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

Spin Hall effect of reflected light at the air-uniaxial crystal interface

Open Access Open Access

Abstract

We theoretically and experimentally study the spin Hall effect of reflected light at an air-uniaxial crystal interface. The spin-dependent nanometer-sized displacements depend not only on the incident polarization and the incident angle of the light beam, but also on the orientation of the crystal optic axis. The experimental results are in perfect agreement with theoretical predictions for the vertical and horizontal polarization incidence.

©2010 Optical Society of America

1. Introduction

The reflection and transmission of a plane wave at a planar interface is presented by Snell’s Law and the Fresnel formulas. A bounded beam, a wave packet containing a set of plane wave components with different wave-vectors, however, may undergo the longitudinal Goos-Hanchen shift or/and the transverse Imbert-Fedorov (IF) shift. For the IF effect, the center of the reflected beam shows a small spatial shift perpendicular to the plane of incidence, relative to the position predicted by geometrical optics [14]. During total-reflection, a linearly polarized incident beam splits into two elliptically polarized rays that displace symmetrically [5]. In a uniaxial anisotropic media, the propagation of a beam has been investigated by taking into account of both the ordinary and the extraordinary rays [69]. Perez [10] analyzed the first-order nonspecular transverse effects of three dimensional Gaussian beams. Fadeyeva et al. [11] have found the asymmetric splitting of a high-order circularly polarized vortex-beam in a uniaxial crystal.

Recently, much attention has been paid to the spin Hall effect of light (SHEL) [12]. SHEL is a spin dependent transverse shift of the wave packet perpendicular to the gradient of the refractive index, which has been extensively studied theoretically [1319] and experimentally [16,2022]. This effect originates from the geometric phase gradient in the spin states, which arises from rotation of the coordinate frames attached to partial plane waves with different planes of incidence [15,21]. In fact, the conservation of the total angular momentum of light plays a key role in the phenomenon [14,23], and SHEL can be described as the result of the spin-orbit interaction [12,21]. Up to now, these researches have mostly focused on the homogenous isotropic planar interface. Here we study, theoretically and experimentally, the spin-dependent nanometer-sized displacements of the reflected light induced by SHEL from an air-uniaxial crystal interface that is extensively used in optics.

We demonstrate SHEL by considering the reflection from a planar interface of the uniaxial crystal sample (LiNbO3). When reflected at the interface, a linearly polarized incident beam splits into two spin components that acquire opposite lateral displacements perpendicular to the incident plane: the component parallel ( = + 1, right-circularly polarized, denoted by) and antiparallel ( = −1, left-circularly polarized, denoted by) to the central wave vector, as shown in Fig. 1 . We use a laboratory Cartesian frame (x, y, z) attached to the interface, and employ the coordinate systems of individual beams, where (xI, y, zI) attached to the incident beam and (xR, y, zR) attached to the reflected beam. The zI, zR axes attach to the directions of the central wave vector of the incident and reflected beam as determined by the Snell’s law, respectively.

 figure: Fig. 1

Fig. 1 Schematic of SHEL at an air–uniaxial crystal interface. θI, incident angle; δy|+>, the displacement of |+> spin component. The subscripts I and R correspond to incident and reflected light, respectively.

Download Full Size | PDF

2. Theoretical analysis

The eigenstates of reflection are the linear p- and s-polarization states. For the air-uniaxial crystal interface, their Fresnel reflection coefficients vary with the ordinary and the extraordinary components, which is more intricate than that in the air-glass interface. Therefore, the displacements induced by SHEL depend not only on the incident polarization of the beam, but also on the orientation of the crystal optic axis.

An incident Gaussian beam can be considered as a wave-packet containing a distribution of wave-vectors k(I)=kI(z^I+κ(I)) centered aroundkIz^I, with κ(I)=(kxIx^I+kyy^)/kI =κxIx^I+κyy^, |κ(I)|<<1, and z^I is a unit vector along zI axis and kI=2π/λ with λ being the wavelength of the light in the incident medium. For the reflected beam: k(R)=kR(z^R+κ(R)) with κ(R) =(kxRx^R+kyy^)/kR=κxRx^R+κyy^, |κ(R)|<<1, and kR = kI for Snell’s Law. For an arbitrary wave-vector, k^(I)z^(z^k^(I))=k^(R)z^(z^k^(R)), connecting κ(I) and κ(R):kxI=kxR and kyI=kyR=ky.

