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

Enhanced displacements in reflected beams at hyperbolic metamaterials

Open Access Open Access

Abstract

We examine the Goos-Hänchen (G-H) shift of a Gaussian beam reflected on a thin slab of Ag/TiO2 hyperbolic multilayer metamaterial (HMM). The HMM is modeled using the effective medium theory which yields the anisotropic dielectric functions of the HMM. The G-H shifts can be very large on the surface of the HMM. It can be about 40 µm which are far bigger than the G-H shifts on the usual materials like metals and dielectrics. The enhancement is due to the excitation of the Brewster modes in HMM. Such Brewster modes in HMM have a well-defined frequency-dependent line shape. We relate the the half width at half maximum of the G-H shift to the imaginary part of the complex frequency of the Brewster mode. Moreover, we also present results for the Imbert-Fedorov shifts as well as the spin Hall effect of light on the surface of a thin HMM slab. We show that the spin Hall effect on the HMM slab is much more pronounced than that on the surface of metal. Thus a thin HMM slab can be used to enhance the lateral displacements, which can have many interesting applications for optical devices.

© 2016 Optical Society of America

1. Introduction

A beam of light reflected off an interface experiences a longitudinal Goos-Hänchen (G-H) shift [1] as well as a transverse Imbert-Fedorov (I-F) shift [2,3]. Both the G-H and I-F shifts have been widely studied in different systems, such as the dielectric [4], metallic [5] and photonic crystal [6] slabs. Shifts at the surface of more complex systems like topological insulators are also studied now [7]. Even the control of the G-H shifts by using a pump beam has been considered [8]. In general, it is difficult to observe the G-H shift at a plane metallic interface in experiments since the shift has the same order as the wavelength of the incident light. However, Merano et al. [9] succeeded in observing the shifts at a bare metallic surface. Various schemes have been proposed to enhance the G-H shift in order to make a direct measurement. These include the metasurfaces [10], thin films [11], and left-handed metamaterials [12]. In recent years, the beam shifts also stimulated the demonstration of weak measurement of shifts [13–15], and led to applications in sensing and switching [16, 17].

As another interesting effect of the lateral displacement, the spin Hall effect of light (SHEL) has been studied both theoretically [18, 19] and experimentally [20–22]. When a linearly polarized beam is reflected, the beam will split into left- and right-circularly polarized components. The splitting results from an effective spin-orbit coupling [18]. The SHEL may help manipulate the photon spin effectively and thus encourage the development of new generations of all-optical devices.

In this paper we propose a scheme using a thin slab made of hyperbolic metamaterial (HMM) to enhance the lateral displacements. Recently, hyperbolic metamaterials (HMMs) have provoked much interest due to their hyperbolic dispersion relation [23]. As a result of the novel property, this metamaterial can support the propagation of high wavevectors. Thus, many interesting properties and applications appeared, including negative refraction [24], sub-diffraction imaging [25, 26], sub-wavelength modes [27, 28], and thermal emission engineering [29]. It is known that there are two most popular HMM structures: a multilayer structure made of alternating metallic and dielectric layers [32], and a nanowire structure fabricated by embedding metallic nanowires in a dielectric medium [33]. In this paper, we focus on multilayer HMM since it’s easier to fabricate and realize.

For the light displacement on HMM, the recent literature [34, 35] mainly discussed the impact of the volume fraction and nonlocal effect on the G-H shifts of reflected light at the surface of infinite multilayer HMM. In this paper, not only have we studied the G-H shift of a reflected Gaussian beam at the surface of a thin HMM slab made of Ag and TiO2, but we have also calculated the I-F shift and the SHEL in the same scheme. The G-H shift, I-F shift and SHEL are calculated, analyzed and compared between slabs made of HMM, Ag and TiO2.

