Next Article in Journal
Crystal Structure and Functional Characterization of an S-Formylglutathione Hydrolase (BuSFGH) from Burkholderiaceae sp.
Previous Article in Journal
Highly Reliable Temperature Sensor Based on p-GaN/AlGaN/GaN Hybrid Anode Diode with Wide Operation Temperature from 73 K to 573 K
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Angle Structural Color Scattering Features from Polymeric Photonic Structures

1
Department of Physics, Prifysgol Aberystwyth University, Aberystwyth SY23 3BZ, UK
2
Minton, Treharne & Davies Ltd., Coryton, Cardiff CF14 7HY, UK
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(4), 622; https://doi.org/10.3390/cryst13040622
Submission received: 10 March 2023 / Revised: 29 March 2023 / Accepted: 31 March 2023 / Published: 4 April 2023
(This article belongs to the Section Crystal Engineering)

Abstract

:
Three-dimensional goniometric measurements of structural color from melt-sheared polymer nanoparticle composites is presented revealing high-angle features arising from Bragg scattering. These features are presented in terms of ‘scattering cones’ from full 180° hemispherical scans showing the spectroscopic and angular properties of these scattering spots. This work identifies the Miller indices of the photonic crystal planes responsible for these features and makes further observations as to the appearance of this scattering in the context of these structures as quasi-ordered systems. We probe structural color metrics such as peak wavelength and the tunability of reflectance intensities. As such, this report contributes towards an improved understanding of Bragg scattering and structural color from structures with inherent disorder. The complexity and specificity of color quality across the scattering hemisphere is an important consideration for practical uses such as in sensing applications, and we suggest that soft photonics, in particular, are strong candidates in high-angle color uses.

1. Introduction

The origins of brilliant structural color from photonic crystals (PCs) are well understood in terms of strong light–matter interactions in the form of photonic bandgaps, initially put forth by Yablonovitch and John [1,2]. These stem from the resonance of light with nanoscale dielectric periodicities [3,4,5]. Sanders [6] was among the first to recognize that the characteristic iridescence of precious opal gemstones stems from the Bragg diffraction of light from the three-dimensional repetition of the constituent silica nanospheres, producing intense coloration that changes with viewing angle. This structural color appears throughout the natural world, with the distinctive tree-like repetitions within the wings of Morpho butterflies [7,8] among the most reported on structures in the photonic crystal literature. This has given rise to ‘bioinspired’ photonic crystal devices aiming to replicate the mechanisms of nature in the development of structurally colored materials. Within the field of PCs, soft photonic systems in particular are of emerging importance [9,10]. As PCs find wider application, polymeric materials are particularly strong candidates as smart materials [11,12,13,14] on account of their robustness, viscoelasticity, and capacity for large-scale fabrication by means of roll-to-roll (R2R) processing [14,15,16], which allows for consistent and inexpensive fabrication over the scale of many m2.
Polymer opal (PO) thin films are, therefore, highly relevant systems in which to understand the crystallographic origins of structural color. These core-interlayer-shell nanoparticle composites are ordered by R2R melt shear fabrication [16], whereby order iterates from the outer surfaces of the films as the particle cores self-assemble into twinned fcc arrangements. During shear, a region of melt flow is continually present in the bulk—resulting in structures with inherent disorder. This is compounded by the lack of next-nearest neighbor interaction [17,18] due to the ‘sticky’ PO shells.
The structural color of POs has been extensively reported on previously [19,20,21,22,23]. However, this has been primarily in terms of the primary Bragg feature occurring from the periodicity of the {111} plane. Other Bragg features have been reported on elsewhere [24]; although, this is specifically in the context of mechanical symmetry-breaking effects [25]. Here we report on additional Bragg features present across the hemisphere of scattering, as measured with three-dimensional reflectance goniophotometry. These features are presented in terms of ‘scattering cones’ showing the spectroscopic reflectivity of the films as a function of ordering depth and viewing angle, and this scattering response is mapped to the Bragg-Snell law in order to ascertain the Miller indices of the plane(s) this scattering originates from. This work builds towards an improved understanding of the structural color of POs and additionally highlights previously unreported high-angle coloration effects from these films.
The nature of structural color stemming from photonic crystals as permanent and resistant to photo bleaching is central to its appeal in aesthetic applications, compared to traditional pigments which lack longevity and tunability. This is in addition to the environmental benefits over potentially toxic dyes, furthermore in terms of photonic crystals as solvent-free systems. In particular, high-angle structural color is pertinent to a range of device, display, and aesthetic applications. For example, Ji and colleagues [26,27] report on high-angle structural color for application in viewing-invariant color filters, and elsewhere geometry dependent photonic crystals have been touted as potential dynamic color filters [28,29]. In the latter works, structural color remained intense up to viewing angles of 60–70°. In terms of other applications, Wu et al. [30] discuss the potential of high-angle structural color PCs as anticounterfeiting mechanisms. They show how the color appearance of silica bilayer crystals was highly saturated up to viewing angles of 55–65°.
However, the vast majority of this existing work relies on multilayer films fabricated by either deposition techniques or ablation methods such as electron beam lithography, neither of which are readily compatible with mass production. Such methods can also be expensive. There remains a lack of literature pertaining to high-angle structural color from materials which can easily be applied with R2R methods, and/or harness the mechanical robustness of flexible photonic devices, which could be applied to a wider range of substrates and surfaces. Furthermore, polymer opals are materials which do not require deposition or bonding to surfaces, as they simply behave like decals or stickers. This makes them able to be installed to surfaces retroactively, and easily removed—with the possibility of being recycled if the film is in good condition. For these reasons, in addition to the inherent disorder of POs, we believe this work represents a step change away from the usual high-angle structural color paradigm in terms of reporting on materials with a notable versatility.