For the horizontally and the vertically polarized incident wave-packets denoted by |H(k(I))> and |V(k(I))>, we can expand everything up to the first order in κ [21,22]:

|H(k(I,R))>=|p(k(I,R))>cotθI,Rκy(I,R)|s(k(I,R))>,|V(k(I,R))>=|s(k(I,R))>+cotθI,Rκy(I,R)|p(k(I,R))>.
The p- and s-polarization eigenstates in Eq. (1) are defined by p^(k(I,R))=s^(k(I,R))×k^(I,R) and , in which 1sin(θI,R(k^(I,R))) 1sinθI,RcotθI,RsinθI,RκxI,xR, where θI is the incident angle of central wave vector related by Snell’s law and θR=πθI. The symbol ∧ denotes a unit vector. Due to the anisotropy of the crystal, the SHEL is dependent on the orientation of the crystal optic axis. In the following, the spin displacement expressions are obtained for three cases where the optic axis is along the x, y, z-axis, respectively.

2.1 Crystal optic axis is along the x-axis

When the crystal optic axis is along the x-axis, the linear p- and s-polarization states of the incident light beam can be written as:

|p(k(I))>=|e(k(I))>+(tanθI+cotθI)κy(I)|o(k(I))>,|s(k(I))>=|o(k(I))>(tanθI+cotθI)κy(I)|e(k(I))>,
where o^(k(I))=k^(I)×x^/sin(ξ(k(I))) and e^(k(I))=o^(k(I))×k^(I) with 1sinξ(k(I)) 1sinξI+cotξIsinξIκxI and ξI=π2θI. Due to the distribution of wave-vectors, the angle ξ(k(I,R)) between the wave vector k(I,R)and the crystal optic axis is around the angle ξI,Rbetween the central wave vector kIz^I and the crystal optic axis. Therefore, the p- or s-polarization state includes e- or o-component with a small amount of o- or e-component.

Using Eqs. (1) and (2), the horizontally and vertically polarized beams are expressed by

|H(k(I))>=|pe(k(I))>(tanθI+cotθI)κy(I)|se(k(I))>+tanθIκy(I)|so(k(I))>,|V(k(I))>=|so(k(I))>+(tanθI+cotθI)κy(I)|po(k(I))>tanθIκy(I)|pe(k(I))>.
After reflection, s- and p-polarizations in Eq. (3) evolve as: |so/e(k(I))> rs,o/s,ex|so/e(k(R))> and |po/e(k(I))>rp,o/p,ex|po/e(k(R))>. rp,o/p,ex and rs,o/s,ex are the Fresnel reflection coefficients at the incident angle of θI for the p- and s-polarization states related to the ordinary or extraordinary rays, respectively, when the crystal optic axis is along x axis. Thus, we can obtain:
|H(k(I))>rp,ex(|H(k(R))>+kyδyxH|V(k(R))>),|V(k(I))>rs,o(|V(k(R))>kyδyxV|H(k(R))>).
In the spin basis set |+>=12(|H>+i|V>) and|>=12(|H>i|V>), Eq. (4) are expressed by
|H(k(I))>rp,ex2[exp(ikyδyx|+>H)|+>+exp(ikyδyx|>H)|>],|V(k(I))>irs,o2[exp(ikyδyx|+>V)|+>exp(ikyδyx|>V)|>],
where

δyx|±>H=cotθIkI(1+rs,exrp,ex1cos2θIrs,orp,extan2θI),
δyx|±>V=cotθIkI(1+rp,ors,o1cos2θIrp,exrs,otan2θI).

Equation (5) shows that two spin components of the reflected light shift oppositely in y axis, with a displacement of δyαH or δyαV for the horizontal or the vertical incident polarization when the crystal optic axis is along the α-axis (α = x, y, z, and α = x in Eqs. (5) and (6)).

