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

Engineered polar magneto-optical Kerr rotation through Wood–Rayleigh anomalies and magnetoplamon resonance coupling

Open Access Open Access

Abstract

Polar magneto-optical Kerr rotation (MOKR) is an important phenomenon in magneto-optical materials. Both intensive and narrow Kerr rotation spectra are needed to ensure the high performance of magneto-optical devices. In this work, by using a nanosphere template and Ag film deposition at different angles, glass/Cr/Ag/Fe/PSS/Ag samples with different Ag nanostructures between polystyrene spheres were fabricated, and different optical modes and their coupling were manipulated. Polar MOKR peaks were demonstrated and analyzed. Narrow and intensive polar MOKR peaks were presented with the Ag film deposited at a glancing angle θ = 30°. The physical mechanism was studied by simulating the electromagnetic-field distribution. It was found that, in the glass/Cr/Ag/Fe/PSS/Ag system with the Ag film deposited at different angles, different optical modes and their interaction are introduced. Polar MOKR peaks due to the coupling between Fabry–Pérot and Bragg plasmons was broad. However, polar MOKR spectra governed by the coupling of WRA and magnetoplasmons demonstrated intensive and narrow peaks for samples with the Ag film deposited at θ = 30°, which would improve the performance of magneto-optical devices with high sensitivity, compact volume, high-density data storage devices, and so on.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

Polar magneto-optical Kerr rotation (MOKR) is defined as the rotation of the polarization plane of a reflective beam with respective to that of a linearly polarized incident beam, when it impinges on samples under an external magnetic field placed perpendicular to the surface of the sample and parallel to the plane of incidence [1]. As one of two parameters of the magneto-optical Kerr effect (MOKE), MOKR occurs in a different reflective plane with Kerr ellipticity in the polar MOKE configuration; thus, the polar MOKE magnetization geometry possess higher sensitivity than transverse and longitudinal magnetization geometries [1,2], which can be used in high-performance biosensors and transducers devices. For this, elaborate efforts have been made to explore Kerr rotation spectra with a narrow bandwidth and high intensity.

By using optical nanostructures including ferromagnetic materials, researchers have made intensive efforts to pursue Kerr rotation spectra with a high intensity and narrow bandwidth. Magnetoplasmons as an universal phenomenon in these nanostructures, combining magnetization and plasmonic properties [3], possess all the properties of surface plasmons such as the trapping of light in subwavelength-scale volumes, interact with other optical modes, and arbitrarily channel an electromagnetic filed [4, 5]. Moreover, it can also be controlled by an external magnetic field [6]. These multi-freedom controllable properties of magnetoplasmons would yield diverse and novel polar MOKR information and high-performance magneto-optical materials can be acquired. The Au-CoxFe3−xO4 core shell embedded in porous alumina membranes demonstrated an s-shaped Kerr rotation signal near Au plasma resonances [7]. The polar MOKR was enhanced by magnetoplasmons in Co/Au multilayer hole arrays made by nanosphere lithography [8]. Belotelov theoretically studied the modulated Kerr rotation in a metal-dielectric system, in which an Au film with subwavelength holes covered a YIG film, high-intensity and resonant Kerr rotation was obtained owing to the extraordinary optical transmission (EOT) phenomenon [9]. In nanostructured ferromagnetic systems, polar MOKR was modulated by both the intensity [10] and the phase [11] of the introduced magnetoplasmons.

Magnetoplasmons have played key roles in the intensity enhancement and resonant frequency of polar MOKR, but they have resulted in a large loss of metals and a strong broadening of polar MOKR spectra. A well-established method for overcoming the disadvantages of magnetoplasmons is coupling with other optical modes [12], such as waveguide modes, Fabry– Pérot modes, and Wood–Rayleigh anomalies (WRA). WRA originate from the diffraction order in plasmonic gratings [13, 14], and abrupt changes occur at the Rayleigh cut-off wavelength in reflection spectra. The abrupt changes due to WRA can highlight the performance of other optical modes through interaction between them. Improved extraordinary optical transimission was achieved through Fabry–Pérot resonance coupling with WRA [15]. The disadvantage of bandwidth broadening due to localized surface plasmons could be overcome with coupling to WRA and extremely narrow plasmon resonances achieved [16]. Studies on the coupling between WRA and magnetoplasmons are scarce [17]. However, regular periodic structures employed in nanostructured magneto-optical material would facilitate WRAs [18, 19]. In a deliberately designed system with magnetoplasmons, WRA resonance, and the interaction between them, high-performance polar MOKR can be achieved.