2. Materials and Methods

2.1. Thin Film Fabrication and Ordering Chracterization

The polymer opals reported here were fabricated from 223 nm diameter particles which comprise a polystyrene (PS) core, a poly(methyl methacrylate) (PMMA) interlayer, and a poly(ethyl acrylate) (PEA) shell in a 34:16:50 volume ratio. These particles were synthesized by starve-fed, multistage seeded emulsion polymerization [31], whereby surfactant quantity can be altered to tune final particle size and, hence, resultant color appearance. Following synthesis, the particle slurry was dried into precursor material, which was then extruded into 1 mm thick ribbons with a Haake MiniLab co–rotating screw extruder. The ribbons were laminated together between PET strips and compressed to form POs (Figure 1a) with a bespoke kit. At this stage, the films are highly disordered and characterized by a milky-blue appearance originating from the structural correlation length [32].
The POs are then ordered by Bending-Induced Oscillatory Shear (BIOS), first proposed by Zhao et al. [16]. The films, of approximately 100–150 µm thickness, are drawn over 1 cm diameter rollers, heated to 100 °C at a shear rate of 1 cm s−1, with the shear direction ĝ indicated. This is repeated 0–40 times in order to obtain improved sample ordering, which is optimal at 40 BIOS iterations (see illustrated inset). The cross-linked PS cores remain rigid as they are forced by the compression to flow through the soft poly(ethyl acrylate) matrix. The end result is rubbery, free-standing films displaying intense structural color which changes with viewing angle and illumination, pictured rightmost in Figure 1a. With the resulting peak wavelength of color appearance dependent on the CIS particle diameter, this process has previously been tuned to engineer POs reflecting in the UV [33] through to red [21] and near-infrared part of the spectrum.
BIOS processed POs have been extensively reported and have been characterized using electron microscopy [20] and small-angle X-ray scattering (SAXS) [34] among other techniques, which have hinted that optical features away from the {111} direction should be visible. Further transmission electron microscopy (TEM) reported in Figure 1b–d validates the expected ordering of these structures as a baseline of efforts to begin probing these other planes. Microscopy was performed with a Thermo Scientific Talos L120C G2 operating at 120 kV in Montage Mode. The films were microtomed into 60–80 µm thickness slices using a Leica EM UC7 & FC7 diamond knife at −60 °C. This was perpendicular to the shear direction to obtain a series of film cross sections, in order to probe the inner ordering mechanics of BIOS where it is well understood that, due to the melt shear nature of the process, a disordered region remains in the center of the film. The slices were stained with RuO4 vapor on copper grids. Figure 1b–d shows the internal structure of 0, 20, and 40 BIOS POs, respectively, with the film surfaces shown at the top of the images. These images have been digitally enhanced for improved clarity to account for staining difficulties. This microscopy clearly shows that structural order begins at the outer surfaces of the films and propagates inwards with successive shear—although, bulk ordering is not achieved (not shown). A high degree of ordering is observed in 40 BIOS opals (Figure 1d), with the PS cores arranged in straight layers. This is seen to occur over the regime of many particles, with ordering depths having previously being shown to extend past the Bragg penetration depth [20]. This is in agreement with separate studies [16] which showed these structures to demonstrate twinned face-centered cubic (fcc) packing [17,18,35].

