Abstract
The effective medium theory (EMT) provides a simplified framework to calculate the electromagnetic responses and is generally considered exact in the all-dielectric system with deep-subwavelength constituents. In this work, we perform the Goos–Hänchen (GH) shift that invalidates the EMT on the multilayered dielectric structures under the common conditions. This breakdown of the EMT arises from the high sensitivity of the GH shift on the phase and magnitude of Fresnel reflection coefficient. The degree of such breakdown shows strong dependence on the polarization angle of incidence and the layer and filling fraction of the structures. Notably, we find that the GH shift is potentially applicable to nano-meter scale thickness sensing, which cannot be displayed based on EMT in some cases. Our findings will provide useful guidance to reduce the calculation errors of the electromagnetic responses and promote the design of precise metrology devices.
1 Introduction
The calculation of the electromagnetic responses in the subwavelength-sized inhomogeneous structures is complex and difficult. To solve this problem, the effective medium theory (EMT) is proposed [1], [2]. Its conceptual process involves replacing the inhomogeneous structures with a homogeneous “effective medium” with uniform properties. This provides the fundamental approximation method to obtain the same response as the complex inhomogeneous structures to the identical excitation. In practice, its two predominant frameworks are the Maxwell–Garnett EMT [3] and Bruggeman EMT [4]. The former is applied in reliable predictions for electro-magnetic properties in a continuous host medium containing sparse inclusions at low volume fractions. In contrast, the latter enables precise modeling of the dielectric permittivity and conductivity of the high-concentration systems near the percolation threshold. To date, the EMT has been widely used in the description of the index of permittivity [5], [6], [7], permeability [8], [9], and surface conductivity [10], [11], [12] of complex materials, to name a few. Importantly, the effective medium can exhibit the extraordinary characteristics such as birefringence [13], Anderson localization [14], negative refractive index [15], [16], [17], and near-zero permittivity [18], [19], and leads to the significant findings like the metasurface [20], [21], [22], cloak [23], [24], and super-lens [25], [26], [27].
Despite its extensive applications, the EMT will be invalid in some cases. Typically, the EMT fails to describe the homogenization of the periodic metal-dielectric structures involving the large wave vectors, especially when the metal plasmons and hyperbolic high-k modes exist [28], [29], [30]. The underlying mechanism behind this phenomenon is ascribed to the inappropriate homogenization of the local structure properties that can strongly influence the overall structure properties. Therefore, the all-dielectric systems that do not support the large wave vectors are believed to be adaptable to the EMT for a long time. In recent years, it has been found that the EMT will also be invalid in the all-dielectric systems [31], [32], [33], [34], [35]. Sheinfux et al. first theoretically proposed the breakdown of EMT in a purely dielectric structure near the critical angle [31], which was later experimentally verified by Zhukovsky et al. [32]. Furthermore, the similar phenomena for the evanescent waves are successively predicted and observed in the one-dimensional disordered deep-subwavelength optical system due to the Anderson localization [33], [34]. Lately, a fundamental breakdown of the EMT by the photonic spin Hall effect in all-dielectric systems has been reported [35]. It stems from the spin–orbit interaction of light that is sensitive to the slight phase accumulations during the wave propagation. These works reveal that, not only the magnitude, but also the phase deviation of Fresnel reflection coefficients calculated by EMT may have great impacts on the final electromagnetic response. The Goos–Hänchen (GH) shift is the longitudinal beam shift in the incident plane [36], and attracted intensive attention due to its potential applications in precise sensing [37], [38], [39]. Besides, it is usually studied in the framework of EMT [40], [41], [42]. Now, an interesting question arises: as the typical electromagnetic response associated with the Fresnel reflection coefficient phase and magnitude, will the GH shift invalidate the EMT in the general all-dielectric system?
To address this question, we calculate the spatial and angular GH shifts on the all-dielectric periodic multilayered structures via the EMT and transfer matrix method (TMM), respectively, in this work. Numerical results show that the breakdown of EMT will emerge under the general conditions even without the critical angle. The physical mechanism is that the GH shift is very sensitive to the phase and magnitude of the Fresnel reflection coefficient. Besides, we examine the effects of the polarization angle of the incidence and the layer and filling fraction of the structures on the degree of EMT breakdown. Finally, we reveal the potential application of the GH shift on the detection of structure defects at the extreme nanoscale, which cannot be demonstrated in the EMT framework in some cases. These results can deepen the understanding of the EMT in various application realms and further the design of the relevant devices.