The self-assembly of 2D nanospheres is a cost-efficient, easily scalable, and variable nanofabrication technique. Combined with glancing-angle film deposition, diverse subwavelength shapes were formed at both the bottom and top of nanospheres, and different optical modes could be manipulated [20, 21]. Furthermore, highly hexaganol arrayed monolayers of uniform sphere combined with metallic film would beneficial for WRA occurring. Inspired by these ideas, in this paper, by using 2D monolayer PSS self-assembly on a glass (substrate)/Cr/Ag/Fe film, with the aid of angle-depended Ag film sputtering, different forms of surface plasmons could be manipulated. Both the polar MOKR and reflection properties could be elucidated by optical modes and their interaction occurred. And, narrow and high-intensity Kerr rotation peaks were assigned to the coupling of WRA and magnetoplasmons.

Two series of glass/Cr/Ag/Fe/PSS/Ag samples were fabricated using angle-dependent Ag film sputtering onto glass/Cr/Ag/Fe/PSS templates. The fabrication details of the templates were reported in ref [22]. Firstly, a Cr(20 nm)/Ag(200 nm)/Fe(10 nm) film was deposited on glass by three-source dc magnetron sputtering from Cr, Ag, and Fe targets in turn, where Cr (20 nm) was deposited as a buffer before Ag to strengthen the adhesion of Ag. Then, 2D arrays of PSSs with different diameters were aligned on Fe films through self-assembly. The schematic plot in Fig. 1 depicts the placement of templates during the deposition of the uppermost-layer Ag film (a) parallel to the sample stage θ = 0° or (d) with a glancing angle θ = 30° relative to the sample stage. The depositions were performed thrice by rotating the sample stage at angle intervals of 60°, and an Ag film of thickness 35 nm was deposited each time. In the experiment, PSSs with diameter d = 280, 300, and 414 nm were used. The morphologies of samples with d = 300 nm were characterized using a commercial atomic force microscopy (AFM) instrument (Slover P47, NT-MDT). Fig. 1 (b) and (e) show top-view AFM images of samples deposited at θ = 0° and θ = 30°, respectively. Clearly, the Ag film deposited at θ = 0°(b) demonstrated a circular shape on top of PSS arrays, whereas that deposited at θ = 30°(e) seemed like trianglar patches on top of PSS arrays. After all measurements (both polar MOKR and reflectivity measurements), PSSs were washed off by immersing the two series of samples with d = 300 nm in chlorobenzene to study the shape of Ag patches at the bottom of PSS arrays. Scanning electron microscopy (SEM) images of samples with d = 300 nm without PSS deposited at θ = 0°(c) and θ = 30°(f) were acquired. The triangle and circle nanostructures at the bottom of PSS could be deduced as in Fig. 1(c) and (f), respectively. This result is in good agreement with ref [20].

 figure: Fig. 1

Fig. 1 Schematic of Ag films deposited onto glass/Cr/Ag/Fe/PSS templates placed at θ = 0° (a) or θ = 30° (d) relative to the sample stage. Top-view AFM images of samples (b,e) and SEM images (c,f) of Ag structures deposited at θ = 0° and θ = 30°.

Download Full Size | PDF

The spectra of polar MOKR θk and optical reflectance R were measured at room temperature by using a home-made Kerr spectrometer from 300 nm to 800 nm and by using a reflectometer from 200 nm to 1200 nm, respectively [22]. The θk measurements were carried out at an incident angle of 5° under an external magnetic field of 1.0 T. In the reflectivity measurements, the data were normalized with respect to the reflectivity of Al films. The reflection spectra and electromagnetic (EM)-field distribution simulation were performed using commercial software FDTD solution based on finite difference time domain (FDTD) method with mesh of 0.1 nm. The data of Ag and Fe came from measured data using ellipsometry.

