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

Extraordinary optical transmission inside a waveguide: spatial mode dependence

Open Access Open Access

Abstract

We study the influence of the input spatial mode on the extraordinary optical transmission (EOT) effect. By placing a metal screen with a 1D array of subwavelength holes inside a terahertz (THz) parallel-plate waveguide (PPWG), we can directly compare the transmission spectra with different input waveguide modes. We observe that the transmitted spectrum depends strongly on the input mode. A conventional description of EOT based on the excitation of surface plasmons is not predictive in all cases. Instead, we utilize a formalism based on impedance matching, which accurately predicts the spectral resonances for both TEM and non-TEM input modes.

© 2016 Optical Society of America

1. Introduction

For more than two decades, the extraordinary optical transmission (EOT) of electromagnetic radiation through subwavelength apertures has been an important research topic [1]. Since the initial report [2], the EOT effect has been studied at frequencies from optical [3,4] to microwave [5,6], and has been considered among the hallmark examples of the significant role of surface waves in light-matter interactions. Nearly all previous studies have involved free-space coupling of waves to the subwavelength structures (or coupling from a uniform dielectric medium), such that both the incident wave and the far-field transmitted wave are transverse electromagnetic (TEM) waves. One might expect different behavior if the subwavelength apertures are placed inside a waveguide, since both input and output fields would be restricted to the lone (or few) allowed waveguide mode(s), which need not necessarily be TEM. Some previous works have demonstrated EOT via subwavelength holes inside hollow metal waveguides [7–9], but to our knowledge, the mode selectivity of the EOT effect – that is, the extent to which the input mode affects the transmission spectrum – has not been investigated. A comparison of TEM and non-TEM mode excitation can illuminate the role of surface waves in mediating the transmission through subwavelength hole arrays.

Here, we use broadband terahertz radiation to study the EOT effect for an array of subwavelength holes in a metal screen that has been embedded inside a metal parallel-plate waveguide (PPWG). Unlike previous reports [7–9], our use of a PPWG allows us to excite the subwavelength hole with a TEM mode (which is not possible in hollow rectangular or circular waveguides) [10], as well as non-TEM modes [11,12], permitting a direct comparison between these two distinct situations for the first time. We also highlight the comparison to the normal-incidence transmission through a similar hole array in free-space. We observe that the conventional explanation which relies on the excitation of spoof surface plasmon polaritons (SPPs) [3,13,14] is not descriptive in all cases, particularly where the incident wave has a longitudinal field component (i.e., where it is not TEM). Previous work has demonstrated EOT without the excitation of SPPs in a circular waveguide [8], but our use of the PPWG allows us to directly compare TEM to non-TEM modes experimentally. We utilize another theoretical framework based on electromagnetic mode-matching and circuit theory [15–17] which more accurately predicts the resonant transmission frequencies for both the TEM and non-TEM cases.

2. Design and theory

Our experimental design is depicted in Fig. 1(a). The waveguide is excited using quasi-optic coupling of broadband THz pulses in a conventional THz time-domain spectroscopy system with photoconductive antennas. Each of two PPWG sections is fabricated out of aluminum with length L = 10 mm, and a variable plate separation b. We show results for selected values of b = 0.75, 1.00, and 1.25 mm. The metal screens are made from stainless steel foil of thickness, t = 25 μm (corresponding to roughly λ/40 where λ is the wavelength corresponding to the fundamental transmission resonance), and have a one-dimensional (1D) array of holes with uniform spacing between holes of either a = 0.75 mm or 1.0 mm. We explore two different hole diameters: the larger diameter holes (380 μm) are fabricated by conventional machining, while the smaller diameter holes (300 μm) are fabricated by laser drilling. The width of the waveguide and the metal foil in the lateral (unconfined) direction, as well as the length of the 1D hole array, are all much larger than the excitation beam diameter, thereby ensuring no edge effects. The screen is sandwiched between the two waveguide sections, as illustrated in Fig. 1(b), forming an effective continuous waveguide with the EOT screen located half-way along the length. For comparison, we also fabricate a two-dimensional (2D) array of holes in stainless steel foil of the same thickness and hole diameter of 380 μm, which we characterize by normal-incidence transmission spectroscopy in free-space (i.e., no waveguide).

 figure: Fig. 1

Fig. 1 Panel (a) shows a 3D view of the PPWG (WG1 and WG2) with an array of holes in the middle, where the waveguide plates are pictured translucent to see the screen, b is distance between the two plates, a is the distance between two holes, d is the diameter of the holes, and L is the length of the waveguide section. Panel (b) shows a side view of the experimental configuration where the THz wave is coupled into WG1 from the left and the screen with holes of thickness t is sandwiched in the middle between WG1 and WG2.