2 Theoretical model and numerical method
We will illustrate our idea by considering an all-dielectric periodic multilayered structure composed of N pairs of TiO2 and Al2O3 with permittivities ɛ1 and ɛ2, and thicknesses d1 and d2, respectively, and the superstrate ZnSe and substrate Si3N4 with permittivities ɛinc and ɛout, as shown in Figure 1(a). The Fresnel reflection and refraction coefficients of such a structure can be rigorously calculated by the TMM as [43]
with the converting M and propagating P matrix expressed as
where
![Figure 1:
Illustrations of the GH shifts and the EMT. (a) Schematic of an all-dielectric periodic multilayered structure containing low-index (L) layers made of alumina (Al2O3), high-index (H) layers made of titania (TiO2), a superstrate made of zinc selenide (ZnSe), and a substrate made of silicon nitride (Si3N4), and the spatial Δ and angular Θ GH shifts of the Gaussian beam with the incident angle θ
i
. (b) Schematic of the rigorous transfer matrix method (TMM) with the different stacking orderings [LH] and [HL], and effective medium theory (EMT).](/document/doi/10.1515/nanoph-2025-0314/asset/graphic/j_nanoph-2025-0314_fig_001.jpg)
Illustrations of the GH shifts and the EMT. (a) Schematic of an all-dielectric periodic multilayered structure containing low-index (L) layers made of alumina (Al2O3), high-index (H) layers made of titania (TiO2), a superstrate made of zinc selenide (ZnSe), and a substrate made of silicon nitride (Si3N4), and the spatial Δ and angular Θ GH shifts of the Gaussian beam with the incident angle θ i . (b) Schematic of the rigorous transfer matrix method (TMM) with the different stacking orderings [LH] and [HL], and effective medium theory (EMT).
For the EMT illustrated in Figure 1(b), such structure can be described as an effective uniaxial slab with relative permittivity
The Fresnel reflection and refraction coefficients between the superstrate and slab are
where the matrices M and P have the similar forms as that in Eq. (2). Δout,eff = keffz/koutz and
Next, we proceed to derive the GH shift for reflection, i.e., the longitudinal centroid shift of the reflected beam. The vector angular spectrum of the incident Gaussian beam is [44]
with
where β is the polarization angle, w0 is the beam waist, and k
ix
and k
iy
are wave vector components along the x
i
and y
i
axes, respectively, in local coordinates
in which k
rx
and k
ry
are wave vector components in local coordinates
The GH shift at any given plane z = const can be given by
with
where β is the polarization angle (β = 0 and β = 90° denote the p- and s-polarization, respectively), and ϕp,s is the phase of rp,s. Notably, the expression of the spatial (Δ) GH shift derived by the stationary-phase analysis applies only when the reflection coefficient phase varies slowly.
3 Results and discussion
In the following numerical simulations, if no otherwise stated, we assume that the permittivities ɛ1 and ɛ2 of TiO2 and Al2O3 are 6.67 [45] and 3.12 [46], respectively, the thicknesses d1 and d2 of TiO2 and Al2O3 are 10 nm, permittivities ɛinc and ɛout of superstrate ZnSe and substrate Si3N4 are 6.65 [47] and 4.04 [48], respectively, the periodicity number of the structure Npair = 10, the stacking orderings of the multilayers is [HL], the polarization angle β = 0, the wavelength of incidence in vacuum is 0.6328 μm, and the beam waist of incidence is 27 μm. Besides, we emphasize that the results calculated by TMM are exact.
