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

Free-surface photopolymerizable recording material for volume holography

Open Access Open Access

Abstract

We present a new organic photosensitive material for volume holographic recording. Advanced material composition and sample preparation eliminates the need for a protective layer and allows layer fabrication with variable thickness and a free surface. Optimized chemical formulation results in a high energetic sensitivity, high angular selectivity and high inducible refractive index contrast. We investigate the photoresponse and nonsinusoidal refractive index profiles. We demonstrate highly resolved optical structuring with up to 8000 lines per mm. Imaging of the holographic phase gratings is accomplished by optical microscopy.

©2013 Optical Society of America

1. Introduction

Photopolymers are highly attractive materials for numerous current and future applications. Based on holographically fabricated structures with a periodic modulation of the refractive index, these include three-dimensional photonic crystals [1,2], optical data storage technologies [3], diffractive and holographic optical elements (DOEs and HOEs) [4], devices in optical networks, such as (De-) Multiplexers [5] or narrow-spectral bandwidth filters [6] and holographic interferometry [7].

The scope of applications for photosensitive polymers would significantly expand with the possibility of combination and integration with diverse materials and substrates. The properties of particular interest for many emerging applications concern layer formatting and sandwiching in multilayer configurations. Free-surface samples that can be adhered to a wide range of substrates allow novel integrated optics by multiple-interface options. In addition, relatively thick layers are required to ensure high performance in terms of diffraction efficiency, angular selectivity, fill factor, and storage capacity. In order to fulfill those requirements, we develop a new material based on epoxy guest-host systems. With similar systems we have demonstrated the generation of single-mode waveguides by UV-mask exposure fabrication [8]. The material composition is also suitable for the incorporation of self-organized materials, such as nanoparticles. This provides the potential for improved systems with additional functionality [911] or for polymeric templates [12].

SU8 is well known for lithography applications [13]. Cost- and time-efficient, it is commonly used for the fabrication of micro- and nano-electromechanical systems (MEMS/NEMS) [14]. It forms layers up to 2 mm thick and enables high resolution and high aspect ratio. With good adhesion on silicon and glass, high chemical resistance and biological compatibility, it shows great promise as a basis for advanced integrated systems [15].

However, with no diffusing component SU8 is not qualified for use in volume holography. By doping with epoxy monomers, we combine the advantages of SU8 with the advantages of a diffusing guest-host system. This enables all-optical processing and avoids the need of post-exposure chemical treatment. Furthermore, the optical properties can be matched to specific requirements [16]. However, the capability of three-dimensional holographic structuring remains challenging. In addition, new material compositions for highly functional components should offer high photosensitivity, sufficient inducible local refractive index change, good dimensional stability and high resolution.

While commonly photopolymer samples and films are fixed in layer thickness and covered with glass plates [17], we apply a spin-coating fabrication technique. This allows a high flexibility in substrate choice, layer thickness and surface design, and opens diverse possibilities for new applications, especially for integrated optics. Among those are multilayer devices and holographic stacks, as for example proposed by [18]. Stacks may consist of diverse photonic materials. Multiple layers may be fabricated in separate processing steps. With functionalities such as wide bandwidth, multilayer configurations might complement multiplexing techniques.

This paper is structured as follows. Section 2 presents details on the material composition and sample preparation. Methods used for holographic recording and optical analysis are described in section 3. This includes the investigation of the material’s dimensional stability as well as the influence of the refractive index contrast and the layer thickness on the maximum diffraction efficiency and on the angular response. Results on the material photoresponse, energetic sensitivity, angular selectivity and resolution are presented in section 4. Finally, we discuss the appearance of nonsinusoidal refractive index profiles as well as the influence of the exposure duration.

2. Material composition and sample preparation

Synthesizing a volume holographic recording material with the attribute of a free surface imposes several requirements on the material composition. Three-dimensional holographic patterning in stable uncovered layers entails antithetic exigencies. A diffusing guest-host system enables a light-induced refractive index modulation. At the same time, the material system must be able to form volumetrically stable and homogeneous layers. Furthermore, the host and guest components are to differ in refractive index as much as possible, nonetheless being chemically compatible with each other. The molecular weight of the guest components must be small enough to allow diffusion, but high enough to form layers. Finally, the reproducibility in chemical formulation and sample preparation has to be ensured.