2. Effective medium theory of HMM

The structure of multilayer HMM is shown in Fig. 1(a). In general, an effective medium theory [36, 37] can be used to describe the optical properties of multilayer or nanowire media. In HMM, the diagonal elements of the permittivity tensor satisfy the hyperbolic relation. Here, the HMM is made of alternating layers of silver and TiO2 used in the experiment [32]. According to the effective medium theory, the effective permittivities of the multilayer structure are given by [38],

ϵ=fϵAg+(1f)ϵTiO2,
ϵ=ϵAgϵTiO2fϵTiO2+(1f)ϵAg,
where f is the filling fraction of silver and the permittivity tensor has the form of diag(ϵ, ϵ, ϵ) with respect to the principal axis. In our case, z-axis is the optical axis of the material. Within the Drude model, the permittivity of the silver is ϵAg=ϵωP2/[ω(ω+iτ1)], with the parameters ϵ = 3.7, the bulk plasmon frequency ωP = 1.4 × 1016rad/s and the relaxation time τ = 0.45 × 10−14s. The permittivity of TiO2 can be well described by the formula ϵTiO2=5.913+0.2441/(λ20.0803), where λ is the wavelength in micrometer [39].

 figure: Fig. 1

Fig. 1 (a) Illustration of a multilayer HMM; (b)-(d) describe the effective permittivities ϵ and ϵ for a Ag/TiO2 multilayer HMM with a silver filling fraction of f = 0.35. (b) plots the real part of ϵ and ϵ; (c) and (d) plot the imaginary part of ϵ and ϵ, respectively. The red and blue lines correspond to ϵ and ϵ, respectively.

Download Full Size | PDF

Figures. 1(b)–1(d) show the real and imaginary parts of the effective permittivities for Ag/TiO2 multilayer HMM. In Fig. 1(b), as the wavelength increases, the medium changes from type I HMM to elliptical medium at 395 nm wavelength. And then at 551 nm wavelength, it changes from elliptical medium to type II HMM. As shown in Fig. 1(c) and 1(d), the imaginary part of ϵ is extremely large at the transition point 395nm, while the imaginary part of ϵ remains small. This is because of the denominator in Eq. (2). When the real part of the denominator becomes zero, then the Im ϵ becomes large.

3. Goos-Hänchen shift on HMM

As shown in Fig. 2(a), a light beam is incident on a thin slab made of HMM with an incident angle θ, and is reflected at the surface of HMM. The reflected beam experiences a short longitudinal shift in the incident plane δGH, known as the Goos-Hänchen (G-H) shift. In order to calculate the shift of the beam, we first resolve the incident beam into its plane wave components [10],

EI(r,ω)=d2κei(kxx+kyy)+iw(kx,ky)zEI(κ,ω).
where κ=kxx^+kyy^,w(kx,ky)=k02kx2ky2 and k0 = 2π/λ. Here, we assume that the center of the field distribution locates at x = 0 in the plane of the interface and the beam propagates with a central wave-vector kx0. The field distribution can be described by the following expression,
EI(kx,ky)=E02σxσyπexp[σx2(kxkx0)24]exp[σy2ky24].
σy is the waist radius of the incident beam which is chosen as 160 µm in our model. As a result of the oblique incidence, σx = σy/cos θ. Then, the amplitude of the reflected field can be expressed as,
ER(r,ω)=d2κei(kxx+kyy)iw(kx,ky)zER(κ,ω).

 figure: Fig. 2

Fig. 2 (a) Geometry of the beam reflection at the surface of a slab made of HMM. The reflected beam will undergo a longitudinal shift δGH and transverse shift δIF. (b), (c) and (d) show the G-H shift δGH at the surface of a slab made of Ag/TiO2 multilayer HMM, Ag and TiO2, respectively. The slabs have the same thickness of 120 nm.

Download Full Size | PDF

Consequently, the expression of the G-H shift is given by [4, 10, 40],

δGH=id2κER(kx,ky)kxER(kx,ky)d2κER(kx,ky)ER(kx,ky)λ2πφθ.
φ is the phase of the reflection coefficient. It is clear to see that the shift reaches its peak when the phase of the reflected field experiences a sharp variation over the incident angle. It should be noted that the incident angle for achieving a maximum shift is close to the Brewster angle. In all the calculations we use the rigorous formula (the middle one in Eq. (6)). We will show one case in Fig. 3(a) which exhibits a major difference between exact and approximate results.

 figure: Fig. 3

Fig. 3 (a) plots the G-H shifts calculated using the exact (solid) and approximate (dashed) results at the surface of HMM near the Brewster angle for 345 nm. (b) plots the values of |rp|, and (c) plots the relative phases φ.