2.2. Reflectance Characterization

Three-dimensional relative reflectance goniophotometry was undertaken with a Reflet 180S in a dark box. This is illustrated in Figure 2 in order to elucidate the co-ordinate geometries. The collection optics can move freely within any values of measurement, or viewing angle θm, and viewing plane angle ϕm in the upper hemisphere of the instrumentation. This allows a number of planes of scattering to be studied and the viewing angle dependency of the resultant color appearance to be fully examined. θm and the incident illumination angle θI are defined as 0° at the zenith, whereas ϕm is defined as ±90° in the plane of illumination. The sample stage height is adjustable to align the focusing optics and incident light with the sample surface.
The spectrometer was calibrated with respect to a 99% Lambertian reflectance standard and dark background. The measured values of light intensity are then applied, whereby the dark background is subtracted from the measured intensity, with this divided by the reflectance profile of the Lambertian standard and then multiplied by 100% to give relative reflectance. To calibrate with respect to a mirror or light source would cause relative reflectance values to be very low outside of the specular direction. Hence, calibration with respect to an omnidirectional scatterer alleviates this issue to bring values into the normal range, important for the consideration of structural color visible at high angles away from the specular. As a caveat, care should be taken with this technique as around the specular direction values become very high and, thus, are removed from the spectra in postprocessing.
The PO films are aligned with the shear direction ĝ parallel to the plane of illumination, ϕm = ±90°. A collimated 3200K halogen tungsten bulb was incident on the films in a spot size of approximately 8 mm, with the angle of illumination given as displacement from the zenith. Scattered light was collected by the detector, with a collection solid angle of 3.3 × 10−4 Sr, which was fiber coupled to a CCD spectrometer. The light source and detector rotate approximately 20–25 cm from the center of the instrument. Measurements were taken over complete viewing hemispheres −90° ≤ θm ≤ 90° in 0.1° angle increments, with spectrometer collection taken over the complete visible wavelength range of 380–780 nm in increments of 1 nm. This was fully automated within the Reflet 180S software which returned this data as a complete matrix of relative reflectance over the viewing hemispheres.

