You are on page 1of 17

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article

Cite This: ACS Omega 2018, 3, 9658−9674 http://pubs.acs.org/journal/acsodf

Achieving Color and Function with Structure: Optical and Catalytic


Support Properties of ZrO2 Inverse Opal Thin Films
Geoffrey I. N. Waterhouse,*,†,‡,§ Wan-Ting Chen,†,‡,§ Andrew Chan,†,‡,§
and Dongxiao Sun-Waterhouse†

School of Chemical Sciences, The University of Auckland, Auckland 1010, New Zealand

The MacDiarmid Institute for Advanced Materials and Nanotechnology, Wellington 6140, New Zealand
§
The Dodd-Walls Centre for Photonic and Quantum Technologies, Dunedin 9054, New Zealand
*
S Supporting Information

ABSTRACT: Taking inspiration from natural photonic


Downloaded by INST TEKNOLOGI BANDUNG at 23:46:45:494 on June 17, 2019

crystal architectures, we report herein the successful


fabrication of zirconia inverse opal (ZrO2 IO) thin-film
photonic crystals possessing striking iridescence at visible
wavelengths. Poly(methyl methacrylate) (PMMA) colloidal
from https://pubs.acs.org/doi/10.1021/acsomega.8b01334.

crystal thin films (synthetic opals) were first deposited on


glass microscope slides, after which the interstitial voids in the
films were filled with a Zr(IV) sol. Controlled calcination of
the resulting composite films yielded iridescent ZrO2 IO thin
films with pseudo photonic band gaps (PBGs) along the
surface normal at visible wavelengths. The PBG position was dependent on the macropore diameter (D) in the inverse opals
(and thus proportional to the diameter of the PMMA colloids in the sacrificial templates), the incident angle of light with
respect to the surface normal (θ), and also the refractive index of the medium filling the macropores, all of which were
accurately described by a modified Bragg’s law expression. Au/ZrO2 IO catalysts prepared using the ZrO2 IO films
demonstrated outstanding performance for the reduction of 4-nitrophenol to 4-aminophenol in the presence of NaBH4, which
can be attributed to the interconnected macroporosity in the films, which afforded a high Au nanoparticle dispersion and also
facile diffusion of 4-nitrophenol to the catalytically active Au sites.

1. INTRODUCTION the display feathers of the peacock (Pavo cristatus) contains 2D


Intriguing, ingenious, and diverse are the mechanisms by which arrays of melanin rods, with the diameter of the rods
nature uses submicron-scale periodic architectures called depending on the location in the eye (rods in the green part
photonic crystals to create structural color.1−14 A characteristic have a diameter of around 140 nm, whereas rods in the orange
feature of structurally colored materials is their iridescence or part have a slightly larger diameter of ∼150 nm).13 Finally,
color change with viewing angle with respect to a fixed white precious opal obtains its color from a face-centered cubic (fcc)
light source (cf. pigment-based colored materials that generally (3D) arrangement of monodisperse spherical silica (SiO2)
look the same color at all illumination and viewing angles). colloids, with colloid diameters typically in the range 200−300
The striking iridescence shown by photonic crystals arises from nm.14 The diffracted wavelengths in opals increase with colloid
the selective diffraction of certain wavelengths of white light on diameter (i.e., at normal incidence, opals made of 200 nm
one-dimensional, two-dimensional (2D), or three-dimensional colloids reflect in the violet, those made from 250 nm colloids
(3D) periodic structures within the material, where typically in the green, and those made from 300 nm colloids in the red).
the periodicity is comparable to visible wavelengths (400−700 Light refraction and diffraction leading to color in opal-based
nm). Figure S1 shows digital photographs for some common photonic crystals are depicted in Figure S2 and have been
naturally occurring structurally colored objects, as well as discussed in depth below.15−24 In each of the examples in
scanning electron microscopy images of the periodic dielectric Figure S1, a dark background provided by protein or a
structures responsible for the iridescent colors of these objects. sandstone matrix enhances the color play of the photonic
These examples serve to highlight key concepts exploited here crystal architecture, making the iridescence more vivid. These
in our research aimed at the fabrication of ZrO2 inverse opal are but a brief snapshot of the diverse architectures used in
thin-film photonic crystals. In Paua shell (Haliotis iris), 400 nature to generate structural color. Physicists and engineers are
nm-thick tiles of calcium carbonate add mechanical strength to
the molluscs shell, but also cause the shell to diffract strongly at Received: June 13, 2018
both near-infrared (first-order diffraction) and visible (second-, Accepted: August 9, 2018
third-, and fourth-order diffraction) wavelengths.12 The eye in Published: August 21, 2018

© 2018 American Chemical Society 9658 DOI: 10.1021/acsomega.8b01334


ACS Omega 2018, 3, 9658−9674
ACS Omega Article

now using “top-down” approaches25−28 and chemists and undergraduate teaching labs in Advanced Physical Chemistry
biologists are using “bottom-up” approaches29−69 to fabricate (CHEM 310) and Materials Chemistry (CHEM 380) at the
photonic crystals operating at near-infrared wavelengths (for University of Auckland. Indeed, 4−12 weeks undergraduate
optical communications, especially waveguides, fiber-optic research projects based on photonic crystal fabrication by the
cables, fast optic switches, etc.) and visible wavelengths colloidal crystal template method have proved extremely
(sensing, high-efficiency laser cavities). Nature serves as a popular with final year BSc students and honors program
valuable teacher for technology breakthroughs in these areas. students, offering a rich tapestry of open-ended learning
Among the methods used by chemists for photonic crystal experiences that encompass, but are not limited to, emulsion
fabrication, the colloidal crystal template approach is arguably polymerization for polymer colloid synthesis, colloid crystal-
the most popular due to the diverse range of photonic crystals lization techniques, colloidal crystal templating strategies, sol−
that can be obtained.42−69 The first step in this approach is the gel chemistry, optical characterization techniques (UV−vis
synthesis of monodisperse polymer or silica colloids via transmittance and reflectance), computational modeling of
surfactant-free emulsion polymerization or the Stö ber photonic band-gap structures, as well as hands-on exposure to
method,30−33,42,48 respectively, followed by the crystallization advanced characterization techniques not normally accessed by
of these colloids to form a colloidal crystal or synthetic opal undergraduates (scanning electron microscopy (SEM)/energy-
(typically a face-centered cubic array of colloids with a solid dispersive spectrometry (EDS), transmission electron micros-
fraction of 74 vol %). Due to the high interfacial surface free copy (TEM), X-ray diffraction (XRD), X-ray photoelectron
energies of submicron-sized spherical colloids, polymer and spectroscopy (XPS), N2 physisorption). Parallel research
silica colloids tend to spontaneously self-assemble to form well- around natural structurally colored materials such as iridescent
ordered colloidal crystals, which assists this fabrication route. beetles, bird feathers, and so forth has allowed students to
Strategies used to form colloidal crystals from aqueous or greatly extend their knowledge of color and how it is generated
ethanoic colloidal dispersions include gravitational sedimenta- (i.e., extending beyond the norms such as pigment-based color,
tion, centrifugation, evaporation, and various “falling meniscus” transition-metal complexes, charge-transfer complexes, fluo-
methods.33−42,70−74 The latter methods are especially ame- rescence and phosphorescence, chemiluminescence), culmi-
nable for colloidal crystal thin-film deposition on planar nating in enhanced student interest in higher postgraduate
substrates (this is preferable if detailed optical analyses are programs. Methodologies introduced here for inverse opal
going to be performed on the colloidal crystals and their thin-film fabrication are designed to be easily adopted
inverted replicas). In the subsequent stage of the colloidal elsewhere to enhance chemistry-based teaching programs.
crystal templating strategy, the colloidal crystal (i.e., opal) is
used as a sacrificial template to fabricate inverse opal replicas, 2. EXPERIMENTAL SECTION
comprising a face-centered cubic array of air spheres 2.1. Materials. Methyl methacrylate (99%, Aldrich), 2,2′-
(macropores) in a solid matrix.42−69 To obtain the inverse azobis(2-methylpropionamidine) dihydrochloride (97%),
opal, the interstitial voids in the colloidal crystal template (26 zirconium(IV) propoxide (70 wt % in 1-propanol), concen-
vol % of the structure) are filled with a solid material via sol− trated HCl (37 wt %), tetrachloroauric acid trihydrate
gel, electrochemical, or nanoparticle infiltration methods, after (≥99.8%), urea (99%), sodium borohydride (%), and 4-
which the original colloidal crystal template is selectively nitrophenol (%) were all obtained from Sigma-Aldrich and
removed by dissolution or calcination (heating in air). A used without further purification. Methanol (99%), ethanol
multitude of inverse opal-based photonic crystal materials have (99.4%), acetone (99%), n-heptane (99.5%), benzene (99.7%),
been synthesized by this strategy, including metals, metal dichloromethane (99%), carbon tetrachloride (99.5%), toluene
oxides, conducting polymers, and carbon materials.29,42−69 The (98%), and bromobenzene (95%) were obtained from local
three-dimensionally ordered macroporous structure of inverse suppliers. Milli-Q water (18.2 MΩ cm resistivity) was used in
opals, together with their inherent photonic crystal properties, all experiments.
creates new opportunities in optical sensor development, 2.2. Synthesis of Monodisperse Poly(methyl meth-
separation, photocatalysis, and catalysis, among others. acrylate) (PMMA) Colloids. Three batches of monodisperse
Particularly valuable in this context are inverse opal thin PMMA colloids with diameters of 317, 360, and 441 nm
films suitable for “on chip” diagnostic or sensing applications, (denoted PMMA #1, PMMA #2, and PMMA #3, respectively,
though reliable design strategies toward large-area inverse opal in the text below) were synthesized by the surfactant-free
thin films are currently limited, prompting further research in emulsion polymerization of methyl methacrylate at temper-
this area. atures between 70 and 80 °C (Table S1, Figure S3), following
In response, our group has been targeting new and approved procedures introduced by Schroden et al. with a few
approaches toward inverse opal thin films, primarily metal modifications.48 Briefly, a volume of methyl methacrylate
oxides42,62−66 and conducting polymers.67−69 Zirconia (ZrO2) (0.3 or 0.4 L depending on the batch) and water (1.6 L) was
attracts particular interest owing to its high thermal stability heated under vigorous mechanical stirring and a N 2
and chemical inertness, leading to end applications in high- atmosphere to the desired reaction temperature (70, 75, or
performance ceramics, catalysis, sensing, and high-temperature 80 °C). Subsequently, 2′-azobis(2-methylpropionamidine)
fuel cell applications.48,49,75−80 Herein, we report the first dihydrochloride (1.5 g) was quickly added to initiate the
successful fabrication and detailed optical characterization of polymerization reaction, which was then continued for 3 h at
ZrO2 inverse opal thin films. Evaluation of the performance of the specified reaction temperature. Finally, the PMMA colloid
the zirconia inverse opal (ZrO2 IO) films for refractive index suspensions were quickly cooled to room temperature, filtered
sensing and catalytic applications was a further aim of the through glass wool, and stored in glass Schott bottles for later
research. At this point, it should be mentioned that most of the use.
methodologies introduced here for the fabrication and 2.3. Fabrication of PMMA Colloidal Crystal Thin
characterization of the ZrO2 IO thin films were developed in Films. PMMA colloidal crystal (PMMA CC) thin films on
9659 DOI: 10.1021/acsomega.8b01334
ACS Omega 2018, 3, 9658−9674
ACS Omega Article

Figure 1. SEM images for the PMMA colloidal crystal films viewed along the [111] direction. (a, b) PMMA CC #1; (c, d) PMMA CC #2; and (e,
f) PMMA CC #3. Images on the left were taken at 20 000× magnification, those on the right at 100 000× magnification. The insets in (a), (c), and
(e) show Fourier transforms of the corresponding image and confirm a high degree of structural order.

glass microscope slides were fabricated using the flow- volume of 500 mL, which was then poured into a vertical-
controlled vertical deposition (FCVD) method outlined in walled glass beaker. Glass microscope slides were then
Figure S4a,b. Briefly, 20−25 mL of the PMMA colloid immersed vertically in the diluted colloid suspension, after
suspensions obtained in Section 2.2 was diluted with water to a which a peristaltic pump slowly removed the colloidal
9660 DOI: 10.1021/acsomega.8b01334
ACS Omega 2018, 3, 9658−9674
ACS Omega Article

Table 1. Summarized Structural and Optical Data for the PMMA Colloidal Crystal Films and Their ZrO2 Inverse Opal
Replicasa
optical data and derived constantsb
−1
sample 2
BET surface area (m g ) DSEM (nm) b
DUV−vis (nm) navg ϕ λmax (nm) θ = 0° Δλ/λmax a/λmax
PMMA CC #1 317 319 1.35 0.72 707 0.073 0.63
PMMA CC #2 360 362 1.36 0.72 802 0.074 0.63
PMMA CC #3 441 439 1.37 0.74 983 0.074 0.64
ZrO2 IO #1 47.7 260 242 1.16 0.16 462 0.125 0.79
ZrO2 IO #2 44.9 287 281 1.15 0.15 526 0.120 0.77
ZrO2 IO #3 47.6 342 345 1.12 0.12 635 0.128 0.77
a
D = center-to-center distance between spheres on fcc (111) planes. bDUV−vis, navg, and ϕ determined from plots of λmax2 versus sin2 θ.

suspension (pump rate about 0.3 mL min−1). As the meniscus dispersion was placed on a holey carbon-coated copper TEM
of the colloid suspension moved down the glass slide, colloidal grid. The ethanol was then allowed to evaporate at room
crystals of thickness 5−7 μm (18−25 layers of colloids) were temperature for several hours before sample imaging.
deposited. In general, the FCVD method proved superior to an X-ray powder diffraction patterns were obtained on a
evaporation method (Figure S4c) also trialed for colloidal Siemens D5000 X-ray diffractometer equipped using a Cu Kα1
crystal fabrication. source (λ = 1.5418 Å, 40 mA, 40 kV). Data were collected over
2.4. Fabrication of Zirconia Inverse Opal Thin Films. the 2θ range of 2−100°, with a step size of 0.02° and a scan
ZrO2 IO thin films were prepared using the colloidal crystal rate of 2° min−1. ZrO2 crystallite sizes (L) were estimated from
templates fabricated above. Briefly, a Zr(IV) sol was first the X-ray powder diffraction data using peak line widths and
prepared by adding Zr(IV) propoxide in 1-propanol (5 mL) to the Scherrer equation.82
5 mL of methanol under vigorous stirring, followed by the XPS data was collected on the soft X-ray Beamline at the
addition of concentrated HCl (2 mL) and water (2 mL). The Australian Synchrotron. The endstation was equipped with a
resulting mixture was stirred for 30 min to give a clear and hemispherical electron energy analyzer and an analysis
homogeneous sol, after which it was diluted 5-fold with chamber of base pressure ∼1 × 10−10 Torr. Spectra were
methanol. For the colloidal crystal infiltration step, the coated excited using monochromatic Al Kα X-rays (1486.69 eV).
glass slides were placed on a slight incline (∼5°), after which Core-level scans were collected over the Zr 3d and Au 4f
3−4 drops of the diluted Zr(IV) sol were placed on the regions with an analyzer pass energy of 20 eV. The binding
uppermost edge of the colloidal crystal film. Gravity and energy scale was calibrated against the C 1s signal at 285.0 eV
capillary action then allowed the sol to infiltrate all the from adventitious hydrocarbons.
interstitial spaces in the film. After hydrolyzing and drying for N2 physisorption isotherms for the ZrO2 inverse opals were
24 h at room temperature, the resulting ZrO2/PMMA CC collected at 77 K on a Micromeritics TriStar 3000. Specific
composites were then calcined using the following protocol: surface areas and pore sizes were calculated from N2
heating from room temperature to 300 °C at 2 °C min−1, physisorption isotherms according to the Brunauer−Em-
holding at 300 °C for 2 h, heating from 300 to 400 °C at 2 °C mett−Teller (BET)83 and Barret−Joyner−Halenda84 methods.
min−1, then holding at 400 °C for 2 h. Finally, the films were Prior to analysis, samples were heated at 180 °C overnight in a
cooled to room temperature and then subjected to detailed vacuum oven.
characterization. UV−vis transmittance spectra for the PMMA colloidal
2.5. Fabrication of Au/ZrO2 IO Catalysts. Gold crystal films and ZrO2 inverse opal films were recorded over
nanoparticles were deposited on ZrO2 IO films at nominal the wavelength range 300−1100 nm on a Shimadzu UV-1700
weight loadings of 5 and 18 wt % using a modified version of spectrometer, equipped with a custom-built sample holder for
the procedure described by Zanella.81 Briefly, the ZrO2 IO UV−vis transmittance measurements at different film angles to
films were immersed in a 100 mL aqueous solution containing the sample beam path.
a certain amount of HAuCl4·3H2O and urea (1 g) and then The photonic band-gap structure diagram for the ZrO2 IO
heated to 80 °C and kept at this temperature for 4 h. By films was created using the MIT photonic-band package. Fully
increasing the amount of HAuCl4·3H2O in the solution, the vectoral eigenmodes of Maxwell’s equations with periodic
amount of gold deposited on the films could be systematically boundary conditions were computed by preconditioned
increased. After reaction for 4 h, the films were rinsed with conjugate-gradient minimization of the block Rayleigh
water, dried in air at 90 °C, and then heated to 300 °C for 1 h quotient in a planewave basis, along the high-symmetry
to reduce adsorbed Au(OH)3 or Au2O3 to metallic Au0 directions U-L-Γ-X-W-K for an fcc first Brilluoin zone.85 The
nanoparticles. additional parameters for ZrO2 calculations are as follows:
2.6. Characterization. SEM micrographs were taken using rsphere = 1.06 and neff = 2.1.
a Philips XL-30S Field Emission Gun scanning electron 2.7. Catalytic Activity Tests. The reduction of 4-
microscope (FEGSEM) operated at an electron accelerating nitrophenol at room temperature with sodium borohydride
voltage of 5 kV in high vacuum. Prior to analysis, specimens (NaBH4) as the reductant was used to test the activity of the
were mounted on black carbon tape and sputter coated with ZrO2 IO and Au/ZrO2 IO catalysts. In a typical experiment,
platinum for 60 s with a Quorum Q150RS. catalyst (2 mg) was added to aqueous 4-nitrophenol (0.15
TEM images were collected on a Tecnai T12 transmission mM, 25 mL) with stirring. Then, 2.75 mL of the resulting
electron microscope operating at an accelerating voltage of 120 dispersion was added to a standard quartz cuvette (path length
kV, equipped with a 4 megapixel camera. Specimens were 1 cm). To start the reaction, 0.25 mL of freshly prepared
dispersed in ethanol, and then a drop of the resulting NaBH4 (0.1 mol L−1) was then added. The reaction progress
9661 DOI: 10.1021/acsomega.8b01334
ACS Omega 2018, 3, 9658−9674
ACS Omega Article

Figure 2. UV−vis transmittance spectra collected in air at incident angles of 0, 5, 10, 15, 20, and 25° with respect to the [111] direction for (a)
PMMA CC #1; (b) PMMA CC #2; and (c) PMMA CC #3. As the incident angle of light increased from 0 to 25°, the [111] PBGs blue-shifted in
accordance with eq 2. (d) Plot of λmax2 versus sin2 θ for the PMMA CC films.

was monitored by collecting UV−vis absorption spectra every the [111] direction for films fabricated from each PMMA
30 s from 250 to 600 nm. The absorbance at 400 nm was colloid batch. In each case, the films crystallized with their fcc
monitored as a function of time, and pseudo-first-order rate (111) planes parallel to the underlying glass substrate. Further,
constants (k′) for the reduction of 4-NP were calculated from each batch was found to be highly monodisperse, with all
the slopes of plots of ln(At/A0) versus t. colloids in each batch having the same size resulting in very
sharp spots in the Fourier transforms performed on the SEM
3. RESULTS images (see insets in Figure 1a,c,e). In the higher magnification
3.1. Structural and Optical Characterization of the images, small necks can be seen between neighboring colloids,
PMMA Colloidal Crystal Films. The colloidal crystal which likely formed during the drying of the colloidal crystals.
template technique relies heavily on the synthesis of This necking adds a degree of mechanical strength to the
monodisperse spherical colloids and their subsequent self- PMMA colloidal crystals, which is beneficial for colloidal
assembly to form well-ordered colloidal crystals. Accordingly, crystal templating (see below). The diameters of the colloids,
preliminary SEM and UV−vis characterization studies were calculated as the center-to-center distance between adjacent
performed on PMMA CC #1, PMMA CC #2, and PMMA CC colloids on the (111) planes, were 317, 360, and 441 nm for
#3 thin films to assess their suitability as templates for ZrO2 PMMA CC #1, PMMA CC #2, and PMMA CC #3,
inverse opal fabrication. Figure 1 shows the SEM images along respectively (Table 1). The colloidal crystal films developed
9662 DOI: 10.1021/acsomega.8b01334
ACS Omega 2018, 3, 9658−9674
ACS Omega Article

Figure 3. SEM images for ZrO2 inverse opal films viewed along the [111] direction; (a, b) ZrO2 IO #1; (c, d) ZrO2 IO #2; and (e, f) ZrO2 IO #3.
Images on the left were taken at 50 000× magnification and those on the right at 100 000× magnification. The insets in (a), (c), and (e) show
digital photographs of the ZrO2 inverse opal thin films in air, illuminated and viewed along the [111] direction.

some cracks during fabrication and SEM imaging (Figure S5a) Colloidal crystal film thicknesses were typically in the range
but generally had very acceptable structural uniformity. of 5−7 μm (Figures S4d and S5b).
9663 DOI: 10.1021/acsomega.8b01334
ACS Omega 2018, 3, 9658−9674
ACS Omega Article

Figure 4. (a−c) UV−vis transmittance spectra collected in air for the ZrO2 inverse opal films at incident angles of 0−25° with respect to the [111]
direction. The spectra were collected in 5° increments. (d) Plot of λmax2 versus sin2 θ for the ZrO2 inverse opal films. (e) Digital photograph of
ZrO2 IO #2 illuminated and viewed along the surface normal, and (f) the same film illuminated along the surface normal but viewed at an angle of
around 30° with respect to the surface normal.

9664 DOI: 10.1021/acsomega.8b01334


ACS Omega 2018, 3, 9658−9674
ACS Omega Article

crystals (Figure S2). For first-order diffraction on any fcc plane


of an opal-based photonic crystal (including inverse opals), the
position of the PBG can be described by the following
equation

λmax = 2dhkl navg 2 − sin 2 θ (1)

where λmax is the PBG position (in nm), dhkl is the interlayer
spacing for a particular (hkl) plane, θ is the incident angle of
light with respect to the normal of the plane, and navg is the
average refractive index of the photonic crystal, typically
calculated as navg = [ϕnsolid + (1 − ϕ)nvoid], where ϕ is the solid
volume fraction and nsolid and nvoid are the refractive indices of
the solid and void components, respectively. Note for a PMMA
colloidal crystal in air with an ideal solid volume fraction of
0.74, navg = (0.74 × 1.492 + 0.26 × 1.00) = 1.364. For first-
order diffraction on fcc (111) planes, eq 1 transforms to eq 2
since d111 = 0.8165D.

λmax = 1.633D navg 2 − sin 2 θ (2)

When θ = 0°, eq 2 reduces further to


λmax = 1.633Dnavg (3)

Thus, the PBG position for first-order diffraction along the


[111] direction should be directly proportional to D (i.e., λmax
= 2.23 × D), in excellent accord with the shift in the PBG
position to longer wavelengths seen on going from PMMA CC
#1 to PMMA CC #3 (Figure 2 and Table 1). The normalized
PBG bandwidths (Δλ/λmax) and PBG angular frequencies (a/
λmax) were near identical for all three different colloidal crystals,
consistent with their identical construction (differing only in
the colloid diameter). As predicted by eq 2, the PBGs along
the [111] direction blue-shifted as the incident angle of light
with respect to the [111] direction increased (Figure 2). Plots
of λmax2 versus sin2 θ yielded straight lines (r2 > 0.995, Figure
Figure 5. (a) Plot of position of the photonic band gap along the 2d). Values of D, navg, and ϕPMMA determined from the slopes
[111] direction and the normalized band-gap width (Δλ/λ) versus (−1.6332D2) and y-axis intercept (1.6332D2navg2) for the
the center-to-center distance between macropores (D) on fcc (111) colloidal crystal films were in excellent accord with expect-
planes in the ZrO2 inverse opal films. (b) Photonic band-gap structure ations (Table 1, columns 4−6).
diagram for a ZrO2 inverse opal, showing a pseudo photonic band gap 3.2. Structural and Optical Characterization of ZrO2
at a/λ = 0.77 (indicated by the blue shaded area) between the second Inverse Opal Films. After the colloidal crystal templating
and third bands along the fcc L → Γ direction. step, brightly colored films were obtained, indicating the
success of the inversion process. Whereas the PMMA colloidal
The PMMA colloidal crystals all showed structural color, crystals (opals) were constructed from solid spheres with air in
especially when viewed at grazing angles that shifted their between, the inverse opal replicas comprised a face-centered
photonic band gaps (PBGs) along the [111] and [200] cubic array of air spheres (macropores) in a ZrO2 matrix
directions into the visible region. UV−vis transmittance spectra (Figure 3). When viewed along the [111] direction, the ZrO2
for the PMMA colloidal crystal films collected along the [111] IO #1, ZrO2 IO #2, and ZrO2 IO #3 films appeared purple,
direction, that is, normal to the (111) planes, are shown in green, and orange, respectively. UV−vis transmittance spectra
Figure 2. PMMA CC #1, PMMA CC #2, and PMMA CC #3 collected along the [111] direction (θ = 0°) revealed band
gave intense PBGs along the [111] direction (θ = 0°) at 707, gaps at 462, 526, and 635 nm (Figure 4), consistent with the
802, and 983 nm, respectively (Table 1). Note that here we reflected color of each film. In agreement with eq 2, the PBG
have collected transmittance spectra rather than reflectance along the [111] direction shifted to shorter wavelengths as the
spectra, and as such the PBGs will appear as sharp dips in incident angle of light was increased with respect to the [111]
transmittance (by convention we still use the term λmax here to direction. Values of D, navg, and ϕZrO2 obtained from plots of
describe the PBG position). The shift in the PBG position λmax2 versus sin2 θ (Figure 4d) are listed in Table 1. Figure 4e,f
(λmax) to longer wavelengths with increasing colloid diameter shows a ZrO2 IO #2 film illuminated and viewed along the
(D), and hence increasing interlayer spacing along the [111] [111] direction and at an angle of 30° with respect to the angle
direction, is in perfect accord with predictions of the modified of illumination, respectively. Note the striking color change
Bragg’s law expressions below,48,63 which consider both with anglegraphic evidence for the successful realization of
refraction and diffraction of light in opal-based photonic structural color in the films.
9665 DOI: 10.1021/acsomega.8b01334
ACS Omega 2018, 3, 9658−9674
ACS Omega Article

Figure 6. (a, b) UV−vis transmittance spectra collected along the [111] direction for different ZrO2 inverse opal thin films in air (n = 1.000) and
solvents of increasing refractive index (ranging from methanol, n = 1.329, to bromobenzene, n = 1.558). For both samples, the PBGs progressively
red-shifted as the refractive index of the solvent filling the macropores in the films increased. Insets in (a) and (b) shows contact angle data for a
water droplet on each ZrO2 IO film. The digital photographs below the spectra show each ZrO2 IO film in air and ethanol, separately.

As with the PMMA colloidal crystal templates, the PBG templates (Figure 3). Further evidence in support of this
position along the [111] direction for the ZrO2 IOs shifted to theory is found in the SEM images of Figure S5c−f, where an
longer wavelengths as D increased (Table 1 and Figure 5a). island of ZrO2 IO is seen (Figure S5e). In addition, some
The normalized PBG bandwidths (Δλ/λmax) and PBG angular cracks and other defects can be seen in film cross sections
frequencies (a/λmax) were also very similar for the three (Figure S5c,d), although these defects seem to have little
different samples, consistent with expectations for ZrO2 IOs of detrimental effect on the optical properties of the ZrO2 IO #2
comparable solid volume fraction (0.12−0.16). The a/λmax film.
value of 0.77 found for ZrO2 IO #2 and ZrO2 IO #3 agreed ZrO2 can exist in three different polymorphic forms,
perfectly with the value predicted in the photonic band-gap monoclinic, tetragonal, and cubic, with the cubic form being
structure diagram for a ZrO2 IO with a solid volume fraction of the most thermodynamically stable.86,87 Powder X-ray
around 0.12 (Figure 5b). In the diagram, a pseudo photonic diffraction patterns (Figure S6) for the ZrO2 IOs confirmed
band gap opens between the second and third bands along the that all were composed of tetragonal ZrO2 (space group P42/
L → Γ or [111] direction at a/λmax = 0.77, as indicated by the nmc), consistent with the literature related to ZrO2 crystals
blue shaded area. obtained by mild calcination of amorphous zirconia
It should be noted here that on transforming the PMMA precursors.86,87 The broadness of the XRD peaks is evidence
colloidal crystals to their inverse opal replicas, the center-to- for the presence of nanocrystalline ZrO2 in the walls of the
center distances between spheres on the fcc (111) planes inverse opals. Using the Scherrer equation,82 ZrO2 crystallite
decreased by ∼18−22%. The theoretical maximum solid sizes were estimated to be 3−5 nm. The ring structures seen in
volume fraction for the inverse opals is 0.26, but the actual the selected area diffraction pattern for ZrO2 IO #2 (Figure 8a)
values were approximately one half of this (ϕZrO2 = 0.12−0.16). validated this size estimate.
The data suggests that the during the template removal 3.3. Refractive Index Sensing Using ZrO2 Inverse
process, the films cracked and shrunk to form small islands on Opal Films. One of the most fascinating and intriguing
the glass microscope slides, though clearly the islands retained features of the ZrO2 IO thin films was their dramatic color
the desirable (111) orientation of the PMMA colloidal crystal change on being wetted by water or other liquids. As indicated
9666 DOI: 10.1021/acsomega.8b01334
ACS Omega 2018, 3, 9658−9674
ACS Omega Article

refractive index (Figure 7). The plots for all three ZrO2 IO
samples were remarkably linear (r2 > 0.999). This linearity can
be understood by substituting [ϕnsolid + (1 − ϕ)nvoid] for navg
in eq 3, which leads to the following equation
λmax = 1.633DϕnZrO2 + 1.633D(1 − ϕ)n void (4)
Here, nvoid represents the refractive index of the medium (i.e.,
solvent) filling the void space in the inverse opals. In
agreement with experimental observations, a plot of λmax
versus nsolvent is predicted to be linear, with slope = 1.633D(1
− ϕ) and y-axis intercept = 1.633Dϕ nZrO2. Using the slopes
and intercepts of the straight lines in Figure 7, we calculated
values of D and ϕZrO2 for the ZrO2 IO #1, ZrO2 IO #2, and
ZrO2 IO #3 films using the following expressions48,63
slope × nZrO2 + intercept
D=
1.633 × nZrO2 (5)
intercept
ϕ=
slope × nZrO2 + intercept (6)

Figure 7. Plot of the PBG position (λmax) along the [111] direction Using nZrO2 = 2.1, the D values obtained were 251, 284, and
versus the refractive index (nsolvent) of the solvent filling the 331 nm, respectively, and the ϕZrO2 values obtained were 0.158,
macropores in the ZrO2 inverse opal thin films. For all samples, a
strong linear relationship was observed (r2 > 0.999), in agreement 0.122, and 0.115, respectively, all in excellent accord with the
with eq 4. values determined for the same parameters using other
methods (Table 1). The very strong linearity of the plots in
Figure 7 suggested the ZrO2 IO films could be used in
in eqs 1−3, the photonic band-gap position (λmax) is highly refractive index sensing. To determine the refractive index of
dependent on the average refractive index of the photonic an unknown liquid, all that would be required is to determine
crystal. On filling the macropores in the ZrO2 IO films with the PBG position along the [111] direction using one of the
liquids of increasing refractive index, navg will increase, and thus ZrO2 IO films. Once the PBG position was determined, the
the PBG along the [111] direction is predicted to shift refractive index of the unknown liquid could be extrapolated
progressively to longer wavelengths (i.e., red shift). Figure 6a,b from the plot in Figure 7 for that same ZrO2 IO or alternatively
shows UV−vis transmittance spectra collected for ZrO2 IO #1 calculated directly using the linear regression equation.
and ZrO2 IO #2 films, respectively, in air and then liquids of To test the sensitivity of the ZrO2 IO films for refractive
increasing refractive index. Liquids used were methanol (n = index sensing, we collected UV−vis spectra for the different
1.329), ethanol (n = 1.361), n-heptane (n = 1.388), films in acetone (n = 1.3587) and ethanol (n = 1.3614), which
dichloromethane (n = 1.4241), carbon tetrachloride (n = differ by only 0.0027 refractive index units. For the ZrO2 IO #1
1.4601), benzene (n = 1.501), and bromobenzene (1.558). As film, a PBG shift to longer wavelengths of 1.2 nm was detected
the refractive index of the liquid increased, the PBG along the on going from acetone to ethanol, with a slightly large shift of
[111] direction shifted progressively to longer wavelengths. 1.4 nm found for the ZrO2 IO #2 film. Based on this result, we
The large change in the color of the ZrO2 IO #1 and ZrO2 IO thus estimate that the ZrO2 IO films should be able to discern
#2 films on going from air to ethanol is apparent in the digital two liquids with a refractive index difference as small as 0.0005.
photos provided in Figure 6. ZrO2 IO #1 changed from purple 3.4. Optical Properties of Gold Nanoparticle-Deco-
to yellow, and ZrO2 IO #2 changed from green to red. Once rated ZrO2 Inverse Opal Films. Fascinated by the optical
the ethanol evaporated from the films, the original purple and properties of the bare ZrO2 inverse opal films, we decided to
green colors of the films returned. explore this aspect further by decorating the films with gold
Surprisingly, the ZrO2 IO films were quite hydrophobic, (Au) nanoparticles. Unlike bulk gold, which is yellow, gold
giving water droplet contact angles ranging from 120 to 124° nanoparticles display a range of size-dependent colors due to
(Figure 6a,b, insets). These values were independent of the localized surface plasmon resonances (LSPRs).63,91,92 The
macropore diameter in the ZrO2 IOs. By comparison, SiO2 IO LSPRs are a collective oscillation of valence electrons on the
and TiO2 IO samples prepared by the same colloidal crystal surface of the gold nanoparticles induced by the electric field of
templating method (discussed briefly below) were hydrophilic light at the resonance frequency. For spherical gold nano-
and easily wetted by water. This result suggests that the particles of size <20 nm in water, the LSPR absorption
combination of inverse opal structure (i.e., very spherical and maximum is typically observed around 520 nm (i.e., the
regularly sized macropores) and the unique physical properties nanoparticles absorb light strongly in the green part of the
of the very small ZrO2 crystallites in the walls of the inverse visible spectrum), and we see the other reflected wavelengths
opals was responsible for the unexpected hydrophobicity (especially the red part of the visible spectrum). For this
displayed by the ZrO2 IO films.88−90 reason, a vial of small spherical Au NPs in water will look red
Returning to the data in Figure 6a,b, we noted the PBG to the human eye. Similarly, stained glass windows containing
positions for each ZrO2 IO in air and each liquid, which were Au nanoparticles appear red to the observer. We were
then used to create plots of PBG position versus solvent particularly interested in decorating the ZrO2 IO #2 film
9667 DOI: 10.1021/acsomega.8b01334
ACS Omega 2018, 3, 9658−9674
ACS Omega Article

Figure 8. (a, b) TEM images for ZrO2 IO #2 and (c, d) 5 wt % Au/ZrO2 IO #2. The images in (c) and (d) reveal well-dispersed Au nanoparticles
of size around 5−7 nm; (e, f) shows digital photographs of the 5 wt % Au/ZrO2 IO #2 film in air and after wetting with ethanol. The inset in (a)
shows a selected area diffraction pattern (scale bar = 0.2 Å−1). The well-developed ring structure confirms the presence of small nanocrystals of
tetragonal ZrO2 (P42/nmc).

with gold nanoparticles since that film has a PBG along the expected to absorb. This offered the opportunity to study two
[111] direction at 526 nm where the gold nanoparticles were competing optical phenomena (selective diffraction and
9668 DOI: 10.1021/acsomega.8b01334
ACS Omega 2018, 3, 9658−9674
ACS Omega Article

small nanocrystalline tetragonal ZrO2 crystallites (<5 nm).


After decoration with Au nanoparticles at a loading of 5 wt %,
small Au nanoparticles of size 5−8 nm could be discerned in
the TEM images. The small size and excellent dispersion of the
Au nanoparticles over the ZrO2 IO support are consistent with
the known strong metal−support interaction between Au and
ZrO2. The XRD pattern for the 5 wt % Au/ZrO2 IO #2 film
(Figure S8) confirmed the successful deposition of Au
nanoparticles, with new broad peaks characteristic for small
Au crystals appearing (especially apparent in the XRD pattern
of the 18 wt % Au/ZrO2 IO #2 film, Figure S8). XPS data in
Figure S9 also confirmed the successful deposition of metallic
Au nanoparticles, with the Au 4f signal intensities increasing in
proportion to the nominal Au loadings. As expected, the Zr 3d
signals were attenuated slightly as the Au loading increased.
When illuminated and viewed along the [111] direction, the
Au/ZrO2 IO #2 film appeared green in air (Figure 8e). Thus,
the photonic crystal was rejecting wavelengths of light typically
used by the Au nanoparticles for absorption (i.e., the PBG
properties of the ZrO2 IO film appeared to dominate the
perceived color along the [111] direction; note: a noninverse
opal Au/ZrO2 specimen would certainly appear purple under
the same illumination conditions). On adding a drop of
ethanol to the film, the PBG for the 5 wt % Au/ZrO2 IO #2
film was red-shifted considerably and decoupled from the Au
LSPR, resulting in a brightly red-colored film (Figure 8f)
whose color is the sum of both PBG diffraction and Au LSPR
absorption.
The LSPRs of Au nanoparticles are known to be sensitive to
the refractive index of the surrounding medium91,92 and have
been used previously to develop refractive index sensors for
liquids. The data in Figures 6 and 7 revealed that the ZrO2 IO
films were extremely effective for refractive index sensing.
Accordingly, it was of novelty to see if the 5 wt % Au/ZrO2 IO
#2 film could be used to get a double determination of liquid
refractive index from a single UV−vis spectrum. Figure 9a
shows UV−vis transmittance spectra for the 5 wt % Au/ZrO2
IO #2 film in air (n = 1.000), methanol (n = 1.329), ethanol (n
= 1.362), n-heptane (n = 1.388), dichloromethane (n = 1.424),
carbon tetrachloride (n = 1.460), and benzene (n = 1.501). In
air, the ZrO2 IO #2 PBG along the [111] direction and the Au
LSPR absorption were almost perfectly overlapped, giving a
Figure 9. (a) UV−vis transmittance spectra collected along the [111]
broad feature centered at 532 nm (i.e., responsible for the
direction for a 5 wt % Au/ZrO2 IO #2 thin film in air (n = 1.000) and green reflected color seen in Figure 8e). On immersion in the
solvents of increasing refractive index (ranging from methanol, n = various organic solvents, the Au LSPR and ZrO2 IO PBG
1.329, to benzene, n = 1.501). Both the PBG and the Au LSPR decoupled, giving two distinct features. The Au LSPR feature
progressively red-shifted linearly as the refractive index of the solvent was centered between 545 and 556 nm, and the ZrO2 IO #2
in the film increased, as shown in (b). The inset in (b) shows how the PBG centered between 663 and 754 nm. In both cases, these
hydrophobicity of the ZrO2 IO #2 film decreased after Au features shifted linearly to longer wavelengths on increasing the
nanoparticle deposition (top = before Au deposition; bottom = refractive index of the liquid the film was immersed in (Figure
after Au deposition).
9b). Accordingly, it can be concluded that the 5 wt % Au/ZrO2
IO #2 film allows two parallel determinations of the refractive
reflection by the ZrO2 IO #2 film and strong absorption at index of a liquid from a single UV−visible transmittance
comparable wavelengths by the Au nanoparticles). spectrum measurement. Using the slope and intercept of the
Figure 8 shows transmission electron microscopy (TEM) PBG regression line in Figure 9b, values of D = 288.4 nm and
images for the ZrO2 IO #2 film (Figure 8a,b) and the 5 wt % ϕZrO2 = 0.11 were calculated using eqs 5 and 6, respectively
Au/ZrO2 IO #2 film (Figure 8c,d). In both cases, the films (nZrO2 = 2.1 was assumed). These values are in good agreement
were scraped off the underlying glass substrate for imaging,
which did not adversely affect the ordered structure of the with data for bare ZrO2 IO #2 (D = 287 nm, ϕZrO2 = 0.15,
films. The selected area diffraction pattern for the ZrO2 IO #2 Table 1), indicating that gold Au nanoparticle deposition did
film showed a characteristic ring structure, which supported not really alter the geometric properties of the ZrO2 IO film
the earlier XRD findings that the films were composed of very significantly.
9669 DOI: 10.1021/acsomega.8b01334
ACS Omega 2018, 3, 9658−9674
ACS Omega Article

Figure 10. UV−vis absorbance spectra for aqueous solutions of 4-nitrophenol (0.15 mM, 2.75 mL) and NaBH4 (0.1 M, 0.25 mL) in the presence
of (a) ZrO2 IO #2; (b) 5 wt % Au/ZrO2 IO #2; and (c) 18 wt % Au/ZrO2 IO #3. (d) Plots of ln(At/A0) versus time.

3.5. Catalytic Activity of the Gold Nanoparticle- where At is the absorbance of the 4-nitrophenol solution in
Decorated ZrO2 Inverse Opal Films for 4-Nitrophenol water at 400 nm at time t, A0 is the initial absorbance of the 4-
Reduction. As a further demonstration of the functionality of nitrophenol solution in water at 400 nm (i.e., at t = 0 min), k′
the Au/ZrO2 IO #2 films fabricated in this study, we decided is the pseudo-first-order rate constant (units min−1), and t is
to evaluate the catalytic activity of the as-prepared films, as well the reaction time (in min). The reaction commences as soon
as films decorated with Au nanoparticles at two different as the reducing agent (NaBH4) is added to the 4-nitrophenol
loadings (5 and 18 wt %), for the reduction of 4-nitrophenol solution containing the catalyst. Figure 10a shows the UV−vis
using NaBH4 at the reducing agent.93,94 In the presence of a absorption spectra for an aqueous solution containing 4-
large excess of NaBH4, the kinetics of this reaction, which is nitrophenol, NaBH4, and ZrO2 IO. A gradual decrease in the
catalyzed by metal nanoparticles, obeys pseudo-first-order absorbance of the 4-nitrophenol peak at 400 nm is seen over
kinetics, as described by the following integrated rate law time (24 min), which can largely be attributed to 4-

ji A zy
nitrophenol adsorption on the ZrO2 IO surface (rather than

lnjjj t zzz = −k′t


catalytic reduction). The rate constant of 0.026 min−1 was
j A0 z
k {
understandably low. After decoration with Au nanoparticles,
(7) known active catalysts for 4-nitrophenol reduction in the
9670 DOI: 10.1021/acsomega.8b01334
ACS Omega 2018, 3, 9658−9674
ACS Omega Article

presence of NaBH4, the rate constant jumped dramatically to position along the [111] direction with increasing solvent
k′ = 0.740 min−1 for the 5 wt % Au/ZrO2 IO #2 sample and refractive index. Further, by functionalizing the ZrO2 IO films
then to k′ = 1.927 min−1 for the 18 wt % Au/ZrO2 IO #2 with gold nanoparticles, a novel refractive index sensing
sample. The evolution of the main reaction product, 4- platform was created that allowed parallel determinations of
aminophenol, during the reaction is evident by the develop- liquid refractive index (one strategy used the small shift in the
ment of the peak at 295 nm in the spectra. In the case of the 18 Au LSPR position, and the other utilized the large shift in the
wt % Au/ZrO2 IO #2 sample, the reaction proceeded so PBG position as the refractive index of the solvent in the Au/
rapidly that it was almost impossible to follow by UV−vis ZrO2 IO films increased). Due to their inherent macroporosity,
spectroscopy (reaching completion in 2 min). The rates ZrO2 IOs represent near-ideal supports for noble metal
reported here accord well with the best literature data for 4- catalysts, which we confirmed through the successful
nitrophenol reduction over the supported gold catalysts.94−97 application of Au/ZrO2 IO films to the reduction of aqueous
This high activity can be understood in terms of the open 4-nitrophenol in the presence of NaBH4. Results of this study
macroporous structure of the ZrO2 IO support that allows are expected to encourage others to utilize the colloidal crystal
facile transport of reagents to the Au active sites. The high gold template approach to fabricate novel photonic crystal thin films
nanoparticle dispersion realized in the Au/ZrO2 IO system for sensing and catalytic applications.
(Figure 8c,d) is also advantageous in this context. Supported
Au nanoparticle catalysts offer many advantages over
unsupported Au nanoparticles in catalytic applications,

*
ASSOCIATED CONTENT
S Supporting Information
especially the stabilization afforded against nanoparticle The Supporting Information is available free of charge on the
aggregation. After storage for 1 year at room temperature in ACS Publications website at DOI: 10.1021/acsome-
closed vials, the 5 wt % Au/ZrO2 IO #2 and 18 wt % Au/ZrO2 ga.8b01334.
IO #2 catalysts demonstrated identical performance as the Photos and SEM data for the selection of natural
freshly prepared catalysts for 4-nitrophenol reduction. photonic crystals; details of the equipment setups used
3.6. Author Commentary on the Versatility of to synthesize the colloidal crystals and colloidal crystal
Colloidal Crystal Template Strategy. The data above thin films; additional SEM data for PMMA colloidal
demonstrates the wonderful structural color palette that is crystal films and their inverse opal replicas; UV−vis
available by fabricating a single type of inverse opal film. By transmittance spectra for the ZrO2 IO #1 and ZrO2 IO
varying the diameter of the PMMA colloids in the colloidal #2 films in acetone and ethanol; XRD patterns for the
crystal templates, it is possible to controllably modify the ZrO2 IO films and Au/ZrO2 IO catalysts; XPS data for
position of PBGs along the [111] direction in the ZrO2 IO
Au/ZrO2 IO catalysts; SEM data for SiO2 and TiO2
films. By rotating the samples with respect to the source of
inverse opals (PDF)


illumination, the PBG along the [111] direction can be shifted
to shorter wavelengths, and by filling the macropores in the
inverse opals with liquids of different refractive indices, it is AUTHOR INFORMATION
possible to redshift the PBG over a large range of wavelengths. Corresponding Author
Through systematic manipulation of all of these parameters in *E-mail: g.waterhouse@auckland.ac.nz. Tel: 64-9-9237212.
the current study, we have been able to recreate, via structural Fax: 64-9-373 7422.
engineering, all of the vivid colors exhibited by the natural ORCID
photonic crystal specimens in Figure S1. Importantly, all of the Geoffrey I. N. Waterhouse: 0000-0002-3296-3093
optical phenomena we have reported relating to the ZrO2 IO Notes
films are accurately captured by the modified Bragg’s law The authors declare no competing financial interest.


expressions described in eqs 1−3. Further, by simply
substituting the Zr(IV) propoxide precursor with tetraethylor- ACKNOWLEDGMENTS
thosilicate or titanium (IV) propoxide and changing the
Funding support from the MacDiarmid Institute for Advanced
solvent from methanol to ethanol, keeping all other fabrication
Materials and Nanotechnology and the Dodd Walls Centre for
procedures the same, amorphous SiO2 IO or nanocrystalline
Photonic and Quantum Technologies is gratefully acknowl-
TiO2 IO films, respectively, can readily be accessed (Figure
edged.


S10). All of the aformentioned aspects make colloidal crystal
templating a fascinating topic for advanced-level research as REFERENCES
well as undergraduate teaching laboratories. We encourage
(1) Vukusic, P.; Sambles, J. R. Photonic Structures in Biology.
others to enter this research space. Nature 2003, 424, 852−855.
(2) Kinoshita, S.; Yoshioka, S. Structural Colors in Nature: The Role
4. CONCLUSIONS of Regularity and Irregularity in the Structure. ChemPhysChem 2005,
ZrO2 inverse opal thin films with striking angle-dependent 6, 1442−1459.
structural color at visible wavelengths were successfully (3) Lenau, T.; Barfoed, M. Colours and Metallic Sheen in Beetle
fabricated by the colloidal crystal template technique. The Shells − A Biomimetic Search for Material Structuring Principles
optical properties of the films were dominated by first-order Causing Light Interference. Adv. Eng. Mater. 2008, 10, 299−314.
(4) Khudiyev, T.; Dogan, T.; Bayindir, M. Biomimicry of
diffraction on fcc (111) planes, with excellent accord Multifunctional Nanostructures in the Neck Feathers of Mallard
established between the optical properties of the films and (Anas platyrhynchos L.) Drakes. Sci. Rep. 2014, 4, No. 4718.
predictions based on a modified Bragg’s law expression. (5) Hsiung, B.-K.; Siddique, R. H.; Jiang, L.; Liu, Y.; Lu, Y.;
Refractive index sensing with a sensitivity estimated around Shawkey, M. D.; Blackledge, T. A. Tarantula-Inspired Noniridescent
∼0.0005 units was demonstrated using the ZrO2 inverse opal Photonics with Long-Range Order. Adv. Opt. Mater. 2017, 5,
thin films, based on the progressive linear shift in the PBG No. 1600599.

9671 DOI: 10.1021/acsomega.8b01334


ACS Omega 2018, 3, 9658−9674
ACS Omega Article

(6) Naleway, S. E.; Porter, M. M.; McKittrick, J.; Meyers, M. A. (31) Razo, D. A. S.; Pallavidino, L.; Garrone, E.; Geobaldo, F.;
Structural Design Elements in Biological Materials: Application to Descrovi, E.; Chiodoni, A.; Giorgis, F. A Version of Stöber Synthesis
Bioinspiration. Adv. Mater. 2015, 27, 5455−5476. Enabling the Facile Prediction of Silica Nanospheres Size for the
(7) Zhang, W.; Gu, J.; Liu, Q.; Su, H.; Fan, T.; Zhang, D. Butterfly Fabrication of Opal Photonic Crystals. J. Nanopart. Res. 2008, 10,
Effects: Novel Functional Materials Inspired from the Wings Scales. 1225−1229.
Phys. Chem. Chem. Phys. 2014, 16, 19767−19780. (32) Bogush, G. H.; Tracy, M. A.; Zukoski, C. F. Preparation of
(8) Galusha, J. W.; Richey, L. R.; Jorgensen, M. R.; Gardner, J. S.; Monodisperse Silica Particles: Control of Size and Mass Fraction. J.
Bartl, M. H. Study of Natural Photonic Crystals in Beetle Scales and Non-Cryst. Solids 1988, 104, 95−106.
Their Conversion into Inorganic Structures via a Sol−Gel Bio- (33) Gu, Z.-Z.; Chen, H.; Zhang, S.; Sun, L.; Xie, Z.; Ge, Y. Rapid
Templating Route. J. Mater. Chem. 2010, 20, 1277−1284. Synthesis of Monodisperse Polymer Spheres for Self-Assembled
(9) Mille, C.; Tyrode, E. C.; Corkery, R. W. Inorganic Chiral 3-D Photonic Crystals. Colloids Surf., A 2007, 302, 312−319.
Photonic Crystals with Bicontinuous Gyroid Structure Replicated (34) Cong, H.; Yu, B.; Tang, J.; Lia, Z.; Liu, X. Current Status and
from Butterfly Wing Scales. Chem. Commun. 2011, 47, 9873−9875. Future Developments in Preparation and Application of Colloidal
(10) Xu, J.; Guo, Z. Biomimetic Photonic Materials with Tunable Crystals. Chem. Soc. Rev. 2013, 42, 7774−7800.
Structural Colors. J. Colloid Interface Sci. 2013, 406, 1−17. (35) Wang, J.; Wen, Y.; Ge, H.; Sun, Z.; Zheng, Y.; Song, Y.; Jiang,
(11) Parker, R. M.; Guidetti, G.; Williams, C. A.; Zhao, T.; L. Simple Fabrication of Full Color Colloidal Crystal Films with
Narkevicius, A.; Vignolini, S.; Frka-Petesic, B. The Self-Assembly of Tough Mechanical Strength. Macromol. Chem. Phys. 2006, 207, 596−
Cellulose Nanocrystals: Hierarchical Design of Visual Appearance. 604.
Adv. Mater. 2018, 30, No. 1704477. (36) Han, M. G.; Shin, C. G.; Jeon, S.-J.; Shim, H. S.; Heo, C. J.; Jin,
(12) Zhao, H.; Guo, L. Nacre-Inspired Structural Composites: H.; Kim, J. W.; Lee, S. Full Color Tunable Photonic Crystal from
Performance-Enhancement Strategy and Perspective. Adv. Mater. Crystalline Colloidal Arrays with an Engineered Photonic Stop-Band.
2017, 29, No. 1702903. Adv. Mater. 2012, 24, 6438−6444.
(13) Zi, J.; Yu, X.; Li, Y.; Hu, X.; Xu, C.; Wang, X.; Liu, X.; Fu, R. (37) Zhang, C.; Sun, S.; Zhang, Y.; Zhan, Y.; Zhao, H.; Chen, J.; Xu,
Coloration Strategies in Peacock Feathers. Proc. Natl. Acad. Sci. U.S.A. L. Template-assisted Chemical Synthesis of Au Opal Photonic Crystal
2003, 100, 12576−12578. Film with Complete Photonic Band Gaps in the Visible. Mater. Lett.
(14) Sanders, J. V. Colour of Precious Opal. Nature 1964, 204, 2014, 131, 272−275.
1151−1153. (38) Zhang, J.; Sun, Z.; Yang, B. Self-Assembly of Photonic Crystals
(15) Yablonovitch, E. Inhibited Spontaneous Emission in Solid-State from Polymer Colloids. Curr. Opin. Colloid Interface Sci. 2009, 14,
Physics and Electronics. Phys. Rev. Lett. 1987, 58, 2059−2062. 103−114.
(16) Yablonovitch, E.; Gmitter, T. J. Photonic Band Structure: The (39) von Freymann, G.; Kitaev, V.; Lotsch, B. V.; Ozin, G. A.
Face-Centered-Cubic Case. Phys. Rev. Lett. 1989, 63, 1950−1953. Bottom-Up Assembly of Photonic Crystals. Chem. Soc. Rev. 2013, 42,
(17) Yablonovitch, E. Photonic Band-Gap Structures. J. Opt. Soc.
2528−2554.
Am. B 1993, 10, 283−295.
(40) Cai, Z.; Liu, Y. J.; Teng, J.; Lu, X. Fabrication of Large Domain
(18) John, S. Strong localization of Photons in Certain Disordered
Crack-Free Colloidal Crystal Heterostructures with Superposition
Dielectric Superlattices. Phys. Rev. Lett. 1987, 58, 2486−2489.
Bandgaps Using Hydrophobic Polystyrene Spheres. ACS Appl. Mater.
(19) Biswas, R.; Sigalas, M. M.; Subramania, G.; Ho, K.-M. Photonic
Interfaces 2012, 4, 5562−5569.
Band Gaps in Colloidal Systems. Phys. Rev. B 1998, 57, 3701−3705.
(41) Hatton, B.; Mishchenko, L.; Davis, S.; Sandhage, K. H.;
(20) Biswas, R.; Sigalas, M. M.; Subramania, G.; Soukoulis, C. M.;
Ho, K.-M. Photonic Band Gaps of Porous Solids. Phys. Rev. B 2000, Aizenberg, J. Assembly of Large-Area, Highly Ordered, Crack-Free
61, 4549−4553. Inverse Opal Films. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 10354−
(21) Busch, K.; John, S. Photonic Band Gap Formation in Certain 10359.
Self-Organizing Systems. Phys. Rev. E 1998, 58, 3896−3908. (42) Waterhouse, G. I. N.; Waterland, M. R. Opal and Inverse Opal
(22) Joannopoulos, J. D.; Villeneuve, P. R.; Fan, S. Photonic Photonic Crystals: Fabrication and Characterization. Polyhedron
Crystals: Putting a New Twist on Light. Nature 1997, 386, 143−149. 2007, 26, 356−368.
(23) Joannopoulos, J. D.; Johnson, S. G.; Winn, J. N.; Meade, R. D. (43) Velev, O. D.; Kaler, E. W. Structured Porous Materials via
Photonic Crystals Molding the Flow of Light, 2nd ed.; Princeton Colloidal Crystal Templating: From Inorganic Oxides to Metals. Adv.
University Press: Princeton, 2008. Mater. 2000, 12, 531−534.
(24) López, C. Three-Dimensional Photonic Bandgap Materials: (44) Jiang, P.; Cizeron, J.; Bertone, J. F.; Colvin, V. L. Preparation of
Semiconductors for Light. J. Opt. A: Pure Appl. Opt. 2006, 8, R1−R14. Macroporous Metal Films from Colloidal Crystals. J. Am. Chem. Soc.
(25) Takahashi, S.; Suzuki, K.; Okano, M.; Imada, M.; Nakamori, T.; 1999, 121, 7957−7958.
Ota, Y.; Ishizaki, K.; Noda, S. Direct Creation of Three-Dimensional (45) Blanco, A.; Chomski, E.; Grabtchak, S.; Ibisate, M.; John, S.;
Photonic Crystals by a Top-Down Approach. Nat. Mater. 2009, 8, Leonard, S. W.; Lopez, C.; Meseguer, F.; Miguez, H.; Mondla, J. P.;
721−725. Ozin, G. A.; Toader, O.; Van Driel, H. M. Large-Scale Synthesis of a
(26) Pang, Y. K.; Lee, J. C. W.; Lee, H. F.; Tam, W. Y.; Chan, C. T.; Silicon Photonic Crystal with a Complete Three-Dimensional
Sheng, P. Chiral Microstructures (Spirals) Fabrication by Holographic Bandgap Near 1.5 Micrometres. Nature 2000, 405, 437−440.
Lithography. Opt. Express 2005, 13, 7615−7620. (46) Suezaki, T.; O’Brien, P. G.; Chen, J. I. L.; Loso, E.; Kherani, N.
(27) Romanato, F.; Businaro, L.; Vaccari, L.; Cabrini, S.; Candeloro, P.; Ozin, G. A. Tailoring the Electrical Properties of Inverse Silicon
P.; De Vittorio, M.; Passaseo, A.; Todaro, M. T.; Cingolani, R.; Opals - A Step Towards Optically Amplified Silicon Solar Cells. Adv.
Cattaruzza.; et al. Fabrication of 3D Metallic Photonic Crystals by X- Mater. 2009, 21, 559−563.
ray Lithography. Microelectron. Eng. 2003, 67−68, 479−486. (47) Míguez, H.; Chomski, E.; García-Santamaría, F.; Ibisate, M.;
(28) Lin, S. Y.; Fleming, J. G.; Hetherington, D. L.; Smith, B. K.; John, S.; López, C.; Meseguer, F.; Mondia, J. P.; Ozin, G. A.; Toader,
Biswas, R.; Ho, K. M.; Sigalas, M. M.; Zubrzycki, W.; Kurtz, S. R.; O.; van Driel, H. M. Photonic Bandgap Engineering in Germanium
Bur, J. A Three-Dimensional Photonic Crystal Operating at Infrared Inverse Opals by Chemical Vapor Deposition. Adv. Mater. 2001, 13,
Wavelengths. Nature 1998, 394, 251−253. 1634−1637.
(29) Moon, J. H.; Yang, S. Chemical Aspects of Three-Dimensional (48) Schroden, R. C.; Al-Daous, M.; Blanford, C. F.; Stein, A.
Photonic Crystals. Chem. Rev. 2010, 110, 547−574. Optical Properties of Inverse Opal Photonic Crystals. Chem. Mater.
(30) Stö ber, W.; Fink, A.; Bohn, E. Controlled Growth of 2002, 14, 3305−3315.
Monodisperse Silica Spheres in the Micron Size Range. J. Colloid (49) Yan, H.; Blanford, C. F.; Holland, B. T.; Smyrl, W. H.; Stein, A.
Interface Sci. 1968, 26, 62−69. General Synthesis of Periodic Macroporous Solids by Templated Salt

9672 DOI: 10.1021/acsomega.8b01334


ACS Omega 2018, 3, 9658−9674
ACS Omega Article

Precipitation and Chemical Conversion. Chem. Mater. 2000, 12, (69) Yasin, M. N.; Brooke, R. K.; Rudd, S.; Chan, A.; Chen, W.-T.;
1134−1141. Waterhouse, G. I. N.; Evans, D.; Rupenthal, I. D.; Svirskis, D. 3-
(50) Holland, B. T.; Blanford, C. F.; Stein, A. Synthesis of Dimensionally Ordered Macroporous PEDOT Ion-Exchange Resins
Macroporous Minerals with Highly Ordered Three-Dimensional Prepared by Vapor Phase Polymerization for Triggered Drug
Arrays of Spheroidal Voids. Science 1998, 281, 538−540. Delivery: Fabrication and Characterization. Electrochim. Acta 2018,
(51) Stein, A.; Wilson, B. E.; Rudisill, S. G. Design and Functionality 269, 560−570.
of Colloidal-Crystal-Templated Materials − Chemical Applications of (70) Norris, D. J.; Arlinghaus, E. G.; Meng, L.; Heiny, R.; Scriven, L.
Inverse Opals. Chem. Soc. Rev. 2013, 42, 2763−2803. E. Opaline Photonic Crystals: How Does Self-Assembly Work? Adv.
(52) Stein, A.; Li, F.; Denny, N. R. Morphological Control in Mater. 2004, 16, 1393−1399.
Colloidal Crystal Templating of Inverse Opals, Hierarchical (71) Zhou, Z.; Zhao, X. S. Flow-Controlled Vertical Deposition
Structures, and Shaped Particles. Chem. Mater. 2008, 20, 649−666. Method for the Fabrication of Photonic Crystals. Langmuir 2004, 20,
(53) Blanford, C. F.; Schroden, R. C.; Al-Daous, M.; Stein, A. 1524−1526.
Tuning Solvent-Dependent Color Changes of Three-Dimensionally (72) Jiang, P.; Bertone, J. F.; Hwang, K. S.; Colvin, V. L. Single-
Ordered Macroporous (3DOM) Materials Through Compositional Crystal Colloidal Multilayers of Controlled Thickness. Chem. Mater.
and Geometric Modifications. Adv. Mater. 2001, 13, 26−29. 1999, 11, 2132−2140.
(54) Yan, H.; Yang, Y.; Fu, Z.; Yang, B.; Xia, L.; Fu, S.; Li, F. (73) Kuai, S.-L.; Hub, X.-F.; Haché, A.; Truong, V.-V. High-Quality
Fabrication of 2D and 3D Ordered Porous ZnO films using 3D Opal Colloidal Photonic Crystals Obtained by Optimizing Growth
Templates by Electrodeposition. Electrochem. Commun. 2005, 7, Parameters in a Vertical Deposition Technique. J. Cryst. Growth
1117−1121. 2004, 267, 317−324.
(55) Vu, A.; Li, X.; Phillips, J.; Han, A.; Smyrl, W. H.; Bühlmann, P.; (74) Wang, X.; Husson, S. M.; Qian, X.; Wickramasinghe, S. R.
Stein, A. Three-Dimensionally Ordered Mesoporous (3DOM) Vertical Cell Assembly of Colloidal Crystal Films with Controllable
Carbon Materials as Electrodes for Electrochemical Double-Layer Thickness. Mater. Lett. 2009, 63, 1981−1983.
Capacitors with Ionic Liquid Electrolytes. Chem. Mater. 2013, 25, (75) Yashima, M.; Kakihana, M.; Yoshimura, M. Metastable-Stable
4137−4148. Phase Diagrams in the Zirconia-Containing Systems Utilized in Solid-
(56) Deng, Y.; Liu, C.; Yu, T.; Liu, F.; Zhang, F.; Wan, Y.; Zhang, L.; Oxide Fuel Cell Application. Solid State Ionics 1996, 86−88, 1131−
Wang, C.; Tu, B.; Webley, P. A.; Wang, H.; Zhao, D. Facile Synthesis 1149.
of Hierarchically Porous Carbons from Dual Colloidal Crystal/Block (76) Zhang, C.; Li, C.; Yang, J.; Cheng, Z.; Hou, Z.; Fan, Y.; Lin, J.
Copolymer Template Approach. Chem. Mater. 2007, 19, 3271−3277. Tunable Luminescence in Monodisperse Zirconia Spheres. Langmuir
(57) Hong, W.; Chen, Y.; Feng, X.; Yan, Y.; Hu, X.; Zhao, B.; Zhang, 2009, 25, 7078−7083.
F.; Zhang, F.; Xu, Z.; Lai, Y. Full-Color CO2 Gas Sensing by an (77) Li, J.; Chen, J.; Song, W.; Liu, J.; Shen, W. Influence of Zirconia
Inverse Opal Photonic Hydrogel. Chem. Commun. 2013, 49, 8229− Crystal Phase on the Catalytic Performance of Au/ZrO2 Catalysts for
8231. Low-Temperature Water Gas Shift Reaction. Appl. Catal., A 2008,
(58) Kuo, C. Y.; Lu, S. Y.; Chen, S.; Bernards, M.; Jiang, S. Stop 334, 321−329.
Band Shift Based Chemical Sensing with Three-Dimensional Opal (78) Menegazzo, F.; Pinna, F.; Signoretto, M.; Trevisan, V.;
and Inverse Opal Structures. Sens. Actuators, B 2007, 124, 452−458. Boccuzzi, F.; Chiorino, A.; Manzoli, M. Quantitative Determination
(59) Fenzl, C.; Hirsch, T.; Wolfbeis, O. S. Photonic Crystals for of Sites Able to Chemisorb CO on Au/ZrO2 Catalysts. Appl. Catal., A
Chemical Sensing and Biosensing. Angew. Chem., Int. Ed. 2014, 53, 2009, 356, 31−35.
3318−3335. (79) Zhang, X.; Wang, H.; Xu, B.-Q. Remarkable Effect of Zirconia
(60) Ko, Y. L.; Tsai, H. P.; Lin, K. Y.; Chen, Y. C.; Yang, H. in Au/ZrO2 Catalyst for CO oxidation. J. Phys. Chem. B 2005, 109,
Reusable Macroporous Photonic Crystal-Based Ethanol Vapor 9678−9683.
Detectors by Doctor Blade Coating. J. Colloid Interface Sci. 2017, (80) Comotti, M.; Li, W.-C.; Spliethoff, B.; Schüth, F. Support Effect
487, 360−369. in High Activity Gold Catalysts for CO Oxidation. J. Am. Chem. Soc.
(61) Wang, H.; Zhang, K.-Q. Photonic crystal structures with 2006, 128, 917−924.
tunable structure color as colorimetric sensors. Sensors 2013, 13, (81) Zanella, R.; Giorgio, S.; Henry, C. R.; Louis, C. Alternative
4192−4213. Methods for the Preparation of Gold Nanoparticles Supported on
(62) Waterhouse, G. I. N.; Metson, J. B.; Idriss, H.; Sun-Waterhouse, TiO2. J. Phys. Chem. B 2002, 106, 7634−7642.
D. Physical and Optical Properties of Inverse Opal CeO2 Photonic (82) Patterson, A. L. The Scherrer Formula for X-Ray Particle Size
Crystals. Chem. Mater. 2008, 20, 1183−1190. Determination. Phys. Rev. 1939, 56, 978−982.
(63) Waterhouse, G. I. N.; Chen, W.-T.; Chan, A.; Jin, H.; Sun- (83) Brunauer, S.; Emmett, P. H.; Teller, E. Adsorption of Gases in
Waterhouse, D.; Cowie, B. C. C. Structural and Optical Properties of Multimolecular Layers. J. Am. Chem. Soc. 1938, 60, 309−319.
γ-Al2O3 Inverse Opals. J. Phys. Chem. C 2015, 119, 6647−6659. (84) Barrett, E. P.; Joyner, L. G.; Halenda, P. P. The Determination
(64) Bauer, E.; Mueller, A. H.; Usov, I.; Suvorova, N.; Janicke, M. of Pore Volume and Area Distributions in Porous Substances. I.
T.; Waterhouse, G. I. N.; Waterland, M. R.; Jia, Q. X.; Burrell, A. K.; Computations from Nitrogen Isotherms. J. Am. Chem. Soc. 1951, 73,
McCleskey, T. M. Chemical Solution Route to Conformal Phosphor 373−380.
Coatings on Nanostructures. Adv. Mater. 2008, 20, 4704−4707. (85) Johnson, S.; Joannopoulos, J. D. Block-Iterative Frequency-
(65) Jovic, V.; Idriss, H.; Waterhouse, G. I. N. Slow Photon Domain Methods for Maxwell’s Equations in a Planewave Basis. Opt.
Amplification of Gas-Phase Ethanol Photo-Oxidation in Titania Express 2001, 8, 173−190.
Inverse Opal Photonic Crystals. Chem. Phys. 2016, 479, 109−121. (86) Srinivasan, R.; Davis, B. H.; Cavin, O. B.; Hubbard, C. R.
(66) Waterhouse, G. I. N.; Wahab, A. K.; Al-Oufi, M.; Jovic, V.; Crystallization and Phase Transformation Process in Zirconia: An In
Anjum, D. H.; Sun-Waterhouse, D.; Llorca, J.; Idriss, H. Hydrogen Situ High-Temperature X-ray Diffraction Study. J. Am. Ceram. Soc.
Production by Tuning the Photonic Band Gap with the Electronic 1992, 75, 1217−1222.
Band Gap of TiO2. Sci. Rep. 2013, 3, No. 2849. (87) Srinivasan, R.; De Angelis, R. J.; Ice, G.; Davis, B. H.
(67) Sharma, M.; Waterhouse, G. I. N.; Loader, S. W. C.; Garg, S.; Identification of Tetragonal and Cubic Structures of Zirconia using
Svirskis, D. High Surface Area Polypyrrole Scaffolds for Tunable Drug Synchrotron X-radiation source. J. Mater. Res. 1991, 6, 1287−1292.
Delivery. Int. J. Pharm. 2013, 443, 163−168. (88) Sato, O.; Kubo, S.; Gu, Z.-Z. Structural Color Films with Lotus
(68) Seyfoddin, A.; Chan, A.; Chen, W.-T.; Rupenthal, I. D.; Effects, Superhydrophilicity, and Tunable Stop-Bands. Acc. Chem. Res.
Waterhouse, G. I. N.; Svirskis, D. Electro-Responsive Macroporous 2009, 42, 1−10.
Polypyrrole Scaffolds for Triggered Dexamethasone Delivery. Eur. J. (89) Nishimura, R.; Hyodo, K.; Sawaguchi, H.; Yamamoto, Y.;
Pharm. Biopharm. 2015, 94, 419−426. Nonomura, Y.; Mayama, H.; Yokojima, S.; Nakamura, S.; Uchida, K.

9673 DOI: 10.1021/acsomega.8b01334


ACS Omega 2018, 3, 9658−9674
ACS Omega Article

Fractal Surfaces of Molecular Crystals Mimicking Lotus Leaf with


Phototunable Double Roughness Structures. J. Am. Chem. Soc. 2016,
138, 10299−10303.
(90) Phillips, K. R.; Vogel, N.; Burgess, I. B.; Perry, C. C.; Aizenberg,
J. Directional Wetting in Anisotropic Inverse Opals. Langmuir 2014,
30, 7615−7620.
(91) Haiss, W.; Thanh, N. T. K.; Aveyard, J.; Fernig, D. G.
Determination of Size and Concentration of Gold Nanoparticles from
UV−Vis Spectra. Anal. Chem. 2007, 79, 4215−4221.
(92) Medda, S. K.; De, S.; De, G. Synthesis of Au Nanoparticle
Doped SiO2−TiO2 films: Tuning of Au Surface Plasmon Band
Position through Controlling the Refractive Index. J. Mater. Chem.
2005, 15, 3278−3284.
(93) Tzounis, L.; Contreras-Caceres, R.; Schellkopf, L.; Jehnichen,
D.; Fischer, D.; Cai, C.; Uhlmann, P.; Stamm, M. Controlled Growth
of Ag Nanoparticles Decorated onto the Surface of SiO2 Spheres: A
Nanohybrid System with Combined SERS and Catalytic Properties.
RSC Adv. 2014, 4, 17846−17855.
(94) Cai, R.; Ellis, P. R.; Yin, J.; Liu, J.; Brown, C. M.; Griffin, R.;
Chang, G.; Yang, D.; Ren, J.; Cooke, K.; Bishop, P. T.; Theis, W.;
Palmer, R. E. Performance of Preformed Au/Cu Nanoclusters
Deposited on MgO Powders in the Catalytic Reduction of 4-
Nitrophenol in Solution. Small 2018, 14, No. 1703734.
(95) Gangula, A.; Podila, R.; Ramakrishna, M.; Karanam, L.;
Janardhana, C.; Rao, A. M. Catalytic Reduction of 4-Nitrophenol
using Biogenic Gold and Silver Nanoparticles Derived from Breynia
rhamnoides. Langmuir 2011, 27, 15268−15274.
(96) Ma, T.; Yang, W.; Liu, S.; Zhang, H.; Liang, F. A Comparison
Reduction of 4-Nitrophenol by Gold Nanospheres and Gold
Nanostars. Catalysts 2017, 7, 38.
(97) Shen, W.; Qu, Y.; Pei, X.; Li, S.; You, S.; Wang, J.; Zhang, Z.;
Zhou, J. Catalytic Reduction of 4-Nitrophenol using Gold Nano-
particles Biosynthesized by Cell-Free Extracts of Aspergillus sp. WL-
Au. J. Hazard Mater. 2017, 321, 299−306.

9674 DOI: 10.1021/acsomega.8b01334


ACS Omega 2018, 3, 9658−9674

You might also like