Download Full Size | PDF

In traditional cases, the G-H shift for s− polarized beam can be ignored since it’s very small. Thus, we only focus on the case of the p− polarized beam. This is due to the fact that we are not considering here the angular G-H shift. As is pointed out by Merano et al. [41] and Dennis et al. [15], if the angular G-H shift is considered also, the s− polarization alone can be appreciable. Based on Eqs. (3)(6), Fig. 2(b) shows the G-H shifts on the Ag/TiO2 multilayer HMM for p− polarized beam with different incident wavelengths. The thickness d of HMM is chosen as 120 nm. Figure. 2(b) shows that as the wavelength of the incident field increases, the G-H shift changes from positive to negative. The maximum positive value of the shift is 38.4 µm at 69.104° for λ = 330 nm, while the maximum negative value is −38.6 µm at 67.022° for λ = 345 nm. As λ continues to increase, the maximum value and critical angle decrease gradually. Comparing with the G-H shift of traditional materials, the HMM is a very good candidate to enhance G-H shifts.

To compare the G-H shifts at HMM and its two constituents: metal and dielectric, we also calculate the shifts on Ag and TiO2 slabs of the same thickness. As shown in Fig. 2(c) and 2(d), with λ changing from 345 nm to 600 nm, the G-H shifts on the Ag slab are smaller than 1 µm, while on the TiO2 slab, the shifts vary from −1 µm to 3 µm. It is clear that the shifts on HMM are significantly enhanced for specific wavelengths ranging from 330 nm to 395 nm.

Now we relate the well-defined peaks in the G-H shift for HMM to the existence of the Brewster modes [42]. For an anisotropic slab, the reflection coefficient r for p−polarized light is given by [43],

rp=(kezϵk0z)(kez+ϵk0z)(kezϵk0z)(kez+ϵk0z)e2ikezd(kez+ϵk0z)(kez+ϵk0z)(kezϵk0z)(kezϵk0z)e2ikezd,
where kez=ϵ(k02κ2/ϵ),k0z=k02κ2. By setting the numerator of rp to be zero, we can easily get the complex Brewster mode and the corresponding complex Brewster angle θB. The real value of the Brewster angle corresponds to the position of the peak in the shift. Moreover, it’s found that the absolute imaginary value of the complex Brewster angle is equal to the half width at half maximum (HWHM) of the shift. This result indicates that the width of the shift is related to the dissipation of the medium, which gives rise to the imaginary part of the angle. For example, the complex Brewster angle at the surface of HMM for λ = 345 nm is 67.022° + 0.052°i. In Fig. 3(a), we show the calculated G-H shifts near this Brewster angle using the exact and approximate results. It shows that the HWHM is 0.104° when the approximate result is used, exactly equal to the imaginary part of the angle. When the exact result is used, the HWHM is 0.130°, a little larger due to the Gaussian distribution of the wavevector. However, for the lossless dielectric slab considered here, the imaginary value vanishes, leading to the disappearance of the peak in the G-H shift, as shown in Fig. 2(d).

It is well known that the large G-H shift originates from the phase change of the reflection coefficient. The absolute value and phase of rp with λ = 345 nm are plotted in Fig. 3(b) and 3(c). For the HMM slab,|rp| is close (but not equal) to zero at θB = 67.022°, accompanied by a steep phase change (see black curves). This corresponds to the position of the peak in Fig. 2(b), indicating the relation between the shift and phase change. For the lossless dielectric slab, the value of rp is exactly zero and the phase has an abrupt jump (see red curves). In this case, the phase change of an infinite slope without reflection has no physical meaning [4]. For the metal slab, no Brewster mode is found due to high reflection (see green curves). This corresponds to small displacements at the metal slab. It can be seen that the combination of multilayer metal and dielectric slabs modifies the effective permittivities, which leads to large shifts due to the change of the reflection coefficients and to the existence of the Brewster modes. This property can not be found in the metal and dielectric slabs.