3. Results and Discussion

Hemispherical scans of relative reflectance are shown in Figure 3 for polymer opals of 0, 20, and 40 BIOS passes, showing the change in structural color properties as a function of increasing structural order and ordering depth [20]. These represent the stacked spectra collected at every viewing angle between ±90°, where the opal is illuminated from the zenith. The scattering features of interest are shown enlarged for each opal, with these rescaled for clarity in Figure 3d–f. Relative reflectance can be seen to exceed 100% due to measurements having been taken with reference to a Lambertian scatterer which scatters omnidirectionally as discussed, compared to polymer opals which by their nature as photonic crystals display preferential scattering directionality. The most intense specular scattering around the 0° viewing direction has been removed, in addition to detector obscuration.
One can observe the high-angle scattering feature to be of low intensity relative to the wider scattering envelope (see Figure 3a) for a 0 BIOS film. However at 20 BIOS, Figure 3b, this feature has fully come into focus, and intensity was seen to peak at 40 BIOS passes (Figure 3c). This feature is visible over viewing angles 70° to 20°, in line with the highest viewing angles of structural color reported in the literature; although, closer to the specular the scattering is rapidly superseded in intensity by a central scattering feature originating from the {111} plane. As shown in Figure 3c, the symmetry of the high-angle scattering feature is shown to degrade as viewed either side of the specular, illustrated by the difference in appearance of scattering around the θm = ±50° directions. This is likely to be the result of the competitive ordering–disordering of the BIOS process, whereby the melt shear process eventually begins to degrade the structural crystallinity. These structural changes also serve to affect the shape of the scattering feature, shown particularly clearly by Figure 3e,f. The peak wavelengths of the high-angle feature and central scattering cone around 0° were extracted in 10° increments, and fitted to the Bragg-Snell law:
m λ = 2 d n e f f 2 sin 2 θ I ,
where the value of effective refractive index neff is taken from the literature to be 1.52 [36]. The plane spacing d111 was calculated from the peak wavelengths extracted from the primary scattering feature shown around the 0° viewing angle from the {111} plane. This is carried out by fitting Voigt functions away from the specular at a viewing angle where both the {111} and high-angle feature have distinct intense peaks in the spectra in order to give the most accurate result.
Extraction of the peak wavelength allowed for calculation of the lattice parameter a:
d = a h 2 + k 2 + l 2
which varies slightly as a function of BIOS as the crystal experiences deformation, taking on calculated values between 343 and 352 nm. For example, for 40 BIOS, the central wavelength of the 1st order reflectance peak from the {111} plane for θm = −40° is 552 nm, which according to the Bragg-Snell in calculating d111:
d 111 = λ 2 n e f f 2 sin 2 θ I = 552   nm 2   1.52 2 sin 2 40 = 200.4   nm   ± 2.3 .
Inserting the calculated value of d111 to calculate the lattice parameter a:
a = d 111 h 2 + k 2 + l 2 = 200.4   nm 3 = 347   nm   ± 4 .
Given the core-shell particle diameter of 233 nm, the calculated values of a are in the expected range for a face-centered cubic crystal. Reversing this process in order to calculate the Miller indices for the high-angle scattering features, for 40 BIOS one extracts the central peak wavelength at the same viewing angle of θm = −40° of 452 nm in order to calculate the plane spacing:
d ? ? ? = 452   nm 2   1.52 2 sin 2 40 = 164.1   nm   ± 2.3 .
The usual crystallographic relationship between lattice parameter a, plane spacing d, and Miller indices was then applied to find the Miller indices corresponding to the peak wavelengths of the high-angle scattering feature. The relationship between d??? and a = 347 nm as calculated previously is given by:
h 2 + k 2 + l 2 = a d = 347 164 = 2.12   ± 0.04   2
with the only allowed combination of h, k, and l for the face-centered cubic lattice case being given by the {002} plane group. In the same manner, peak wavelengths were extracted prior with the fitting of Voigt functions; this is carried out for viewing angle values from −80° to 80° in 10° increments, where possible. This fits to values expected in accordance with the Bragg–Snell law, as shown by Figure 4. The central specular feature given by the {111} plane is denoted in red, with the {002} plane group feature shown in blue, and the dashed lines corresponding to the Bragg-Snell law behavior. Error is seen to be somewhat greater for the {002} feature due to the difficulty in accurate peak wavelength extraction where intensity is markedly lower. Nevertheless, these calculated values of h, k, and l show a good fit to the Bragg-Snell law and this data conclusively evidences the Bragg scattering origins of this high-angle structural color. Although the level of fit is seen to depreciate past approximately 50°, the fit remains acceptable up to around 70° for 20 and 40 BIOS. Our previous work [20] has highlighted that good structural ordering is seen to reach the Bragg penetration depth of around 10 µm for approximately 10 BIOS, and this is the primary suggestion as to why the fit with Bragg-Snell is particularly poor for 5 BIOS. This is most pertinently the case around the specular direction, illustrated by a dip in the peak wavelengths (Figure 4a). Here, the measured light penetrates deepest into the sample before reflection, so this corresponds with the viewing direction where the measured scattering is most sensitive to the crystalline disorder that persists in the film at this stage in processing. This is, furthermore, evidenced by that seen for 10 BIOS, where around the specular viewing direction the fit to Bragg-Snell is markedly improved in line with reasoning that ordering has now permeated down to the Bragg penetration depth. This figure additionally elucidates other structural changes that occur as BIOS advances—for example, the Bragg-Snell fit for the {002} plane is shown to reach an optimal point at 10 BIOS. Asymmetry in peak wavelength is clearly evident for 20 BIOS at viewing angles of ±20°, and this area of the fit continues to depreciate for 40 BIOS over the range ±10–30°. This is likely caused by the innate disordering mechanisms of shear [16] beginning to affect the outer layers of the crystal, because one can also observe how the {111} fit drifts from the expected behavior, albeit slightly. Indeed, this suggestion is also evidenced by the discrepancies in the calculated values of the lattice parameter indicate deformation competing with self-assembly. Further work is necessary in optimizing the ordering depth obtained from BIOS in line with desired application, and mapping this in further detail to other structural color metrics such as reflectance intensity and color purity, in addition to quantification of the color appearance.
Figure 5 provides further insight regarding the interplay of the two scattering features, with spectroscopic data shown for the (a) θm = 70° and (b) −10° viewing angles. It was not possible to observe the specular direction for illumination from the zenith due to obscuration by the detector. With successive structural ordering, the {111} reflectance peak is shown to become the dominant scattering feature, and reference to Figure 3 indicates this occurs around viewing angles of approximately 20–30° displaced from the zenith. The wavelength FWHM of the {111} seen to sharpen as order permeates the structure, in line with expectation. Whilst direct evaluation of the {002} peak FWHM above the background continuum is difficult, the reflectance intensity increases significantly with viewing angle towards the center of the feature at approximately 50°. Additional understanding of the tuning of this peak could be used as a potential ordering parameter for the system, for example, in terms of FWHM or in terms of comparison with the {111} peak. Further work in optimizing the BIOS process could lead to the enhancement of the {002} reflectance relative to the background, which would evidently be a necessary factor in polymer opals finding high-angle applications.

