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

Metasurface holographic movie: a cinematographic approach

Open Access Open Access

Abstract

Animation for a metasurface hologram was achieved using a cinematographic approach. Time-lapsed images were reconstructed using sequentially arranged metasurface hologram frames. An Au rectangular nanoaperture was adopted as a meta-atom pixel and arrayed to reproduce the phase distribution based on the help of a Pancharatnam–Berry phase. We arrayed 48 hologram frames on a 2-cm2 substrate and measured and assessed the retardation of fabricated meta-atoms to reconstruct the holographic image, successfully demonstrating the movie with a frame rate of 30 frames per second.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Metasurface is a planar branch of metamaterials consisting of metallic or dielectric unit cells (meta-atoms) with a subwavelength dimension. It can change the amplitude, polarization, and/or phase of an incident electromagnetic wave. Moreover, it can be used to modify and tailor optical wavefronts into designed distribution [1], for example, the parabolic phase for focusing as a lens or complex diffraction pattern for metasurface holography. To date, various applications have been studied, including very thin metalenses [24], waveplates [57], and polarization converters [8].

The phase delay mechanisms of meta-atoms can be categorized into metallic and dielectric from the viewpoint of the material. Metallic metasurfaces have sufficient phase shift, reaching up to π rad, based on plasmon resonance with a very thin antenna pattern. However, their efficiency is limited [9] due to the intrinsic ohmic loss. The dielectric meta-atoms have superior efficiency owing to their lossless nature, and fabrication difficulty arises when achieving a relatively high aspect ratio.

Holography is a promising metasurface application, especially in the visible wavelength range. It can minimize visual fatigue compared with other three-dimensional (3D) imaging methods using the stereoscopic approach, for example, virtual reality glasses and 3D cinema [1012]. Metasurface holograms consisting of metallic or dielectric meta-atom arrangements are actively studied owing to their compactness and wide viewing angle [13,14] when reconstructing a holographic image. The viewing angle $\phi $ of a hologram is defined as:

$$\phi = 2\; \textrm{sin}^{{ - 1}}\left( {\frac{{\lambda }}{{{2p}}}} \right),$$
where λ denotes the wavelength of the incident light, and p denotes the period of the hologram pixel. Hence, by adopting smaller pixels, we can achieve a wider viewing angle [15], and when the period is smaller than the wavelength, the viewable range covers the whole hemisphere. Therefore, the subwavelength nature of metasurface is an advantage for holographic applications.

The holographic movie is one of the most expected frontiers. Several methods have been proposed to achieve a holographic movie based on spatial light modulator devices, for example, the acousto-optic modulator, liquid crystal display, liquid crystal on silicon, digital mirror device, and photoreactive materials [16]. However, the balance among the pixel period, the modulation depth of the phase or intensity, and the achievable frame rate is the key issue. Regarding the holographic movie, one of the expected solutions is the use of reconfigurable metasurfaces. Moreover, several methods have been reported, including polarization switching, chemical activation of meta-atoms, and strain multiplexing by stretchable substrates. In polarization switching, polarization-sensitive metasurfaces are used, and two holographic image frames can be switched by selecting one of the illuminating orthogonal polarization pairs (vertical/horizontal linear polarizations or right/left circular polarizations) [1720]. Chemical activation uses a reversible shift of plasmonic responses of pixel material involved in the molecular phase change between the metallic and dielectric state caused by gas exposure. This method enables the on/off control of the reconstructed image, although the response time is limited regarding the frame rate [21]. In strain multiplexing, holograms fabricated on a stretchable substrate are used, and three hologram frames can be selectively reconstructed with substrate strains of 0%, 12%, and 30% [22]. Thus, as the number of reconstructable frames is limited, metasurface holographic movies consisting of more than 4 frames have not yet been reported.

Another advantage of metasurface holography is the compactness of the hologram size, which is enabled by the smallness of meta-atoms at the given resolution. This enables both wide projection width and high resolution holography because the sampling size Ls and sampling pitch s of the Fourier plane are expressed by Ls = λf/p, and s = λf/D, respectively [23,24]. Here f is the focal length of the Fourier transform lens, and D = np is the whole size of the hologram, where n is the pixel number. This compactness also allows sequential playback with both substantial image resolution and frame rate. Therefore, in this research, we have adopted an alternative, cinematographic approach toward a holographic movie. The cinematograph is the first instrument used to record and project a movie [25] and to play back the image frames exposed on a film sequentially. Although this approach have also been studied using holographic film [26,27], image quality of this method is limited because subwavelength pixel pitch cannot be achieved due to optical exposure. Therefore, in this research, we have adopted cinematographic approach based on metasurface holography. A set of 48 metasurface hologram frames are fabricated on a 2-cm2 substrate, and the movie was played back by sequentially reconstructing each frame with a substantial frame rate of 30 frames per second (fps).