Download Full Size | PDF

To assess the influence of the incident mode on the EOT spectrum, we compare the transmission through a 2D array in free-space to that of a 1D array inside a PPWG for both TEM and TE1 input waveguide modes, displayed in Fig. 2. In this comparison, the 2D array has holes in a square lattice with a spacing of 1.0 mm in both directions. The 1D array inside the PPWG also has a 1.0 mm spacing. For the 2D array in free-space, measurements with vertical and horizontal polarizations produce identical results due to the four-fold rotational symmetry of the array as shown in Fig. 2(a). For the 1D array inside the PPWG, shown in Fig. 2(b), we observe that excitation with the TEM waveguide mode (polarization perpendicular to the linear hole array) produces the same resonant frequency as that of the free-space measurements. However, for TE1 input mode, where the electric field polarization axis is parallel to the 1D hole array, shown in Fig. 2(c), the resonance clearly shifts to a higher frequency. If the EOT physics only depended on the electric field polarization, we would expect the same results for all three of these measurements. Since the amplitude variation of the mode across the hole is small in both cases, we conclude that the presence of a longitudinal magnetic field component (as in the TE1 waveguide mode) has a significant influence on the EOT resonant frequency.

 figure: Fig. 2

Fig. 2 Power transmission at normal incidence for (top) a 2D array of holes in free-space illuminated by vertical polarization (black) or horizontal polarization (gray), (middle) a 1D array of holes inside a PPWG in TEM mode (blue), (bottom) a 1D array of holes inside a PPWG in TE1 mode (red). The dashed purple and oranges lines (overlapped) show prediction of RWA minimum and SPP maximum, respectively. This minimum value agrees well with 2D array in free-space and 1D array in the TEM PPWG. Clearly the resonant frequency is shifted for the 1D array in the TE1 PPWG.

Download Full Size | PDF

In the most commonly discussed theoretical descriptions of EOT, the transmission maxima are determined [13] by the excitation of SPPs at the dielectric-metal interface:

fmax,SPP=cεd+εmεdεm(mLx)2+(nLy)2,
while transmission minima are attributed to Rayleigh-Wood’s anomalies (RWA):
fmin,RWA=cεd(mLx)2+(nLy)2.
Here, εd is the dielectric constant of the surrounding medium (in this case air), εm is the dielectric constant of the metal, Lx and Ly are the periodicity in their respective dimensions, and m and n are integers (not both zero). This description has been used in the optical range; we note that EOT has also been demonstrated at microwave frequencies where strongly surface-localized SPPs are not supported because of the extremely large conductivity of most metals. Nevertheless, periodic corrugations on metals can support spoof surface plasmons [14] at any frequency, so that this SPP explanation of the EOT effect has still been used. Others have demonstrated the EOT effect in subwavelength hole arrays in free-space at THz frequencies, relying on an explanation involving the excitation of SPPs [18,19]. At THz frequencies, the conductivity of most metals is high, but not as high as at lower frequencies. Thus, it is important to note that metals at THz frequencies can support SPPs even in the absence of surface corrugations [20].

We compare these predictions to our situation with the subwavelength hole array situated inside a waveguide. The values of the SPP maxima and RWA minima predicted by Eqs. (1) and (2), shown as dashed lines in Fig. 2, nearly coincide since in the limit that εm >> εd the dielectric prefactors in Eqs. (1) and (2) are approximately equal. In earlier works at other frequencies, these predictions have been used with some success, although often they exhibit an unexplained offset from the measured transmission features [1,3]. In our case, there is good agreement between the SPP prediction of the transmission minima and the experimental results for both the 2D free-space array and the 1D hole array when excited by the TEM waveguide mode. However, our data demonstrate that the free-space SPP prediction does not apply when the hole array is excited by the TE1 waveguide mode (see Fig. 2(c)). We emphasize that, in the comparison between the TEM and TE1 modes, the electric field polarization lies entirely in the plane of the EOT screen’s surface in both cases (i.e., the propagating wave is incident normally, and the electric field vector is transverse, in all of the cases discussed here).