We meet the requirements with a material composition based on an epoxy oligomer. Both the host and guest molecules have epoxy functional groups. The corresponding mechanism of polymerization is a cationic ring-opening polymerization (CROP). The oligomer host-system features an aromatic bond of the epoxy group. In case of the monomer guest, the epoxy group is bound aliphaticly. The refractive indices of the host and guest components at the sodium-d-line are nhost>1.58 and nguest1.46, respectively. Guest and host components are compatible with each other in all proportions. A sensitized photo-acid-generator (PAG) is used, causing crosslinking by cationic polymerization at 405 nm [15].

Figure 1 shows the wavelength-dependent attenuation of the material in solution. The photopolymer samples feature excellent transmission in the visible spectrum.

 figure: Fig. 1

Fig. 1 The spectral optical attenuation of the material solution features excellent transmission in the visible spectrum.

Download Full Size | PDF

To fabricate free-surface samples without a protective layer but with sufficient dimensional stability, we limit the dope concentration. As a result, we receive tack-free coatings up to a guest content of 17% by weight. The viscosity for spin coating (24.000 m Pa s) is adjusted by dilution with solvent. In contrast to other systems [19], the solvent does not function as diffusing component. This allows us to prepare samples in form of dry film layers. Rotation speed of 500 - 4000 min-1enables the creation of defect-free layers with thickness ranging from 50 µm up to more than 250 µm. We identify the glass transition temperature to be TG=25°C. This ensures sufficient mobility for monomer diffusion while crosslinking reduces the mobility of the matrix. Fixation by UV flood exposure (dose 350 mJ/cm2) and hard bake up to 140°C results in TG=130°C.

3. Optical structuring and characterization techniques

Optical techniques we use to investigate the fundamental properties and behavior of the developed photopolymer system are primarily based on the recording and analysis of one-dimensional, plane-wave volume holograms. Both transmission- and reflection-type holographic gratings are created and detected for the purpose of optical characterization. Volumetric hologram recording is done by two freely propagating beams within a holographic exposure setup. The light source is a linearly s-polarized 405 nm diode laser. Exposure time is controlled by an electromechanical shutter. Laser power is adjusted between 0.4 and 1 mW per beam, while the beam diameter is 2 mm. The exposure time varies from 1 to 20 seconds. After exposure and a setting time of about 20 minutes the samples are fixed by UV flood cure.

Recorded holographic gratings are analyzed by observing the diffracted light in a rotation-scan setup, as for example in [20]. The transmitted signal of a 543 nm HeNe laser is detected while the hologram under test is rotated. Angular resolution is below 0.01 degrees. The probe beam used for detection is significantly smaller in diameter than the recording beam. Using smaller probe beams, for example, with a diameter of 0.2 mm, entail less signal variation and deliver more precise results [21].

From the angular resolved transmission we derive the efficiencies of particular diffraction orders. The angular response includes information about the grating period via

Λ(λ,θ)=λ2nsinθ.
We determine volume shrinkage in the direction of the grating vector, comparing the grating period Λ(λp,θp), derived from the rotation scan, with the period of the exposure interference pattern Λ(λexp,θexp). Herein λexp and λp are the exposure and probe beam wavelength, respectively. θexp and θp are the differences of the Bragg angles, given by the exposure geometry and derived from the rotation scan, respectively. The volumetric shrinkage derived from the angular response is 2% in average. This ensures good dimensional stability [21].

By comparison of the angular resolved diffraction efficiency with a rigorous solution of the coupled wave theory (RCWT) [22], we derive the layer thickness d and the refractive index contrastΔn. The diffraction efficiency η for a probe wavelength λ and the probe angle ϑp is given by:

η=sin2(πdΔn2λcos(ϑp)).
Figure 2 shows the impact of layer thickness and refractive index contrast on the angular response. Both gratings displayed feature the same index contrast Δn=3.6103. A variation in layer thickness of only 20% has a strong effect on the angular selectivity. Within the possibilities of material preparation to set arbitrary layer thickness from 50 to 250 μm, this enables the adjustment of the angular response over a wide range.

 figure: Fig. 2

Fig. 2 Interdependency of index modulation and layer thickness: measurement (dots) and RCW theory (line). Within constant index contrast a variation in layer thickness has a strong impact on angular response.