4. Conclusions

In conclusion this work presents new insights regarding high-angle scattering features from polymer opals, melt-shear-ordered photonic crystals. These features are shown to appear up to 70° removed from the specular viewing direction. Evidence is presented as to their origin associated with the {002} plane group of the twinned fcc structure, determined from the good agreement with the Bragg-Snell law across the viewing plane. The relative reflectance intensity is seen to increase as structural order is gradually introduced into the films. Further data is presented showing the interplay of the {111} and {002} reflectance peaks which features as a function of structural order, and results from the complex ordering–disordering mechanisms stemming from melt shearing. An improved understanding of Bragg scattering from structures with some inherent disorder, and the complexity of structural color quality across the scattering hemisphere, lends itself to such materials being developed as a low-cost, readily scalable, soft photonics approach finding applications as diverse as antiforgery and sensing devices. This is important on account of the understanding of systems with an inherent compatibility with roll-to-roll processes.

Author Contributions

Conceptualization, C.E.F. and G.R.; methodology, G.R. and C.E.F.; software, M.G. and G.R.; validation, C.E.F.; formal analysis, C.E.F. and G.R.; investigation, G.R.; resources, C.E.F.; data curation, G.R. and C.E.F.; writing—original draft preparation, G.R. and C.E.F.; writing—review and editing, M.G., M.B. and J.J.T.; visualization, G.R. and C.E.F.; supervision, C.E.F., M.G. and M.B.; project administration, C.E.F.; funding acquisition, C.E.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded in part by Llywodraeth Cymru KESS-2 (European Social Fund), SPARC-II (European Regional Development Fund).

Data Availability Statement

Data supporting reported results can be found at the Aberystwyth University PURE depository.