2.2 Crystal optic axis is along the y-axis

In this situation, 1sinξ(k(I))1sinξI+cotξIsinξIκxI(I)=1 with ξI=π2. Using the expressions: o^(k(I))=y^×k^(I) and e^(k(I))=k^(I)×o^(k(I)), we have

|p(k(I))>=|o(k(I))>+cotθIκy(I)|e(k(I))>,|s(k(I))>=|e(k(I))>cotθIκy(I)|o(k(I))>.
After reflection, the horizontally and vertically polarized beams evolve into:
|H(k(I))>rp,o2[exp(ikyδyy|+>H)|+>+exp(ikyδyy|>H)|>],|V(k(I))>irs,ey2[exp(ikyδyy|+>V)|+>exp(ikyδyy|>V)|>],
where

δyy|±>H=cotθIkI(1+rs,orp,o),
δyy|±>V=cotθIkI(1+rp,eyrs,ey).

2.3 Crystal optic axis is along the z-axis

In this case, the linear s- and p-polarization states can be simply expressed as follows:

|p(k(I)))=|e(k(I))>,|s(k(I))>=|o(k(I))>,
witho^(k(I))=z^×k^(I)/sin(ξ(k(I))) and e^(k(I))=o^(k(I))×k^(I), where ξ(k(I))=θI(k(I)). And we can obtain:
|H(k(I))>rp,ez2[exp(ikyδyz|+>H)|+>+exp(ikyδyz|>H)|>],|V(k(I))>irs,o2[exp(ikyδyz|+>V)|+>exp(ikyδyz|>V)|>],
where

δyz|±>H=cotθIkI(1+rs,orp,ez),
δyz|±>V=cotθIkI(1+rp,ezrs,o).

Equations (5), (8) and (11) present the analytical expressions for the evolution of the horizontally or the vertically polarized Gaussian beam after reflection from an air-uniaxial crystal interface. The reflected light beam experiences spin-dependent displacements along the y-axis. The amount of the splitting between the two spin components are given by Eqs. (6), (9) and (12). It shows that the displacement induced by SHEL depends not only on the incident polarization and the incident angle, but also on the orientation of the crystal optic axis. In fact, the p- or s-polarization state in Eqs. (2), (7) and (10) includes e- or o-component with a small amount of o- or e-component. If replacing the air-uniaxial crystal interface with an air-glass one, Eqs. (6), (9) and (12) reduce to the Eqs. (1) and (2) in the Ref [22]. since the difference between the ordinary and extraordinary rays disappears.

3. Experimental results and discussion

The spin-dependent shifts are measured by the experimental setup that is similar to that in Refs [21,22]. As shown in Fig. 2 , a Gaussian beam at 632.8 nm is generated by a He–Ne laser, passing through the first lens L1 with focal length of 25 mm and a Glan polarizer P1 to produce an initially linearly polarized focused beam. The beam splits into its two spin components when reflected at the interface of LiNbO3 with the size of 10 × 5 × 4 mm, whose crystal optic axis is along the side of 4 mm. The refractive index for the ordinary and extraordinary rays are no = 2.232, ne = 2.156 at 632.8 nm, respectively. Because the displacement is at the nanometer scale, the intensity distribution on the cross section of the reflected beam along the propagation direction is nearly unchanged. The second polarizer P2 is nearly orthogonal to the central wave vector of the reflected light. After P2, the two components of the splitting beam interfere. As a result, the light intensity redistributes and the position of the light spot moves [17]. This movement of the spot corresponds to an amplification of the relative displacement between the two splitting spin components. The second lens L2 with focal length of 125 mm is used to collimate the beam. Then, we use a position-sensitive detector (PSD) to measure the amplified displacement after L2 for calculating the original displacement induced by SHEL. As a matter of fact, this experimental setup, which is only sensitive to spatial displacements of the wave packets [22], converts the position displacements caused by SHEL into a momentum shift.

 figure: Fig. 2

Fig. 2 Experimental setup. The He–Ne laser generates a Gaussian beam at 632.8 nm; HWP, half-wave plate for attenuating the intensity after P1 to prevent the position-sensitive detector (PSD) from saturating; L1 and L2, lenses with 25 and 125 mm focal lengths, respectively; P1 and P2, Glan polarizers.