Also, it can be seen that the phase transition point from type I HMM to elliptical medium is 395 nm. This wavelength is not equal to transition point of the G-H shifts from positive to negative (i.e., 335 nm). This indicates that the sign of the shift has no direct relevance to the phase of the medium. In order to explain this point, we need to consider the reflection coefficient rp which determines the shift. Equation. (7) shows that both ϵ and ϵ affect the reflection coefficient, indicating that the phase of the medium may determine the shift of the reflection beam. However, as mentioned above, Fig. 1(b) and 1(c) show that ϵ always has a large real or imaginary part. In this case, κ2/ϵ is very small and thus can be ignored, leading to kezϵk0. That’s why the shift doesn’t show the similar property of a sharp transition as ϵ exhibits. In addition, the reason why the shift and critical angle of the shift change smoothly as λ increases, is that the ϵ changes smoothly.

4. Imbert-Fedorov shift on HMM

The geometry of the I-F shift is also shown in Fig. 2(a). Contrary to the longitudinal G-H shift, a circularly (or elliptically) polarized incident beam experiences a transverse shift perpendicular to the incident plane on reflection. This is known as the Imbert-Fedorov (I-F) shift. In our calculation, we consider the initial state to be right circularly polarized, |+=(H^+iV^)/2. The horizontally and vertically polarized incident beams Ĥ and V^ can be expressed in terms of unit polarization vectors ŝ and p^ [20],

H^=p^wkykxks^21+(wkykxk)2,
V^=s^+wkykxkp^21+(wkykxk)2.

The unit ŝ and p^ polarization vectors are defined by s^=z^×κ/κ and p^=s^×k/k. Similar to the G-H shift, the I-F shift can be expressed as [8, 40],

δIF=id2κER(kx,ky)kyER(kx,ky)d2κER(kx,ky)ER(kx,ky).

Assuming the slow variation of the reflection coefficients, we get the well-known expression for the I-F shift,

δIF=cotθk[1+2cos(φsφp)|rp||rs||rp|2+|rs|2],
where φs, φp are the phases of the reflection coefficients rs and rp, respectively.

Figures. 4(a)–(c) depict I-F shifts on the HMM, Ag and TiO2 slabs. The negative value indicates that the shift is along the −y direction at the interface. As λ increases, the displacement increases and reaches a peak, and then decreases to zero as the incident angle increases to 90 degrees. it can be seen that as λ increases from 350 nm to 600 nm, the I-F shifts on Ag and TiO2 slabs are all under 40 nm. For type I HMM (λ less than 395 nm) and type II HMM (λ larger than 551 nm), the I-F shifts are around 40 nm. The similar results of the shifts on metal, dielectric and HMM slabs implies that the HMM doesn’t affect the transverse shifts much. However, for λ = 500 nm which is in the phase of elliptical medium of the multilayer slab, the I-F shift has a large value of 135 nm. This is enhanced by around 3 times larger than on the metal or dielectric slabs at the same λ. Though this enhancement has no relation to the property of hyperbolic dispersion, this still offers us a way of enhancing the transverse I-F shift by using the multilayer dielectric/metal slab.

 figure: Fig. 4

Fig. 4 (a), (b) and (c) show the I-F shift δIF at the surface of a slab made of Ag/TiO2 multilayer HMM, Ag and TiO2, respectively. The slabs have the same thickness of 120 nm. The black, red, green and blue lines denote the wavelengths of 350 nm, 400 nm, 500 nm and 600 nm, respectively.

Download Full Size | PDF

5. Spin Hall effect of light on HMM

The geometry of the SHEL is shown in Fig. 5(a). As described in the introduction, when reflected at the interface, a linearly polarized incident beam splits into two spin components. The two components acquire opposite displacements perpendicular to the incident plane. In the spin basis, the horizontal and vertical states |H〉 and |V〉 can be expressed as |H=|++|2 and |V=|+|2i. Upon reflection, the two incident states |H〉 and |V〉 evolve as [20, 44],

|Hrp+iwkykxkrs1iwkykxk|+2+rpiwkykxkrs1+iwkykxk|2
|Virs+wkykxkrp1iwkykxk|+2+irswkykxkrp1+iwkykxk|2

 figure: Fig. 5

Fig. 5 (a) Scheme of the SHEL on a HMM slab. (b) and (c) show the δ|+H at the slab of HMM and Ag, respectively; (d), (e) and (f) show the δ|+V at the slab of HMM, Ag and TiO2, respectively. The black, red, green and blue curves denote the wavelengths of 350 nm, 400 nm, 500 nm and 600nm, respectively.