Download Full Size | PDF

4. Material response and performance

The photoresponse of the developed material to exposure fluence is widely linear. As shown in Fig. 3 , there is a rapid increase in the material response with increasing energy density. The response curve corresponds to an exposure time of 15 seconds.

 figure: Fig. 3

Fig. 3 Linear material response to deposited energy shows a rapid increase of the refractive index contrast with increasing energy density.

Download Full Size | PDF

In the case of cationic initiators, it is expected from theory to find the induced index change to be proportional to the absorbed energy [23]. Exposure duration should not have an impact on the degree of polymerization. Figure 4 shows the refractive index contrast versus exposure time for various exposure fluences. We observe a weak dependence on exposure time at low energetic fluence, whereas there is a stronger dependence at higher energetic fluence.

 figure: Fig. 4

Fig. 4 Refractive index contrast versus exposure time for various exposure fluences shows a constant material response at low energetic fluence.

Download Full Size | PDF

In contrast, the energetic sensitivity S shows an influence of the exposure duration. S is defined as the ratio of the refractive index contrast to deposited energy It:

S=ΔnIt.
Progressions for various exposure fluences are shown in Fig. 5 . We derived optimum exposure duration that is independent on the energetic fluence. The optimum exposure time topt is in the order of 10 seconds. The material shows high photosensitivity as compared with other epoxy-resin photopolymers [24]. The maximum energetic sensitivity, read from Fig. 5, is Smax = S(topt) = 11 cm2/J.

 figure: Fig. 5

Fig. 5 Energetic sensitivity versus exposure time for various exposure fluence shows: Optimal exposure time is independent on the deposited energy.

Download Full Size | PDF

Figure 5 also reveals a weak dependence of the sensitivity on exposure intensity I. For short exposure times, the maximal sensitivity is achieved at low intensity. This is similar to characteristics of nanocomposite materials [9]. In the case of longer exposure duration, the behavior is reversed; the sensitivity increases with exposure intensity.

The angularly resolved diffraction efficiency of the first diffraction order is displayed in Fig. 6 . The measured diffracted efficiency is obviously in good agreement with the rigorous coupled wave theory. We reach diffraction efficiency close to 100% with a layer thickness of 200 μm. Angular selectivity better than 1 degree is obtained at a layer thickness of about 250 μm. Among other applications, sharp angular selectivity is very important for multiplexing [25] and all related applications.

 figure: Fig. 6

Fig. 6 Angular response shows high selectivity: measurement (dots) and RCW theory (line). Exposure time was 6s, fluence 260 mJ/cm2. Layer thickness is 280 μm, grating period Λ = 1.7 μm and index contrastΔn=1.8103.

Download Full Size | PDF

Higher diffraction orders appear in case of overmodulation, but may also be caused by saturation effects [26]. This can be described by the parameter:

R=|K|2DF,
where D is the diffusion coefficient, F half the maximum polymerization rate and K the grating vector. A small R parameter results in deviations from the sinusoidal modulation of the refractive index [16]. This may be caused by high polymerization rate, slow diffusion and/or big grating periods. Considering higher orders of the refractive index contrast, those effects can be included in coupled wave theory [27].

Figure 7 shows the transmitted light in a corresponding measurement and the derived index profile. In this case, the saturation effects are caused by very long dark storage times, resulting in high viscosity and consequently slow diffusion rates. By comparison with RCWT simulations, we derive the first, second, and third order refractive index contrast to be: Δn(0)=4.5103, Δn(1)=1.5103 and Δn(2)=0.5103. The corresponding index profile over the lateral position x, displayed in Fig. 7, is calculated via

 figure: Fig. 7

Fig. 7 Saturation effects. Left: angular resolved measurement (dots) and RCW theory (line). Exposure time was 4.5 s, fluence 700 mJ/cm2. Layer thickness is 140μm, grating period Λ = 2.7μm and index contrast Δn(0)=4.5103, Δn(1)=1.5103 and Δn(2)=0.5103. Right: corresponding nonsinusoidal refractive index profile (line) in comparison with sinusoidal interference pattern (dots).