Download Full Size | PDF

Figure 3 describes the displacement δ|+>V of the |+> spin component induced by SHEL as a function of the incident angle, in the case of vertical polarization (the central electric-field-vector E//y). The blue, red, green curves indicate the theoretical prediction when the crystal optic axis is along x, y, z axis, respectively, and the dots, circles, triangles are the experimental results. The theory is nicely confirmed by the experimental data. Limited by the large holders of the optical elements and the small area of the interface, displacements at small and large incident angles were not measured. The negative displacement indicates that the |+> spin component shifts along the −y direction at the interface. The displacement increases and reaches the maximum, then subsequently decreases to zero as the incident angle increases to 90 deg. The red and green curves almost overlap due to the similarity of the ratios rp,ey/rs,ey and rp,ez/rs,o.

 figure: Fig. 3

Fig. 3 Displacements of the |+> spin component, δ|+>V, as a function of incidence angle, θI, in the case of vertical polarization. The dots, circles, triangles correspond to the cases that the crystal optic axis is along x, y, z axis respectively. The solid curves represent the theoretical predictions.

Download Full Size | PDF

We also measured the displacements for the horizontal polarization incidence. In this case, the displacement δ|+>H increases with the incident angle θI, and rises rapidly when approaching Brewster angle θB (~65, 66 and 66 degree for the three different direction of crystal optic axis in our experiment), as illustrated in Fig. 4(a) . The inset shows the results for 0<θI <90°. When θI>θB, the | + > spin component shifts oppositely and the displacement decreases withθI as shown in Fig. 4(b). The details in the dashed frame are shown in the inset. There are obviously deviations from the theory in the proximity of θB. The deviations may be caused by two reasons [22]: the Eqs. (6), (9) and (12) need to be modified near Brewster angle (rp0) and the intensity is too weak to accurately determine the position for PSD. When the incident angle is far from the Brewster angle, the experimental data are in agreement with the theory. For the horizontal incidence, the blue curve is overlapped with the red one in the most range because rs,exor rs,o are almost identical to rp,ex or rp,o.

 figure: Fig. 4

Fig. 4 Displacements δ|+>H as a function of incidence angleθI, in the case of horizontal polarization. The blue, red, green curves indicate the theoretical prediction when the crystal optic axis is along x, y, z axis, respectively, and the dots, circles, triangles are the corresponding experimental data. (a) θI< Brewster angleθB. The inset shows the results for 0<θI<90°; (b) θI> θB. The inset shows the enlarged image for 71° <θI<75°.

Download Full Size | PDF

4. Conclusions

In conclusion, we have calculated and measured the displacements induced by the spin Hall effect of the reflected light for an air-uniaxial crystal interface. Experimental results are in good agreement with the theory taking into account of the anisotropy of the crystal. Because the displacements depend not only on the incident polarization and the incident angle of the light beam, but also on the orientation of the crystal optic axis, it offers a flexible way to control the amount of the displacement.

Acknowledgments

The authors acknowledge financial support from the National Natural Science Foundation of China under Grant No 10821062, and the National Basic Research Program of China under Grant Nos. 2006CB921601 and 2007CB307001.

References and links

1. C. Imbert, “Calculation and experimental proof of transverse shift induced by total internal reflection of a circularly polarized-lightbeam,” Phys. Rev. D Part. Fields 5(4), 787–796 (1972).

2. V. G. Fedoseyev, “Conservation-Laws and transverse motion of energy on reflection and transmission of electromagnetic-waves,” J. Phys. A 21(9), 2045–2059 (1988). [CrossRef]  

3. M. A. Player, “Angular-Momentum balance and transverse shifts on reflection of light,” J. Phys. A 20(12), 3667–3678 (1987). [CrossRef]  

4. K. Y. Bliokh, I. V. Shadrivov, and Y. S. Kivshar, “Goos-Hänchen and Imbert-Fedorov shifts of polarized vortex beams,” Opt. Lett. 34(3), 389–391 (2009). [CrossRef]   [PubMed]  

5. N. N. Punko and V. V. Filippov, “Beam splitting due to the finite size of the medium during total reflection,” JETP Lett. 39, 20–23 (1984).