To accurately predict the transmission resonances for both TEM and non-TEM modes, we turn to an alternate description originating from frequency selective surface (FSS) analysis applied to EOT [17]. Instead of invoking SPPs, one can apply a circuit theory perspective, in which transmission is mediated by impedance matching through the subwavelength aperture. As discussed in [17,21], the structure of a 2D array of subwavelength holes in free-space can be divided into unit cells, each containing one aperture, with side lengths corresponding to the periodicity of the array. Then, with the appropriate boundary conditions, each unit cell can be viewed as a fictitious quasi-rectangular waveguide containing a screen with a single aperture. The aperture is a discontinuity in the waveguide that permits excitation of higher order evanescent modes. The transmission process is mediated by matching the impedance of the quasi-rectangular waveguide to that of the aperture. Since the aperture is subwavelength in size and therefore has a large admittance [7], this matching criterion requires that the waveguide mode also have a large admittance. Asffcm (where fcis the cutoff frequency of the corresponding m,n mode), TM waveguide modes have a diverging admittance whereas the admittance of TE waveguide modes vanishes. Thus, the evanescent TM modes facilitate impedance matching and exhibit a transmission maximum as the frequency approaches fcmnand a minimum atfcmn.

We apply this analysis in a new context to the PPWG system, where we can directly compare TEM and TE1 waveguide modes to the case of a 2D array in free-space. Each unit cell of our subwavelength hole array is viewed as a fictitious rectangular waveguide with dimensions determined in the x-direction by the hole spacing and in the y-direction by the plate separation. Based on the symmetry of a hole centered in the waveguide, for TEM input excitation, all TM2p,2q (with m = 2p and n = 2q) modes show this divergent nature of the admittance as cutoff is approached. In contrast, for TE1 input excitation, it is the TM2p + 1,2q (with m = 2p + 1 and n = 2q) modes [7]. The relevant cutoff frequencies are given by the standard results for a rectangular waveguide, where in our axis labeling, b (the plate separation) corresponds to the mode index m, and a (the hole spacing) corresponds to the mode index n:

fmin,TEM=fcTM2p,2q=c2πεd(mπb)2+(nπa)2=cεd(pb)2+(qa)2,
fmin,TE1=fcTM2p+1,2q=c2πεd(mπb)2+(nπa)2=cεd(2p+1b)2+(qa)2,

The relevant mode facilitating transmission is the lowest order evanescent TM waveguide mode, which for TEM is the TM20 mode, and for TE1 is the TM12 mode. It is important to note that, for the TEM waveguide mode, this cutoff relationship in Eq. (3) is the same as Eq. (2), which is the transmission minimum predicted by RWA. Thus, for the TEM waveguide mode, this impedance matching description gives the same prediction for the transmission minimum as does the SPP description. However, for the case of excitation by the TE1 waveguide mode, Eq. (4) does not reduce to the same result as Eq. (2) for the TM12 mode due to the difference of the spatial mode pattern. We emphasize that this analysis only applies to arrays and not to a single isolated hole since the justification for the fictitious rectangular waveguide arises from the periodicity of the array using appropriate boundary conditions [21].

Based on this approach, the transmission can be calculated analytically as a function of frequency using the coupled-mode method (CMM). As described above, we model the system as a rectangular waveguide with a single subwavelength hole. To apply the CMM, we expand the incident field into eigenmodes, then calculate the scattering coefficients from incident to transmitted field by setting appropriate continuity conditions [15,16]. Previously, the CMM in the context of EOT has only been applied to free-space and so would use plane waves as the input and output fields. Here, we adapt the method to use the supported waveguide modes (i.e. TE mode field patterns) for the input (Region 1) and output (Region 3). For the array of holes (Region 2), we employ the discrete waveguide modes of a subwavelength hole, even though they are all evanescent, and match the conditions across these three regions. We use the three least evanescent aperture modes directly excited by the incident mode for good convergence.