First, we check the invalidity of the EMT in the structure shown in Figure 1(a), via the performance of spatial and angular GH shifts. Under this scenario, the critical angle 51.23° exists at the interface between the superstrate and effective medium. Figure 2 illustrates breakdown of the EMT by the spatial and angular GH shifts on the all-dielectric periodic multilayered structures with different stacking orderings [LH] and [HL], in case of ɛinc = 6.65. According to the EMT, such a multilayered structure of both [LH] and [HL] will be homogenized to the same effective medium, which means the
![Figure 2:
Breakdown of the EMT by the spatial (Δ) and angular (Θ) GH shifts on the all-dielectric periodic multilayered structures with different stacking orderings [LH] and [HL], in the case of ɛinc = 6.65. (a) Reflectivity |r
p
|2, (b) Fresnel reflection coefficient phase ϕ
p
, (c) spatial GH shifts (Δ), and (d) angular GH shifts (Θ).](/document/doi/10.1515/nanoph-2025-0314/asset/graphic/j_nanoph-2025-0314_fig_002.jpg)
Breakdown of the EMT by the spatial (Δ) and angular (Θ) GH shifts on the all-dielectric periodic multilayered structures with different stacking orderings [LH] and [HL], in the case of ɛinc = 6.65. (a) Reflectivity |r p |2, (b) Fresnel reflection coefficient phase ϕ p , (c) spatial GH shifts (Δ), and (d) angular GH shifts (Θ).
In terms of the previous works [31], [32], [33], [34], the existence of a critical angle is a mandatory condition for the EMT breakdown in an all-dielectric system. Does it mean the GH shift calculated by the EMT is always exact in the all-dielectric structures without the critical angle? To tackle this issue, we replace the superstrate ZnSe prime in the setup shown in Figure 1(a) by the free space ɛinc = 1, and review the performances of reflectivity |r
p
|2, the reflection coefficient phase ϕ
p
, and GH shifts Δ and Θ in Figure 3. Interestingly, the discrepancies of the spatial and angular GH shifts persist as well in the angle range
![Figure 3:
Breakdown of the EMT by the spatial (Δ) and angular (Θ) GH shifts on the all-dielectric periodic multilayered structures with different stacking orderings [LH] and [HL], in the case of ɛinc = 1. (a) Reflectivity |r
p
|2, (b) Fresnel reflection coefficient phase ϕ
p
, (c) spatial GH shifts (Δ), and (d) angular GH shifts (Θ).](/document/doi/10.1515/nanoph-2025-0314/asset/graphic/j_nanoph-2025-0314_fig_003.jpg)
Breakdown of the EMT by the spatial (Δ) and angular (Θ) GH shifts on the all-dielectric periodic multilayered structures with different stacking orderings [LH] and [HL], in the case of ɛinc = 1. (a) Reflectivity |r p |2, (b) Fresnel reflection coefficient phase ϕ p , (c) spatial GH shifts (Δ), and (d) angular GH shifts (Θ).
Now, we proceed to examine the breakdown of the EMT by the spatial and angular GH shifts with different polarization angle β of incidence from Figure 4. As we can see, for the case of ɛinc = 6.65, the discrepancies in both spatial and angular GH shifts decrease with the growing polarization angle in regions far from the critical angle, whereas these discrepancies are robust against the variations of polarization angle near the critical angle. If the superstrate is free space, the errors of spatial and angular GH shifts under p-polarization have a pronounced peak near the incident angle θ i ≈ 64°. But these errors rapidly diminish to near-zero values at larger polarization angles. These findings imply that the EMT achieves higher accuracy for larger polarization angles in the all-dielectric systems.

Influence of the polarization angle β on the performance of the EMT breakdown by the spatial (Δ) and angular (Θ) GH shifts, in the cases of (a), (b) ɛinc = 6.65, and (c), (d) ɛinc = 1.
Figure 5 depicts the breakdown of the EMT by the spatial and angular GH shifts with different filling fraction g = [0.2, 0.4, 0.6, 0.8] of the structure. The g is defined as
![Figure 5:
Influence of the filling fractions g = [0.2, 0.4, 0.6, 0.8] of structure on the performance of the EMT breakdown by the spatial (Δ) and angular (Θ) GH shifts, in the cases of (a), (b) ɛinc = 6.65, and (c), (d) ɛinc = 1. The filling fraction
g
=
d
1
/
d
1
+
d
2
$g={d}_{1}/\left({d}_{1}+{d}_{2}\right)$
, and d1 + d2 = 20 nm is always satisfied.](/document/doi/10.1515/nanoph-2025-0314/asset/graphic/j_nanoph-2025-0314_fig_005.jpg)
Influence of the filling fractions g = [0.2, 0.4, 0.6, 0.8] of structure on the performance of the EMT breakdown by the spatial (Δ) and angular (Θ) GH shifts, in the cases of (a), (b) ɛinc = 6.65, and (c), (d) ɛinc = 1. The filling fraction
Furthermore, we investigate the dependence of EMT breakdown induced by GH shifts on the unit-cell thickness, as shown in Figure 6. The results demonstrate that the EMT breakdown persists across a broad range of unit-cell thickness in the deep-subwavelength scale. For the case of ɛinc = 6.65, the discrepancies in both spatial and angular GH shifts become larger with the growing unit-cell thickness. In contrast, when ɛinc = 1, such deviations display a decreasing trend as the unit-cell thickness increases.