2. Experimental

2.1 Design

Figures 1(a) and (b) present the schematic illustration of a metasurface holographic movie, which is inspired by the sequential playback of cinematography. A He–Ne laser with a 633-nm wavelength was used as the light source. The 48 frames of a metasurface hologram were arranged on a substrate to form a sequential loop actuated by an external two-axis stage. Each frame consists of rectangular metallic aperture meta-atoms. The wide viewing angle, high resolution, and fast frame rate are essential requirements for a holographic movie. We adopted the pixel period p of 300 nm to achieve whole hemisphere coverage at 633-nm illumination. The resolution of each frame was determined as 2048 × 2048 pixels to achieve high image quality beyond the 1080 p full HD (1920 × 1080 pixels) and the applicability to the fast Fourier transform. Therefore, the size of each frame is 0.6 × 0.6 mm2. Because we used a Gaussian beam He–Ne laser with a full-width half-maximum beam diameter of 0.6 mm, the frame period on the glass substrate was calculated as 1 mm. Thus, the stage actuation speed of 30 mm/s is required to achieve 30 fps.

 figure: Fig. 1.

Fig. 1. (a) Schematic illustration of cinematography-inspired metasurface holographic movie. (b) Device design. 48 hologram frames are arranged on a substrate. Aperture array patterns for characterization with constant rotation angles θ to the x-axis of the substrate at 0°, 45°, 90°, and 135° were also fabricated. (c) Schematic of a metallic rectangular aperture meta-atom.

Download Full Size | PDF

We selected 48 images of the earth from different angles to express the holographic movie. All images were captured from a public domain video, which was downloaded from a website [28]. These images were converted into the phase distribution using a free software library of iterative Fourier transform algorithm (IFTA) (WFL ver. 3.5.0), which was developed by Matsushima [29]. We assumed the homogeneous intensity distribution of planar incident light and used a random phase distribution as the initial phase input.

An Au rectangular nanoaperture meta-atom with the length, width, and thickness of 200, 100, and 120 nm, respectively, was adopted (Fig. 1(c)) [30]. This meta-atom should produce a π phase shift between linear polarizations parallel and perpendicular to the length direction and work as a half-wave plate (HWP). To achieve full 0–2π phase coverage, we adopted the Pancharatnam–Berry phase (P-B phase) or geometric phase [30]. The output Jones vector Eout from an incident left circular polarization (LCP) Ein = (1, i) through an HWP with the rotation angle θ to the x-axis is expressed as:

$${\boldsymbol{E}_{\textrm{out}}} = \textbf{HWP}_{\theta} \;{\boldsymbol{E}_{\textrm{in}}} = \left( {\begin{array}{cc} {\textrm{cos} \;2{\theta }}&{\textrm{sin} \;2{\theta }}\\ {\textrm{sin}\; 2{\theta }}&{-\textrm{cos} \;2{\theta }} \end{array}} \right)\left( {\begin{array}{c} {1}\\ {i} \end{array}} \right) = {{e}^{{2i}{\theta }}}\left( {\begin{array}{c} {1}\\ {{- i}} \end{array}} \right),$$
where HWPθ denotes the Jones matrix of the HWP. As presented in this equation, the incident LCP was converted into the right circular polarization (RCP) with a P-B phase delay of 2θ. Therefore, a full 2π phase coverage can be achieved by rotating meta-atoms with the rotation angles between 0 and π. When a retarder with a phase delay of Δ and rotation angle θ is considered instead of HWP, Eout for the LCP incident can be expressed as:
$${\boldsymbol{E}_{\textrm{out}}} = \textbf{RET}_{{\varDelta}, {\theta}} \;{\boldsymbol{E}_{\textrm{in}}} = \textrm{cos}\frac{{\varDelta }}{{2}}{{e}^{{i} \frac{{\varDelta }}{{2}}}}\left( {\begin{array}{c} {1}\\ {i} \end{array}} \right) + \textrm{sin}\frac{{\varDelta }}{{2}}{{e}^{{i} \frac{{\varDelta }}{{2}}{ - i}\frac{{\pi }}{{2}}}}{{e}^{{2i}{\theta }}}\left( {\begin{array}{c} {1}\\ {{ - i}} \end{array}} \right){,}$$
where RETΔ,θ denotes the Jones matrix of the retarder. Note that only the RCP component exhibits rotation dependence, and the polarization intensity ratio $\eta $ of the hologram tends to $\eta = (\textrm{sin}\frac{\varDelta}{2}/\textrm{cos}\frac{\varDelta}{2})^{2} = \textrm{tan}^{2}\frac{\varDelta}{2}. $ Thus, according to the above IFTA-calculated phase distribution, the meta-atom pixels are mapped with each rotation angle and converted into GDSII CAD data using a Python package (Gdspy). The average data size per frame consisting of 2048 × 2048 pixels was approximately 200 MB. To measure Δ on the fabricated sample, a 0.6 × 0.6 mm2 aperture array pattern with constant rotation angles (θ = 0°, 45°, 90°, and 135°) was arranged at the upper left of the chip with the hologram frames (Fig. 1(b)).

