Perspective pubs.acs.org/cm Materials Chemistry in 3D Templates for Functional Photonics Paul V. Braun* Department of Materials Science and Engineering, Frederick Seitz Materials Research Laboratory, Beckman Institute for Advanced Science and Technology, University of Illinois at Urbana−Champaign, Urbana, Illinois 61801, United States ABSTRACT: This Perspective overviews many of the developments in templated porous three-dimensional photonics, with a particular focus on functional architectures, and provides suggestions for future opportunities for research. A significant diversity of 3D structures is available today with characteristic dimensions appropriate for providing strong light−matter interactions, in no small part due to recent advances in 3D patterning techniques. However, the optical functionality of these structures has generally remained limited. Advances in materials chemistry have the opportunity to dramatically increase the function of templated 3D photonics, and a few examples of highly functional templated 3D photonics for sensing, solar energy harvesting, optical metamaterials, and light emission are presented as first examples of success. KEYWORDS: photonic crystal, photonic band gap, self-assembly, three-dimensional, optoelectronics, metamaterials benefits.12 Real and meaningful advances in materials chemistry, which yield device quality materials when performed within complex 3D templates, would provide a substantive advance in the state-of-the-art in functional 3D photonics for a number of these applications. For the purposes of this Perspective, I take a rather narrow definition of “functional photonics” to include structures or devices which enable energy transduction between photons and electrons, significantly narrow thermal emission, enhance photocatalysis, enable chemical sensing, or enable manipulation of light over small dimensions. I will not discuss passive optical elements, which function primarily as dielectric stacks, optical filters, or any similar element, unless the form and function is considerably different than dielectric lenses and other comparable optical elements. The rationale for not including these passive optical elements is that, aside from transformation optics, most three-dimensionally templated solids are likely to have greater losses due to optical scattering than traditional dielectric elements (e.g., lenses and dielectric stacks) formed via conventional routes. Recently, there have been several exciting developments in templated 3D functional photonics that demonstrate templating can be used to form optically functional structures, which may provide inspiration for additional research. Via hydrogen- 1. INTRODUCTION Over the past 20 years, the diversity of templated porous threedimensional photonics has exploded, as has the diversity of functional photonics (e.g., light-emitting diodes (LEDs), solar cells, optical modulators, optical detectors); however, the intersection of these fields has remained rather surprisingly small. In this Perspective, I discuss some of the recent advances in templated photonics, why perhaps the number of functional 3D photonic systems has remained limited, and opportunities for the future. Fundamentally, the limitation in functional 3D systems appears to be very much due to limitations in materials chemistry. Most of the materials chemistries which operate within a 3D template do not yield what I define as “functional” materials. Rather, with only a few exceptions, they yield polycrystalline or amorphous, often highly defective, materials that are not acceptable for high performance optics. While this may be acceptable for some fields, e.g., energy storage,1,2 this is not acceptable for the field of optics, where the materials often need to be optoelectronically active and both the real and imaginary parts of the refractive index need to be precisely controlled. Many 3D photonic crystal devices have been proposed, including zero-threshold lasers,3,4 low-loss waveguides,5−7 high-efficiency light-emitting diodes (LEDs), and solar cells,8−10 but such devices have generally not been realized because of material limitations. Exciting concepts in metamaterials, including negative refraction and cloaking, could be made practical using 3D structures formed from high quality metals or, as has been suggested, entirely new materials to minimize optical losses.11 Metamaterial devices which incorporate electrically pumped gain elements to balance the inherent optical loss of such devices would provide additional © 2013 American Chemical Society Special Issue: Celebrating Twenty-Five Years of Chemistry of Materials Received: July 13, 2013 Revised: October 1, 2013 Published: October 11, 2013 277 dx.doi.org/10.1021/cm4023437 | Chem. Mater. 2014, 26, 277−286 Chemistry of Materials Perspective plasma passivation, chemical vapor grown (CVD) polycrystalline silicon inverse opals were formed with intrinsic properties similar to conventional polycrystalline silicon, and both p- and n-type doping was demonstrated.13,14 Growth of single crystal GaAs 15 and GaN 16 within a 3D template has been demonstrated. For the GaAs system, modulation of electrically pumped light emission by the photonic band structure of the structure was observed. The colloidal crystal templated growth of 3D structures with significant thermal emission spectral narrowing has been proposed17 and fabricated.18−20 In the field of sensors, a number of systems have been demonstrated21,22 and may represent the most mature of the functional photonic systems. Competing with templating methods to form 3D photonic structures are semiconductor processing methods such as deep reactive ion etching, which have been used to form 3D photonic crystals.23 Such strategies have the advantage that they start with a high quality semiconductor grade substrate, but the number of symmetries that can be formed is limited, as is the thickness of the resultant 3D structure. Figure 2. SEM images of templates for 3D photonic structures. (a) Self-assembled silica opal (Reprinted with permission from ref 41. Copyright 2001 Nature Publishing Group.); (b) polymer woodpile formed by direct ink writing (Reprinted with permission from ref 42. Copyright 2004 Nature Publishing Group.); (c) polymer woodpile formed by laser direct writing, schematic of the body-centered tetragonal (bct) (left) and face-centered cubic (fcc) (right) unit cell is shown in the inset (Reprinted with permission from ref 43. Copyright 2012 The Optical Society.); (d) top and side views of a polymer photonic crystal formed via multibeam interference lithography (Reprinted with permission from ref 44. Copyright 2009 Royal Society of Chemistry.). 