Acknowledgments

The authors thank Varichem Ltd., UK, and Jeremy Baumberg (University of Cambridge) for supply of materials. They also acknowledge the contribution of Qibin Zhao and group at Shanghai Jiao Tong University for assistance with electron microscopy imaging.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yablonovitch, E. Inhibited spontaneous emission in solid-state physics and electronics. Phys. Rev. Lett. 1987, 58, 2059. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. John, S. Strong localization of photons in certain disordered dielectric superlattices. Phys. Rev. Lett. 1987, 58, 2486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Yablonovitch, E.; Gmitter, T. Photonic band structure: The face-centered-cubic case. Phys. Rev. Lett. 1989, 63, 1950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Leung, K.-M.; Liu, Y. Full vector wave calculation of photonic band structures in face-centered-cubic dielectric media. Phys. Rev. Lett. 1990, 65, 2646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Parker, A.R. 515 million years of structural colour. J. Opt. Pure Appl. Opt. 2000, 2, R15. [Google Scholar] [CrossRef]
  6. Sanders, J.V. Colour of precious opal. Nature 1964, 204, 1151–1153. [Google Scholar] [CrossRef]
  7. Kinoshita, S.; Yoshioka, S.; Fujii, Y.; Okamoto, N. Photophysics of structural color in the Morpho butterflies. Forma 2002, 17, 103–121. [Google Scholar]
  8. Vukusic, P.; Sambles, J.; Lawrence, C.; Wootton, R. Quantified interference and diffraction in single Morpho butterfly scales. Proc. R. Soc. Lond. Ser. B Biol. Sci. 1999, 266, 1403–1411. [Google Scholar] [CrossRef] [Green Version]
  9. Kolle, M.; Lee, S. Progress and Opportunities in Soft Photonics and Biologically Inspired Optics. Adv. Mater. 2018, 30, 1702669. [Google Scholar] [CrossRef]
  10. Lyu, Q.; Li, M.; Zhang, L.; Zhu, J. Bioinspired Supramolecular Photonic Composites: Construction and Emerging Applications. Macromol. Rapid Commun. 2022, 43, 2100867. [Google Scholar] [CrossRef]
  11. Zeng, S.; Liu, Y.; Li, S.; Shen, K.; Hou, Z.; Chooi, A.P.; Smith, A.T.; Chen, Z.; Sun, L. Smart Laser-Writable Micropatterns with Multiscale Photo/Moisture Reconstructible Structure. Adv. Funct. Mater. 2021, 31, 2009481. [Google Scholar] [CrossRef]
  12. Qiao, Y.; Meng, Z.; Wang, P.; Yan, D. Research Progress of Bionic Adaptive Camouflage Materials. Front. Mater. 2021, 8, 79. [Google Scholar] [CrossRef]
  13. Dong, X.; Wu, P.; Schaefer, C.G.; Zhang, L.; Finlayson, C.E.; Wang, C. Solvatochromism based on structural color: Smart polymer composites for sensing and security. Mater. Des. 2018, 160, 417–426. [Google Scholar] [CrossRef]
  14. Shen, X.; Wu, P.; Schäfer, C.G.; Guo, J.; Wang, C. Ultrafast assembly of nanoparticles to form smart polymeric photonic crystal films: A new platform for quick detection of solution compositions. Nanoscale 2019, 11, 1253–1261. [Google Scholar] [CrossRef] [PubMed]
  15. Li, H.; Wu, P.; Zhao, G.; Guo, J.; Wang, C. Fabrication of industrial-level polymer photonic crystal films at ambient temperature Based on uniform core/shell colloidal particles. J. Colloid Interface Sci. 2021, 584, 145–153. [Google Scholar] [CrossRef] [PubMed]
  16. Zhao, Q.; Finlayson, C.E.; Snoswell, D.R.; Haines, A.; Schäfer, C.; Spahn, P.; Hellmann, G.P.; Petukhov, A.V.; Herrmann, L.; Burdet, P. Large-scale ordering of nanoparticles using viscoelastic shear processing. Nat. Commun. 2016, 7, 11661. [Google Scholar] [CrossRef] [Green Version]
  17. Finlayson, C.E.; Spahn, P.; Snoswell, D.R.; Yates, G.; Kontogeorgos, A.; Haines, A.I.; Hellmann, G.P.; Baumberg, J.J. 3D Bulk Ordering in Macroscopic Solid Opaline Films by Edge-Induced Rotational Shearing. Adv. Mater. 2011, 23, 1540–1544. [Google Scholar] [CrossRef] [Green Version]
  18. Pursiainen, O.L.J.; Baumberg, J.J.; Winkler, H.; Viel, B.; Spahn, P.; Ruhl, T. Shear-Induced Organization in Flexible Polymer Opals. Adv. Mater. 2008, 20, 1484–1487. [Google Scholar] [CrossRef]
  19. Pursiainen, O.L.; Baumberg, J.J.; Winkler, H.; Viel, B.; Spahn, P.; Ruhl, T. Nanoparticle-tuned structural color from polymer opals. Opt. Express 2007, 15, 9553–9561. [Google Scholar] [CrossRef] [Green Version]
  20. Rosetta, G.; An, T.; Zhao, Q.; Baumberg, J.J.; Tomes, J.J.; Gunn, M.D.; Finlayson, C.E. Chromaticity of structural color in polymer thin film photonic crystals. Opt. Express 2020, 28, 36219–36228. [Google Scholar] [CrossRef]