Download Full Size | PDF

When |rs||rp|wkykxk1, which is satisfied as long as the incident angle is not too close to the Brewster angle, Eqs. (12) and (13) have the approximate form [44, 45],

|Hrp2[exp(ikyδ|+H)|++exp(ikyδ|H)|],
|Virs2[exp(ikyδ|+V)|+exp(ikyδ|V)|].

The opposite displacements of the two spin components depend on the input polarization state, resulting from the polarization-dependent Fresnel reflections at the interface. For |H〉 and |V〉 input polarizations, the displacements for different spins δ|±H and δ|±V are given by the approximate expression [45],

δ|±H=λ2π[1+|rs||rp|cos(φsφp)]cotθ,
δ|±V=λ2π[1+|rp||rs|cos(φpφs)]cotθ.

Figures. 5(b) and 5(c) describe the displacements δ|+H for the case of horizontal polarization on the HMM and Ag slabs, respectively. Figure. 5(b) shows that the shift rises rapidly as the incident angle approaches to the Brewster angle. As the incident angle surpasses the Brewster angle, the shift becomes opposite and decreases rapidly. It should be noted that in the phase of type I HMM (λ less than 395 nm), the displacements is larger than 2 µm when the incident angle is close to the Brewster angle. Compared with the shifts on the metal, the shifts on the HMM slab are remarkably enhanced for λ less than 400 nm. This attributes to the change of the effective permittivities due to the combination of multilayer metal and dielectric slabs. In the case of TiO2 slab, Eq. (14) is not satisfied because of the small reflection coefficient for p polarization.

Finally, we calculated the displacements of the |+〉 spin component δ|+V in the case of vertical polarization. In Figs. 5(d)–5(f) we plot δ|+V on the HMM, Ag and TiO2 slabs, respectively. When the incident wavelengths change from 350 nm to 600 nm, the shifts on Ag and TiO2 slabs are smaller than 60 nm. For type I HMM and type II HMM, the SHEL shifts are around 40 nm. This is similar to the case of I-F effect, which also implies that the HMM also doesn’t have a big impact on the displacements for the initial vertical polarization. However, for a wavelength 500 nm which is in the elliptical phase of the multilayer slab, the displacement is about 120 nm, which is enhanced by almost 2 (4) times than the dielectric (metal) slab at the same wavelength. Similar to the case of I-F shift, this also offers us a method of enhancing the SHEL for the initial vertical polarization by using the multilayer dielectric/metal slab.

6. Conclusion

To summarize, we first studied the Goos-Hänchen shift of a reflected Gaussian beam at the surface of a thin slab made of Ag/TiO2 multilayer HMM. The effective medium theory was used to describe the anisotropic permittivities of HMM. We presented a rigorous calculation of the shifts without any approximations. The G-H shift on HMM slab can be enhanced to a value of about 40 µm, which is much larger than the G-H shifts at either metallic or dielectric surfaces. The enhanced shifts are explained in terms of the characteristics of the Brewster modes in HMM. We demonstrated how the calculation of G-H shifts using approximate formula can give quite erroneous result when the Brewster modes are excited. We also presented the first calculations of the Imbert-Fedorov (I-F) effect as well as the spin Hall effect of light (SHEL) at the surface of a thin HMM slab. We have found that for specific wavelength the I-F shift at HMM can be enhanced by almost 3 times than the result for metal or dielectric slabs. For the SHEL, compared with the shifts on the metal, the shifts on the HMM slab are found to be enhanced remarkably for wavelength less than 400 nm. Thus a thin HMM slab can be used as a candidate to enhance the lateral displacements.

Funding

Joint Fund of the National Natural Science Foundation of China (Grant No. U1330203); National Natural Science Foundation of China (Grants No. 11274242, No. 11474221, No. 11574229, No. 11104075 and No. 11504272); 973 program (Grant No. 2013CB632701); Shanghai Science and Technology Committee (Grant No. 15XD1503700, No. 15YF1412400); International Exchange Program for Graduate Students of Tongji University; Shanghai Education Commission Foundation.