Download Full Size | PDF

n(x)=n0+Δn(0)cos(|K|x)+Δn(1)cos(2|K|x)+Δn(2)cos(3|K|x).

The material resolution enables three-dimensional periodic structuring with up to 8000 lines per mm. However, it is expected to find an optimum spatial frequency response between low spatial frequency roll-off and high spatial frequency cut-off [28]. We achieved the optimum material response with grating periods in the micrometer range, resulting in about 500 lines per mm. Nonetheless, the material is designed to record a wide range of pattern spacings, enabling both transmission and reflection holograms [29]. Figure 8 shows light-microscopic images of a transmission grating (left) and reflection grating (right).

 figure: Fig. 8

Fig. 8 Microscopic imaging: transmission grating (TG, left) and reflection grating (RG, right) in cross-sectional view. The grating periods are Λ = 1.3 μm (TG) and Λ = 130 nm (RG), respectively.

Download Full Size | PDF

Imaging of holographic gratings entails high standards on image contrast and resolution. Phase gratings do not feature surface modulations to be mapped by scanning probe microscopy and without conductive species [11] they give no contrast in electron microscopy. With the resolution limit of conventional optical microscopy imaging of reflection gratings seems impossible. To overcome this we implemented imaging in sectional view. The reflection grating planes are nearly parallel to the image plane. The focus plane determines the sectional plane. Altering the adjustment of the focus plane moves the stripes along the sectional-grating vector. The slant angle of the grating planes against the focus plane φ dictates the sectional-grating period via

Λs=Λsinφ.
Λs = 30 μm was read from the image (Fig. 8, right). This corresponds to Λ = 130 nm and a slant angle of φ = 0.25 deg.

5. Conclusions

With epoxy based material composition and limited dope proportions we have developed and optimized a novel photopolymer system for holographic recording, which does not require a protective layer. It exhibits high refractive index contrast (in the order of 5103) at exposure energy densities of about 0.5 J/cm2 with good dimensional stability. Diffraction efficiency close to 100% is achieved with 200 μm thick layers. Sharp angular selectivity better than 1 degree is obtained. We demonstrate holographic recording of both transmission and reflection gratings with up to 8000 lines per mm. All-optical processing in combination with additional freedom degrees in functionality makes this material system interesting for a wide range of applications, particularly for integrated optical and optoelectronic components. We now have at our disposal a highly sensitive material for volume holographic recording and three-dimensional optical structuring with the advantage of adhesion on various substrates and widely variable layer thickness. The realization of volumetrically stable, free-surface samples allows the integration with numerous photonic materials in diverse multilayer configurations.

Acknowledgments

We acknowledge support by the Helmholtz Research School on Security Technologies and by the Cobios Foundation. Microscopic recordings are provided by the Zentraleinrichtung Elektronenmikroskopie (ZELMI). T. S. thanks Michael Zschocher for fundamental discussions and assistance.

References and links

1. M. Campbell, D. N. Sharp, M. T. Harrison, R. G. Denning, and A. J. Turberfield, “Fabrication of photonic crystals for the visible spectrum by holographic lithography,” Nature 404(6773), 53–56 (2000). [CrossRef]   [PubMed]  

2. S. Orlic, C. Müller, and A. Schlösser, “All optical fabrication of three-dimensional photonic crystals in photopolymers by multiplex-exposure holographic recording,” Appl. Phys. Lett. 99(13), 131105 (2011). [CrossRef]  

3. S. Orlic, E. Dietz, T. Feid, S. Frohmann, and C. Müller, “Optical investigation of photopolymer systems for microholographic storage,” J. Opt. A, Pure Appl. Opt. 11(2), 024014 (2009). [CrossRef]  

4. A. V. Lukin, “Holographic optical elements,” J. Opt. Technol. 74(1), 65–70 (2007). [CrossRef]  

5. T. Kim, S. Chung, S. Han, and B. Lee, “Photopolymer-based demultiplexers with superposed holographic gratings,” IEEE Photon. Technol. Lett. 17(3), 618–620 (2005). [CrossRef]  

6. R. T. Ingwall and D. A. Waldman, “Holographic filter with a wide angular field of view and a narrow spectral bandwidth,” U.S. patent 7,480,084-B2 (20 January 2009).