  21. Rosetta, G.; Butters, M.; Tomes, J.J.; Little, J.; Gunn, M.D.; Finlayson, C.E. Quantifying the saturation of structural color from thin film polymeric photonic crystals. In Proceedings of the Photonic and Phononic Properties of Engineered Nanostructures X, San Francisco, CA, USA, 1–6 February 2020; p. 112890B. [Google Scholar]
  22. Schäfer, C.G.; Gallei, M.; Zahn, J.T.; Engelhardt, J.; Hellmann, G.t.P.; Rehahn, M. Reversible light-, thermo-, and mechano-responsive elastomeric polymer opal films. Chem. Mater. 2013, 25, 2309–2318. [Google Scholar]
  23. Schlander, A.M.B.; Gallei, M. Temperature-Induced Coloration and Interface Shell Cross-Linking for the Preparation of Polymer-Based Opal Films. ACS Appl. Mater. Interfaces 2019, 11, 44764–44773. [Google Scholar] [CrossRef] [PubMed]
  24. Schäfer, C.G.; Smolin, D.A.; Hellmann, G.P.; Gallei, M. Fully reversible shape transition of soft spheres in elastomeric polymer opal films. Langmuir 2013, 29, 11275–11283. [Google Scholar]
  25. Kontogeorgos, A.; Snoswell, D.R.E.; Finlayson, C.E.; Baumberg, J.J.; Spahn, P.; Hellmann, G.P. Inducing Symmetry Breaking in Nanostructures: Anisotropic Stretch-Tuning Photonic Crystals. Phys. Rev. Lett. 2010, 105, 233909. [Google Scholar] [CrossRef] [Green Version]
  26. Ji, C.; Lee, K.-T.; Guo, L.J. High-color-purity, angle-invariant, and bidirectional structural colors based on higher-order resonances. Opt. Lett. 2019, 44, 86–89. [Google Scholar] [CrossRef] [PubMed]
  27. Lee, K.T.; Ji, C.; Banerjee, D.; Guo, L.J. Angular-and polarization-independent structural colors based on 1D photonic crystals. Laser Photonics Rev. 2015, 9, 354–362. [Google Scholar] [CrossRef] [Green Version]
  28. Zhao, Y.; Zhao, Y.; Hu, S.; Lv, J.; Ying, Y.; Gervinskas, G.; Si, G. Artificial Structural Color Pixels: A Review. Materials 2017, 10, 944. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Cho, E.-H.; Kim, H.-S.; Cheong, B.-H.; Oleg, P.; Xianyua, W.; Sohn, J.-S.; Ma, D.-J.; Choi, H.-Y.; Park, N.-C.; Park, Y.-P. Two-dimensional photonic crystal color filter development. Opt. Express 2009, 17, 8621–8629. [Google Scholar] [CrossRef]
  30. Wu, S.; Liu, T.; Tang, B.; Li, L.; Zhang, S. Structural Color Circulation in a Bilayer Photonic Crystal by Increasing the Incident Angle. ACS Appl. Mater. Interfaces 2019, 11, 10171–10177. [Google Scholar] [CrossRef]
  31. Winkler, H.; Ruhl, T. Moulded Bodies Consisting of Core-Shell Particles. U.S. Patent 20050142343A1, 2005. [Google Scholar]
  32. Ossi, P. Disordered Materials: An Introduction; Springer: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  33. Rosetta, G.; Gunn, M.; Tomes, J.J.; Butters, M.; Pieschel, J.; Hartmann, F.; Gallei, M.; Finlayson, C.E. Transparent Polymer Opal Thin Films with Intense UV Structural Color. Molecules 2022, 27, 3774. [Google Scholar] [CrossRef]
  34. Shen, Z.; Shi, L.; You, B.; Wu, L.; Zhao, D. Large-scale fabrication of three-dimensional ordered polymer films with strong structure colors and robust mechanical properties. J. Mater. Chem. 2012, 22, 8069–8075. [Google Scholar] [CrossRef]
  35. Finlayson, C.E.; Baumberg, J.J. Polymer opals as novel photonic materials. Polym. Int. 2013, 62, 1403–1407. [Google Scholar] [CrossRef] [Green Version]
  36. Finlayson, C.E.; Baumberg, J.J. Generating bulk-scale ordered optical materials using shear-assembly in viscoelastic media. Materials 2017, 10, 688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. (a) Polymer opals are fabricated from composite nanoparticles of polystyrene (PS) cores coated in a poly(ethyl acrylate) (PEA) shell, grafted to the core by a poly(methyl methacrylate) (PMMA) interlayer. These begin as amorphous films characterized by a milky appearance, before being sheared into arrays giving saturated structural color of wavelength corresponding to core diameter. Transmission electron microscopy (TEM) of film slices of (b) 0 BIOS, (c) 20 BIOS, and (d) 40 BIOS are shown on a common scale, with the film edges shown at the top of the figures. Structural order is seen to iteratively improve and permeate into the films with further shear as the PS cores self-assemble within the PEA matrix. A value of 40 BIOS POs are shown to demonstrate excellent order across areas of 10 s of particles, with the PS cores arranged neatly over many layers. Some figure elements have previously appeared in [20] and have been digitally enhanced.