In addition to this analytical approach, we also perform numerical computations based on the finite-element method (FEM). We treat the metal waveguide plates as perfect electric conductors (PEC) and the screen as a real (finite conductivity) metal using complex permittivity values from Ref. 20 (although there is little difference if it is treated as a PEC).

3. Results

In Figs. 3 and 4, we compare the experiment, FEM simulation, and CMM theory for a hole diameter of d = 380 μm and various plate separations b for TEM, shown in Fig. 3, and TE1, shown in Fig. 4, input waveguide modes. Figures 3(a)–(c) and Figs. 4(a)–(c) show the results for a hole spacing of a = 0.75 mm. For comparison, we also display the results for a different hole spacing, a = 1.0 mm, shown in Figs. 3(d) and 4(d). We find good agreement between experiment and prediction for the peak position of the resonant frequencies and minima for both TEM and TE1 excitation modes. The data showing the larger hole spacing further emphasizes the significance of the impedance matching mediated by TMm,n modes. In the TEM case, since the relevant mode is TM20, we anticipate no dependence on the hole spacing, as reflected in the lack of a spectral shift between Figs. 3(a) and 3(d). In contrast, for the TE1 input mode where the relevant mode is TM12, a spectral shift with a change in a is anticipated, and correctly predicted (vertical dotted lines).

 figure: Fig. 3

Fig. 3 Power transmission for a fixed hole separation of (a-c) a = 0.75 mm and (d) a = 1.00 mm for the plate separations of b = 0.75 mm (black, thin line), b = 1.00 mm (red), and b = 1.25 mm (blue, thick line) with TEM mode PPWG excitation. The dashed lines indicate the cutoff of the corresponding TM20 mode.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Power transmission for a fixed hole separation of (a-c) a = 0.75 mm and (d) a = 1.00 mm for the plate separations of b = 0.75 mm (black, thin line), b = 1.00 mm (red), and b = 1.25 mm (blue, thick line) with TE1 mode PPWG excitation. The dashed lines indicate the cutoff of the corresponding TM12 mode. For experiment and theory with a = 0.75 mm and b = 1.25 mm, we see many oscillations which can be attributed to the excitation of higher order propagating waveguide modes (i.e. cutoff of TE3 is 350GHz).

Download Full Size | PDF

This waveguide EOT configuration has potential interest for narrowband tunable filters. In most cases, the geometry of the EOT structure is fixed (i.e., the periodicity of the hole array is not variable), so frequency tuning is challenging. However, in our case the waveguide plate spacing can be easily varied, thus dynamically tuning the transmission resonance. Such tunable filters may find future use in THz communications systems where the PPWG platform has been discussed in the context of various signal processing functionalities [22,23].

4. Conclusion

In conclusion, we describe the first comprehensive study of the role of the incident wave’s spatial mode on the EOT phenomenon. By using a PPWG (in contrast to the rectangular or circular waveguides more typically employed at lower frequencies), we are able to perform a direct comparison between the cases of TEM and non-TEM input modes. This work is reminiscent of recent studies of near-perfect transmission through structures with effective permittivity near zero [24,25]. Here, we focus more specifically on the role of higher order evanescent waveguide modes, which is more closely related to discussions involving surface plasmon-mediated enhanced transmission. We find that the conventional description based on the excitation of surface plasmon modes (or spoof surface plasmons) does not accurately capture the physics of the two distinct mode cases investigated. That is, it does not correctly predict the transmission frequencies in the case when the incident wave is not a TEM wave. Instead, we adapt a formalism based on impedance matching, which more accurately captures the details of the spectra in both cases.

Funding

This research has been supported in part by the National Science Foundation (NSF).

