Pasupathy, ACS Nano 13, 2558 (2019).
  • (76) S. W. LaGasse, and C. D. Cress, Nano Lett. 20, 6623 (2020).
  • (77) Y. Xie, Y. Tan, and A. W. Ghosh, Phys. Rev. B 96, 205151 (2017).
  • (78) Y. Li, Q. Wan, Y. Peng, G. Wang, Z. Qian, G. Zhou, and M. B. A. Jalil, Sci. Rep. 5, 18458 (2015).
  • (79) In graphene, the electronic GH shift is spin and valley independent. However, with magnetic barrier and/or considering small intrinsic spin-orbit coupling or extrinsic Rashba-type spin-orbit coupling, the GH shift will depend on spin and valley.