6. A. Ciattoni, B. Crosignani, and P. Di Porto, “Vectorial theory of propagation in uniaxially anisotropic media,” J. Opt. Soc. Am. A 18(7), 1656–1661 (2001). [CrossRef]  

7. A. Ciattoni and C. Palma, “Optical propagation in uniaxial crystals orthogonal to the optical axis: paraxial theory and beyond,” J. Opt. Soc. Am. A 20(11), 2163–2171 (2003). [CrossRef]  

8. F. Flossmann, U. T. Schwarz, M. Maier, and M. R. Dennis, “Polarization singularities from unfolding an optical vortex through a birefringent crystal,” Phys. Rev. Lett. 95(25), 253901 (2005). [CrossRef]   [PubMed]  

9. E. Brasselet, Y. Izdebskaya, V. Shvedov, A. S. Desyatnikov, W. Krolikowski, and Y. S. Kivshar, “Dynamics of optical spin-orbit coupling in uniaxial crystals,” Opt. Lett. 34(7), 1021–1023 (2009). [CrossRef]   [PubMed]  

10. L. I. Perez, “Nonspecular transverse effects of polarized and unpolarized symmetric beams in isotropic-uniaxial interfaces,” J. Opt. Soc. Am. A 20(4), 741–752 (2003). [CrossRef]  

11. T. A. Fadeyeva, A. F. Rubass, and A. V. Volyar, “Transverse shift of a high-order paraxial vortex-beam induced by a homogeneous anisotropic medium,” Phys. Rev. A 79(5), 053815 (2009). [CrossRef]  

12. M. Onoda, S. Murakami, and N. Nagaosa, “Hall effect of light,” Phys. Rev. Lett. 93(8), 083901 (2004). [CrossRef]   [PubMed]  

13. 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]   [PubMed]  

14. K. Y. Bliokh and Y. P. Bliokh, “Polarization, transverse shifts, and angular momentum conservation laws in partial reflection and refraction of an electromagnetic wave packet,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 75(6 ), 066609 (2007). [CrossRef]   [PubMed]  

15. K. Y. Bliokh, Y. Gorodetski, V. Kleiner, and E. Hasman, “Coriolis effect in optics: unified geometric phase and spin-Hall effect,” Phys. Rev. Lett. 101(3), 030404 (2008). [CrossRef]   [PubMed]  

16. K. Y. Bliokh, A. Niv, V. Kleiner, and E. Hasman, “Geometrodynamics of spinning light,” Nat. Photonics 2(12), 748–753 (2008). [CrossRef]  

17. A. Aiello and J. P. Woerdman, “Role of beam propagation in Goos-Hänchen and Imbert-Fedorov shifts,” Opt. Lett. 33(13), 1437–1439 (2008). [CrossRef]   [PubMed]  

18. C. M. Krowne, “Spin moments of particles detected using the spin Hall effect of light through weak quantum measurements,” Phys. Lett. A 373(4), 466–472 (2009). [CrossRef]  

19. H. L. Luo, S. C. Wen, W. X. Shu, Z. X. Tang, Y. H. Zou, and D. Y. Fan, “Spin Hall effect of a light beam in left-handed materials,” Phys. Rev. A 80(4), 043810 (2009). [CrossRef]  

20. M. Merano, A. Aiello, M. P. van Exter, and J. P. Woerdman, “Observing angular deviations in the specular reflection of a light beam,” Nat. Photonics 3(6), 337–340 (2009). [CrossRef]  

21. O. Hosten and P. Kwiat, “Observation of the spin hall effect of light via weak measurements,” Science 319(5864), 787–790 (2008). [CrossRef]   [PubMed]  

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

23. V. G. Fedoseyev, “Conservation laws and angular transverse shifts of the reflected and transmitted light beams,” Opt. Commun. 282(7), 1247–1251 (2009). [CrossRef]  

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 (4)