7. M. Květoň, V. Ledl, A. Havranek, and P. Fiala, “Photopolymer for optical holography and holographic interferometry,” Macromol. Symp. 295(1), 107–113 (2010). [CrossRef]  

8. H. Sieber, H.-J. Boehm, U. Hollenbach, J. Mohr, U. Ostrzinski, K. Pfeiffer, M. Szczurowski, and W. Urbanczyk, “Low-loss single mode waveguides in polymer,” Proc. SPIE 8431, 84311R, 84311R-10 (2012). [CrossRef]  

9. L. Criante, R. Castagna, F. Vita, D. E. Lucchetta, and F. Simoni, “Nanocomposite polymeric materials for high density optical storage,” J. Opt. A, Pure Appl. Opt. 11(2), 024011 (2009). [CrossRef]  

10. E. Hata, K. Mitsube, K. Momose, and Y. Tomita, “Holographic nanoparticle-polymer composites based on step-growth thiol-ene photopolymerization,” Opt. Mater. Express 1(2), 207–222 (2011). [CrossRef]  

11. T. N. Smirnova, L. M. Kokhtych, A. S. Kutsenko, O. V. Sakhno, and J. Stumpe, “The fabrication of periodic polymer/silver nanoparticle structures: in situ reduction of silver nanoparticles from precursor spatially distributed in polymer using holographic exposure,” Nanotechnology 20(40), 405301 (2009). [CrossRef]   [PubMed]  

12. L. de Sio, S. Ferjani, G. Strangi, C. Umeton, and R. Bartolino, “Universal soft matter template for photonic applications,” Soft Matter 7(8), 3739–3743 (2011). [CrossRef]  

13. A. Erdmann, T. Fühner, F. Shao, and P. Evanschitzky, “Lithography simulation: modeling techniques and selected applications,” Proc. SPIE 7390, 739002, 739002-17 (2009). [CrossRef]  

14. R. A. Singh, N. Satyanarayana, T. S. Kustandi, and S. K. Sinha, “Tribo-functionalizing Si and SU8 materials by surface modification for application in MEMS/NEMS actuator-based devices,” J. Phys. D Appl. Phys. 44(1), 015301 (2011). [CrossRef]  

15. V. J. Cadarso, K. Pfeiffer, U. Ostrzinski, J. B. Bureau, G. A. Racine, A. Voigt, G. Gruetzner, and J. Brugger, “Direct writing laser of high aspect ratio epoxy microstructures,” J. Micromech. Microeng. 21(1), 017003 (2011). [CrossRef]  

16. V. L. Colvin, R. G. Larson, A. L. Harris, and M. L. Schilling, “Quantitative model of volume hologram formation in photopolymers,” J. Appl. Phys. 81(9), 5913–5923 (1997). [CrossRef]  

17. D. A. Waldmann, R. T. Ingwall, P. K. Dahl, M. G. Horner, E. S. Kolb, H.-Y. S. Li, R. A. Minns, and H. G. Schild, “Cationic ring-opening photopolymerization methods for volume hologram recording,” Proc. SPIE 2689, 127–141 (1996). [CrossRef]  

18. K. Yu, K. S. Yin, and J. E. Wreede, “Multiple layer holograms,” U.S. patent 5,282,066 (25 January 1994).

19. M. Szczurowski, W. Urbanczyk, M. Napiorkowski, P. Hlubina, U. Hollenbach, H. Sieber, and J. Mohr, “Differential Rayleigh scattering method for measurement of polarization and intermodal beat length in optical waveguides and fibers,” Appl. Opt. 50(17), 2594–2600 (2011). [CrossRef]   [PubMed]  

20. C. Neipp, S. Gallego, M. Ortuno, A. Marquez, A. Belendez, and I. Pascual, “Characterization of a PVA/acrylamide photopolymer. Influence of a cross-linking monomer in the final characteristics of the hologram,” Opt. Commun. 224(1-3), 27–34 (2003). [CrossRef]  

21. G. J. Steckman and F. Havermeyer, “High spatial resolution measurement of volume holographic gratings,” Proc. SPIE 6136, 613602, 613602-9 (2006). [CrossRef]  

22. M. G. Moharam and T. K. Gaylord, “Rigorous coupled-wave analysis of planar-grating diffraction,” J. Opt. Soc. Am. 71(7), 811–818 (1981). [CrossRef]  