The measured spectra of polar MOKR θk are displayed in Fig. 2 for the samples with θ = 0° (a–c) and θ = 30° (d–f). The diameters of PSSs were d = 280 (a, d), 300 (b, e), and 414 (c, f) nm. For clarity, the MOKR peaks are marked in Fig. 2. Evidently, the polar MOKR spectra of these two series demonstrated different behaviors. For samples deposited at θ = 0°, prominent but broad 1st peaks were revealed at λ = 600 (a), 640 (b), and 755 (c) nm in Fig. 2. The 2nd peaks located at λ = 358 (a), 380 (b), and 438 (c) nm were shallow and hardly discernible. However, for the samples deposited at θ = 30°, in addition to the 1st broad peaks located at λ = 625 (d), 654 (e), 785 (f) nm, pronounced and narrow 2nd peaks were observed at λ = 367 (d), 389 (e), and 460 (f) nm, respectively.

 figure: Fig. 2

Fig. 2 θk spectra as functions of the wavelength of samples with deposition angle θ = 0° (a–c) and θ = 30°(d–f). The PSS diameter d = 280 (a, d), 300 (b, e), and 414 (c, f) nm.

Download Full Size | PDF

This tremendous discrepancy was not found in the reflection spectra (measured without external magnetic field) of the two series of samples in Fig. 3, where both the measured (black squares) and calculated (red triangles) reflection spectra of the two series of samples were layered out. The agreement between the measured and calculated reflection spectra in Fig. 3 consolidated our material research base and method reliability. The reflection spectra of samples with Ag deposited at θ = 0° (a–c) and θ = 30°(d–f) showed the same behavior, and no significant discrepancy was observed in the reflection spectra of these two series of samples as in MOKR spectra, except that the 2nd dips in Fig. 3 (d–f) were slightly deeper and narrower than those in Fig. 3 (a–b). Evidently, the reflection dips in Fig. 3 corresponded to MOKR peaks in Fig. 2 well. For samples deposited at θ = 0°, broad 1st dips were presented at λ = 596 (a), 622 (b), and 756 (c) nm, and the 2nd narrow ones located at λ = 361 (a), 381 (b), and 448 (c) nm. For the samples deposited at θ = 30°, broad 1st dips were near λ = 630 (d), 669 (e), 796 (f) nm, and the narrow 2nd ones were observed at λ = 368 (d), 397 (e), and 466 (f) nm, respectively.

 figure: Fig. 3

Fig. 3 Reflectivity spectra as functions of the wavelength of samples with deposition angle θ = 0° (a–c) and θ = 30°(d–f). The PSS diameter d = 280 (a, d), 300 (b, e) and 414 (c, f) nm.

Download Full Size | PDF

By combining the MOKR and reflection spectra in Fig. 2 and Fig. 3, we analyze these novel phenomena based on the theory of diffraction and surface plasmon plaritons (SPPs). When light perpendicularly impinges on the 2D hexagonal PSS arrays covered by Ag film from air, SPPs occur at Ag-PSS and PSS-Fe-Ag hybrid interface [23], the diffraction or SPP resonance order can be calculated by [13,14,24]

kd/spp=ki±mGx±nGy,
Where k⃗i = n0 sin θ2π/λ is the wave vectors of light incident, and k⃗d/spp = neff 2π/λ represent the wave vectors of diffracted beam or SPP from the 2D hexagonal arrays. Gx=Gy=4π3d are the reciprocal vectors of the 2D hexagonal grating, n0 = 1.0 is the refractive index of air, and neff is the effective refractive index of the Ag/Fe/PSS/Ag system [25]. Different optical modes are denoted by neff. In this subwavelength condition, the Rayleigh anomalous wavelength can be written as [24]
mλRA=Λ(nisinθ+neff),
Where Λ is the reciprocal lattice vector, θ is the incident angle, and at normal incidence, θ = 0°. Clearly, the location of the 2nd MOKR peaks/reflection dips is near to that of the Rayleigh cut-off wavelength [13–17,24]. As the tremendous intensity discrepancy occurred in 2nd MOKR peaks but not occurred in 2nd reflection dips, it was deduced that the 2nd MOKR peaks in samples with up-layer Ag deposited at θ = 30° in Fig. 2(d–f) might be attributed to RWA coupled with magnetoplasmons at the PSS-Fe interfaces, which could be controlled by the external magnetic field, and thus could prompt polar MOKR signal. Although different Ag nanostructure shapes underneath the PSSs were demonstrated as indicated in Fig. 1(c) and (f), the impact of nanostructure shape on the 2nd MOKR peaks was not as that in Co0.12Al2O3-Au system [26]. The 1st MOKR peaks of the two series of samples demonstrated similar behavior, furthermore, the resonant wavelength is larger that of the period, which suggests that they might be related to the Fabry–Pérot (F-P) mechanism for deep enough of the 2D grating at the middle of PSS [15,27].