2. TEMPLATING The inherent value of templating is to decouple patterning strategies from the final material. For two-dimensional systems, direct lithographic patterning of functional materials is common, and the demand for templating strategies is limited. However, lithographic patterning of materials on the nanometer to micrometer scale in three-dimensions is quite challenging, in particular for structures with complex or broken symmetry, and there is a limited subset of available materials which can be effectively patterned. Templating strategies enable the three-dimensional structure to be formed in the most patternable material. Then, the structure of this template is imprinted into a functional but difficult to directly pattern second material (Figure 1). assembly of silica or polystyrene colloids or by the aforementioned laser or ink direct writing of polymers. The structural quality of readily available 3D templates has now become quite impressive. Via direct laser writing (DLW), nearly completely arbitrary 3D structures containing minimum feature sizes as small as ∼100 nm have been formed using advanced stimulated emission depletion (STED)-based methods (Figure 3), and slightly larger features can be obtained on Figure 1. Schematic of the templating process, in this case starting with a self-assembled colloidal crystal, which is filled with a secondary material (green), followed by template removal, providing a final structure which is the inverse of the starting template. Figure 3. Oblique-view SEM image of a 52 layer woodpile photonic crystal with a lattice constant of 350 nm fabricated with STED-DLW. Reprinted with permission from ref 46. Copyright 2011 The Optical Society. The most common templates for 3D photonic structures are provided by colloidal self-assembly silica or polymer opals,24−29 direct laser writing of photoresist,30−32 direct ink writing,32,33 and multibeam (holographic) interference lithography,34−37 although other templates are certainly possible (Figure 2). In the aggregate, these template generation methods provide extraordinary structural diversity but with a rather limited materials diversity. The template materials available consist mostly of polymers (photoresists and common polymers such as polystyrene), the ceramics titania and silica, and isolated metals. While there are examples of other template materials including carbon colloids38 and photopatternable chalcogenide glasses,39,40 the quality and structural regularity of these templates is not nearly as good as those formed by self- commercial systems.32 Via interference lithography, large-area periodic 3D structures with features on the order of 200 nm have been formed by a number of groups,35−37 although it is not yet clear if the minimum feature size provided by interference lithography can be as small as that provided by direct laser writing with STED, and formation of aperiodic structures by interference lithography is challenging. Colloidal templating also enables features with quite small dimensions (on the order of 10% of the colloid diameter),29 but there are 278 dx.doi.org/10.1021/cm4023437 | Chem. Mater. 2014, 26, 277−286 Chemistry of Materials Perspective deposition strategies based on chemical vapor deposition (CVD), which was first demonstrated through the growth of Si 3D photonic crystals using a colloidal crystal template,41,49 and the related technique of atomic layer deposition (ALD).50 The diversity of material which can now be grown via CVD and ALD is impressive, including many metals, polymers, semiconductors, and oxides, and is continually increasing.50,51 Both CVD and ALD have two similar challenges. Because the growth is conformal, any 3D structure which contains a narrow throat cannot be fully filled. As material deposits, the channels required for the precursors to diffuse into the structure simply pinch-off.52 The second challenge is that the deposited material is polycrystalline or amorphous, which significantly reduces the possibility that the material can be optoelectronically active. Under the right conditions, both ALD and CVD can deposit epitaxial thin films but only on single crystal or highly textured substrates. However, since the initial 3D templates currently available are polycrystalline or amorphous, the resulting deposited film is also polycrystalline or amorphous. I do not, however, want to discount the importance of ALD and CVD. Both techniques are being increasingly used due to the many optically interesting materials which can be formed, potential high purity of the deposited materials, and very good control over the thickness of the deposited layer. For a few systems, it has been shown that a templated polycrystalline material can have optoelectronic properties which approach that of a conventional polycrystalline thin film.13,14,53 CVD can be used to grow an epitaxial thin film on a single crystal substrate, in which case the process is often termed metal−organic vapor-phase epitaxy (MOVPE) or metal organic CVD (MOCVD).54 MOCVD is widely used on the semiconductor growth industry, and its importance for making semiconductor devices (e.g., semiconductor lasers based on GaAs) cannot be overstated. Until recently, epitaxial MOCVD had not been used in conjunction with a 3D template. Two reports15,16 now indicate the possibility of using epitaxial MOCVD as a bottom-up template infilling method and will be discussed in greater detail in Section 3 of this Perspective. A highly versatile generally bottom-up