2.2 Fabrication and characterization

Figure 2 presents a schematic of the fabrication process. At first, Cr, Au, and Si layers were deposited on a glass substrate by using a facing target sputtering apparatus (FTS Co., NFTS-3S-R0601). The Cr, Au, Si, and glass substrates were 2-nm, 120-nm, 20-nm, and 500-µm thick, respectively. The hologram pattern was then drawn by electron beam (EB) lithography (JEOL Co., JBX-6300SL). The drawing time of the sample (including 48 hologram frames, aperture array patterns, and global alignment marks) was 6.5 hours. We used a positive e-beam resist (ZEON Co., ZEP520A-7) with a thickness of 300 nm and performed oxygen plasma ashing after resist development for descam. Then, the resist pattern was transferred to the Si layer by using SF6 reactive ion etching (RIE), followed by Au layer milling by Ar plasma with an RIE apparatus (Tateyama Machine Co., TEP-Xd). Finally, the Si mask layer was removed by SF6 RIE again.

 figure: Fig. 2.

Fig. 2. Fabrication process for Au nanoaperture

Download Full Size | PDF

Figure 3 presents the SEM images of the fabricated hologram and aperture array patterns. Table 1 presents the average dimensions of these meta-atoms. The height H of the apertures is measured using a surface profiler (Bruker, Dektak XT) after Au deposition. The averaged dimensional error of fabricated apertures to the design was up to 7%.

 figure: Fig. 3.

Fig. 3. Scanning electron microscopy (SEM) images of fabricated sample. (a) A hologram frame and aperture array patterns for retardation measurement with the fast axis angles θ of (b) 0°, (c) 45°, (d) 90°, and (e) 135°.

Download Full Size | PDF

Tables Icon

Table 1. Dimensional parameters of apertures measured from a hologram frame and aperture array patterns [nm]

3. Results and discussion

3.1 Optical characterization

We used the spectroscopic rotating analyzer method (RAM) to characterize the optical performances of fabricated metasurface [7]. Figure 4(a) presents the measurement setup constructed on polarization microscopy. The aperture array pattern sample was set on the stage of microscopy, of which the fast axis was fixed at an orientation of φ = 45° at each pattern by rotating the sample substrate. The sample was sandwiched between the polarizer and analyzer at orientations of 90° and ψ (0°–180°), respectively. The light transmitted through the sample was expanded and guided to the USB spectrometer (QE65000, Ocean Optics) through an objective lens (TU Plan Fluor 10×/0.3A Pol) and an optical fiber (P50-2-IV-VIS) with a 50-µm core diameter.

 figure: Fig. 4.

Fig. 4. (a) The optical system for the rotating analyzer method. Aperture array patterns were measured. Each pattern was directed to φ = 45° by rotating the sample substrate. (b) Transmittance spectra of aperture array patterns with the analyzer angle ψ of 135° and the fast axis angles θ of 0°, 45°, 90°, and 135°. (c) Angular plot of measured intensity (counts) of each array pattern rotated toward the φ = 45° direction at the incident wavelength of 633 nm with 90° incident polarization. (d) Derived retardation spectra of aperture array patterns with the fast axis angles of θ = 0°, 45°, 90°, and 135°. Dashed lines in (b) and (d) shows corresponding electromagnetic simulation results.

Download Full Size | PDF

The Stokes vector S of the transmitted light is expressed by Eq. (4):

$$\textbf{S} = \textbf{LP}_{\psi } \cdot {\textbf{X}}_{45^{\circ}} \cdot \textbf{LP}_{90^{\circ}} \cdot {\textbf{S}}_{\textrm{unp}},$$
where $\textbf{LP}$, $\textbf{X}$, and ${\textbf{S}}_{\textrm{unp}}$ denote the Mueller matrices of a linear polarizer, the aperture array pattern sample, and the Stokes vector of unpolarized light source, respectively. $\textbf{X}$ was modeled as $\textbf{X} = \textbf{PP} \cdot \textbf{RET}$, where $\textbf{PP}$ and $\textbf{RET}$ denote partial polarizer and retarder. The intensity of the transmitted light S0 is given by:
$${S}_{0} = \frac{{{{p}_{1}}^{2}{ + }{{p}_{2}}^{2} - {2}{{p}_{1}}{{p}_{2}}\;\textrm{cos}\;\varDelta \;\textrm{cos} \;2{\psi }+ ({{p}_{1}}^{2} - {{p}_{2}}^{2})\textrm{sin} \;2{\psi}}}{4},$$
where ${p_1}$ and ${p_2}$ denote the amplitude transmittances of the partial polarizer for the electric field perpendicular and parallel to the aperture fast axis, respectively. Δ denotes the retardation of the retarder.

The transmittance spectra of the aperture array patterns were measured by rotating the analyzer angle ψ with a glass substrate reference. The retardation Δ is then derived by parameter fitting to the transmittance spectra using Eq. (5).

Figure 4(b) presents the transmittance spectra. Figure 4(c) presents the angular plot of the transmittance at a wavelength of 633 nm. Figure 4(d) presents retardation spectra. Dashed lines in Figs. 4(b) and 4(d) represents electromagnetic simulation results calculated with a commercial finite element method software COMSOL Multiphysics 5.1 (COMSOL Inc., USA), with the measured aperture dimensions summarized in Table 1. Simulation results agree well to the measurements, although some peak shift in transmittance spectra and peak broadening in retardation spectra were observed. These shifts are considered to be due to fabrication error including cross-sectional tapering and connection of adjacent apertures. From Fig. 4(c), we calculated the ellipticities of the transmitted light for aperture array patterns of θ = 0°, 45°, 90°, and 135° as 0.28, 0.54, 0.36, and 0.55, respectively. The averaged transmittance of aperture array pattern samples is approximately 5%, and the retardation performance peaks were observed at the wavelength range from 630 to 700 nm, which correlates well with the previous research [30].

3.2 Projection of holographic movie

Figure 5 presents the experimental setup for the holographic movie projection. The fabricated substrate was fixed on a two-axis stage (Chuo Seiki, MMU-60X) and illuminated by left circular-polarized light. A He–Ne laser (λ = 633 nm) was used with a linear polarizer and a quarter-wave plate. A PC can control this stage and can be moved in the x-y plane with a maximum speed of 30 mm/s. The beam diameter was adjusted by an iris with a diameter of 1.1 mm to eliminate stray light. Then, each hologram frame is sequentially reconstructed on white cardboard as a screen, and a holographic movie was presented. Figure 6 presents clips (frame numbers of #10, #21, #35, and #45) of a movie frames taken using iPhone 6s Plus. The movie is also available online (see Visualization 1). The image angle of the globe was 62.6°. As the reconstructed image was taken at an angle as shown in Fig. 5, the appears to shrink horizontally.

 figure: Fig. 5.

Fig. 5. Experimental setup for projection

Download Full Size | PDF

 figure: Fig. 6.

Fig. 6. Captured images of reconstructed hologram, with the frame numbers of (a) #10, (b) #21, (c) #35, and (d) #45.