To further study the physical mechanism of the MOKR peaks in Fig. 2, the simulated EM-field distribution in the xz-plane (denoted in Fig. 1) of the 1st and the 2nd reflection dips (corresponding to the 1st and the 2nd KR peaks) of samples with top layer Ag deposited at θ = 0°(a–c) and θ = 30°(d–f) is presented in Fig. 4 and Fig. 5, respectively.

 figure: Fig. 4

Fig. 4 Field-amplitude distribution in the xz-plane of the 1st reflection dips for samples deposited at θ = 0°(a–c) and θ = 30°(d–f). The location of EM-field is near λ = 596 (a), 622 (b), 756 (c), 630 (d), 669 (e) and 796 (f) nm.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Field-amplitude distribution in the xz-plane of the 2nd reflection dips for samples deposited at θ = 0°(a–c) and θ = 30°(d–f), the location of EM-field is near λ = 361 (a), 381 (b), 448 (c), 368 (d), 397 (e), and 466 (f) nm.

Download Full Size | PDF

The location of EM-field of the 1st dips in Fig. 3 is shown in Fig. 4 for samples with top layer Ag deposited at θ = 0°(a–c) and θ = 30°(d–f), respectively, which exhibit similar behavior. The EM-field of the 1st reflection dips samples with d= 280 and 300 nm in Fig. 4(a, d) and (b, e) were strongly distributed in the upper part of the interior of the nanospheres and weakly distributed in free space. However, the EM-field distribution for samples with d = 414 nm in Fig. 4 (c, f) was mainly localized in the upper spaces between nanospheres. But they all demonstrated diffraction pattern in free space, and we could ascribe the 1st MOKR peaks in Fig. 2(a–c) and (d–f) to the coupling between F-P resonance and Bragg plasmons at the upper capped Ag-PSS interface.

The EM-field distribution of the 2nd reflection dips is illustrated in Fig. 5 for samples with top layer Ag deposited at θ = 0°(a–c) and at θ = 30°(d–f), respectively. The white dashed line shows the Ag-PSS and PSS-Fe interfaces. Obviously, distinct differences were revealed in the EM-field distribution of the 2nd reflection dips between the two series samples. The EM-field of the 2nd dips of samples deposited at θ = 0°(a–c) were located at the PSS-Ag interface and the bottom of the inside of nanospheres. Whereas the EM-field of the 2nd dips of samples deposited at θ = 30°(d–f) were strongly localized in the complicated Ag-PSS and PSS-Fe interfaces, in addition to the bottom of the inside of nanospheres. Such localized magnetoplasmons coupled with WRA would enhance the interaction of the EM-field with ferromagnetic materials and lead to intensive and narrow polar MOKR, as could be seen in Fig. 2 (d–f). In contrast, for the EM-field located at the bottom of the inside of nanospheres and PSS-Ag interface, only weak polar MOKR peaks were observed in Fig. 2 (a–c). This distinct difference was not found in reflection spectra, because the localized EM-field in both PSS-Ag and PSS-Fe interface would all lead to similar reflection results without external magnetic field, as indicated by the 2nd dips in reflection spectra in Fig. 3 (a–c) and (d–f).

In conclusion, by using a nanosphere template and glancing-angle film deposition, we presented a facile method to investigate the impact of coupling between different optical modes on polar MOKR. Intensive and narrow polar MOKR spectra were demonstrated for samples with the deposition angle θ = 30°. Through a numerical analysis and simulation of EM-field distribution, the polar MOKRs of glass/Cr/Ag/Fe/PSS/Ag systems were assigned to modulation by the coupling between WRA, F-P and different surface plasmons. The coupling between WRA and magnetoplasmons on the PSS-Fe interfaces led to intensive and narrow polar MOKR peaks. The study of narrow and high-intensity polar MOKR spectra might pave the way to the discovery of new physics and design of high-performance MO devices.