References and links

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

2. F. I. Fedorov, “K teorii polnovo otrazenija,” Dokl. Akad. Nauk. SSR 105, 465 (1955).

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

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

5. P. T. Leung, C. W. Chen, and H. -P. Chiang, “Large negative Goos-Hänchen shift at metal surfaces,” Opt. Commun. 281(5), 1312–1313 (2008). [CrossRef]  

6. T. Paul, C. Rockstuhl, C. Menzel, and F. Lederer, “Resonant Goos-Hänchen and Imbert-Fedorov shifts at photonic crystal slabs,” Phys. Rev. A 77(5), 053802 (2008). [CrossRef]  

7. F. Liu, J. Xu, G. Song, and Y. Yang, “Goos-Hänchen and Imbert-Fedorov shifts at the interface of ordinary dielectric and topological insulator,” J. Opt. Soc. Am. B 30(5), 1167–1172 (2013). [CrossRef]  

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

9. M. Merano, A. Aiello, G.W. ’t Hooft, M. P. van Exter, E. R. Eliel, and J. P. Woerdman, “Observation of Goos-Hänchen shifts in metallic reflection,” Opt. Express 15(24), 15928–15934 (2007). [CrossRef]   [PubMed]  

10. V. J. Yallapragada, A. P. Ravishankar, G. L. Mulay, G. S. Agarwal, and V. G. Achanta, “Observation of giant Goos-Hänchen and angular shifts at designed metasurfaces,” Sci. Rep. 6, 19319 (2016). [CrossRef]  

11. C. Li and X. Yang, “Thin-film enhanced Goos-Hänchen shift in total internal reflection,” Chin. Phys. Lett. 21(3), 485–488 (2004). [CrossRef]  

12. I. V. Shadrivov, A. A. Zharov, and Y. S. Kivshar, “Giant Goos-Hänchen effect at the reflection from left-handed metamaterials,” Appl. Phys. Lett. 83(13), 2713–2715 (2003). [CrossRef]  

13. G. Jayaswal, G. Mistura, and M. Merano, “Weak measurement of the Goos-Hänchen shift,” Opt. Lett. 38(8), 1232–1234 (2013). [CrossRef]   [PubMed]  

14. S. Goswami, M. Pal, A. Nandi, P. K. Panigrahi, and N. Ghosh, “Simultaneous weak value amplification of angular Goos-Hänchen and Imbert-Fedorov shifts in partial reflection,” Opt. Lett. 39(21), 6229–6232 (2014). [CrossRef]   [PubMed]  

15. M. R. Dennis and Jörg B Götte, “The analogy between optical beam shifts and quantum weak measurements,” New J. Phys. 14(7), 073013 (2012). [CrossRef]  

16. X. Yin and L. Hesselink, “Goos-Hänchen shift surface plasmon resonance sensor,” Appl. Phys. Lett. 89, 261108 (2006). [CrossRef]  

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

18. 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]  

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

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

21. Y. Gorodetski, A. Niv, V. Kleiner, and E. Hasman, “Observation of the spin-based plasmonic effect in nanoscale structures,” Phys. Rev. Lett. 101(4), 043903 (2008). [CrossRef]   [PubMed]  

22. N. Hermosa, A. M. Nugrowati, A. Aiello, and J. P. Woerdman, “Spin Hall effect of light in metallic reflection,” Opt. Lett. 36(16), 3200–3202 (2011). [CrossRef]   [PubMed]  

23. L. Ferrari, C. Wu, D. Lepage, X. Zhang, and Z. Liu, “Hyperbolic metamaterials and their applications,” Prog. Quantum Electron. 40, 1–40 (2015). [CrossRef]  

24. A. J. Hoffman, L. Alekseyev, S. S. Howard, K. J. Franz, D. Wasserman, V. A. Podolskiy, E. E. Narimanov, D. L. Sivco, and C. Gmachl, “Negative refraction in semiconductor metamaterials,” Nat. Mater. 6(12), 946–950 (2007). [CrossRef]   [PubMed]  

25. Z. Jacob, L. V. Alekseyev, and E. Narimanov, “Optical hyperlens: Far-field imaging beyond the diffraction limit,” Opt. Express 14(18), 8247–8256 (2006). [CrossRef]   [PubMed]  

26. Z. Liu, H. Lee, Y. Xiong, C. Sun, and X. Zhang, “Far-field optical hyperlens magnifying sub-diffraction-limited objects,” Science 315(5819), 1686 (2007). [CrossRef]   [PubMed]  