infilling method is electrodeposition (if the template is conductive, the growth can be conformal rather than bottom-up). This approach can be used to infill templates with almost any material that can be electroplated, which includes many metals and semiconductors, and a few conductive oxides and polymers.19,20,26,36,55,56 Figure 5 shows a schematic of the electrochemical process for templating the growth of CdSe with a self-assembled colloidal crystal, as well as SEM images of colloidal crystal templated CdS and Ni structures after removal of the colloidal crystal template. A challenge in templated electrodeposition is that the template needs to be on top of a conducting substrate which limits some of the optical template fabrication methods (e.g., holography on a reflective substrate is challenging), and the growth front of the electrochemically deposited material must remain planar which can require the use of complicated pulse sequences.20,45 The most trivial method to completely infill a template is to simply fill the template with a monomer solution which is subsequently polymerized. Then, if desired, the colloidal particles are removed. Despite the simplicity, this remains a powerful approach for forming chemical and pressure sensors,21,22,57 as it enables the direct patterning of a soft responsive polymer into a 3D optically active photonic crystal. If the dimensions or refractive index of the structure changes in some limitations. It is difficult to crystallize colloids into anything except a face-centered-cubic structure, colloidal templates always contain some defects, and the defect density tends to increase significantly as the colloid diameter decreases below ∼250 nm. Templates can also be created via electrochemical etching of silicon or aluminum, which provides a disordered pore network in the x−y plane, and the possibility of modulating the pore size and volume fraction in the z-direction. However, more complex highly ordered structures are challenging.45 As expected, there is a trade-off between complexity and template fabrication speed. Serial techniques such as direct laser writing have fabrication rates limited by the available laser power and beam rastering optics. Holographic strategies can pattern larger areas in a single laser exposure but with limitations on the structural complexity. The self-assembly methods can cover nearly limitless areas but only with a small set of crystal symmetries, and as the rate of self-assembly increases, the level of random defects also tends to increase. The set of materials chemistries available to be templated still remains limited. Although a diversity of materials can be deposited as thin films via gas-phase techniques, many of the deposition methods do not effectively deposit material deep within a three-dimensional template. Any technique that deposits material line-of-sight (e.g., sputtering or molecular beam epitaxy) will only significantly coat the top or top few layers of the template. Any technique in which the volume of the material changes significantly during materials deposition, for example, sol−gel-based template infilling strategies, will have limited use for photonic structures due to the defects formed by the volume change. Nanoparticle infilling has been effectively used to imbed fluorescent emitters within 3D photonic crystal crystals,47 but this approach is only somewhat effective to form 3D templated structures (Figure 4).48 Even Figure 4. SEM cross-section of a Ge nanoparticle inverse opal. The spaces between the nanoparticles have been filled with polymer to increase the effective refractive index of the inverse opal to 2.05. The glass substrate is at the bottom of the image. Reprinted with permission from ref 48. Copyright 2007 American Chemical Society. when the nanoparticles fully fill the template, the resulting structure suffers from a lower than might be expected effective refractive index due to the maximum volume fraction that can be occupied by packed particles (there is still considerable void volume between the nanoparticles),48 and the presence of a very high density of interfaces due to the large number of nanoparticle−nanoparticle contacts limits the possibility of good electrical performance. The most commonly applied approaches for template infilling are the highly successful gas-phase conformal 279 dx.doi.org/10.1021/cm4023437 | Chem. Mater. 2014, 26, 277−286 Chemistry of Materials Perspective Figure 5. (a) Schematic of the conditions used to electrodeposit CdSe inside a self-assembled colloidal crystal. Reprinted with permission from ref 26. Copyright 1999 Nature Publishing Group. (b) Electrodeposited CdS inverse opal templated by a silica colloidal crystal. Reprinted with permission from ref 26. Copyright 1999 Nature Publishing Group. The colloidal crystal has only been partially removed. (c) Ni inverse opal formed by electrodeposition through a colloidal crystal. After removal of the colloidal template, the Ni structure was electropolished to increase the structural openness. Reprinted with permission from ref 20. Copyright 2007 John Wiley & Sons, Inc. Figure 6. (a) Optical micrograph of a glucose responsive polymer inverse opal. The lines are cracks in the structure that were present in the initial colloidal template. (b) SEM image of the top surface of the polymer inverse opal. Reprinted with permission from ref 22. Copyright 2004 American Chemical Society. the presence of an analyte or the photonic crystal is mechanically deformed, the resulting change in the lattice constant of the photonic crystal can be read out optically. In one example, we formed a glucose sensor using this approach (Figure 6). Here, the polymer was functionalized with a boronic acid which, when it bonds with glucose, changes its state of charge, which in turn results in swelling of the polymer with water. Because the structure is porous, glucose can diffuse rapidly into the interior of the photonic crystal, which is important if a fast response (order of minutes) is required. While not the subject of this Perspective, the reader is also encouraged