Influence of the unit-cell thickness on the performance of the EMT breakdown by the spatial (Δ) and angular (Θ) GH shifts, in the cases of (a) ɛinc = 6.65 and (b) ɛinc = 1.
In the preparation of multilayered structures, the losses of materials are inevitable. That is to say, the relative permittivities of dielectric components are generally the complex numbers with tiny imaginary parts in the experiment. In Figure 7, we examine the effects of the losses of multilayered components on the EMT. It can be seen that, the spatial and angular GH shifts still invalidate the EMT for the loss cases, and the slight losses

Influence of the multilayered components loss on the performance of the EMT breakdown by the spatial (Δ) and angular (Θ) GH shifts, in the cases of (a) ɛinc = 6.65 and (b) ɛinc = 1.
According to Ref. [31], the performance of EMT breakdown rely on the periodicity number Npair, and the severe breakdown phenomena happen only when Npair ≫ 1. Here, we perform the breakdown of the EMT for case of Npair = 1 by the spatial and angular Θ GH shifts with the stacking ordering [HL] and ɛinc = 1 in Figure 8. Although on the ultrathin slab with only 30∼33 nm thickness, the spatial and angular GH shifts still contradict to the conventional EMT. Importantly, only 1 nm variation in layer thickness can make very large distinctions in GH shifts. Whereas, such ultra-sensitivity on the thickness does not exhibit via the EMT. These differences can be explained by the reflectivity |r p |2 and reflection coefficient phase ϕ p . The 1 nm thickness change can hardly affect the |r p |2 and ϕ p in the framework of EMT, but can be demonstrated via the TMM. It indicates that the homogenization in EMT may overlook the extreme small property variations of mediums. These phenomena indicate the GH shift is a powerful tool in the detection of dielectric structure defects at the extreme nanoscale, which has been overlooked before in the framework of EMT.
![Figure 8:
Breakdown of the EMT by the GH shifts with the stacking ordering [HL] in case of ɛinc = 1 and Npair = 1, and the detection of dielectric structure defects at the extreme nanoscale. (a) spatial (Δ) GH shifts, (b) angular (Θ) GH shifts, (c) reflectivity |r
p
|2, and (d) Fresnel reflection coefficient phase ϕ
p
.](/document/doi/10.1515/nanoph-2025-0314/asset/graphic/j_nanoph-2025-0314_fig_008.jpg)
Breakdown of the EMT by the GH shifts with the stacking ordering [HL] in case of ɛinc = 1 and Npair = 1, and the detection of dielectric structure defects at the extreme nanoscale. (a) spatial (Δ) GH shifts, (b) angular (Θ) GH shifts, (c) reflectivity |r p |2, and (d) Fresnel reflection coefficient phase ϕ p .
To experimentally validate the proposed concept, we employ the weak measurement [49] techniques in our setup. The experimental configuration begins with a Gaussian beam generated by the He–Ne laser passing through a Glan laser polarizer (GLP1) and a focusing lens (L1) before incident on the all-dielectric multilayered structures. The reflected beam then propagates through a quarter-wave plate (QWP), followed by a half-wave plate (HWP) and another Glan laser polarizer (GLP2), before being focused by a second lens (L2) onto a CCD detector for measurement. The breakdown of EMT can be experimentally demonstrated by comparing the measured GH shifts from all-dielectric multilayered structures with different stacking sequences. The weak measurement protocol enables significant amplification of the GH shift through careful selection of nearly orthogonal post-selection and preselection states. The multilayered structures, composed of alternating alumina (Al2O3) and titania (TiO2) layers with designed stacking orders, are fabricated using atomic layer deposition (ALD) to ensure precise thickness control and interfacial quality.
