Open Access
25 August 2016 Review of roles for photonic crystals in solar fuels photocatalysis
Jeremy J. Pietron, Paul A. DeSario
Author Affiliations +
Abstract
We cover the intersection of nanophotonics, materials such as photonic crystals (PC), but also irregular templated light scattering interfaces, with their application to solar fuels photocatalysis. We describe the fundamental principles of adapting nanophotonics to photocatalysis, particularly slow photon effects and how appropriate choice of stop band and edge position of the PC can be exploited for light harvesting. We also review several representative examples of nanophotonic design applied to photocatalytic semiconductor materials. We include the most heavily investigated photocatalytic materials (such as TiO2), as well as inherently visible light active semiconductors, and materials sensitized with semiconductor nanocrystals or plasmonic metal nanoparticles. Finally, we review alternative scattering interfaces useful for improving the performance of solar fuels photocatalysis.

1.

Introduction

1.1.

Solar Fuel Photocatalysis at Semiconductors and the Challenge of Improving Light Harvesting

Direct solar water splitting is in principle one of the most sustainable pathways to produce clean fuels such as hydrogen (H2) from cheap and abundant resources; however, the paucity of photocatalysts that efficiently drive water splitting under solar irradiation presently limits its viability as a route to renewable fuels. Functional photocatalysts must be cheap, stable in water, and able to perform the three fundamental functions in the photocatalytic process: (1) absorb solar radiation to produce reactive charge carriers (i.e., electron–hole pairs); (2) efficiently transfer charge carriers to reactive sites at the semiconductor surface while minimizing electron–hole recombination; and (3) drive chemical reactions of interest with high selectivity and yield of the desired product. Semiconductor photocatalysts with band edges satisfying the thermodynamic requirements for both water oxidation and water reduction, such as titania (TiO2),1 are limited by poor sunlight harvesting as a consequence of their large optical bandgap. Narrower bandgap semiconductors that absorb in the visible region, such as BiVO4 and Fe2O3, do not photogenerate the required thermodynamic potentials for water splitting.24

Multiple strategies have been investigated to improve light utilization by wide bandgap semiconducting photocatalysts, with the most common including doping to shift band potentials, and sensitization with plasmonic metal nanoparticles, quantum dots, or dyes.57 Alternatively, the physical arrangement of the semiconductor can be controlled at the nanometer-to-micrometer scale to improve light harvesting. In the past decade, photonic crystals (PCs) of various morphologies have been investigated as a means of enhancing light utilization in semiconductor-based photocatalysts. We discuss below the recent trends in photocatalyst design that exploit the unique optical properties of PCs to improve sunlight utilization and highlight examples of how these photocatalyst structures are being used to enhance photochemical splitting of water.

1.2.

Slow Photons and Photonic Crystals

To clarify how PCs can enhance photocatalytic processes, including H2 generation, some background information on PCs and how they interact with light is necessary. Whether in pure or sensitized photocatalyst materials, the phenomenon characteristic of PCs that has been most commonly exploited to enhance light absorption in semiconductor photocatalysts is the “slow photon effect”—a slowing of light propagation in photonic materials at wavelengths near the photonic bandgap. PCs are ordered, periodic structures with alternating regions of high and low refractive index that satisfy Bragg diffraction conditions.811 A “forbidden band” of radiation with wavelengths on the same order as the periodicity of the PC structure cannot propagate through the PC, leading to a strong reflectance. The center of this band of radiation is known as the stop band or photonic bandgap. At the high- (blue) and low-energy (red) regions of the stop band, photons are multiply scattered in the lower refractive index and higher refractive index portion of the PC, respectively (Fig. 1),12 thus effectively lowering the group velocity of light at the edges of the stop band.13 The propagation of slow photons increases their optical path length within the semiconductor or sensitizer and enhances the probability of generating charge carriers.

Fig. 1

Simplified optical band structure of a PC. At photon energies approaching a full bandgap or a stop band from the red side, the group velocity of light decreases and light can be increasingly described as a sinusoidal standing wave that has its highest amplitude in the high-refractive index part of the structure. At energies above the band gap or stop band, the standing wave is predominantly localized in the low-index part of the photonic crystal. Reproduced with permission from Ref. 12. Copyright 2003 American Chemical Society.

JPE_7_1_012007_f001.png

Although more exotic structures exist, inverse opals (IOs) are the most common PC structure utilized in photocatalysis. These structures are typically prepared by templating monodisperse polystyrene spheres in an opal arrangement and infiltrating the voids of the face-centered cubic (fcc) array with oxide precursors before removing the template by calcination12,1418 (Fig. 2). The photonic bandgap (λp) for the IO films obeys a modified version of Bragg’s law

Eq. (1)

λp=1.6Dnp2f+ns2(1f)sinθ,
where D is the diameter of the spherical voids in the fcc structure, f is the volume fraction of the spherical pore volume, np is the refractive index of the pore-filling fluid, ns is the refractive index of the solid framework, and θ is the angle between incident light and (111) plane of the fcc structure.15 The exact frequency of the photonic bandgap is highly sensitive to the refractive indices of the solid and voids and will redshift as the refractive index of the pore-filling fluid increases (Fig. 3). The critical spatial dimension, D, is controlled directly during IO synthesis; the stop band is synthetically tuned as a function of the spherical template diameter (Fig. 4).19

Fig. 2

Scanning electron micrograph (SEM) of (a) colloidal crystal template of polystyrene spheres (200 nm) and (b) three-dimensional ordered macromesoporous architecture of Mo:BiVO4. Reproduced with permission from Ref. 18. Copyright 2014 American Chemical Society.

JPE_7_1_012007_f002.png

Fig. 3

Transmission UV–visible spectra of a TiO2 IO structure with 169-nm voids and filled with solvents of different refractive index and insert showing band position as a function of refractive index of the void-filling solvent. The photonic stop band redshifts as the refractive index of the pore-filling fluid increases. The refractive index of the TiO2 solid framework was estimated to be 2.40. Reproduced with permission from Ref. 15. Copyright 1999 American Institute of Physics.

JPE_7_1_012007_f003.png

Fig. 4

Transmission spectra measured at normal incidence of opal structures deriving from three different polystyrene sphere diameters. Reproduced with permission from Ref. 19. Copyright 2011 American Chemical Society.

JPE_7_1_012007_f004.png

2.

Coupling Photonic Crystals to Photocatalysis—Tuning the Red Edge of the Stop Band to the Photocatalyst Absorption Maximum

2.1.

Tuning the Photonic Crystal Stop Band Around the Semiconductor Band Edge

Precise tuning of the stop band is necessary to exploit the slow photon effect for enhanced photocatalysis. Findings by Chen et al.2022 established the ground rules for slow photon-enhanced photocatalysis. They applied anatase TiO2 IO films with stop bands centered at wavelengths between 280 and 500 nm to the photocatalytic degradation of methylene blue under monochromatic UV radiation at 370±10  nm.20 The most photocatalytically active anatase PCs featured stop bands at 345 nm–not 370 nm—because it is at the red edge of the stop band that light is most substantially slowed in the high index material. Monochromatic light of wavelengths matched to the stop band does not propagate in the PC. Anatase IO films with stop bands at 345 nm had a low-energy edge coincident with the 370-nm illumination (near the anatase band edge), enabling them to enhance degradation rates by a factor of 1.9. PC films with stop bands at 370 nm rejected the incident radiation and were less catalytically active than nonphotonic nanocrystalline TiO2 controls. As predicted by Eq. (1), the positions of the PC film stop bands were sensitive to the incident angle of radiation; tilt angles of 10 deg, 20 deg, and 45 deg substantially shifted the stop bands of all films, changing the films that were most photocatalytically active.

When white-light was used to irradiate the films, the order of the reactivity for the series of stop bands changed: PCs with stop bands at 300 nm exhibited the most photocatalytic activity enhancement (with an enhancement factor of 2.3 relative nanocrystalline TiO2). Under polychromatic radiation, the effects of reflection compete with the slow light effect at PCs. The wavelengths at which reflection occurs comprise a larger portion of the stop band than those wavelengths at the edges of the stop band, at which the slow photon effects dominate. Thus, positioning the stop band directly at the absorption edge can result in diminished photocatalytic activity under broad spectrum illumination.20 However, absorption by the dielectric material strongly diminishes photonic effects.23 Thus, positioning the stop band at energies higher (shorter wavelengths) than the TiO2 band edge enhances band-edge excitation and at the same time minimizes reflective losses occurring at wavelengths between the stop band and its high-energy edge.

2.2.

Tuning the Photonic Crystal Stop Band Around Sensitizer Excitation Maxima

The rules for matching the red edge of the stop band to excitation energy apply to sensitized PCs as well, whether they are sensitized with dyes, quantum dots, plasmonic metal nanoparticles, or atomic dopants; slow photon edges—rather than the stop band itself—should be matched to absorption of the sensitizer, with care taken to avoid reflective losses from the stop band. For example, slow photon effects when carefully matched to the 550-nm absorption of Ti3+-doped TiO2 PCs enhanced the broadband visible light photocatalytic activity for methylene blue degradation by about a factor of 2 compared to the equivalently Ti3+-doped nonphotonic counterpart; doped TiO2 PCs with more poorly matched stop bands exhibited much more modest enhancement.24

Curti et al.25 prescribe some standard experimental guidelines that will help clarify the extent to which the slow photon effect is responsible for changes in photocatalytic activity. They recommend that the photocatalytic performance of IOs be directly compared to disordered structures of the same material with comparable surface area and porosity to isolate the effects of surface area. Broadband excitation should be avoided to allow one to discriminate between photogenerated carriers originating from slow photons and those originating from normal bandgap excitation. Under these controlled experimental conditions, one can quantify the level of photocatalytic enhancement attributed to the slow photon effect and utilize the ground rules established in the seminal work by Chen et al.2022 and others19,26 to engineer the stop band position for maximum photocatalytic activity.

3.

Hydrogen Generation Photocatalysis at Photonic Crystals

Using the design principles described above, many researchers have exploited the slow photon effect to enhance H2 generation by photochemical or photoelectrochemical water splitting in photocatalysts including TiO2,2730 WO3,19 Fe2O3,31 Bi2WO6,32 and BiVO418,33 (results and references are summarized in Table 1). At Pt-modified TiO2 PCs,27 the photocatalytic efficiency of H2 generation as a function of the stop band position obeyed the same trends as those observed for dye degradation reactions by Chen et al.20 The largest enhancement factor under broadband excitation of 2.5 occurred when the stop band was tuned to 342 nm—where the red edge overlapped the TiO2 band edge.27

Table 1

Diverse examples of photonic enhancement of photocatalytic and photoelectrochemical water splitting/H2 generation reactions. Note that conditions that may impact the enhancement factor, such as excitation conditions, use of hole scavengers in reaction slurry/electrolyte, and photoanode bias in photoelectrochemical measurements, are omitted for the sake of brevity.

Material/morphologyReactionEnhancement factorCommentReference number
Pt/TiO2 IOH2 generation in methanol:water2.5Slow photon enhancement of H2 generation photocatalysis27
TiO2 and Au/TiO2 IOPhotoelectrochemical water splitting2.07 without Au and 2.86 with AuSlow photon enhancement and plasmonic enhancement28
Pt/TiO2 IOH2 generation in formate:water<1 per surface area under UV excitationEnhancement by IOs only observed at select excitation wavelengths; surface-area-normalized enhancement only marginally better than a disordered control29
Ti3+-doped TiO2 NT array with top PC layerPhotoelectrochemical water splitting10 under visible lightCombination of slow photon enhancement and visible light sensitization with dopant Ti3+30
α-Fe2O3 nanoparticles deposited on graphene IOPhotoelectrochemical water splitting1.4Reduced electron–hole recombination and fast electron transfer31
WO3 IOPhotoelectrochemical water splitting3 near excitation wavelength of activity maximum; >7 at low-activity wavelengthsExtensive angle-dependent study of slow photon enhancement; enhancement due to IO structure more pronounced at wavelengths where WO3 is initially less active19
Bi2WO4 IOPhotoelectrochemical water splitting3Slow photon effects and reduced bulk recombination32
Mo:BiVO4 and Au-modified Mo:BiVO4 IOPhotoelectrochemical water splitting8 over entire potential range; additional 55% increase with Au nanoparticlesCombination of slow photon enhancement, increased surface area and improved charge transfer, and SPR enhancement33
Mo:BiVO4 IOPhotoelectrochemical water splitting3 to 7, depending on electrochemical potentialCombination of cation-substituted composite oxide chemistry and PC-enhanced photocatalysis; control size of mesopores in addition to size of macropores18
C quantum dots/BiVO4 IOPhotoelectrochemical water splitting6Combination of slow photon effects and visible light sensitization from C QDs34
CdS/ZrO2 IOH2 generation in methanol:water4.7CdS activity enhanced at blue edge of stop band of passive support35
Au on TiO2 NTs array with TiO2 PC top layerPhotoelectrochemical water splitting2.5 compared to sample with no PC top layerMaximum enhancement when Au SPR matched to PC photonic bandgap36
Au/TiO2 IOPhotoelectrochemical water splitting2Coupling of plasmonic TiO2 sensitization with slow-photon effects37
Pt/TiO2 templated from butterfly wingPhotoelectrochemical water splitting2.3Enhancement of photocatalysis by aperiodic structure (unclear if enhancement is surface area or photonic effects)38
N-doped TiO2 templated from leaf replicaPhotocatalytic water splitting/H2 generation3.3 to 8Demonstration of viability of structural approach (unclear if enhancement due to surface area or photonic effects)39

While narrower bandgap alternatives to TiO2 for photocatalytic water splitting24 may absorb more visible light, none features 100% photon-to-carrier conversion efficiency, and thus all can still benefit from improved light utilization. Alternatives to TiO2 also suffer from drawbacks involving other parts of the photocatalytic process, such as valence/conduction band energies that are inefficient for water-splitting chemistry, and from fast electron–hole recombination kinetics. In particular, Bi2WO6 and BiVO4 suffer from high bulk recombination rates and thus benefit from the shortened carrier path lengths to the semiconductor surface featured in the IO architecture.18,3234 Nan et al.34 show particularly strong enhancement factors of both surface-normalized photocurrents and photocatalytic oxygen and H2 generation at the IO form of BiVO4 (Fig. 5). While the red edge of the stop band in their BiVO4 IO structures (450  nm) overlaps well with the band edge of BiVO4, their observed enhancement is probably due to more than the slow light effect alone. The simultaneous improvement of light management and electron–hole separation with IO forms of Bi2WO6 and BiVO4 enhances photoelectrochemical water splitting relative to non-PC forms of the semiconductors to a significantly greater extent than that achieved when expressing TiO2 in PC form (about 3× to 8× for Bi2WO6 and BiVO4 compared to about 2× to 2.5× for TiO2, Table 1). Note that while the improvement is more substantial in the PC forms of Bi2WO6 and BiVO4, the overall photoefficiencies are still typically lower than that achievable with TiO2.

Fig. 5

Chronoamperometry measurements for (1) IO-BiVO4 films modified with carbon quantum dots; (2) unmodified IO-BiVO4 films; (3) Mo-doped BiVO4 films; undoped BiVO4 films. (a) Chronoamperometry measurements at an external potential of 0.9  V versus the reversible H2 electrode (RHE) with repeated on/off cycles under UV/visible irradiation; (b) linear sweep voltammetry measurements at 50  mVs1 under UV/visible irradiation—the dashed lined represents the dark current of the undoped BiVO4 film; (c) photoconversion efficiency as a function potential versus RHE; (d) surface-area-normalized H2 and O2 generation rates under UV/visible irradiation. Reproduced with permission from Ref. 34. Copyright 2015 American Institute of Physics.

JPE_7_1_012007_f005.png

Zhang et al.33 compared the photoelectrochemical water-splitting activity of IO and planar Mo:BiVO4 photoanodes. When the stop band was tuned to 513 nm, near the 520-nm electronic bandgap of the oxide, photoelectrochemical activity was enhanced by about 8×. Photoelectrochemical activity was still enhanced—but to a much lesser extent—when the stop band was tuned to a frequency at which slow photons did not overlap the electronic bandgap excitation of BiVO4. The enhanced activity in the absence of the slow photon effect is attributed to improved charge separation imparted by the PC architectures. Zhou et al.18 used transient photocurrent decay to measure the extent to which IO architectures enhanced charge migration during photoelectrochemical water splitting in BiVO4 PCs. Carrier lifetimes were extended in BiVO4 photoanodes with a macroporous IO structure compared to those observed in disordered films. The longest lifetimes were measured in IO structures containing both macro and mesoporosity; photoelectrochemical water-splitting activity was enhanced by roughly 3× to 7× in such films.

Mitchell et al.35 applied a somewhat different approach to PC-enhanced photocatalysis, in which CdS nanocrystals (NCs) supported in a photocatalytically inactive zirconium oxide (ZrO2) PC performed photocatalysis directly, instead of acting as a sensitizer. They observed a maximum 5× enhancement for H2 generation under visible light illumination relative to nonphotonic CdS/ZrO2 at the “blue edge” of the PC stop band, suggesting that the supported CdS nanoparticles exploit slow photons in the lower-index voids in the PC—in contrast to the case in which the PC support itself is photocatalytic. Their results—combined with the myriad results that demonstrate the opposite case in which the PC itself acts as the photocatalyst material and enhancement of photocatalysis occurs via localization of light coincident with the red edge of the stop band in the PC—convincingly demonstrate that the localization of the different band edges in PC-derived photocatalysts can be effectively tailored to application via materials design.

4.

Combining Photonic Crystal Architectures with Plasmonics

Plasmonic metal nanoparticles, in particular Au and Ag, have recently received substantial attention as sensitizers for solar water-splitting catalysts due to their chemical stability and surface plasmon–resonant frequencies that overlap the solar spectrum.4043 Whereas PCs improve carrier generation at or near the electronic band edge, plasmonic sensitizers can extend the useful range of photocatalysts much farther into the red (i.e., to lower energy, longer-wavelength radiation). Properly applied, the combination of PCs and plasmonic sensitizers enhances photochemical/photoelectrochemical activity to a greater extent than either strategy alone.33,36,37,44,45 For example, Zhang et al.33 measured a roughly 26% enhancement in photoelectrochemical water splitting when they modified planar BiVO4 films with 20-nm plasmonic Au nanoparticles, but measured nearly 55% enhancement when they modified PC BiVO4 with the same plasmonic Au nanoparticles. When plasmonic Au was present, the photocurrent response was extended beyond the electronic bandgap of BiVO4 at 530 to 560 nm.

Analogous to other cases of combining photonics with sensitizers, the optimum enhancements from PCs + plasmonic sensitizers are achieved when the slow photon regions are positioned to overlap the surface plasmon resonance (SPR) frequency maximum while simultaneously avoiding reflective losses in electronic bandgap excitation.36,37,44,45 A large increase in optical extinction occurs when the slow photon region of a PC is well-matched to the SPR.36,37 Zhang et al.36 coupled TiO2 PC structures to TiO2 nanotube (NT) arrays, creating TiO2 NT/PC (NTPC) assemblies and subsequently deposited SPR-active Au NCs of different sizes onto the TiO2 NTPCs. The intensity of the SPR response of Au NCs with SPR maxima well-coupled to the PC photonic bandgap was more strongly enhanced than those with more poorly matched SPR maxima (Fig. 6). The SPR-driven photoelectrochemical response (Fig. 7) and H2 generation were similarly more strongly enhanced at the well-matched Au/TiO2 NTPCs. Zhang et al.37 achieved a maximum 2× enhancement in photoelectrochemical water splitting when the red edge of the stop band of PC TiO2 was positioned at 518 nm, close to the SPR frequency maximum of the 10-nm Au particles, but at a position where the photonic stop band (at 450 nm) would not interfere with bandgap excitation of the TiO2. When the Bragg reflection of the PC overlapped the Au SPR, the plasmonic enhancement was significantly suppressed.

Fig. 6

Diffuse refelectance UV–visible absorption spectrum of Au NCs on TiO2 NTs and the TiO2 NTPC assembly (with a stop band centered 552  nm) with the spectrum of TiO2 subtracted for clarity of the Au SPR features. Reproduced with permission from Ref. 36. Copyright 2013 American Chemical Society.

JPE_7_1_012007_f006.png

Fig. 7

Photoelectrochemical properties of the TiO2 NTPC and Au/TiO2 NTPC. (a) Amperometric It curves at an applied potential of 1.23  V versus RHE under illumination of visible light with wavelength 420  nm with 60 s light on/off cycles; (b) incident photon-to-current conversion efficiency plots in the range of 400 to 700 nm at 1.23  V versus RHE. Reproduced with permission from Ref. 36. Copyright 2013 American Chemical Society.

JPE_7_1_012007_f007.png

5.

Butterflies and Leaves: Alternate, Bioinspired, Photonic Crystal Geometries for Photocatalysis

Nature provides examples of structures that have unique interactions with light—whether they are naturally occurring PCs or disordered structures that multiply scatter light—that have inspired structural photocatalyst mimics. Butterfly wings38,46,47 and artificial leaves39 have both recently been used as structural templates for photocatalysts. Butterfly wing-templated TiO2 materials that feature regular, bicontinuous, gyroid structures with clear photonic bandgaps have been fabricated.46

Although not a PC, the Papilio helenus Linnaeus butterfly wing contains structural elements that improve light utilization, including honeycomb-like hole arrays that multiply scatter light, “ridges” that guide light into the hole array, and a bottom reflective layer that recaptures light escaping from the honeycomb arrays. Platinum-modified TiO2 replicas of P. helenus Linnaeus butterfly wings can function as H2-generating photocatalysts.38 The wing-pattered photocatalyst strongly enhances ultrabandgap light absorption and shows a 7× improvement in H2 generation relative to nanoparticulate TiO2, although it was unclear how much of this enhancement could be attributed solely to surface area. Demonstrating the hierarchical possibilities of biostructure-derived photocatalytic architectures, Yan and coworkers47 deposited plasmonic gold nanorods onto bismuth vanadate (BVO) structures templated from wings of the butterfly Paplio nirius (Fig. 8) and achieved substantial photocatalytic activity at red and near-infrared light illumination. Zhou et al.39 replicated the hierarchical structure of Anemone vitifolia Buch leaves, which contain various components that serve to focus, scatter or guide light and enhance light–matter interactions. Using a two-step infiltration process, they produced an N-doped TiO2 inorganic replica of the leaf structure that showed a roughly 3× to 8× enhancement in H2 generation compared to nanoparticulate TiO2; the activity was enhanced by another order of magnitude when modified with Pt nanoparticles.

Fig. 8

SEM images of (a) top view of original P. nirius wing scale, (b) top view of templated BVO replica of the P. nirius wing scale, (c) cross section of the original scale, (d) cross section of the BVO wing, (e) TEM image of the gold nanorods (AuNR), (f) HRTEM image of a single AuNR, (g) AuNRs-loaded BVO wing with insets showing the simplified model (upper) and an enlarged image (lower), (h) HRTEM image of the AuNRs-loaded BVO wing, and (i) schematic diagram of the controlled assembly of AuNRs onto the BVO unit. Reproduced from Ref. 47 (open access). Copyright 2016 Nature Publishing Group.

JPE_7_1_012007_f008.png

6.

Conclusions and Comments

Exploitation of the slow photon effect of photonic bandgap-containing materials requires rigorous matching of the red edge of the stop band with the desired excitation wavelength while avoiding absorption losses due to nonpropagation of light at the stop band. When PCs are correctly applied, photocatalytic activity enhancement factors of about 2 to 3 can be readily achieved, and in some cases enhancement of 5× or more may be possible. It is also possible to achieve light-utilization enhancement via the more disorganized process of internal light scattering, which is less demanding in structural precision, but also more difficult to design in a rational way.

A couple of important caveats should be mentioned. First, application of the slow photon effect to photocatalysis is illumination angle dependent, and as such can only be effectively applied in film geometries in which the illumination angle can be well-controlled. Practical photocatalysts are often used as fluid-dispersed (suspended) powders to maximize both mass transport and light utilization. The reactor geometry limitation inherent to PCs (i.e., a thin-film form-factor) means that careful consideration must be given to mass transport and light harvesting when designing the interfaces. Additionally, solar fuels photocatalysis involves many challenging photophysical and photochemical steps. Light harvesting is critical, but even with substantial improvements in light harvesting, many photocatalyst materials are still 2 to 3 orders of magnitude (or more)—rather than a factor of 2 to 3—from practical viability. One important class of solar fuels photocatalytic reactions omitted from this section involve reduction of carbon dioxide (CO2) to hydrocarbons as its final step, rather than protons to H2. Photochemical and electrochemical reductions of CO2 comprise an enormous research challenge in its own right. Improvements on all photocatalytic fronts, however, continue apace; a number of composite solar fuels photocatalyst materials are approaching activity competitive with alternative routes to solar fuels and photonics could conceivably contribute enough additional efficiency to make them viable.

Acknowledgments

The authors would like to thank the Office of Naval Research for its support of this work. We also would like to thank Debra Rolison of the U.S. Naval Research Laboratory for helpful comments on this paper.

References

1. 

A. Fujishima and K. Honda, “Electrochemical photolysis of water at a semiconductor electrode,” Nature, 238 (5358), 37 –38 (1972). http://dx.doi.org/10.1038/238037a0 Google Scholar

2. 

R. Abe, “Recent progress on photocatalytic and photoelectrochemical water splitting under visible light,” J. Photochem. Photobiol. C Photochem. Rev., 11 (4), 179 –209 (2010). http://dx.doi.org/10.1016/j.jphotochemrev.2011.02.003 1389-5567 Google Scholar

3. 

A. Kudo and Y. Miseki, “Heterogeneous photocatalyst materials for water splitting,” Chem. Soc. Rev., 38 (1), 253 –278 (2009). http://dx.doi.org/10.1039/B800489G CSRVBR 0306-0012 Google Scholar

4. 

M. Hernández-Alonso et al., “Development of alternative photocatalysts to TiO2: challenges and opportunities,” Energy Environ. Sci., 2 (12), 1231 –1257 (2009). http://dx.doi.org/10.1039/B907933E EESNBY 1754-5692 Google Scholar

5. 

M. Anpo and M. Takeuchi, “The design and development of highly reactive titanium dioxide photocatalysts operating under visible light irradiation,” J. Catal., 216 (1–2), 505 –516 (2003). http://dx.doi.org/10.1016/S0021-9517(02)00104-5 JCTLA5 0021-9517 Google Scholar

6. 

S. Banerjee et al., “New insights into the mechanism of visible light photocatalysis,” J. Phys. Chem. Lett., 5 (15), 2543 –2554 (2014). http://dx.doi.org/10.1021/jz501030x JPCLCD 1948-7185 Google Scholar

7. 

H. Xu et al., “Recent advances in TiO2-based photocatalysis,” J. Mater. Chem. A, 2 (32), 12642 –12661 (2014). http://dx.doi.org/10.1039/c4ta00941j Google Scholar

8. 

Y. N. Xia et al., “Monodispersed colloidal spheres: old materials with new applications,” Adv. Mater., 12 (10), 693 –713 (2000). http://dx.doi.org/10.1002/(SICI)1521-4095(200005)12:10<693::AID-ADMA693>3.0.CO;2-J ADVMEW 0935-9648 Google Scholar

9. 

C. López, “Materials aspects of photonic crystals,” Adv. Mater., 15 (20), 1679 –1704 (2003). http://dx.doi.org/10.1002/adma.200300386 ADVMEW 0935-9648 Google Scholar

10. 

K. Busch et al., “Periodic nanostructures for photonics,” Phys. Rep., 444 (3–6), 101 –202 (2007). http://dx.doi.org/10.1016/j.physrep.2007.02.011 PRPLCM 0370-1573 Google Scholar

11. 

H. R. Li et al., “Photonic crystal coupled plasmonic nanoparticle array for resonant enhancement of light harvesting and power conversion,” Phys. Chem. Chem. Phys., 14 (41), 14334 –14339 (2012). http://dx.doi.org/10.1039/C2CP42438J PPCPFQ 1463-9076 Google Scholar

12. 

S. Nishimura et al., “Standing wave enhancement of red absorbance and photocurrent in dye-sensitized titanium dioxide photoelectrodes coupled to photonic crystals,” J. Am. Chem. Soc., 125 (20), 6306 –6310 (2003). http://dx.doi.org/10.1021/ja034650p JACSAT 0002-7863 Google Scholar

13. 

A. Imhof et al., “Large dispersive effects near the band edges of photonic crystals,” Phys. Rev. Lett., 83 (15), 2942 –2945 (1999). http://dx.doi.org/10.1103/PhysRevLett.83.2942 PRLTAO 0031-9007 Google Scholar

14. 

A. Stein, B. E. Wilson and S. G. Rudisill, “Design and functionality of colloidal-crystal-templated materials—chemical applications of inverse opals,” Chem. Soc. Rev., 42 (7), 2763 –2803 (2013). http://dx.doi.org/10.1039/c2cs35317b CSRVBR 0306-0012 Google Scholar

15. 

S. Nishimura et al., “Fabrication technique for filling-factor tunable titanium colloidal crystal replicas,” Appl. Phys. Lett., 81 (24), 4532 –4534 (2002). http://dx.doi.org/10.1063/1.1524693 APPLAB 0003-6951 Google Scholar

16. 

L. I. Halaoui, N. M. Abrams and T. E. Mallouk, “Increasing the conversion efficiency of dye-sensitized TiO2 photoelectrochemical cells by coupling to photonic crystals,” J. Phys. Chem. B, 109 (13), 6334 –6342 (2005). http://dx.doi.org/10.1021/jp044228a JPCBFK 1520-6106 Google Scholar

17. 

C. W. Cheng et al, “Quantum-dot sensitized TiO2 inverse opals for photoelectrochemical hydrogen generation,” Small, 8 (1), 37 –42 (2012). http://dx.doi.org/10.1002/smll01660 SMALBC 1613-6810 Google Scholar

18. 

M. Zhou et al., “Photoelectrodes based upon Mo:BiVO4 inverse opals for photoelectrochemical water splitting,” ACS Nano, 8 (7), 7088 –7098 (2014). http://dx.doi.org/10.1021/nn501996a ANCAC3 1936-0851 Google Scholar

19. 

X. Q. Chen et al., “Enhanced incident photon-to-electron conversion efficiency of tungsten trioxide photoanodes based on 3D-photonic crystal design,” ACS Nano, 5 (6), 4310 –4318 (2011). http://dx.doi.org/10.1021/nm200100v ANCAC3 1936-0851 Google Scholar

20. 

J. I. L. Chen et al., “Amplified photochemistry with slow photons,” Adv. Mater., 18 (14), 1915 –1919 (2006). http://dx.doi.org/10.1002/adma.200600588 Google Scholar

21. 

J. I. L. Chen et al., “Effect of disorder on the optically amplified photocatalytic efficiency of titania inverse opals,” J. Am. Chem. Soc., 129 (5), 1196 –1202 (2007). http://dx.doi.org/10.1021.ja066102a JACSAT 0002-7863 Google Scholar

22. 

J. I. L. Chen et al., “Synergy of slow photon and chemically amplified photochemistry in platinum nanocluster-loaded inverse titania opals,” J. Am. Chem. Soc., 130 (16), 5420 –5421 (2008). http://dx.doi.org/10.1021/ja800288f JACSAT 0002-7863 Google Scholar

23. 

G. von Freymann et al., “Tungsten inverse opals: the influence of absorption on the photonic band structure in the visible spectral region,” Appl. Phys. Lett., 84 (2), 224 –226 (2004). http://dx.doi.org/10.1063/1.1639941 Google Scholar

24. 

D. Y. Qi et al., “Enhanced photocatalytic performance of TiO2 based on synergistic effect of Ti3+ self-doping and slow light effect,” Appl. Catal. B, 160–161 621 –628 (2014). http://dx.doi.org/10.1016/j.apcatb.2014.06.020 ACBEE3 0926-3373 Google Scholar

25. 

M. Curti et al., “Inverse opal photonic crystals as a strategy to improve photocatalysis: underexplored questions,” J. Phys. Chem. Lett., 6 (19), 3903 –3910 (2015). http://dx.doi.org/10.1021/acs.jpclett.5b01353 JPCLCD 1948-7185 Google Scholar

26. 

H. Xie et al., “Facile fabrication of 3D-ordered macroporous nanocrystalline iron oxide films with highly efficient visible light induced photocatalytic activity,” J. Phys. Chem. C, 114 (21), 9706 –9712 (2010). http://dx.doi.org/10.1021/jp102525y Google Scholar

27. 

J. Liu et al., “Enhancement of photochemical hydrogen evolution of Pt-loaded hierarchical titania photonic crystals,” Energy Environ. Sci., 3 (10), 1503 –1506 (2010). http://dx.doi.org/10.1039/c0ee00116c EESNBY 1754-5692 Google Scholar

28. 

K. Kim et al., “Optimization for visible light photocatalytic water splitting: gold-coated and surface-textured TiO2 inverse opal nano-networks,” Nanoscale, 5 (14), 6254 –6260 (2013). http://dx.doi.org/10.1039/c3nr01552a NANOHL 2040-3364 Google Scholar

29. 

F. Sordello and C. Minero, “Photocatalytic hydrogen production on Pt-loaded TiO2 inverse opals,” Appl. Catal. B, 163 452 –458 (2015). http://dx.doi.org/10.1016/j.apcatb.2014.08.028 ACBEE3 0926-3373 Google Scholar

30. 

Z. H. Zhang et al., “Microwave-assisted self-doping of TiO2 photonic crystals for efficient photoelectrochemical water splitting,” Appl. Mater. Interfaces, 6 (1), 691 –696 (2014). http://dx.doi.org/10.1021/am404848n Google Scholar

31. 

K.-Y. Yoon et al., “Hematite-based photoelectrochemical water splitting supported by inverse opal structures of graphene,” Appl. Mater. Interfaces, 6 (24), 22634 –22639 (2014). http://dx.doi.org/10.1021/am506721a Google Scholar

32. 

L. W. Zhang et al., “Bi2WO6 inverse opals: facile fabrication and efficient visible-light-driven photocatalytic and photoelectrochemical water-splitting activity,” Small, 7 (19), 2714 –2720 (2011). http://dx.doi.org/10.41002/smll.201101152 SMALBC 1613-6810 Google Scholar

33. 

L. W. Zhang et al., “Plasmonic enhancement in BiVO4 photonic crystals for efficient water splitting,” Small, 10 (19), 3970 –3978 (2014). http://dx.doi.org/10.1002/smll.201400970 SMALBC 1613-6810 Google Scholar

34. 

F. Nan et al., “Carbon quantum dots coated BiVO4 inverse opals for enhanced photoelectrochemical hydrogen generation,” Appl. Phys. Lett., 106 153901 (2015). http://dx.doi.org/10.1063/1.4918290 APPLAB 0003-6951 Google Scholar

35. 

R. Mitchell, R. Brydson and R. E. Douthwaite, “Enhancement of hydrogen production using photoactive nanoparticles on a photochemically inert photonic macroporous support,” Phys. Chem. Chem. Phys., 17 (1), 493 –499 (2015). http://dx.doi.org/10.1039/c4cp04333b PPCPFQ 1463-9076 Google Scholar

36. 

Z. H. Zhang et al., “Plasmonic gold nanocrystals coupled with photonic crystal seamlessly on TiO2 nanotube photoelectrodes for efficient visible light photoelectrochemical water splitting,” Nano Lett., 13 (1), 14 –20 (2013). http://dx.doi.org/10.1021/nl3029202 NALEFD 1530-6984 Google Scholar

37. 

X. Zhang et al., “Coupling surface plasmon resonance of gold nanoparticles with slow-photon-effect of TiO2 photonic crystals for synergistically enhanced photoelectrochemical water splitting,” Energy Environ. Sci., 7 (4), 1409 –1419 (2014). http://dx.doi.org/10.1039/c3ee43278e EESNBY 1754-5692 Google Scholar

38. 

H. H. Liu et al., “Hydrogen evolution via sunlight water splitting on an artificial butterfly wing architecture,” Phys. Chem. Chem. Phys., 13 (23), 10872 –10876 (2011). http://dx.doi.org/10.1039/c1cp20787c PPCPFQ 1463-9076 Google Scholar

39. 

H. Zhou et al., “Artificial inorganic leafs for efficient photochemical hydrogen production inspired by natural photosynthesis,” Adv. Mater., 22 (9), 951 –956 (2010). http://dx.doi.org/10.1002/adma20092039 ADVMEW 0935-9648 Google Scholar

40. 

Z. W. Liu et al., “Plasmon resonant enhancement of photocatalytic water splitting under visible illumination,” Nano Lett., 11 (3), 1111 –1116 (2011). http://dx.doi.org/10.1021/nl104005n NALEFD 1530-6984 Google Scholar

41. 

C. G. Silva et al., “Influence of excitation wavelength (UV or visible light) on the photocatalytic activity of titania containing gold nanoparticles for the generation of hydrogen or oxygen from water,” J. Am. Chem. Soc., 133 (3), 595 –602 (2011). http://dx.doi.org/10.1021/ja1086358 JACSAT 0002-7863 Google Scholar

42. 

H. M. Chen et al., “Plasmon inducing effects for enhanced photoelectrochemical water splitting: X-ray absorption approach to electronic structures,” ACS Nano, 6 (8), 7362 –7372 (2012). http://dx.doi.org/10.1021/nn3024877 ANCAC3 1936-0851 Google Scholar

43. 

P. A. DeSario et al., “Plasmonic enhancement of visible-light water splitting with AuTiO2 composite aerogels,” Nanoscale, 5 (17), 8073 –8083 (2013). http://dx.doi.org/10.1039/c3nr01429k NANOHL 2040-3364 Google Scholar

44. 

Z. Y. Chen et al., “Inverse opal structured Ag/TiO2 plasmonic photocatalyst prepared by pulsed current deposition and its enhanced visible light photocatalytic activity,” J. Mater. Chem. A, 2 (3), 824 –832 (2014). http://dx.doi.org/10.1039/c3ta13985a Google Scholar

45. 

Z. Y. Cai et al., “In situ gold-loaded titania photonic crystals with enhanced photocatalytic activity,” J. Mater. Chem. A., 2 (2), 545 –553 (2014). http://dx.doi.org/10.1039/c3ta13878j Google Scholar

46. 

C. Mille, E. C. Tyrode and R. W. Corkery, “3D titania photonic crystals replicated from gyroid structures in butterfly wing scales: approaching full band gaps at visible wavelengths,” RSC Adv., 3 (9), 3109 –3117 (2013). http://dx.doi.org/10.1039/c2ra22506a Google Scholar

47. 

R. Yan et al., “Bio-inspired plasmonic nanoarchitectured hybrid system towards enhanced far red-to-near infrared solar photocatalysis,” Sci. Rep., 6 20001 (2016). http://dx.doi.org/10.1038/srep20001 SRCEC3 2045-2322 Google Scholar

Biography

Jeremy J. Pietron received his PhD in analytical chemistry from the University of North Carolina, Chapel Hill, in 1998. After completing a National Research Council Postdoctoral Fellowship at the U.S. Naval Research Laboratory (NRL), Washington, DC, from 1999 to 2002, he joined the laboratory as a full-time staff scientist in 2003. His primary research interests are nanostructured materials, electrocatalysis, plasmonic nanomaterials, heterogeneous catalysis, and solar fuels photocatalysis.

Paul A. DeSario received his PhD in environmental engineering from Northwestern University in 2011. He joined the Surface Chemistry Branch of the U.S. NRL, Washington, DC, as a National Research Council Postdoctoral Fellow from 2011 to 2014 before joining the laboratory as a full-time staff scientist in 2015. His primary research interests are in the nanoscale design of catalytic materials for applications in energy conversion and remediation.

© 2016 Society of Photo-Optical Instrumentation Engineers (SPIE)
Jeremy J. Pietron and Paul A. DeSario "Review of roles for photonic crystals in solar fuels photocatalysis," Journal of Photonics for Energy 7(1), 012007 (25 August 2016). https://doi.org/10.1117/1.JPE.7.012007
Published: 25 August 2016
Lens.org Logo
CITATIONS
Cited by 14 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Photonic crystals

Photocatalysis

Information operations

Gold

Photocatalytic water splitting

Plasmonics

Semiconductors

Back to Top