to investigate the subject of polymerized colloidal crystal arrays (PCCAs) if interested in stimuli-responsive photonic crystals.58−61 PCCAs consist of a 3D nontouching periodic array of colloidal particles suspended in a polymerized matrix. When the matrix expands or contracts, the wavelength of light shifts accordingly, providing a facile read-out mechanism. It is particularly important to understand the chemistry of the template. The template must withstand the conditions required for infilling (e.g., the high temperatures often required for CVD or the often corrosive conditions of an electroplating bath), and generally, it is desirable for there to remain a path to remove the template. Except for colloidal-based templates, which can be formed from thermally and chemically stable materials such as silica or carbon, and a few of the direct ink written templates, which can be formed from metals or ceramics, the majority of the templates are based on polymers (e.g., photoresists), which are not thermally stable and have limited chemical stability. A common approach to overcome this apparent shortcoming is a double templating (double inversion) strategy.37,40,62 In this strategy, the template is used to pattern a thermally and/or chemically robust material which can be formed at lower temperatures, for example, CVD or ALD grown silica or alumina. Then, the template is removed, often via oxygen plasma, leaving behind an inverse of the template. Now, this new and more robust template is used to impart a structure into a material which can only be formed at high temperatures. This double inversion strategy has been used to convert structures formed by direct ink writing,62,63 laser direct writing,64 and optical interference lithography37 into semiconductors such as silicon (Figures 7 and 8). The product of the double inversion is a direct replica of the starting structure. Figure 7. SEM images of a silicon woodpile fabricated using DLW and double-inversion. (a) Top view. (b) FIB cross-section. (c) and (d) compare the double inverted Si woodpile and the original SU-8 woodpile and show that both the structure and low surface roughness are maintained. Reprinted with permission from ref 64. Copyright 2006 John Wiley & Sons, Inc. 280 dx.doi.org/10.1021/cm4023437 | Chem. Mater. 2014, 26, 277−286 Chemistry of Materials Perspective dimensions of the polymer change as the glucose concentration changes due to a change in the water volume fraction in the polymer.21,22 Although the field of polymer inverse opals is reasonably mature, greater challenges exist when deterministic polymeric templates are used (e.g., via interference lithography or direct laser writing), as these templates are often cross-linked by the writing process, making it challenging to remove them without degrading the templated polymer. Development of photoresists appropriate for 3D patterning that can be removed with solvent (the best resists today for 3D patterning are crosslinked and insoluble) or the double inverse process could be used to overcome these issues. There have been a number of promising reports on templated 3D photonic structures for solar energy harvesting. There are now examples of dye sensitized solar cells where the templated structure provides both the porosity required for electrolyte transport and enhancements in the photocollection efficiency (Figure 9).65 There has been a steady interest in Figure 8. (a) Schematic illustration (inset is a low-magnification scanning electron microscopy (SEM) image of a 250 μm × 250 μm woodpile structure) and (b) SEM image of direct-write-assembled polymer woodpile structure. Polymer woodpile is a 16-layer structure. (c) Process sequence for templated assembly of such structures. (d) Low- and (e) high-magnification SEM images of a Si/SiO2/Si hollowwoodpile structure after focused ion beam milling (8-layer structure). Contrast is enhanced in the inset in (e) to reveal the trilayer Si/SiO2/ Si tube wall. Reprinted with permission from ref 62. Copyright 2006 John Wiley & Sons, Inc. Figure 9. (a) Schematic of a DSSC, with a porous silicon photonic crystal attached to the back side of the light harvesting layer. The light enters from the top of the structure. (b) Cross-sectional SEM image of the structure. The titania nanoparticle layer is at the top and the photonic crystal at the bottom. Panel b is reprinted with permission from ref 65. Copyright 2011 John Wiley & Sons, Inc. 3. FUNCTIONAL TEMPLATED PHOTONICS: STATE-OF-THE-ART Despite the great strides that have been made in template-based fabrication methods for 3D periodic and aperiodic structures, the number of functional photonics (by my narrow definition) is rather small. The primary reason is not due to limitations in template fabrication. After all, via direct laser writing, 3D structures with ∼100 nm features are possible, via holographic methods, patterning of large areas with similarly high resolution features is possible, and self-assembly of submicrometer colloidal particles is practiced by many groups. Rather, the fundamental stumbling block is how to fill templates with functional materials. As a result, the number of publications on functional photonics has been rather limited, and the current state-of-the-art is limited to a small, yet promising, set of examples which are highlighted here. As previously mentioned, the most mature of the templated functional photonics is in the area of polymer-based sensors. There are many examples of chemically responsive polymers, and as long as the constituent monomers, or the polymer itself, can be infilled into a 3D template, there exists the possibility to form a photonic-based sensor. Of course, the polymer must respond in such a way to enable an optical signal. This could be either a change in its optical constants or a change in dimension of the polymer, both of which will result in a movement of the diffraction peak generated by the 3D structure. In the glucose sensor example, both the refractive index and the physical thermophotovoltaic (TPV) systems based on 3D metallodielectric structures,17−19,66 and as discussed, templating is a particularly effective way to realize these structures (Figure 10).66 It remains to be seen, however, whether structures with the combination of the required optical and thermal properties can be realized. The required operating temperatures may be as high as 2000 K, and the best structures today are only stable to about 1500−1700 K.66 There is also considerable work on 2D TPV structures,67 and it remains to be determined whether the potentially enhanced optical response of the 3D structures over a 2D structure will outweigh the required additional complexity of a 3D system. This will very much depend on the materials which can be templated into 3D structures that meet the required thermal and optical requirements. Both the dye sensitized and thermophotovoltaic solar systems are rather interesting, if only because unlike many other functional photonics (e.g., LEDs and lasers), where semiconductor-grade material is required, it is acceptable for the material comprising the photonic crystal to be rather defective. In fact, defects, e.g., polycrystallinity and dopants, may actually serve to enhance the optical properties and thermal stability of some materials. The most sophisticated of the functional photonics published to date have been formed by epitaxial MOCVD.15,16 These are the only examples to date of the templated growth of a single crystal, and in one case, even electrically pumped light emission was observed.15 Both examples rely on what is termed selective area epitaxy (SAE).54 Initially, it may not seem possible to grow 281 dx.doi.org/10.1021/cm4023437 | Chem. Mater. 2014, 26, 277−286 Chemistry of Materials Perspective Figure 11. Schematic of the fabrication of a single crystal GaN logpile photonic crystal from a Si/SiO2 logpile template. Reprinted with permission from ref 16. Copyright 2011 American Chemical Society. materials which not only have the potential for optoelectronic activity but also can be grown using processes compatible with forming optoelectronically active devices. Epitaxial MOCVD is particularly attractive in this regard. Without removing the template from the growth chamber, it is possible to deposit not only p- and n-type doped materials but also materials with different bandgaps which can serve to trap or guide electron− hole pairs to enable efficient electrically stimulated photoemission. Figure 13 shows a schematic of a device structure formed via this process, as well as images of the electrically pumped emission and the spectra of the emission under various conditions. Figure 10. (a) SEM micrograph and (b) reflectance properties of tungsten (W) inverse colloidal crystal after annealing to 1000 °C for 12 h. The red spectrum is the measured reflectance, and the dashed red spectrum is the reflectance calculated using the FDTD method. Reprinted with permission from ref 66. Copyright 2013 Nature Publishing Group. 4. NEAR-TERM OPPORTUNITIES Some of the near-term materials opportunities for templated functional photonics are rather obvious and are reasonably linear extensions of the current examples in the areas of sensors, solar energy harvesting, and light emitting structures. Clearly, sensors with increased structural and chemical sophistication can be formed via templating. Any material which can be filled as a monomer solution into a 3D template and subsequently polymerized can in principle be templated, and the number of templates available is continually increasing. For the templated architecture to be effective as an optically responsive sensor, it must respond by either changing dimensions or changing its optical constants. In the sensing arena, the key question to ask is if the design will be more effective than a simple 1D or 2D photonic structure, 3D photonic sensors formed from polymerized colloidal crystal arrays,61 or an already commercially available sensor. If the answer to all parts of this question is yes, there are few if any fundamental issues preventing template-based sensor fabrication methodologies. I also expect there will be continuing improvements in templated architectures for solar energy harvesting for both DSSCs and TPVs. For DSSCs, research on optimization of the photonic response, including through the incorporation of new absorber materials,68 while maintaining the required ion and/or chargecarrier transport characteristics, continues, and for TPV-based solar cells, structures designed for optimized optical properties could certainly be applied using the current templating methodologies. However, for both DSSC and TPV systems, there are a number of critical needs that are not as simple as a semiconductor from the gas phase starting at the substrate− template interface. After all, this requires the semiconductor precursors to diffuse through the thickness of the template without reacting with the template and then efficiently nucleate and grow in an epitaxial fashion on the underlying template. In fact, in a somewhat modified way (using a much simpler template), this is exactly what is already performed in the semiconductor industry, where an oxide stripe is patterned on a semiconductor substrate and then the growth conditions are set such that the growing semiconductor only nucleates and grows on the exposed semiconductor and not on the oxide stripe.54 Under the appropriate conditions, many CVD reactants will only nucleate on specific substrates and will not nucleate on an oxide even after multiple hours in the CVD reactor. This is exactly how the growth of both GaAs and GaN single crystals was performed. The GaN 3D photonic crystal was grown on a GaN substrate, and the GaAs photonic crystal was grown on a GaAs substrate. An example of the process flow for the GaN system is shown in Figure 11. A similar process was used for growing the GaAs system, although the process involved a number of additional steps since both p- and n-type GaAs was grown, as well as an InGaAs quantum well layer, an AlGaAs capping layer to minimize surface recombination, and the necessary electrodes for driving light emission. Images of the GaAs structure at various steps in the process and the final GaN structure are shown in Figure 12. Function is more than just structure. For many applications, function requires the conversion of electrical energy into light or the conversion of light into electrical energy. This requires 282 dx.doi.org/10.1021/cm4023437 | Chem. Mater. 2014, 26, 277−286 Chemistry of Materials Perspective Figure 12. (a) Schematic and SEM of a 3D photonic crystal template partially filled with single crystal GaAs by epitaxy. Inset highlights the bottomup growth of the GaAs. Schematic and focused ion-beam cross-section SEM images of (b) GaAs filled template and (c) inverted (template removed) GaAs structure. Panels a−c reprinted with permission from ref 15. Copyright 2011 Nature Publishing Group. (d) Cross-section TEM image of a five layer single crystal GaN logpile structure. Inset on the right-hand side shows an electron diffraction pattern from the region indicated by the dotted red circle, and the left-hand side inset shows the diffraction pattern from the white dotted circle region. (e) Cross-section SEM image of a nine layer GaN logpile. The top right-hand side image shows an enlarged top view, and the bottom image shows an enlarged section of the perspective view of regions indicated by the dotted black box. Panels d and e reprinted with permission from ref 16. Copyright 2011 American Chemical Society. Figure 13. (a) Schematic of a GaAs 3D photonic crystal (blue) containing an InGaAs light-emitting layer (red). The structure is lithographically patterned into the form of a cylindrical mesa with a ring electrode on the top surface (gold). (b) An optical micrograph of a device under white light illumination showing the Ti/Au ring electrode and the mesa surrounded by the etched GaAs. (c−e) Upon current injection (c, 2 mA; d, 4 mA; e, 6 mA) light is emitted. The light output increases with current. (f) Electroluminescence (EL) spectra from 3D photonic crystal (PhC) LEDs with lattice constants of 735 and 1030 nm, respectively. The shape of the EL spectrum from the 735 nm lattice constant structure does not change when the pores are filled with dodecane, whereas the EL spectrum from the 1030 nm lattice constant structure changes significantly when the pores are filled with o-xylene. The EL spectra from the 1030 nm lattice constant structure are shifted up for clarity. (g) EL spectra from a 3D PhC LED where the emission is not modified by the 3D structure (a = 735 nm). The EL intensity increases linearly with current. Reprinted with permission from ref 15. Copyright 2011 Nature Publishing Group. structural optimization using current materials, which will be discussed in the following section. For the epitaxially grown optoelectronically active 3D structures, the selective area epitaxy process has been demonstrated for a few semi283 dx.doi.org/10.1021/cm4023437 | Chem. Mater. 2014, 26, 277−286 Chemistry of Materials Perspective increase? TPV systems require even more significant materials developments. A viable TPV material will probably need to maintain its structure at temperatures well in excess of 1500 K for months or even years. Today, refractory metals are the most popular emitter materials for TPV systems, but must this be so? Are there other refractory compounds which have the necessary optical properties and better thermal stabilities? If so, can these materials be filled into a 3D template? 3D optical metamaterials is an area where some work has been performed to date, but there remain significant opportunities for advances. Starting from DLW structures, several groups have now formed chiral optical metamaterials by coating the structure with metal via electrodeposition.71−74 However, functional 3D metamaterials are quite challenging. Issues of optical loss are significant, and the required structural perfection is much greater than that required for 3D dielectric structures. Pathways to new materials that have acceptable loss, elements which provide optical gain to compensate loss, and structures with the required perfection are all necessary but currently not available. Advanced materials chemistries will play an important role to realize advanced functional 3D optical metamaterials. Finally, functional devices contain many elements. While development of discrete components is important, real systems require integration of many elements, and all the processing steps must be compatible. 3D porous structures are particularly challenging, as they are not readily compatible with conventional photoresist processing, and are more susceptible to oxidation and other environmental constraints due to their high surface area. Characterization can also be challenging, in particular when the 3D templated material is integrated into a larger system since the material to be characterized is often deeply imbedded within the structure. conductors, but from our experience, GaAs was significantly easier to epitaxially grow than for example AlGaAs; so, expanding this list may not be trivial. The AlGaAs tended to nucleate on the oxide phase, limiting successful epitaxial template infilling. Expanding the diversity of materials available through better process controls should be possible. The number of templates could also be significantly expanded. To date, only a silica colloidal crystal, a holographically defined polymer, and a lithographically defined silica template have been used. No attempts have been made to introduce optically active defects, e.g., waveguides or optical cavities, nor have the structures been optimized for optical function. Both these directions should be possible using direct laser writing template fabrication strategies. 5. CRITICAL NEEDS AND LONG-TERM OPPORTUNITIES For templated functional photonics to grow in significance, a number of critical needs need to be addressed. There is the very limited subset of materials which can be templated with defect densities rivaling conventional electronic materials. Currently, only two semiconductor materials have been templated in single crystal form, GaAs and GaN, and only the GaAs structure was demonstrated to be electrically active. The X-ray pole figure of the GaAs structure indicated it contained stacking faults and other defects, and considerable reduction in the defect density is probably required for practical functional devices. Significant expansion of the list of materials will also require major advances. While better process control might enable growth of some materials using known precursors, e.g., AlGaAs, for fast and effective selective area epitaxy through 3D templates, new precursors, which are particularly robust against secondary nucleation on the template, will probably be needed. It is also a question if selective area epitaxy is the best route for forming large single crystal 3D templated structures. Most large single crystal semiconductors are formed via crystallization from the melt (the standard approach for forming single crystal silicon) and not gas-phase growth, which is generally seen as more appropriate for fabrication of thin films. Intrinsically then, might it be possible to directionally solidify a semiconductor from the melt within a template? The challenges would be significant. Not only would the template have to withstand the required temperatures (1414 °C for silicon), but also the template would need to be chemically stable to the molten semiconductor. For many applications, there would need to be a path to remove the template without damaging the semiconductor. Finally, there is the specific challenge of getting the semiconductor into the template. Unless the surface energy of the template is greater than the surface energy of the molten semiconductor, the required pressures will be significant. While low melting temperature, low surface tension semiconductors such as selenium have been imbibed into a colloidal template,69,70 a similar process using a material such as silicon would be challenging. Some semiconductors can be directly electrodeposited into the template,56 but it is not clear if the semiconductor will remain in the template once it is melted. For solar energy harvesting, both DSSCs and TPV systems continue to receive attention, and the efficiency of both may be enhanced by 3D structuring. For DSSC systems, might templating enable the current collector and the solar absorber to be the same material? Or, if the current collector could be templated in single crystal form, while maintaining the high surface area required for high dye loading, might the efficiency 6. CONCLUSIONS Developments in fabrication of templated porous threedimensional structures now enable formation of an exceptional diversity of structures from a wide subset of materials. However, the optical functionality of these structures still remains limited, in large part due to limitations in filling structures with functional materials. Advances in template infilling strategies, which provide very low defect density high purity semiconductors, chemically responsive materials, and thermally stable materials, are required to advance the optical functionality of these structures. Advances in template fabrication, which yield structurally complex templates which are chemically stable against infilling yet are easily removed, are also needed. Recent advances in materials chemistry have enabled a few specific successful examples of highly functional templated 3D photonics for sensing, solar energy harvesting, optical metamaterials, and light emission, but with further materials chemistry developments, there are opportunities for many additional and significant advances. ■ AUTHOR INFORMATION Corresponding Author *E-mail: pbraun@illinois.edu. Notes The authors declare no competing financial interest. Biography Paul V. Braun is the Ivan Racheff Professor of Materials Science and Engineering at the University of Illinois at Urbana−Champaign and a 284 dx.doi.org/10.1021/cm4023437 | Chem. Mater. 2014, 26, 277−286 Chemistry of Materials Perspective (30) Cumpston, B. H.; Ananthavel, S. P.; Barlow, S.; Dyer, D. L.; Ehrlich, J. E.; Erskine, L. L.; Heikal, A. A.; Kuebler, S. M.; Lee, I. Y. S.; McCord-Maughon, D.; Qin, J.; Rockel, H.; Rumi, M.; Wu, X.-L.; Marder, S. R.; Perry, J. W. Nature 1999, 398, 51. (31) Deubel, M.; von Freymann, G.; Wegener, M.; Pereira, S.; Busch, K.; Soukoulis, C. M. Nat. Mater. 2004, 3, 444. (32) Fischer, J.; Wegener, M. Laser Photonics Rev. 2013, 7, 22. (33) Ahn, B. Y.; Duoss, E. B.; Motala, M. J.; Guo, X.; Park, S.-I.; Xiong, Y.; Yoon, J.; Nuzzo, R. G.; Rogers, J. A.; Lewis, J. A. Science 2009, 323, 1590. (34) Campbell, M.; Sharp, D. N.; Harrison, M. T.; Denning, R. G.; Turberfield, A. J. Nature 2000, 404, 53. (35) Chen, Y. C.; Geddes, J. B., III; Lee, J. T.; Braun, P. V.; Wiltzius, P. Appl. Phys. Lett. 2007, 91, 241103. (36) Miyake, M.; Chen, Y. C.; Braun, P. V.; Wiltzius, P. Adv. Mater. 2009, 21, 3012. (37) Shir, D.; Nelson, E. C.; Chen, Y. C.; Brzezinski, A.; Liao, H.; Braun, P. V.; Wiltzius, P.; Bogart, K. H. A.; Rogers, J. A. Appl. Phys. Lett. 2009, 94, 011101. (38) Goodman, M. D.; Arpin, K. A.; Mihi, A.; Tatsuda, N.; Yano, K.; Braun, P. V. Adv. Opt. Mater. 2013, 1, 300. (39) Feigel, A.; Kotler, Z.; Sfez, B.; Arsh, A.; Klebanov, M.; Lyubin, V. Appl. Phys. Lett. 2000, 77, 3221. (40) Wong, S.; Deubel, M.; Perez-Willard, F.; John, S.; Ozin, G. A.; Wegener, M.; von Freymann, G. Adv. Mater. 2006, 18, 265. (41) Vlasov, Y. A.; Bo, X. Z.; Sturm, J. C.; Norris, D. J. Nature 2001, 414, 289. (42) Gratson, G. M.; Xu, M. J.; Lewis, J. A. Nature 2004, 428, 386. (43) Bruser, B.; Staude, I.; von Freymann, G.; Wegener, M.; Pietsch, U. Appl. Opt. 2012, 51, 6732. (44) Brzezinski, A.; Chen, Y.-C.; Wiltzius, P.; Braun, P. V. J. Mater. Chem. 2009, 19, 9126. (45) Sailor, M. J. Porous Silicon in Practice: Preparation, Characterization and Applications; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, 2011. (46) Fischer, J.; Wegener, M. Opt. Mater. Express 2011, 1, 614. (47) Leistikow, M. D.; Mosk, A. P.; Yeganegi, E.; Huisman, S. R.; Lagendijk, A.; Vos, W. L. Phys. Rev. Lett. 2011, 107, 193903. (48) Shimmin, R. G.; Vajtai, R.; Siegel, R. W.; Braun, P. V. Chem. Mater. 