References

1. C. Genet and T. W. Ebbesen, “Light in tiny holes,” Nature 445(7123), 39–46 (2007). [CrossRef]   [PubMed]  

2. T. W. Ebbesen, H. J. Lezec, H. F. Ghaemi, T. Thio, and P. A. Wolff, “Extraordinary optical transmission through sub-wavelength hole arrays,” Nature 391(6668), 667–669 (1998). [CrossRef]  

3. W. L. Barnes, W. A. Murray, J. Dintinger, E. Devaux, and T. W. Ebbesen, “Surface plasmon polaritons and their role in the enhanced transmission of light through periodic arrays of subwavelength holes in a metal film,” Phys. Rev. Lett. 92(10), 107401 (2004). [CrossRef]   [PubMed]  

4. H. Gao, J. M. McMahon, M. H. Lee, J. Henzie, S. K. Gray, G. C. Schatz, and T. W. Odom, “Rayleigh anomaly-surface plasmon polariton resonances in palladium and gold subwavelength hole arrays,” Opt. Express 17(4), 2334–2340 (2009). [CrossRef]   [PubMed]  

5. A. P. Hibbins, M. J. Lockyear, I. R. Hooper, and J. R. Sambles, “Waveguide arrays as plasmonic metamaterials: transmission below cutoff,” Phys. Rev. Lett. 96(7), 073904 (2006). [CrossRef]   [PubMed]  

6. B. Hou, Z. H. Hang, W. Wen, C. T. Chan, and P. Sheng, “Microwave transmission through metallic hole arrays: Surface electric field measurements,” Appl. Phys. Lett. 89(13), 131917 (2006). [CrossRef]  

7. Y. Pang, A. N. Hone, P. P. M. So, and R. Gordon, “Total optical transmission through a small hole in a metal waveguide screen,” Opt. Express 17(6), 4433–4441 (2009). [CrossRef]   [PubMed]  

8. F. Medina, J. Ruiz-Cruz, F. Mesa, J. M. Rebollar, J. R. Montejo-Garai, and R. Marqués, “Experimental verification of extraordinary transmission without surface plasmons,” Appl. Phys. Lett. 95(7), 071102 (2009). [CrossRef]  

9. F. Medina, F. Mesa, J. A. Ruíz-Cruz, J. M. Rebollar, and J. R. Montejo-Garai, “Study of extraordinary transmission in a circular waveguide system,” IEEE Trans. Microw. Theory Tech. 58(6), 1532–1542 (2010). [CrossRef]  

10. R. Mendis and D. Grischkowsky, “Undistorted guided-wave propagation of subpicosecond terahertz pulses,” Opt. Lett. 26(11), 846–848 (2001). [CrossRef]   [PubMed]  

11. R. Mendis and D. M. Mittleman, “Comparison of the lowest-order transverse-electric (TE1) and transverse-magnetic (TEM) modes of the parallel-plate waveguide for terahertz pulse applications,” Opt. Express 17(17), 14839–14850 (2009). [CrossRef]   [PubMed]  

12. R. Mendis and D. M. Mittleman, “An investigation of the lowest-order transverse-electric (TE1) mode of the parallel-plate waveguide for THz pulse propagation,” J. Opt. Soc. Am. B 26(9), A6–A13 (2009). [CrossRef]  

13. H. F. Ghaemi, T. Thio, D. E. Grupp, T. W. Ebbesen, and H. J. Lezec, “Surface plasmons enhance optical transmission through subwavelength holes,” Phys. Rev. B 58(11), 6779–6782 (1998). [CrossRef]  

14. J. B. Pendry, L. Martín-Moreno, and F. J. Garcia-Vidal, “Mimicking surface plasmons with structured surfaces,” Science 305(5685), 847–848 (2004). [CrossRef]   [PubMed]  

15. L. Martín-Moreno and F. J. García-Vidal, “Minimal model for optical transmission through holey metal films,” J. Phys. Condens. Matter 20(30), 304214 (2008). [CrossRef]  

16. F. J. Garcia-Vidal, L. Martin-Moreno, T. W. Ebbesen, and L. Kuipers, “Light passing through subwavelength apertures,” Rev. Mod. Phys. 82(1), 729–787 (2010). [CrossRef]  

17. F. Medina, F. Mesa, and R. Marques, “Extraordinary transmission through arrays of electrically small holes from a circuit theory perspective,” IEEE Trans. Microw. Theory Tech. 56(12), 3108–3120 (2008). [CrossRef]  

18. D. Qu, D. Grischkowsky, and W. Zhang, “Terahertz transmission properties of thin, subwavelength metallic hole arrays,” Opt. Lett. 29(8), 896–898 (2004). [CrossRef]   [PubMed]  