Finally, it is worthy to note that the EMT fails to predict the reflection phase in many cases, making it ineffective for GH shifts and other phase-sensitive physical phenomena. Therefore, the careful consideration of phase-related effects is essential when evaluating the applicability of EMT. It holds significant implications for the design of novel micro-nano optoelectronic devices.
4 Conclusions
In conclusion, we have reported the breakdown of EMT by the spatial and angular GH shifts on the periodic deep-subwavelength multilayered dielectric structures under the general conditions. Through comparing the GH shifts calculated by the EMT and TMM, it is found that the differences between them are very large in some cases commonly discussed. The underlying mechanism behind this phenomenon is ascribed to the high sensitivity of the GH shift on the Fresnel reflection coefficient phase and magnitude. Meanwhile, the numerical results have shown that the degree of the EMT breakdown can be controlled via adjusting the polarization angle of incidence and the layer and filling fraction of the structures. Finally, we have revealed the potential application of the GH shift on nano-meter scale thickness sensing, which may not be demonstrated within the EMT framework. This work will be helpful for improving the calculating accuracy of the electromagnetic responses and promoting the development of precise metrology devices.
Funding source: Innovation Fund of Xidian University
Funding source: Hunan Provincial Major Sci-Tech Program
Award Identifier / Grant number: 2023ZJ1010
Funding source: Fundamental Research Funds for the Central Universities
Award Identifier / Grant number: YJSJ25020
Funding source: National Natural Science Foundation of China
Award Identifier / Grant number: 12374273
Award Identifier / Grant number: 12421005
Award Identifier / Grant number: 92464205
Funding source: Basic and Applied Basic Research Foundation of Guangdong Province
Award Identifier / Grant number: 2023A1515030152
-
Research funding: This work was supported by the National Natural Science Foundation of China (12374273, 12421005, 92464205), the Basic and Applied Basic Research Foundation of Guangdong Province (2023A1515030152). Hunan Provincial Major Sci-Tech Program (2023ZJ1010), Fundamental Research Funds for the Central Universities (YJSJ25020), and Innovation Fund of Xidian University.
-
Author contributions: WG: writing-original draft, investigation, data curation. YS: writing-original draft, writing-review & editing data, investigation. ZL: software. CZ: investigation. ZC: supervision, validation. YC: supervision, formal analysis. XZ: supervision, project administration, conceptualization. All authors have accepted responsibility for the entire content of this manuscript and consented to its submission to the journal, reviewed all the results and approved the final version of the manuscript.
-
Conflict of interest: Authors state no conflicts of interest.
-
Data availability: Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.
References
[1] T. C. Choy, Effective Medium Theory: Principles and Applications, 2nd ed. Oxford, Oxford University Press, 2015.10.1093/acprof:oso/9780198705093.001.0001Suche in Google Scholar
[2] J. A. Kong, Electromagnetic Wave Theory, Cambridge, EMW Publishing, 2008.Suche in Google Scholar
[3] J. C. M. Garnett, “Colours in metal glasses and in metallic films,” Philos. Trans. R. Soc. Lond. A, vol. 203, pp. 385–420, 1904.10.1098/rsta.1904.0024Suche in Google Scholar
[4] D. A. G. Bruggeman, “Calculation of various physics constants in heterogenous substances I dielectricity constants and conductivity of mixed bodies from isotropic substances,” Ann. Phys., vol. 416, no. 7, pp. 636–664, 1935.Suche in Google Scholar
[5] J. Huang, S. J. Chen, Z. Xue, W. Withayachumnankul, and C. Fumeaux, “Wideband endfire 3-D-printed dielectric antenna with designable permittivity,” IEEE Antennas Wirel. Propag. Lett., vol. 17, no. 11, pp. 2085–2089, 2018. https://doi.org/10.1109/lawp.2018.2857497.Suche in Google Scholar
[6] J. Xin, J. Zong, J. Gao, Y. Wang, Y. Song, and X. Zhang, “Extraction and control of permittivity of hyperbolic metamaterials with optical nonlocality,” Opt. Express, vol. 29, no. 12, pp. 18572–18586, 2021. https://doi.org/10.1364/oe.426746.Suche in Google Scholar PubMed
[7] T. Hou and H. Chen, “Criterion for photonic topological transition in two-dimensional heterostructures,” Opt. Lett., vol. 47, no. 20, pp. 5433–5436, 2022. https://doi.org/10.1364/ol.474505.Suche in Google Scholar
[8] N. Wang, R.-Y. Zhang, C. T. Chen, and G. P. Wang, “Effective medium theory for a photonic pseudospin -1/2 system,” Phys. Rev. B, vol. 102, no. 9, p. 094312, 2020. https://doi.org/10.1103/physrevb.102.094312.Suche in Google Scholar
[9] F. Liu, P. Peng, Q. Du, and M. Ke, “Effective medium theory for photonic crystals with quadrupole resonances,” Opt. Lett., vol. 46, no. 18, pp. 4597–4600, 2021. https://doi.org/10.1364/ol.438083.Suche in Google Scholar
[10] A. Vakil and N. Engheta, “Transformation optics using graphene,” Science, vol. 332, no. 6035, pp. 1291–1294, 2011. https://doi.org/10.1126/science.1202691.Suche in Google Scholar PubMed
[11] B. Rahmani, A. Bagheri, A. Khavasi, and K. Mehrany, “Effective medium theory for graphene-covered metallic gratings,” J. Opt., vol. 18, no. 10, p. 105005, 2016. https://doi.org/10.1088/2040-8978/18/10/105005.Suche in Google Scholar
[12] I. O. Thomas and G. P. Sriastava, “Extension of the modified effective medium approach to nanocomposites with anisotropic thermal conductivities,” Phys. Rev. B, vol. 98, no. 9, p. 094201, 2018. https://doi.org/10.1103/physrevb.98.094201.Suche in Google Scholar
[13] F. Xu, R.-C. Tyan, P.-C. Sun, Y. Fainman, C.-C. Cheng, and A. Scherer, “Fabrication, modeling, and character- ization of form-birefringent nanostructures,” Opt. Lett., vol. 20, no. 24, pp. 2457–2459, 1995. https://doi.org/10.1364/ol.20.002457.Suche in Google Scholar PubMed
[14] M. Segev, Y. Silberberg, and D. N. Christodoulides, “Anderson localization of light,” Nat. Photonics, vol. 7, no. 3, pp. 197–204, 2013. https://doi.org/10.1038/nphoton.2013.30.Suche in Google Scholar
[15] V. M. Shalaev, et al.., “Negative index of refraction in optical metamaterials,” Opt. Lett., vol. 30, no. 24, pp. 3356–3358, 2005. https://doi.org/10.1364/ol.30.003356.Suche in Google Scholar PubMed
[16] X. Zhang, M. Davanço, Y. Urzhumov, G. Shvets, and S. R. Forrest, “From scattering parameters to Snell’s Law: A subwavelength near-infrared negative-index metamaterial,” Phys. Rev. Lett., vol. 101, no. 26, p. 267401, 2008. https://doi.org/10.1103/physrevlett.101.267401.Suche in Google Scholar PubMed
[17] V. M. Shalaev, “Optical negative-index metamaterials,” Nat. Photonics, vol. 1, no. 1, pp. 41–48, 2007. https://doi.org/10.1038/nphoton.2006.49.Suche in Google Scholar
[18] R. J. Pollard, et al.., “Optical nonlocalities and additional waves in epsilon-near-zero metamaterials,” Phys. Rev. Lett., vol. 102, no. 12, p. 127405, 2009. https://doi.org/10.1103/physrevlett.102.127405.Suche in Google Scholar
[19] R. Maas, J. Parsons, N. Engheta, and A. Polman, “Experimental realization of an epsilon-near-zero metamaterial at visible wavelengths,” Nat. Photonics, vol. 7, no. 11, pp. 907–912, 2013. https://doi.org/10.1038/nphoton.2013.256.Suche in Google Scholar
[20] S. Sun, Q. He, S. Xiao, Q. Xu, X. Li, and L. Zhou, “Gradient-index meta-surfaces as a bridge linking propagating waves and surface waves,” Nat. Mater., vol. 11, no. 5, pp. 426–431, 2012. https://doi.org/10.1038/nmat3292.Suche in Google Scholar PubMed
[21] A. V. Kildishev, A. Boltasseva, and V. M. Shalaev, “Planar photonics with metasurfaces,” Science, vol. 339, no. 6125, p. 1232009, 2013. https://doi.org/10.1126/science.1232009.Suche in Google Scholar PubMed
[22] X. Yin, Z. Ye, J. Rho, Y. Wang, and X. Zhang, “Photonic spin Hall effect at metasurfaces,” Science, vol. 339, no. 6126, pp. 1405–1407, 2013. https://doi.org/10.1126/science.1231758.Suche in Google Scholar PubMed
[23] J. Valentine, J. Li, T. Zentgraf, G. Bartal, and X. Zhang, “An optical cloak made of dielectrics,” Nat. Mater., vol. 8, no. 7, pp. 568–571, 2009. https://doi.org/10.1038/nmat2461.Suche in Google Scholar PubMed
[24] T. Cai, et al.., “Experimental realization of a superdispersion-enabled ultrabroadband terahertz cloak,” Adv. Mater., vol. 34, no. 38, p. 2205053, 2022. https://doi.org/10.1002/adma.202209380.Suche in Google Scholar PubMed
[25] J. B. Pendry, “Negative refraction makes a perfect lens,” Phys. Rev. Lett., vol. 85, no. 18, p. 3966, 2000. https://doi.org/10.1103/physrevlett.85.3966.Suche in Google Scholar PubMed
[26] X. Zhang and Z. Liu, “Superlenses to overcome the diffraction limit,” Nat. Mater., vol. 7, no. 6, pp. 435–441, 2008. https://doi.org/10.1038/nmat2141.Suche in Google Scholar PubMed
[27] N. Fang, H. Lee, C. Sun, and X. Zhang, “Sub-diffraction-limited optical imaging with a silver superlens,” Science, vol. 308, no. 5721, pp. 534–537, 2005. https://doi.org/10.1126/science.1108759.Suche in Google Scholar PubMed
[28] F. Wang and Y. R. Shen, “General properties of local plasmons in metal nanostructures,” Phys. Rev. Lett., vol. 97, no. 20, p. 206806, 2006. https://doi.org/10.1103/physrevlett.97.206806.Suche in Google Scholar
[29] O. Kidwai, S. V. Zhukovsky, and J. E. Sipe, “Effective-medium approach to planar multilayer hyperbolic metamaterials: Strengths and limitations,” Phys. Rev. A, vol. 85, no. 5, p. 053842, 2012. https://doi.org/10.1103/physreva.85.053842.Suche in Google Scholar
[30] M. Mahmoodi, S. H. Tavassoli, O. Takayama, J. Sukham, R. Malureanu, and A. V. Lavrinenko, “Existence conditions of high-k modes in finite hyperbolic metamaterials,” Laser Photonics Rev., vol. 13, no. 3, p. 1800253, 2019. https://doi.org/10.1002/lpor.201800253.Suche in Google Scholar
[31] H. H. Sheinfux, I. Kaminer, Y. Plotnik, G. Bartal, and M. Segev, “Subwavelength multilayer dielectrics: Ultrasensitive transmission and breakdown of effective-medium theory,” Phys. Rev. Lett., vol. 113, no. 24, p. 243901, 2014. https://doi.org/10.1103/physrevlett.113.243901.Suche in Google Scholar
[32] S. V. Zhukovsky, et al.., “Experimental demonstration of effective medium approximation breakdown in deeply subwavelength all-dielectric multilayers,” Phys. Rev. Lett., vol. 115, no. 17, p. 177402, 2015. https://doi.org/10.1103/physrevlett.115.177402.Suche in Google Scholar
[33] H. H. Sheinfux, I. Kaminer, A. Z. Genack, and M. Segev, “Interplay between evanescence and disorder in deep subwavelength photonic structures,” Nat. Commun., vol. 7, no. 1, p. 12927, 2016. https://doi.org/10.1038/ncomms12927.Suche in Google Scholar PubMed PubMed Central
[34] H. H. Sheinfux, Y. Lumer, G. Ankonina, A. Z. Genack, G. Bartal, and M. Segev, “Observation of Anderson localization in disordered nanophotonic structures,” Science, vol. 356, no. 6341, pp. 953–956, 2017. https://doi.org/10.1126/science.aah6822.Suche in Google Scholar PubMed
[35] S. Yuan, et al.., “Breakdown of effective-medium theory by a photonic spin Hall effect,” Sci. China Phys. Mech. Astron., vol. 66, no. 11, p. 114212, 2023. https://doi.org/10.1007/s11433-023-2177-3.Suche in Google Scholar
[36] F. Goos and H. Hachen, “Ein neuer und fundamentaler Versuch zur Totalreflexion,” Ann. Phys., vol. 1, nos. 7–8, pp. 333–346, 1947.10.1002/andp.19474360704Suche in Google Scholar
[37] M. Luo and F. Wu, “Refractive-index sensing based on large negative Goos-Hänchen shifts of a wavy dielectric grating,” Phys. Rev. Appl., vol. 22, no. 1, p. 01450, 2024.10.1103/PhysRevApplied.22.014050Suche in Google Scholar
[38] Y. Xi, L. Wu, and L. K. Ang, “Surface exciton polariton enhanced Goos-Hänchen and Imbert-Fedorov shifts and their applications in refractive index sensing,” Opt. Express, vol. 32, no. 7, pp. 11171–11181, 2024. https://doi.org/10.1364/oe.517418.Suche in Google Scholar
[39] H. Pourhassan, E. Safari, M. R. Tohidkia, and A. Aghanejad, “Oxygen sensing of hemoglobin states by Goos-Hänchen effect,” Opt. Laser Tech., vol. 148, p. 107756, 2022, https://doi.org/10.1016/j.optlastec.2021.107756.Suche in Google Scholar
[40] C. Xu, J. Xu, G. Song, C. Zhu, Y. Yang, and G. S. Agarwal, “Enhanced displacements in reflected beams at hyperbolic metamaterials,” Opt. Express, vol. 24, no. 19, pp. 21767–21776, 2016. https://doi.org/10.1364/oe.24.021767.Suche in Google Scholar
[41] E. N. Afshar and A. Namdar, “Temperature dependence of the Goos-Hänchen shift in the nonlinear metal-dielectric nanocomposites,” Opt. Quant. Electron., vol. 51, no. 8, p. 258, 2019. https://doi.org/10.1007/s11082-019-1976-8.Suche in Google Scholar
[42] M. Zoghi, “Reflection shifts in a graphene-coated dielectric–metal composite of non-spherical particles,” Opt. Commun., vol. 451, pp. 160–167, 2019, https://doi.org/10.1016/j.optcom.2019.06.013.Suche in Google Scholar
[43] P. Yeh, Optical Wave in Layered Media, New York, Wiley, 1988.Suche in Google Scholar
[44] Z. Xiao, H. Luo, and S. Wen, “Goos-Hanchen and Imbert-Fedorov shifts of vortex beams at air-left-handed-material interfaces,” Phys. Rev. A, vol. 85, no. 5, p. 053822, 2012. https://doi.org/10.1103/physreva.85.053822.Suche in Google Scholar
[45] J. R. Devore, “Refractive indices of rutile and sphalerite,” J. Opt. Soc. Am. A, vol. 41, no. 6, pp. 416–419, 1951. https://doi.org/10.1364/josa.41.000416.Suche in Google Scholar
[46] I. H. Malitson, F. V. Murphy, and W. S. Rodney, “Refractive index of synthetic sapphire,” J. Opt. Soc. Am. A, vol. 48, no. 1, pp. 72–73, 1958. https://doi.org/10.1364/josa.48.000072.Suche in Google Scholar
[47] D. T. F. Marple, “Refractive index of ZnSe, ZnTe, and CdTe,” J. Appl. Phys., vol. 35, no. 3, pp. 539–542, 1964. https://doi.org/10.1063/1.1713411.Suche in Google Scholar
[48] T. Bååk, “Silicon oxynitride: A material for GRIN optics,” Appl. Opt., vol. 21, no. 6, pp. 1069–1072, 1982. https://doi.org/10.1364/ao.21.001069.Suche in Google Scholar PubMed
[49] S. Chen, C. Mi, L. Cai, M. Liu, H. Luo, and S. Wen, “Observation of the Goos-Hänchen shift in graphene via weak measurements,” Appl. Phys. Lett., vol. 110, no. 3, p. 031105, 2017. https://doi.org/10.1063/1.4974212.Suche in Google Scholar
© 2025 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.