Figure 1. (a) Polymer opals are fabricated from composite nanoparticles of polystyrene (PS) cores coated in a poly(ethyl acrylate) (PEA) shell, grafted to the core by a poly(methyl methacrylate) (PMMA) interlayer. These begin as amorphous films characterized by a milky appearance, before being sheared into arrays giving saturated structural color of wavelength corresponding to core diameter. Transmission electron microscopy (TEM) of film slices of (b) 0 BIOS, (c) 20 BIOS, and (d) 40 BIOS are shown on a common scale, with the film edges shown at the top of the figures. Structural order is seen to iteratively improve and permeate into the films with further shear as the PS cores self-assemble within the PEA matrix. A value of 40 BIOS POs are shown to demonstrate excellent order across areas of 10 s of particles, with the PS cores arranged neatly over many layers. Some figure elements have previously appeared in [20] and have been digitally enhanced.
Crystals 13 00622 g001
Figure 2. The three-dimensional reflectance goniometer instrumentation, with co-ordinate nomenclature inset. The polymer opal is situated in the center of the dark box environment, while the position of light source and collection optics with neutral density filter (NDF) are remotely controlled, allowing the relative reflectance to be measured under a range of viewing conditions. θI defines the incident angle, ϕm the viewing plane, and θm the viewing angle within the plane.
Figure 2. The three-dimensional reflectance goniometer instrumentation, with co-ordinate nomenclature inset. The polymer opal is situated in the center of the dark box environment, while the position of light source and collection optics with neutral density filter (NDF) are remotely controlled, allowing the relative reflectance to be measured under a range of viewing conditions. θI defines the incident angle, ϕm the viewing plane, and θm the viewing angle within the plane.
Crystals 13 00622 g002
Figure 3. Hemispherical scans of relative reflectance on the z-axis for (a) 0, (b) 20, and (c) 40 BIOS pass polymer opals. The high-angle structural color reflectance hotspots are shown for these opals, respectively, in (df). Relative reflectance is seen to exceed 100% due to measurements being taken with reference to a Lambertian scatterer.
Figure 3. Hemispherical scans of relative reflectance on the z-axis for (a) 0, (b) 20, and (c) 40 BIOS pass polymer opals. The high-angle structural color reflectance hotspots are shown for these opals, respectively, in (df). Relative reflectance is seen to exceed 100% due to measurements being taken with reference to a Lambertian scatterer.
Crystals 13 00622 g003
Figure 4. Extracted peak wavelengths are fit against the Bragg-Snell law for polymer opals of (a) 5, (b) 10, (c) 20, and (d) 40 BIOS shear passes, where the peak wavelengths are shown to display good agreement with origins from the {111} plane and {002} plane groups.
Figure 4. Extracted peak wavelengths are fit against the Bragg-Snell law for polymer opals of (a) 5, (b) 10, (c) 20, and (d) 40 BIOS shear passes, where the peak wavelengths are shown to display good agreement with origins from the {111} plane and {002} plane groups.
Crystals 13 00622 g004
Figure 5. Spectroscopic data of relative reflectance for polymer opals of 0–40 BIOS passes from the viewing angles (a) θm = 70° and (b) θm = −10°, with the peaks originating from the {002} and {111} planes indicated.
Figure 5. Spectroscopic data of relative reflectance for polymer opals of 0–40 BIOS passes from the viewing angles (a) θm = 70° and (b) θm = −10°, with the peaks originating from the {002} and {111} planes indicated.
Crystals 13 00622 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Rosetta, G.; Tomes, J.J.; Butters, M.; Gunn, M.; Finlayson, C.E. High-Angle Structural Color Scattering Features from Polymeric Photonic Structures. Crystals 2023, 13, 622. https://doi.org/10.3390/cryst13040622

AMA Style

Rosetta G, Tomes JJ, Butters M, Gunn M, Finlayson CE. High-Angle Structural Color Scattering Features from Polymeric Photonic Structures. Crystals. 2023; 13(4):622. https://doi.org/10.3390/cryst13040622

Chicago/Turabian Style

Rosetta, Giselle, John J. Tomes, Mike Butters, Matthew Gunn, and Chris E. Finlayson. 2023. "High-Angle Structural Color Scattering Features from Polymeric Photonic Structures" Crystals 13, no. 4: 622. https://doi.org/10.3390/cryst13040622

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop