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

POLICRYPS-based electrically switchable Bragg reflector

Open Access Open Access

Abstract

The formation and characterization of a switchable volume reflective element fabricated from a polymer liquid crystal (LC) polymer slice (POLICRYPS) structure by holographic photopolymerization at high temperature (65 °C) using a photosensitive/nematic liquid crystal prepolymer mixture is reported. The submicron Bragg structure formed consists of periodic continuous polymeric walls separated by periodic LC channels. The phase separated NLC self-aligns in a homeotropic alignment between the polymer walls as indicated by polarizing optical microscopy analysis (Maltese cross). The resulting periodic grating structure results in a Bragg reflection notch upon illumination with white light due to the periodic variation in refractive index. Electro-optical experiments realized through in-plane electrodes and temperature experiments confirm that the multilayer structure acts as a Bragg mirror whose reflection efficiency can be controlled by either a small (~3V/µm) electric field or temperature.

© 2015 Optical Society of America

1. Introduction

Holographic photopolymerization (HP) is a simple, fast and attractive way to fabricate one-, two- and three-dimensional (1D, 2D, 3D) complex structures with length scale periodicity commensurate with the optical regime [1–4]. For these applications, the recording material must be optimized for efficient diffraction with low scattering and maximum information storage. Photopolymer materials are inexpensive, easily processed, can record low-loss, high-diffraction-efficiency optical elements, and have large data storage capability [5,6]. Combining the properties of HP and liquid crystals (LC) by mixing before photopolymerization enables the fabrication of switchable periodic structures. Upon photopolymerization induced by a periodic intensity pattern, two distinct phases, polymer and LC-rich can be formed whose ultimate local and periodic morphology is related to the chemistry, wavelength of irradiation, and interference geometry. The final morphology is dictated by a complex balance between polymerization kinetics, phase separation, and diffusion. The Bragg diffraction properties arising due to the periodic refractive index profile can be modulated with an electric field. Two different kinds of switchable gratings grew from this premise including holographic polymer dispersed liquid crystals (H-PDLCs) [7–10], composed of periodic LC nano/micro-domains (droplets) separated by cross linked polymer regions and POLICRYPS [11–13], slices of pure NLC (instead of droplets) separated by polymer walls. The molecular director of the latter case can be uniformly aligned relative to the polymeric walls during the fabrication process and thus much larger refractive index contrast can be obtained. In general a large surface to volume ratio is introduced by the small droplets [14] in the HPDLC case which leads to large field strengths (15-20 V/micron) needed to switch the optical properties. Improvements to the field strengths can be obtained by introducing specialty monomers and additives into the system [14]. A second drawback is light scattering in the blue portion of the visible spectrum due to the droplet morphology. It is worth mentioning that another approach for reducing the scattering losses is to align the LC during formation by utilizing reactive mesogenic monomers [15,16]. The absence of individual NLC droplets (present in H-PDLC) in POLICRYPS structures enables field strengths of a few V/μm to reorient the NLC director on a millisecond timescale. The confinement and subsequent alignment of NLC molecules into uniform channels also prevents light scattering and allows fabrication of structures with high optical quality as diffraction efficiencies as high as 98% have been measured. H-PDLCs have been extensively studied for the realization of both Bragg transmission and reflection gratings [17–20]. For transmission gratings, extensive concepts exploiting the ability to dynamically diffract between two states have been explored. For reflection gratings, it is possible to take advantage of the strong wavelength selectivity and efficient diffraction of a single beam for applications ranging from high resolution reflective spatial light modulators to dynamic filtering. POLICRYPS transmission gratings have been extensively studied and have shown relevance to switchable phase modulators [21], all-optical diffraction gratings [22] and tunable optical filters [23]. This geometry offers a versatile suite of wall geometries, either planar or curved, allows different orientation states (perpendicular or parallel) depending on chemistry and cure conditions, and can be utilized to produce multi-dimension high resolution patterning [24]. To date, however, reflection mode POLICRYPS have not been obtained due to the large differences in Bragg spacing between transmission and reflection geometries and photochemical complexities. Reported here is the first demonstration of a Bragg reflector utilizing the POLICRYPS fabrication technique.

2. Materials and Method

NOA61 (Norland Optical Adheives), LC BL038 (Merck) and a visible photoinitiator (Irgacure 784) were utilized in the weight ratio of 70%, 28%, and 2%.

The holographic setup illustrated in Fig. 1 uses a CW laser emitting at λ = 532nm where the beam is expanded, collimated by means of a telescopic beam expander (spatial filter + collimating lens) and divided into two parts of almost equal intensity by a beam splitter. The two beams (1 and 2) are reflected from two mirrors (M1 and M2) and overlap in a counterpropagating geometry at the entrance plane of the sample cell giving rise to an interference pattern (the interference fringes are perpendicular to the cell surfaces) whose spatial period is approximately 200 nm. The reactive mixture was sandwiched between glass slides (without any chemical treatment) spaced 10 µm apart using glass spacers and exposed to the interference pattern for 10 minutes at 170mW/cm2. Unlike the typical HPDLC process, photopolymerization is performed at elevated temperature (70 °C) to exploit the higher diffusivity of the NLC molecules in the isotropic state. This enables a final morphology devoid of individual small, random dispersed LC droplets. The temperature is controlled by a hot-stage which is used to slowly cool (0.2 °C/min) the exposed reflection grating to room temperature by means of a controlled ramp once the curing process has ended.

 figure: Fig. 1

Fig. 1 Holographic setup for the realization of POLICRYPS structures in reflection geometry: BS, beam splitter; M1, M2 mirrors, HS hot stage; S, sample.

Download Full Size | PDF

3. Optical and elecro-optical characterization

As shown in the schematic representation of Fig. 2(a), the reflective POLICRYPS structure is composed of continuous polymeric layers interspersed between phase separated homeotropically aligned NLC layers.

 figure: Fig. 2

Fig. 2 Schematic of the NLC configuration in reflective POLICRYPS structures (a), a conoscopic image of the sample (b), and an image looking through the sample (c)

Download Full Size | PDF

This geometry was confirmed by inspection of the sample under conoscopic imaging conditions realized by placing a conventional microscope between crossed polarizers and including a condensing and a Bertrand lens (BL). At the back-focal plane of the BL, a dark Maltese cross [Fig. 2(b)] was observed indicating that the NLC molecules were homeotropically aligned within the phase separated LC layers. As further evidence [Fig. 2(c)], the grating area (red circle) shows no scatter and is transparent indicating the absence of a scattering NLC droplet morphology. The periodic structure formed acts as a Bragg mirror with a modulation of the refractive index between pure polymer (np≈1.54) and the ordinary index of the NLC (no≈1.5). The Bragg condition can be expressed as:

2neffΛsinϑ=mλB
where neff is the effective refractive index of the structure, Ʌ is the periodicity, ϑ is the angle between the propagating beam and the sample, m is an integer and λB is the Bragg wavelength. In our case since ϑ = 90° (normal incidence), neff≈1.52 and m = 1, Eq. (1) can be expressed as 2 neff Ʌ = λ.

The reflection spectrum of the sample [Fig. 3(a)] shows a sharp and narrow (≈7nm) peak centered at λB = 523nm suggesting a periodicity of Ʌ≈170nm. The grating diffraction efficiency (η) can be explained by using the coupled-wave theory, for both transmission and reflection gratings, employing an effective medium approach to model the refractive index modulation. Anisotropic coupled-wave theory has yielded excellent agreement with experimental polarization and electro-optical properties of these gratings [25]. In this framework, η depends mainly on grating thickness L, probe wavelength λ, electric field E and temperature T and refractive index contrast of the grating, Δng. In our case, the reflection efficiency of the sample is quite modest (≈16%) since Δngr = np-no = 1.54-1.5 = 0.04 of the structure is small. To confirm this behavior, the sample was heated above the NLC clearing temperature (70 °C). In this condition, the NLC (isotropic phase) exhibits a refractive index of ~1.53 and consequently Δngr is reduced to 0.01. As a consequence, the amplitude of the reflection peak is attenuated and exhibits a reflection efficiency of ~4% [Fig. 3(b)]. We have performed some early explorations on the stability of such a system and have not observed any degradation in electro-optical or optical behavior when subjected to temperature cycles across the nematic-isotropic phase transition (70 °C).

 figure: Fig. 3

Fig. 3 Reflection spectrum of the sample at room temperature (a) and beyond (b) the NLC clearing temperature (70°C) (b).

Download Full Size | PDF

The sample does not exhibit electric-field switching as initially formed as the LC molecules (positive dielectric anisotropy) are already aligned parallel to the field direction across the cell (electric field applied through the bulk of the cell by means of ITO coated glass substrates). To further prove this LC orientation and to show that e-field modulation is possible (proof-of-concept), a sample (same experimental conditions) was fabricated with in-plane electrodes. The schematic of the electrode configuration is reported in Fig. 4a. The electrodes are aluminum (width = z = 5µm; distance = s = 20µm; interdistance = w = 60µm) formed through a standard photolithographic process reported elsewhere [26].

 figure: Fig. 4

Fig. 4 Schematic of the electrodes in the in-plane switching configuration (a). Sketch of the NLC molecules under the influence of the in-plane electric field (b) along with a detail of the electric field distribution affecting only the NLC molecules in proximity to the electrodes(c).

Download Full Size | PDF

Figs. 5(a)-5(d) show polarized optical microscope (POM) images of the sample at different values of the electric field. Bright stripes with increasing intensity (from a-d) confirm that the application of an external voltage (square voltage, 1KHz) induces a reorientation of the NLC molecules along the electric field distribution. The electric field affects only the NLC molecules in proximity to the electrodes (see schematic in Figs. 4 (b) and 4 (c)). This can be explained by considering that the strength of the electric field is proportional to s−2 (where s is the distance between the electrodes). In this case, s = 20µm, and for this reason the quite modest electric field (3 V/ µm) affects only the NLC molecules in proximity to the electrodes as the strength of the electric field is not strong enough to reorient the NLC molecules in the middle between the two electrodes (dark area or homeotropic alignment between the two close bright strips, Figs. 5(a)-5(d). The unaffected area can be reoriented by increasing 2-3 times the amplitude of the electric field. This condition affects irreversibly the electro-optical properties of the sample. Optimization of the electrode pattern for in-plane switching is well known in LCD technologies. Related opportunities include FFS (Fringing Field Switching) techniques but optimization of the In-Plane structure and approach were not the focus of the current study. We simply used the geometry to independently prove that the NLC molecules were in a hometropic orientation.

 figure: Fig. 5

Fig. 5 POM images of the sample under the influence of an in-plane electric field.

Download Full Size | PDF

For impinging white (non-collimated) light, a noticeable difference between bright [Fig. 6(a)] and dark [Fig. 6(d)] states is obtained by rotating the sample between crossed polarizers while keeping the external voltage (3V/µm) applied. This observation is clear proof that the NLC confined between the polymer layers does not exhibit a droplet-based morphology (e.g H-PDLC) but instead is experimental evidence that the NLC exhibits a pure nematic phase and the sample behaves as a uniaxial retardation plate with its optical axis aligned parallel to the polymer layers.

 figure: Fig. 6

Fig. 6 POM images of the structure under the influence of an in-plane electric field (3V/µm) for different angles between the analyzer and the optical axis of the sample.

Download Full Size | PDF

In this geometry, Fig. 7 shows that an electric field induces an enhancement of the reflection peak amplitude (≈10%) due to an increase of Δngr. Noteworthy, despite the reflection efficiency is not very high, the spontaneous vertical alignment of the NLC represents a tremendous opportunity for new applications such as vertically aligned displays. As such, the low reflection efficiency in this geometry confirms the specific orientation. According to the anisotropic coupled-wave theory (Ref. 25), using a Δngr = 0.2 (achievable, for example, with simple E7 NLC) there are means to obtain higher reflection efficiencies. Alternative methods to potentially increase the diffraction efficiency include using a NLC with a negative dielectric anisotropy or by chemically modifying the pre-polymer in order to force the NLC molecules to stay planar aligned in the initial state. Study of both approaches is currently underway. The focus of this paper wasn’t on maximizing the specific reflection efficiency, but instead to show it is possible to use the POLICRYPS methodology to fabricate a reflection grating system (of appropriate periodicity).

 figure: Fig. 7

Fig. 7 Reflection spectrum of the sample without (red curve) and with (blue curve) an applied electric field.

Download Full Size | PDF

4. Conclusion

To summarize, an electrically switchable Bragg reflector has been fabricated using a composite mixture of pre-polymer and nematic liquid crystal through a holographic photopolymerization process. The periodic grating structure is composed of polymeric layers with phase separated, homeotropically aligned NLC layers (POLICRYPS) formed through this one-step in situ process. Electro-optical characterization of these switchable reflective POLICRYPS structures, performed through an in-plane electrode geometry indicates control of the reflection efficiency with a moderate electric field (3 V/µm). Exploration of reflective POLICRYPS samples prepared with different LC materials and phases (liquid crystalline polymers, cholesteric, ferroelectric, azo-LC) is currently underway.

Acknowledgments

The research was supported by the Air Force Office of Scientific Research (AFOSR), Air Force Research Laboratory (AFRL), U.S. Air Force, under grant FA9550-14-1-0050 (P.I. L. De Sio, EOARD 2014/2015) and the Materials and Manufacturing Directorate, AFRL.

References and links

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

2. T. Bunning, L. Natarajan, V. Tondiglia, and R. Sutherland, “Holographic polymer dispersed liquid crystals (H-PDLCs),” Annu. Rev. Mater. Sci. 30(1), 83–115 (2000). [CrossRef]  

3. G. Crawford, “Electrically switchable Bragg gratings,” Opt. Photonics News 14(4), 54–59 (2003). [CrossRef]  

4. R. Mayoral, J. Requena, J. S. Moya, C. López, A. Cintas, H. Miguez, F. Meseguer, L. Vázquez, M. Holgado, and A. Blanco, “3D long-range ordering in an SiO2 submicrometer-sphere sintered superstructure,” Adv. Mater. 9, 257–260 (1997). [CrossRef]  

5. S. L. Rosen, Fundamental Principles of Polymeric Materials (John Wiley & Sons, 1993).

6. J. V. Crivello and E. Reichmanis, “Photopolymer materials and processes for advanced technologies,” Chem. Mater. 26(1), 533–548 (2014). [CrossRef]  

7. V. P. Tondiglia, L. V. Natarajan, R. L. Sutherland, D. Tomlin, and T. J. Bunning, “Holographic formation of electro-optical polymer-liquid crystal photonic crystals,” Adv. Mater. 14(3), 187–191 (2002). [CrossRef]  

8. K. K. Vardanyan, J. Qi, J. N. Eakin, M. De Sarkar, and G. P. Crawford, “Polymer scaffolding model for holographic polymer-dispersed liquid crystals,” Appl. Phys. Lett. 81(25), 4736–4738 (2002). [CrossRef]  

9. M. Jazbinšek, I. Drevensek-Olenik, M. Zgonik, A. K. Fontecchio, and G. P. Crawford, “Characterization of holographic polymer dispersed liquid crystal transmission gratings,” J. Appl. Phys. 90(8), 3831–3837 (2001). [CrossRef]  

10. D. E. Lucchetta, L. Criante, and F. Simoni, “Optical characterization of polymer dispersed liquid crystals for holographic recording,” J. Appl. Phys. 93(12), 9669–9674 (2003). [CrossRef]  

11. R. Caputo, L. De Sio, A. Veltri, C. Umeton, and A. V. Sukhov, “Development of a new kind of switchable holographic grating made of liquid-crystal films separated by slices of polymeric material,” Opt. Lett. 29(11), 1261–1263 (2004). [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. L. De Sio and N. Tabiryan, “Self-aligning liquid crystals in polymer composite systems,” J. Polym. Sci., Part B, Polym. Phys. 52(3), 158–162 (2014). [CrossRef]  

14. M. Schulte, S. Clarson, L. Natarajan, D. Tomlin, and T. Bunning, “The effect of fluorine-substituted acrylate monomers on the electro-optical and morphological properties of polymer dispersed liquid crystals,” Liq. Cryst. 27(4), 467–475 (2000). [CrossRef]  

15. V. K. Thakur and M. R. Kessler, Liquid Crystalline Polymers I (Springer, 2015).

16. V. K. Thakur and M. R. Kessler, Liquid Crystalline Polymers II (Springer, 2015).

17. M. S. Park, Y. H. Cho, B. K. Kim, and J. S. Jang, “Fabrication of reflective holographic gratings with polyurethane acrylate (PUA),” Curr. Appl. Phys. 2(3), 249–252 (2002). [CrossRef]  

18. L. V. Natarajan, C. K. Shepherd, D. M. Brandelik, R. L. Sutherland, S. Chandra, V. P. Tondiglia, D. Tomlin, and T. J. Bunning, “Switchable holographic polymer-dispersed liquid crystal reflection gratings based on thiol-ene photopolymerization,” Chem. Mater. 15(12), 2477–2484 (2003). [CrossRef]  

19. R. A. Ramsey, S. C. Sharma, and K. Vaghela, “Holographically formed Bragg reflection gratings recorded in polymer-dispersed liquid crystal cells using a He-Ne laser,” Appl. Phys. Lett. 88(5), 051121 (2006). [CrossRef]  

20. F. Vita, D. E. Lucchetta, R. Castagna, O. Francescangeli, L. Criante, and F. Simoni, “Detailed investigation of high-resolution reflection gratings through angular-selectivity measurements,” J. Opt. Soc. Am. B 24(3), 471–476 (2007). [CrossRef]  

21. L. De Sio, N. Tabiryan, R. Caputo, A. Veltri, and C. Umeton, “POLICRYPS structures as switchable optical phase modulators,” Opt. Express 16(11), 7619–7624 (2008). [CrossRef]   [PubMed]  

22. L. De Sio, S. Serak, N. Tabiryan, and C. Umeton, “Mesogenic versus non-mesogenic zzo dye confined in a soft-matter template for realization of optically switchable diffraction gratings,” J. Mater. Chem. 21(19), 6811–6814 (2011). [CrossRef]  

23. D. Donisi, L. De Sio, R. Beccherelli, M. A. Caponero, A. d’Alessandro, and C. Umeton, “Optical interrogation system based on holographic soft matter filter,” Appl. Phys. Lett. 98(15), 151103 (2011). [CrossRef]  

24. L. De Sio, N. Tabiryan, and T. J. Bunning, “Spontaneous radial liquid crystals alignment on curved polymeric surfaces,” Appl. Phys. Lett. 104(22), 221112 (2014). [CrossRef]  

25. G. Montemezzani and M. Zgonik, “Light diffraction at mixed phase and absorption gratings in anisotropic media for arbitrary geometries,” Phys. Rev. E Stat. Phys. Plasmas Fluids Relat. Interdiscip. Topics 55(1), 1035–1047 (1997). [CrossRef]  

26. L. Lu, J. Y. Hwang, and L. C. Chien, “Effect of IPS cell structure on the electro-optical property of a room-temperature blue phase liquid crystal,” Symposium Digest of Technical Papers40, 1608–1610 (2009). [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 (7)

Fig. 1
Fig. 1 Holographic setup for the realization of POLICRYPS structures in reflection geometry: BS, beam splitter; M1, M2 mirrors, HS hot stage; S, sample.
Fig. 2
Fig. 2 Schematic of the NLC configuration in reflective POLICRYPS structures (a), a conoscopic image of the sample (b), and an image looking through the sample (c)
Fig. 3
Fig. 3 Reflection spectrum of the sample at room temperature (a) and beyond (b) the NLC clearing temperature (70°C) (b).
Fig. 4
Fig. 4 Schematic of the electrodes in the in-plane switching configuration (a). Sketch of the NLC molecules under the influence of the in-plane electric field (b) along with a detail of the electric field distribution affecting only the NLC molecules in proximity to the electrodes(c).
Fig. 5
Fig. 5 POM images of the sample under the influence of an in-plane electric field.
Fig. 6
Fig. 6 POM images of the structure under the influence of an in-plane electric field (3V/µm) for different angles between the analyzer and the optical axis of the sample.
Fig. 7
Fig. 7 Reflection spectrum of the sample without (red curve) and with (blue curve) an applied electric field.

Equations (1)

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

2 n eff Λsinϑ=m λ B
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.