Fig. 1
Fig. 1 Schematic of SHEL at an air–uniaxial crystal interface. θ I , incident angle; δ y | + > , the displacement of | + > spin component. The subscripts I and R correspond to incident and reflected light, respectively.
Fig. 2
Fig. 2 Experimental setup. The He–Ne laser generates a Gaussian beam at 632.8 nm; HWP, half-wave plate for attenuating the intensity after P1 to prevent the position-sensitive detector (PSD) from saturating; L1 and L2, lenses with 25 and 125 mm focal lengths, respectively; P1 and P2, Glan polarizers.
Fig. 3
Fig. 3 Displacements of the | + > spin component, δ | + > V , as a function of incidence angle, θ I , in the case of vertical polarization. The dots, circles, triangles correspond to the cases that the crystal optic axis is along x, y, z axis respectively. The solid curves represent the theoretical predictions.
Fig. 4
Fig. 4 Displacements δ | + > H as a function of incidence angle θ I , in the case of horizontal polarization. The blue, red, green curves indicate the theoretical prediction when the crystal optic axis is along x, y, z axis, respectively, and the dots, circles, triangles are the corresponding experimental data. (a) θ I < Brewster angle θ B . The inset shows the results for 0< θ I <90°; (b) θ I > θB. The inset shows the enlarged image for 71° < θ I <75°.

Equations (15)

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

| H ( k ( I , R ) ) > = | p ( k ( I , R ) ) > cot θ I , R κ y ( I , R ) | s ( k ( I , R ) ) > , | V ( k ( I , R ) ) > = | s ( k ( I , R ) ) > + cot θ I , R κ y ( I , R ) | p ( k ( I , R ) ) > .
| p ( k ( I ) ) > = | e ( k ( I ) ) > + ( tan θ I + cot θ I ) κ y ( I ) | o ( k ( I ) ) > , | s ( k ( I ) ) > = | o ( k ( I ) ) > ( tan θ I + cot θ I ) κ y ( I ) | e ( k ( I ) ) > ,
| H ( k ( I ) ) > = | p e ( k ( I ) ) > ( tan θ I + cot θ I ) κ y ( I ) | s e ( k ( I ) ) > + tan θ I κ y ( I ) | s o ( k ( I ) ) > , | V ( k ( I ) ) > = | s o ( k ( I ) ) > + ( tan θ I + cot θ I ) κ y ( I ) | p o ( k ( I ) ) > tan θ I κ y ( I ) | p e ( k ( I ) ) > .
| H ( k ( I ) ) > r p , e x ( | H ( k ( R ) ) > + k y δ y x H | V ( k ( R ) ) > ) , | V ( k ( I ) ) > r s , o ( | V ( k ( R ) ) > k y δ y x V | H ( k ( R ) ) > ) .
| H ( k ( I ) ) > r p , e x 2 [ exp ( i k y δ y x | + > H ) | + > + exp ( i k y δ y x | > H ) | > ] , | V ( k ( I ) ) > i r s , o 2 [ exp ( i k y δ y x | + > V ) | + > exp ( i k y δ y x | > V ) | > ] ,
δ y x | ± > H = cot θ I k I ( 1 + r s , e x r p , e x 1 cos 2 θ I r s , o r p , e x tan 2 θ I ) ,
δ y x | ± > V = cot θ I k I ( 1 + r p , o r s , o 1 cos 2 θ I r p , e x r s , o tan 2 θ I ) .
| p ( k ( I ) ) > = | o ( k ( I ) ) > + cot θ I κ y ( I ) | e ( k ( I ) ) > , | s ( k ( I ) ) > = | e ( k ( I ) ) > cot θ I κ y ( I ) | o ( k ( I ) ) > .
| H ( k ( I ) ) > r p , o 2 [ exp ( i k y δ y y | + > H ) | + > + exp ( i k y δ y y | > H ) | > ] , | V ( k ( I ) ) > i r s , e y 2 [ exp ( i k y δ y y | + > V ) | + > exp ( i k y δ y y | > V ) | > ] ,
δ y y | ± > H = cot θ I k I ( 1 + r s , o r p , o ) ,
δ y y | ± > V = cot θ I k I ( 1 + r p , e y r s , e y ) .
| p ( k ( I ) ) ) = | e ( k ( I ) ) > , | s ( k ( I ) ) > = | o ( k ( I ) ) > ,
| H ( k ( I ) ) > r p , e z 2 [ exp ( i k y δ y z | + > H ) | + > + exp ( i k y δ y z | > H ) | > ] , | V ( k ( I ) ) > i r s , o 2 [ exp ( i k y δ y z | + > V ) | + > exp ( i k y δ y z | > V ) | > ] ,
δ y z | ± > H = cot θ I k I ( 1 + r s , o r p , e z ) ,
δ y z | ± > V = cot θ I k I ( 1 + r p , e z r s , o ) .
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.