Download Full Size | PDF

Considering the maximum stage speed of 30 mm/s and frame pitch of 1 mm, we achieved a frame rate of 30 fps, although some frame lagging can be observed in the corner of the frame arrangement. This may be improved by optimizing the stage load or adopting rotational operation. Each reconstructed frame has enough resolution to recognize the earth (Fig. 6), and the 0th diffraction spot can be observed at the center of the reconstructed images. We can consider this spot as the non-modulated LCP component, whereas diffracted images correspond to the RCP component. Using a laser power meter, the intensities of LCP and RCP components were measured as 0.0748 and 0.2442 mW, respectively. From these values and the tan2Δ/2 tendency as described above, the average retardation Δ of frame #1 was estimated at 122°. Retardations of aperture array patterns were measured in this way and summarized in Table 2 together with the RAM results, and conversion efficiency (defined as RCP/(LCP + RCP)×100% [31]). The measured retardation and conversion efficiency of the hologram frame were the minimum, respectively. This is considered to be due to the connections between adjacent apertures as shown in Fig. 3, and the efficiency may be improved by optimizing fabrication parameters to avoid tapering and connection.

Tables Icon

Table 2. Summary of measured retardation Δ [deg] and conversion efficiency [%]

Contrary to the needs of the synchronization between sliding film frames and a blinking light in the traditional cinematograph, this holographic movie does not require light source blinking. Figure 7 presents the intermediate image when the light shines on the border between frames #3 and #4. Note that the two frames are superimposed and can be observed at the same time. Thanks to this characteristic, the holographic movie can be reconstructed by using a continuous wave light source.

 figure: Fig. 7.

Fig. 7. The intermediately reconstructed image with illumination between the frames #3 and #4.

Download Full Size | PDF

On this occasion we have drawn 48 frames on a 2-cm-square chip. We believe this chip size is enough for a very short and looped movie. Considering the current drawing throughput of 6.5 hours, if a whole 4-inch wafer was used, approximately 10,000 frames (corresponding to a movie-playing time of 333 seconds with 30 fps) can be drawn within 833 hours. This throughput might be improved by using character-projection type EB lithography apparatus or stepper photolithography.

4. Summary

We have demonstrated a metasurface holographic movie based on a cinematographic approach. We used a metallic rectangular aperture meta-atom with an average retardation of 122° to express 2048 × 2048 pixels hologram frames, and a movie consisting of 48 frames was successfully reconstructed with a frame rate of 30 fps.

Funding

Japan Society for the Promotion of Science (17H02754).

Acknowledgments

The authors wish to thank Prof. K. Matsushima at Kansai University for the development of free software library for IFTA (WFL ver. 3.5.0).

Disclosures

The authors declare no conflicts of interest.

References

1. H.-T. Chen, A. J. Taylor, and N. Yu, “A review of metasurfaces: physics and applications,” Rep. Prog. Phys. 79(7), 076401 (2016). [CrossRef]  

2. F. Aieta, P. Genevet, M. A. Kats, N. Yu, R. Blanchard, Z. Gaburro, and F. Capasso, “Aberration-Free Ultrathin Flat Lenses and Axicons at Telecom Wavelengths Based on Plasmonic Metasurfaces,” Nano Lett. 12(9), 4932–4936 (2012). [CrossRef]  

3. X. Ni, S. Ishii, A. V. Kildishev, and V. M. Shalaev, “Ultra-thin, planar, Babinet-inverted plasmonic metalenses,” Light: Sci. Appl. 2(4), e72 (2013). [CrossRef]  

4. B. Memarzadeh and H. Mosallaei, “Array of planar plasmonic scatterers functioning as light concentrator,” Opt. Lett. 36(13), 2569–2571 (2011). [CrossRef]  

5. N. Yu, F. Aieta, P. Genevet, M. A. Kats, Z. Gaburro, and F. Capasso, “A Broadband, Background-Free Quarter-Wave Plate Based on Plasmonic Metasurfaces,” Nano Lett. 12(12), 6328–6333 (2012). [CrossRef]  

6. M. Ishii, K. Iwami, and N. Umeda, “An Au nanofin array for high efficiency plasmonic optical retarders at visible wavelengths,” Appl. Phys. Lett. 106(2), 021115 (2015). [CrossRef]  

7. M. Ishii, K. Iwami, and N. Umeda, “Highly-efficient and angle-independent zero-order half waveplate at broad visible wavelength based on Au nanofin array embedded in dielectric,” Opt. Express 24(8), 7966–7976 (2016). [CrossRef]  