Funding

Natural Science Foundation of Shandong Province (ZR2015AM024); National Science Fund for Young Scholars (11704219); Doctoral Scientific Research Start-up Funding from Qufu Normal University (BSQD20130152); Natural Science Foundation of Shandong Province (ZR2016AQ09).

References

1. W. Reim and J. Schoenes, Chapter 2 Magneto-optical spectroscopy of f-electron systems (Elsevier B.V., 1990).

2. A. K. Zvezdin and V. A. Kotov, Modern Magnetooptics and Magnetooptical Materials: Studies in Condensed Matter (IOP Publishing Ltd, 1997). [CrossRef]  

3. K. W. Chiu and J. J. Quinn, “Magneto-plasma surface waves in solids,” IL NUOVO CIMENTO 10, 1 (1972). [CrossRef]  

4. E. Ozbay, “Plasmonics: Merging Photonics and Electronics at Nanoscale Dimensions,” Science 311, 189 (2006). [CrossRef]   [PubMed]  

5. N. I. Zheludev, “Obtaining optical properties on demand,” Science 348, 973 (2015). [CrossRef]   [PubMed]  

6. K. Lodewijks, N. Maccaferri, T. Pakizeh, R. K. Dumas, I. Zubritskaya, J. Åkerman, P. Vavassori, and A. Dmitriev, “Magnetoplasmonic Design Rules for Active Magneto-Optics,” Nano Lett. 14, 7207 (2014). [CrossRef]   [PubMed]  

7. G. Armelles, A. Cebollada, A. Garcíamartín, J. M. Monteromoreno, M. Waleczek, and K. Nielsch, “Magneto-optical Properties of Core-Shell Magneto-plasmonic Au-CoxFe3-xO4$_4$ Nanowires,” Langmuir ACS J. Surfaces Colloids 28, 9127 (2012). [CrossRef]  

8. E. T. Papaioannou, T. Meyer, and B. Hillebrands, “Extraordinary Magneto-Optical Effects and Transmission through Metal-Dielectric Plasmonic Systems,” J. Surfaces Interfaces Mater. 2, 40 (2014).

9. V. I. Belotelov, L. L. Doskolovich, and A. K. Zvezdin, “Surface plasmons and magneto-optic activity in hexagonal Ni anti-dot arrays,” Phys. Rev. Lett. 98, 077401 (2007). [CrossRef]  

10. A. Garciamartin, B. Hjörvarsson, E. Melander, E. T. Papaioannou, G. Ctistis, M. Giersig, P. Fumagalli, P. Patoka, S. D. Pappas, and V. Kapaklis, “Surface plasmons and magneto-optic activity in hexagonal Ni anti-dot arrays,” Opt. Express 19, 23867 (2011). [CrossRef]  

11. N. Maccaferri, A. Berger, S. Bonetti, V. Bonanni, M. Kataja, Q. H. Qin, D. S. Van, Z. Pirzadeh, A. Dmitriev, and J. Nogués, “Tuning the Magneto-Optical Response of Nanosize Ferromagnetic Ni Disks Using the Phase of Localized Plasmons,” Phys. Rev. Lett. 111, 167401 (2013). [CrossRef]  

12. N. Liu and H. Giessen, “Coupling Effects in Optical Metamaterials,” Angewandte Chemie Int. Ed. 49, 9838 (2015).

13. R. W. Wood, “Anomalous Diffraction Gratings,” Phys. Rev. 48, 928 (1935). [CrossRef]  

14. L. Rayleigh, “On the Dynamical Theory of Gratings,” Proc. Royal Soc. Lond. 79, 399 (1907). [CrossRef]  

15. A. T. Rahman, P. Majewski, and K. Vasilev, “Extraordinary optical transmission: coupling of the Wood–Rayleigh anomaly and the Fabry–Perot resonance,” Opt. Lett. 37, 1742 (2015). [CrossRef]  

16. V. G. Kravets, F. Schedin, and A. N. Grigorenko, “Extremely Narrow Plasmon Resonances Based on Diffraction Coupling of Localized Plasmons in Arrays of Metallic Nanoparticles,” Phys. Rev. Lett. 101, 087403 (2008). [CrossRef]  

17. A. V. Chetvertukhin, A. A. Grunin, A. V. Baryshev, T. V. Dolgova, H. Uchida, M. Inoue, and A. A. Fedyanin, “Magneto-optical Kerr effect enhancement at the Wood's anomaly in magnetoplasmonic crystals,” J. Magn. Magn. Mater. 324, 3516 (2012). [CrossRef]  

18. M. Kataja, T. K. Hakala, A. Julku, M. J. Huttunen, S. V. Dijken, and P. Törmä, “`Surface lattice resonances and magneto-optical response in magnetic nanoparticle arrays,” Nat. Commun. 6, 7072 (2011). [CrossRef]  

19. H. Fang, B. Caballero, E. M. Akinoglu, E. T. Papaioannou, A. Garcíamartín, J. C. Cuevas, M. Giersig, and P. Fumagalli, “Observation of a hole-size-dependent energy shift of the surface-plasmon resonance in Ni antidot thin films,” Appl. Phys. Lett. 106, 153104 (2015).

20. A. Kosiorek, W. Kandulski, P. Chudzinski, K. Kempa, and M. Giersig, “Shadow Nanosphere Lithography: Simulation and Experiment,” Nano Lett. 4, 1359 (2004). [CrossRef]  

21. T. Wang, T. Fu, Y. Chen, and Z. Zhang, “Circular dichroism of a tilted U-shaped nanostructure,” Opt. Lett. 42, 2842 (2017). [CrossRef]   [PubMed]  

22. X. Zhang, J. Li, Y. Xia, K. Cao, X. Zhang, Y. Y. Wang, and X. Sun, “Influence of Nanovoids Depth on the Surface Magneto-Optical Kerr Effect Modulations in Magnetic Photonic Crystal Slab,” IEEE Photonics J. 9, 5300309 (2017).

23. H. Raether, Surface Plasmons on Smooth and Rough Surfaces and on Gratings (SpringerBerlin Heidelberg, 1988). [CrossRef]  

24. J. Maraedjouda, R. Caputo, N. Mahi, G. Lévêque, A. Akjouj, P. M. Adam, and T. Maurer, “Angular plasmon response of gold nanoparticles arrays: approaching the Rayleigh limit,” Nanophotonics 6, 279 (2016).

25. J. Strong, “Effect of Evaporated Films on Energy Distribution in Grating Spectra,” Phys. Rev. 49, 291 (1936). [CrossRef]  

26. A. F. Kravets, T. I. Borodinova, and V. G. Kravets, “Strong plasmon enhancement of magneto-optical Kerr rotation in Co–AlO nanogranular films coated with gold nanoparticles,” J. Opt. Soc. Am. B 33, 302 (2016). [CrossRef]  

27. N. Garcia and M. Nietovesperinas, “Theory of electromagnetic wave transmission through metallic gratings of subwavelength slits,” J. Opt. A Pure Appl. Opt. 9, 490 (2007). [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 Schematic of Ag films deposited onto glass/Cr/Ag/Fe/PSS templates placed at θ = 0° (a) or θ = 30° (d) relative to the sample stage. Top-view AFM images of samples (b,e) and SEM images (c,f) of Ag structures deposited at θ = 0° and θ = 30°.
Fig. 2
Fig. 2 θk spectra as functions of the wavelength of samples with deposition angle θ = 0° (a–c) and θ = 30°(d–f). The PSS diameter d = 280 (a, d), 300 (b, e), and 414 (c, f) nm.
Fig. 3
Fig. 3 Reflectivity spectra as functions of the wavelength of samples with deposition angle θ = 0° (a–c) and θ = 30°(d–f). The PSS diameter d = 280 (a, d), 300 (b, e) and 414 (c, f) nm.
Fig. 4
Fig. 4 Field-amplitude distribution in the xz-plane of the 1st reflection dips for samples deposited at θ = 0°(a–c) and θ = 30°(d–f). The location of EM-field is near λ = 596 (a), 622 (b), 756 (c), 630 (d), 669 (e) and 796 (f) nm.
Fig. 5
Fig. 5 Field-amplitude distribution in the xz-plane of the 2nd reflection dips for samples deposited at θ = 0°(a–c) and θ = 30°(d–f), the location of EM-field is near λ = 361 (a), 381 (b), 448 (c), 368 (d), 397 (e), and 466 (f) nm.

Equations (2)

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

k d / s p p = k i ± m G x ± n G y ,
m λ RA = Λ ( n i sin θ + n eff ) ,
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.