19. H. Cao and A. Nahata, “Resonantly enhanced transmission of terahertz radiation through a periodic array of subwavelength apertures,” Opt. Express 12(6), 1004–1010 (2004). [CrossRef]   [PubMed]  

20. S. Pandey, S. Liu, B. Gupta, and A. Nahata, “Self-referenced measurements of the dielectric properties of metals using terahertz time-domain spectroscopy via the excitation of surface plasmon-polaritons,” Photon. Res. 1(4), 148–153 (2013). [CrossRef]  

21. R. Gordon, “Bethe’s aperture theory for arrays,” Phys. Rev. A 76(5), 053806 (2007). [CrossRef]  

22. N. J. Karl, R. W. Mckinney, Y. Monnai, R. Mendis, and D. M. Mittleman, “Frequency-division multiplexing in the terahertz range using a leaky-wave antenna,” Nat. Photonics 9(11), 717–720 (2015). [CrossRef]  

23. K. S. Reichel, R. Mendis, and D. M. Mittleman, “A broadband terahertz waveguide T-junction variable power splitter,” Sci. Rep. 6, 28925 (2016). [CrossRef]   [PubMed]  

24. M. Silveirinha and N. Engheta, “Tunneling of electromagnetic energy through subwavelength channels and bends using ε-near-zero materials,” Phys. Rev. Lett. 97(15), 157403 (2006). [CrossRef]   [PubMed]  

25. B. Edwards, A. Alù, M. E. Young, M. Silveirinha, and N. Engheta, “Experimental verification of epsilon-near-zero metamaterial coupling and energy squeezing using a microwave waveguide,” Phys. Rev. Lett. 100(3), 033903 (2008). [CrossRef]   [PubMed]  

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 Panel (a) shows a 3D view of the PPWG (WG1 and WG2) with an array of holes in the middle, where the waveguide plates are pictured translucent to see the screen, b is distance between the two plates, a is the distance between two holes, d is the diameter of the holes, and L is the length of the waveguide section. Panel (b) shows a side view of the experimental configuration where the THz wave is coupled into WG1 from the left and the screen with holes of thickness t is sandwiched in the middle between WG1 and WG2.
Fig. 2
Fig. 2 Power transmission at normal incidence for (top) a 2D array of holes in free-space illuminated by vertical polarization (black) or horizontal polarization (gray), (middle) a 1D array of holes inside a PPWG in TEM mode (blue), (bottom) a 1D array of holes inside a PPWG in TE1 mode (red). The dashed purple and oranges lines (overlapped) show prediction of RWA minimum and SPP maximum, respectively. This minimum value agrees well with 2D array in free-space and 1D array in the TEM PPWG. Clearly the resonant frequency is shifted for the 1D array in the TE1 PPWG.
Fig. 3
Fig. 3 Power transmission for a fixed hole separation of (a-c) a = 0.75 mm and (d) a = 1.00 mm for the plate separations of b = 0.75 mm (black, thin line), b = 1.00 mm (red), and b = 1.25 mm (blue, thick line) with TEM mode PPWG excitation. The dashed lines indicate the cutoff of the corresponding TM20 mode.
Fig. 4
Fig. 4 Power transmission for a fixed hole separation of (a-c) a = 0.75 mm and (d) a = 1.00 mm for the plate separations of b = 0.75 mm (black, thin line), b = 1.00 mm (red), and b = 1.25 mm (blue, thick line) with TE1 mode PPWG excitation. The dashed lines indicate the cutoff of the corresponding TM12 mode. For experiment and theory with a = 0.75 mm and b = 1.25 mm, we see many oscillations which can be attributed to the excitation of higher order propagating waveguide modes (i.e. cutoff of TE3 is 350GHz).

Equations (4)

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

f max,SPP =c ε d + ε m ε d ε m ( m L x ) 2 + ( n L y ) 2 ,
f min,RWA = c ε d ( m L x ) 2 + ( n L y ) 2 .
f min,TEM = f c T M 2p,2q = c 2π ε d ( mπ b ) 2 + ( nπ a ) 2 = c ε d ( p b ) 2 + ( q a ) 2 ,
f min,T E 1 = f c T M 2p+1,2q = c 2π ε d ( mπ b ) 2 + ( nπ a ) 2 = c ε d ( 2p+1 b ) 2 + ( q a ) 2 ,
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.