8. K. Iwami, M. Ishii, Y. Kuramochi, K. Ida, and N. Umeda, “Ultrasmall radial polarizer array based on patterned plasmonic nanoslits,” Appl. Phys. Lett. 101(16), 161119 (2012). [CrossRef]  

9. G. Lee, J. Sung, and B. Lee, “Recent advances in metasurface hologram technologies (Invited paper),” ETRI J. 41(1), 10–22 (2019). [CrossRef]  

10. D. M. Hoffman, A. R. Girshick, and M. S. Banks, “Vergence – accommodation conflicts hinder visual performance and cause visual fatigue,” J. Vis. 8(3), 33 (2008). [CrossRef]  

11. J. A. Aznar-Casanova, A. Romeo, A. T. Gómez, and P. M. Enrile, “Visual fatigue while watching 3D stimuli from different positions,” J. Optom. 10(3), 149–160 (2017). [CrossRef]  

12. P. V. Johnson, J. Kim, and M. S. Banks, “Stereoscopic 3D display technique using spatiotemporal interlacing has improved spatial and temporal properties,” Opt. Express 23(7), 9252–9275 (2015). [CrossRef]  

13. G.-Y. Lee, G. Yoon, S.-Y. Lee, H. Yun, J. Cho, K. Lee, H. Kim, J. Rho, and B. Lee, “Complete amplitude and phase control of light using broadband holographic metasurfaces,” Nanoscale 10(9), 4237–4245 (2018). [CrossRef]  

14. G. Zheng, H. Mühlenbernd, M. Kenney, G. Li, T. Zentgraf, and S. Zhang, “Metasurface holograms reaching 80% efficiency,” Nat. Nanotechnol. 10(4), 308–312 (2015). [CrossRef]  

15. Q. Jiang, G. Jin, and L. Cao, “When metasurface meets hologram: principle and advances,” Adv. Opt. Photonics 11(3), 518–576 (2019). [CrossRef]  

16. Y. Pan, J. Liu, X. Li, and Y. Wang, “A Review of Dynamic Holographic Three-Dimensional Display: Algorithms, Devices, and Systems,” IEEE Trans. Ind. Inf. 12(4), 1599–1610 (2016). [CrossRef]  

17. A. Arbabi, Y. Horie, M. Bagheri, and A. Faraon, “Dielectric metasurfaces for complete control of phase and polarization with subwavelength spatial resolution and high transmission,” Nat. Nanotechnol. 10(11), 937–943 (2015). [CrossRef]  

18. J. P. Balthasar Mueller, N. A. Rubin, R. C. Devlin, B. Groever, and F. Capasso, “Metasurface Polarization Optics: Independent Phase Control of Arbitrary Orthogonal States of Polarization,” Phys. Rev. Lett. 118(11), 113901 (2017). [CrossRef]  

19. J. Guo, T. Wang, B. Quan, H. Zhao, C. Gu, J. Li, X. Wang, G. Situ, and Y. Zhang, “Polarization multiplexing for double images display,” Opto-Electron. Adv. 2(7), 18002901–18002906 (2019). [CrossRef]  

20. F. Zhang, M. Pu, X. Li, P. Gao, X. Ma, J. Luo, H. Yu, and X. Luo, “All-Dielectric Metasurfaces for Simultaneous Giant Circular Asymmetric Transmission and Wavefront Shaping Based on Asymmetric Photonic Spin-Orbit Interactions,” Adv. Funct. Mater. 27(47), 1704295 (2017). [CrossRef]  

21. J. Li, S. Kamin, G. Zheng, F. Neubrech, S. Zhang, and N. Liu, “Addressable metasurfaces for dynamic holography and optical information encryption,” Sci. Adv. 4(6), eaar6768 (2018). [CrossRef]  

22. S. C. Malek, H.-S. Ee, and R. Agarwal, “Strain Multiplexed Metasurface Holograms on a Stretchable Substrate,” Nano Lett. 17(6), 3641–3645 (2017). [CrossRef]  

23. W. Qu, H. Gu, and Q. Tan, “Holographic projection with higher image quality,” Opt. Express 24(17), 19179–19184 (2016). [CrossRef]  

24. D. Mas, J. Garcia, C. Ferreira, L. M. Bernardo, and F. Marinho, “Fast algorithms for free-space diffraction patterns calculation,” Opt. Commun. 164(4-6), 233–245 (1999). [CrossRef]  

25. L. Lumiere, “The Lumiègre Cinematograph,” J. Soc. Motion Pict. Eng. 27(6), 640–647 (1936). [CrossRef]  

26. M. Yamaguchi, H. Sugiura, T. Honda, and N. Ohyama, “Automatic recording method for holographic three-dimensional animation,” J. Opt. Soc. Am. A 9(7), 1200–1205 (1992). [CrossRef]  

27. M. Yamaguchi, T. Koyama, H. Endoh, N. Ohyama, S. Takahashi, and F. Iwata, “Development of a prototype full-parallax holoprinter,” Proc. SPIE 2406, 50–56 (1995). [CrossRef]  

28. Videvo, “Globe Rotate Matte Loop,” https://www.videvo.net/video/globe-rotate-matte-loop/48/.

29. K. Matsushima, “WaveField Tools,” http://www.laser.ee.kansai-u.ac.jp/WaveFieldTools/.

30. X. Zhang, J. Jin, Y. Wang, M. Pu, X. Li, Z. Zhao, P. Gao, C. Wang, and X. Luo, “Metasurface-based broadband hologram with high tolerance to fabrication errors,” Sci. Rep. 6(1), 19856 (2016). [CrossRef]  

31. B. Wang, F. Dong, Q.-T. Li, D. Yang, C. Sun, J. Chen, Z. Song, L. Xu, W. Chu, Y.-F. Xiao, Q. Gong, and Y. Li, “Visible-Frequency Dielectric Metasurfaces for Multiwavelength Achromatic and Highly Dispersive Holograms,” Nano Lett. 16(8), 5235–5240 (2016). [CrossRef]  

Supplementary Material (1)

NameDescription
Visualization 1       Metasurface holographic movie

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 (7)

Fig. 1.
Fig. 1. (a) Schematic illustration of cinematography-inspired metasurface holographic movie. (b) Device design. 48 hologram frames are arranged on a substrate. Aperture array patterns for characterization with constant rotation angles θ to the x-axis of the substrate at 0°, 45°, 90°, and 135° were also fabricated. (c) Schematic of a metallic rectangular aperture meta-atom.
Fig. 2.
Fig. 2. Fabrication process for Au nanoaperture
Fig. 3.
Fig. 3. Scanning electron microscopy (SEM) images of fabricated sample. (a) A hologram frame and aperture array patterns for retardation measurement with the fast axis angles θ of (b) 0°, (c) 45°, (d) 90°, and (e) 135°.
Fig. 4.
Fig. 4. (a) The optical system for the rotating analyzer method. Aperture array patterns were measured. Each pattern was directed to φ = 45° by rotating the sample substrate. (b) Transmittance spectra of aperture array patterns with the analyzer angle ψ of 135° and the fast axis angles θ of 0°, 45°, 90°, and 135°. (c) Angular plot of measured intensity (counts) of each array pattern rotated toward the φ = 45° direction at the incident wavelength of 633 nm with 90° incident polarization. (d) Derived retardation spectra of aperture array patterns with the fast axis angles of θ = 0°, 45°, 90°, and 135°. Dashed lines in (b) and (d) shows corresponding electromagnetic simulation results.
Fig. 5.
Fig. 5. Experimental setup for projection
Fig. 6.
Fig. 6. Captured images of reconstructed hologram, with the frame numbers of (a) #10, (b) #21, (c) #35, and (d) #45.
Fig. 7.
Fig. 7. The intermediately reconstructed image with illumination between the frames #3 and #4.

Tables (2)

Tables Icon

Table 1. Dimensional parameters of apertures measured from a hologram frame and aperture array patterns [nm]

Tables Icon

Table 2. Summary of measured retardation Δ [deg] and conversion efficiency [%]

Equations (5)

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

ϕ = 2 sin 1 ( λ 2 p ) ,
E out = HWP θ E in = ( cos 2 θ sin 2 θ sin 2 θ cos 2 θ ) ( 1 i ) = e 2 i θ ( 1 i ) ,
E out = RET Δ , θ E in = cos Δ 2 e i Δ 2 ( 1 i ) + sin Δ 2 e i Δ 2 i π 2 e 2 i θ ( 1 i ) ,
S = LP ψ X 45 LP 90 S unp ,
S 0 = p 1 2 + p 2 2 2 p 1 p 2 cos Δ cos 2 ψ + ( p 1 2 p 2 2 ) sin 2 ψ 4 ,
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.