23. R. R. McLeod, A. J. Daiber, M. E. McDonald, T. L. Robertson, T. Slagle, S. L. Sochava, and L. Hesselink, “Microholographic multilayer optical disk data storage,” Appl. Opt. 44(16), 3197–3207 (2005). [CrossRef]   [PubMed]  

24. Y.-C. Jeong, S. Lee, and J.-K. Park, “Holographic diffraction gratings with enhanced sensitivity based on epoxy-resin photopolymers,” Opt. Express 15(4), 1497–1504 (2007). [CrossRef]   [PubMed]  

25. Y. H. Cho, C. W. Shin, N. Kim, B. K. Kim, and Y. Kawakami, “High-performance transmission holographic gratings via different polymerization rates of dipentaerythritol acrylates and siloxane-containing epoxides,” Chem. Mater. 17(25), 6263–6271 (2005). [CrossRef]  

26. C. Neipp, A. Beléndez, S. Gallego, M. Ortuño, I. Pascual, and J. Sheridan, “Angular responses of the first and second diffracted orders in transmission diffraction grating recorded on photopolymer material,” Opt. Express 11(16), 1835–1843 (2003). [CrossRef]   [PubMed]  

27. S. F. Su and T. K. Gaylord, “Calculation of arbitrary-order diffraction efficiencies of thick gratings with arbitrary grating shape,” J. Opt. Soc. Am. 65(1), 59–64 (1975). [CrossRef]  

28. M. R. Gleeson, J. Guo, and J. T. Sheridan, “Optimisation of photopolymers for holographic applications using the Non-local Photo-polymerization Driven Diffusion model,” Opt. Express 19(23), 22423–22436 (2011). [CrossRef]   [PubMed]  

29. T. J. Trout, J. J. Schmieg, W. J. Gambogi, and A. M. Weber, “Optical photopolymers: design and applications,” Adv. Mater. (Deerfield Beach Fla.) 10(15), 1219–1224 (1998). [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 (8)

Fig. 1
Fig. 1 The spectral optical attenuation of the material solution features excellent transmission in the visible spectrum.
Fig. 2
Fig. 2 Interdependency of index modulation and layer thickness: measurement (dots) and RCW theory (line). Within constant index contrast a variation in layer thickness has a strong impact on angular response.
Fig. 3
Fig. 3 Linear material response to deposited energy shows a rapid increase of the refractive index contrast with increasing energy density.
Fig. 4
Fig. 4 Refractive index contrast versus exposure time for various exposure fluences shows a constant material response at low energetic fluence.
Fig. 5
Fig. 5 Energetic sensitivity versus exposure time for various exposure fluence shows: Optimal exposure time is independent on the deposited energy.
Fig. 6
Fig. 6 Angular response shows high selectivity: measurement (dots) and RCW theory (line). Exposure time was 6s, fluence 260 mJ/cm2. Layer thickness is 280 μm, grating period Λ = 1.7 μm and index contrast Δn=1.8 10 3 .
Fig. 7
Fig. 7 Saturation effects. Left: angular resolved measurement (dots) and RCW theory (line). Exposure time was 4.5 s, fluence 700 mJ/cm2. Layer thickness is 140μm, grating period Λ = 2.7μm and index contrast Δ n ( 0 ) =4.5 10 3 , Δ n ( 1 ) =1.5 10 3 and Δ n ( 2 ) =0.5 10 3 . Right: corresponding nonsinusoidal refractive index profile (line) in comparison with sinusoidal interference pattern (dots).
Fig. 8
Fig. 8 Microscopic imaging: transmission grating (TG, left) and reflection grating (RG, right) in cross-sectional view. The grating periods are Λ = 1.3 μm (TG) and Λ = 130 nm (RG), respectively.

Equations (6)

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

Λ( λ,θ )= λ 2nsinθ .
η=si n 2 ( πdΔn 2λcos( ϑ p ) ).
S= Δn It .
R= | K | 2 D F ,
n(x)= n 0 +Δ n ( 0 ) cos( | K |x )+Δ n ( 1 ) cos( 2| K |x )+Δ n ( 2 ) cos( 3| K |x ).
Λ s = Λ sinφ .
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.