27. A. A. Govyadinov and V. A. Podolskiy, “Metamaterial photonic funnels for subdiffraction light compression and propagation,” Phys. Rev. B 73(15), 155108 (2006). [CrossRef]  

28. J. Yao, X. Yang, X. Yin, G. Bartal, and X. Zhang, “Three-dimensional nanometer-scale optical cavities of indefinite medium,” Proc. Natl. Acad. Sci. 108(28), 11327 (2011). [CrossRef]   [PubMed]  

29. Y. Guo, C. L. Cortes, S. Molesky, and Z. Jacob, “Broadband super-Planckian thermal emission from hyperbolic metamaterials,” Appl. Phys. Lett. 101(13), 131106 (2012). [CrossRef]  

30. S-A. Biehs, M. Tschikin, and P. Ben-Abdallah, “Hyperbolic metamaterials as an analog of a blackbody in the near field,” Phys. Rev. Lett. 109(10), 104301 (2012). [CrossRef]   [PubMed]  

31. I. S. Nefedov and C. R. Simovski, “Giant radiation heat transfer through micron gaps,” Phys. Rev. B 84(19), 195459 (2011). [CrossRef]  

32. H. N. S. Krishnamoorthy, Z. Jacob, E. Narimanov, I. Kretzschmar, and V. M. Menon, “Topological transitions in metamaterials,” Science 336, 205–209 (2012). [CrossRef]   [PubMed]  

33. C. R. Simovski, P. A. Belov, A. V. Atrashchenko, and Y. S. Kivshar, “Wire metamaterials: Physics and applications,” Adv. Mater. 24(31), 4229–4248 (2012). [CrossRef]   [PubMed]  

34. J. Zhao, H. Zhang, X. Zhang, D. Li, H. Lu, and M. Xu, “Abnormal behaviors of Goos-Hänchen shift in hyperbolic metamaterials made of aluminum zinc oxide materials,” Photon. Res. 1(4), 160–163 (2013). [CrossRef]  

35. C. Chen, T. Bian, H. Chiang, and P. T. Leung, “Nonlocal optical effects on the Goos-Hänchen shifts at multilayered hyperbolic metamaterials,” J. Opt. 18(2), 025104 (2016). [CrossRef]  

36. S. Rytov, “Electromagnetic properties of a finely stratified medium,” Sov. Phys. JETP 2, 466–475 (1956).

37. B. Wood, J. B. Pendry, and D. P. Tsai, “Directed subwavelength imaging using a layered metal-dielectric system,” Phys. Rev. B 74(11), 115116 (2006). [CrossRef]  

38. S.-A. Biehs, V. M. Menon, and G. S. Agarwal, “Long-range dipole-dipole interaction and anomalous Förster energy transfer across a hyperbolic metamaterial,” Phys. Rev. B 93(24), 245439 (2016). [CrossRef]  

39. J. R. Devore, “Refractive indices of rutile and sphalerite,” J. Opt. Soc. Am. 41(6), 416 (1951). [CrossRef]  

40. C. Li, “Unified theory for Goos-Hänchen and Imbert-Fedorov effects,” Phys. Rev. A 76(1), 013811 (2007). [CrossRef]  

41. M. Merano, N. Hermosa, A. Aiello, and J. P. Woerdman, “Demonstration of a quasi-scalar angular Goos-Hänchen effect,” Opt. Lett. 35(21), 3562–3564 (2010). [CrossRef]   [PubMed]  

42. G. S. Agarwal, “New method in the theory of surface polaritons,” Phys. Rev. B 8(10), 4768 (1973). [CrossRef]  

43. Z. Liu, L. Hu, and Z. Lin, “Enhancing photon tunnelling by a slab of uniaxially anisotropic left-handed material,” Phys. Lett. A. 308(4), 294–301 (2003). [CrossRef]  

44. Y. Qin, Y. Li, X. Feng, Z. Liu, H. He, Y. Xiao, and Q. Gong, “Spin Hall effect of reflected light at the air-uniaxial crystal interface,” Opt. Express 18(16), 16832–16839 (2010). [CrossRef]   [PubMed]  

45. X. Zhou, Z. Xiao, H. Luo, and S. 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]  

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 (5)

Fig. 1
Fig. 1 (a) Illustration of a multilayer HMM; (b)-(d) describe the effective permittivities ϵ and ϵ for a Ag/TiO2 multilayer HMM with a silver filling fraction of f = 0.35. (b) plots the real part of ϵ and ϵ; (c) and (d) plot the imaginary part of ϵ and ϵ, respectively. The red and blue lines correspond to ϵ and ϵ, respectively.
Fig. 2
Fig. 2 (a) Geometry of the beam reflection at the surface of a slab made of HMM. The reflected beam will undergo a longitudinal shift δGH and transverse shift δIF. (b), (c) and (d) show the G-H shift δGH at the surface of a slab made of Ag/TiO2 multilayer HMM, Ag and TiO2, respectively. The slabs have the same thickness of 120 nm.
Fig. 3
Fig. 3 (a) plots the G-H shifts calculated using the exact (solid) and approximate (dashed) results at the surface of HMM near the Brewster angle for 345 nm. (b) plots the values of |rp|, and (c) plots the relative phases φ.
Fig. 4
Fig. 4 (a), (b) and (c) show the I-F shift δIF at the surface of a slab made of Ag/TiO2 multilayer HMM, Ag and TiO2, respectively. The slabs have the same thickness of 120 nm. The black, red, green and blue lines denote the wavelengths of 350 nm, 400 nm, 500 nm and 600 nm, respectively.
Fig. 5
Fig. 5 (a) Scheme of the SHEL on a HMM slab. (b) and (c) show the δ | + H at the slab of HMM and Ag, respectively; (d), (e) and (f) show the δ | + V at the slab of HMM, Ag and TiO2, respectively. The black, red, green and blue curves denote the wavelengths of 350 nm, 400 nm, 500 nm and 600nm, respectively.

Equations (17)

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

ϵ = f ϵ Ag + ( 1 f ) ϵ TiO 2 ,
ϵ = ϵ Ag ϵ TiO 2 f ϵ TiO 2 + ( 1 f ) ϵ Ag ,
E I ( r , ω ) = d 2 κ e i ( k x x + k y y ) + i w ( k x , k y ) z E I ( κ , ω ) .
E I ( k x , k y ) = E 0 2 σ x σ y π exp [ σ x 2 ( k x k x 0 ) 2 4 ] exp [ σ y 2 k y 2 4 ] .
E R ( r , ω ) = d 2 κ e i ( k x x + k y y ) i w ( k x , k y ) z E R ( κ , ω ) .
δ GH = i d 2 κ E R ( k x , k y ) k x E R ( k x , k y ) d 2 κ E R ( k x , k y ) E R ( k x , k y ) λ 2 π φ θ .
r p = ( k e z ϵ k 0 z ) ( k e z + ϵ k 0 z ) ( k e z ϵ k 0 z ) ( k e z + ϵ k 0 z ) e 2 i k e z d ( k e z + ϵ k 0 z ) ( k e z + ϵ k 0 z ) ( k e z ϵ k 0 z ) ( k e z ϵ k 0 z ) e 2 i k e z d ,
H ^ = p ^ w k y k x k s ^ 2 1 + ( w k y k x k ) 2 ,
V ^ = s ^ + w k y k x k p ^ 2 1 + ( w k y k x k ) 2 .
δ IF = i d 2 κ E R ( k x , k y ) k y E R ( k x , k y ) d 2 κ E R ( k x , k y ) E R ( k x , k y ) .
δ IF = cot θ k [ 1 + 2 cos ( φ s φ p ) | r p | | r s | | r p | 2 + | r s | 2 ] ,
| H r p + i w k y k x k r s 1 i w k y k x k | + 2 + r p i w k y k x k r s 1 + i w k y k x k | 2
| V i r s + w k y k x k r p 1 i w k y k x k | + 2 + i r s w k y k x k r p 1 + i w k y k x k | 2
| H r p 2 [ exp ( i k y δ | + H ) | + + exp ( i k y δ | H ) | ] ,
| V i r s 2 [ exp ( i k y δ | + V ) | + exp ( i k y δ | V ) | ] .
δ | ± H = λ 2 π [ 1 + | r s | | r p | cos ( φ s φ p ) ] cot θ ,
δ | ± V = λ 2 π [ 1 + | r p | | r s | cos ( φ p φ s ) ] cot θ .
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.