2007, 19, 2102. (49) Blanco, A.; Chomski, E.; Grabtchak, S.; Ibisate, M.; John, S.; Leonard, S. W.; Lopez, C.; Meseguer, F.; Miguez, H.; Mondia, J. P.; Ozin, G. A.; Toader, O.; van Driel, H. M. Nature 2000, 405, 437. (50) George, S. M. Chem. Rev. 2010, 110, 111. (51) Yagüe, J. L.; Coclite, A. M.; Petruczok, C.; Gleason, K. K. Macromol. Chem. Phys. 2013, 214, 302. (52) Moon, J. H.; Xu, Y.; Dan, Y.; Yang, S. M.; Johnson, A. T.; Yang, S. Adv. Mater. 2007, 19, 1510. (53) Suezaki, T.; Yano, H.; Hatayama, T.; Ozin, G. A.; Fuyuki, T. Appl. Phys. Lett. 2011, 98, 072106. (54) Stringfellow, G. B. Organometallic Vapor-Phase Epitaxy: Theory and Practice, 2nd ed.; Academic Press: San Diego, CA, 1998. (55) Braun, P. V.; Wiltzius, P. Curr. Opin. Colloid Interface Sci. 2002, 7, 116. (56) Pandey, R. K.; Sahu, S. N.; Chandra, S. Handbook of Semiconductor Electrodeposition; Marcel Dekker: New York, 1996. (57) Arsenault, A. C.; Kitaev, V.; Manners, I.; Ozin, G. A.; Mihi, A.; Miguez, H. J. Mater. Chem. 2005, 15, 133. (58) Asher, S. A.; Holtz, J.; Liu, L.; Wu, Z. J. Am. Chem. Soc. 1994, 116, 4997. (59) Holtz, J. H.; Asher, S. A. Nature 1997, 389, 829. (60) Holtz, J. H.; Holtz, J. S. W.; Munro, C. H.; Asher, S. A. Anal. Chem. 1998, 70, 780. (61) Muscatello, M. M. W.; Stunja, L. E.; Asher, S. A. Anal. Chem. 2009, 81, 4978. (62) Gratson, G. M.; Garcia-Santamaria, F.; Lousse, V.; Xu, M. J.; Fan, S. H.; Lewis, J. A.; Braun, P. V. Adv. Mater. 2006, 18, 461. (63) Garcia-Santamaria, F.; Xu, M. J.; Lousse, V.; Fan, S. H.; Braun, P. V.; Lewis, J. A. Adv. Mater. 2007, 19, 1567. Guest Professor at the University of Stuttgart, Germany. Prof. Braun received his B.S. degree with distinction from Cornell University and his Ph.D. in Materials Science and Engineering from Illinois. Following a postdoctoral appointment at Bell Laboratories, Lucent Technologies, he joined the faculty at Illinois in 1999. ■ ACKNOWLEDGMENTS The recent work from my group reported in this Perspective was primarily supported by the US Department of Energy “Light Material Interactions in Energy Conversion” Energy Frontier Research Center under Grant DE-SC0001293. I would also like to acknowledge the many excellent students, postdocs, and colleagues I have had the opportunity to work with on the subject of templated 3D photonics. ■ REFERENCES (1) Stein, A.; Wilson, B. E.; Rudisill, S. G. Chem. Soc. Rev. 2013, 42, 2763. (2) Zhang, H. G.; Yu, X. D.; Braun, P. V. Nat. Nanotechnol. 2011, 6, 277. (3) John, S. Phys. Rev. Lett. 1987, 58, 2486. (4) Yablonovitch, E. Phys. Rev. Lett. 1987, 58, 2059. (5) Lousse, V.; Fan, S. Opt. Express 2006, 14, 866. (6) Johnson, S. G.; Villeneuve, P. R.; Fan, S.; Joannopoulos, J. D. Phys. Rev. B 2000, 62, 8212. (7) Mekis, A.; Chen, J. C.; Kurland, I.; Fan, S.; Villeneuve, P. R.; Joannopoulos, J. D. Phys. Rev. Lett. 1996, 77, 3787. (8) Green, M. A. Nanotechnology 2000, 11, 401. (9) Kim, D.-H.; Cho, C.-O.; Roh, Y.-G.; Jeon, H.; Park, Y. S.; Cho, J.; Im, J. S.; Sone, C.; Park, Y.; Choi, W. J.; Park, Q. H. Appl. Phys. Lett. 2005, 87, 203508. (10) Yu, Z.; Raman, A.; Fan, S. Proc. Natl. Acad. Sci. 2010, 107, 17491. (11) Khurgin, J. B.; Boltasseva, A. MRS Bull. 2012, 37, 768. (12) Zheludev, N. I. Science 2010, 328, 582. (13) Suezaki, T.; O’Brien, P. G.; Chen, J. I. L.; Loso, E.; Kherani, N. P.; Ozin, G. A. Adv. Mater. 2009, 21, 559. (14) Suezaki, T.; Chen, J. I. L.; Hatayama, T.; Fuyuki, T.; Ozin, G. A. Appl. Phys. Lett. 2010, 96, 242102. (15) Nelson, E. C.; Dias, N. L.; Bassett, K. P.; Dunham, S. N.; Verma, V.; Miyake, M.; Wiltzius, P.; Rogers, J. A.; Coleman, J. J.; Li, X. L.; Braun, P. V. Nat. Mater. 2011, 10, 676. (16) Subramania, G.; Li, Q. M.; Lee, Y. J.; Figiel, J. J.; Wang, G. T.; Fischer, A. J. Nano Lett. 2011, 11, 4591. (17) Han, S. E.; Stein, A.; Norris, D. J. Phys. Rev. Lett. 2007, 99, 053906. (18) Denny, N. R.; Han, S. E.; Norris, D. J.; Stein, A. Chem. Mater. 2007, 19, 4563. (19) Arpin, K. A.; Losego, M. D.; Braun, P. V. Chem. Mater. 2011, 23, 4783. (20) Yu, X. D.; Lee, Y. J.; Furstenberg, R.; White, J. O.; Braun, P. V. Adv. Mater. 2007, 19, 1689. (21) Lee, Y. J.; Braun, P. V. Adv. Mater. 2003, 15, 563. (22) Lee, Y. J.; Pruzinsky, S. A.; Braun, P. V. Langmuir 2004, 20, 3096. (23) van den Broek, J. M.; Woldering, L. A.; Tjerkstra, R. W.; Segerink, F. B.; Setija, I. D.; Vos, W. L. Adv. Funct. Mater. 2012, 22, 25. (24) Wijnhoven, J. E. G. J.; Vos, W. L. Science 1998, 281, 802. (25) Holland, B. T.; Blanford, C. F.; Stein, A. Science 1998, 281, 538. (26) Braun, P. V.; Wiltzius, P. Nature 1999, 402, 603. (27) Subramania, G.; Constant, K.; Biswas, R.; Sigalas, M. M.; Ho, K. M. Appl. Phys. Lett. 1999, 74, 3933. (28) Subramania, G.; Biswas, R.; Constant, K.; Sigalas, M. M.; Ho, K. M. Phys. Rev. B 2001, 63, 235111. (29) von Freymann, G.; Kitaev, V.; Lotschz, B. V.; Ozin, G. A. Chem. Soc. Rev. 2013, 42, 2528. 285 dx.doi.org/10.1021/cm4023437 | Chem. Mater. 2014, 26, 277−286 Chemistry of Materials Perspective (64) Tétreault, N.; von Freymann, G.; Deubel, M.; Hermatschweiler, M.; Pérez-Willard, F.; John, S.; Wegener, M.; Ozin, G. A. Adv. Mater. 2006, 18, 457. (65) Mihi, A.; Zhang, C. J.; Braun, P. V. Angew. Chem., Int. Ed. 2011, 50, 5711. (66) Arpin, K. A.; Losego, M. D.; Cloud, A. N.; Ning, H.; Mallek, J.; Sergeant, N. P.; Zhu, L.; Yu, Z.; Kalanyan, B.; Parsons, G. N.; Girolami, G. G.; Abelson, J. R.; Fan, S.; Braun, P. V. Nat. Commun. 2013, 4, 2630. (67) Rinnerbauer, V.; Yeng, Y. X.; Chan, W. R.; Senkevich, J. J.; Joannopoulos, J. D.; Soljacic, M.; Celanovic, I. Opt. Express 2013, 21, 11482. (68) Burschka, J.; Pellet, N.; Moon, S.-J.; Humphry-Baker, R.; Gao, P.; Nazeeruddin, M. K.; Gratzel, M. Nature 2013, 499, 316. (69) Braun, P. V.; Zehner, R. W.; White, C. A.; Weldon, M. K.; Kloc, C.; Patel, S. S.; Wiltzius, P. Europhys. Lett. 2001, 56, 207. (70) Braun, P. V.; Zehner, R. W.; White, C. A.; Weldon, M. K.; Kloc, C.; Patel, S. S.; Wiltzius, P. Adv. Mater. 2001, 13, 721. (71) Gansel, J. K.; Latzel, M.; Frolich, A.; Kaschke, J.; Thiel, M.; Wegener, M. Appl. Phys. Lett. 2012, 100, 101109. (72) Gansel, J. K.; Thiel, M.; Rill, M. S.; Decker, M.; Bade, K.; Saile, V.; von Freymann, G.; Linden, S.; Wegener, M. Science 2009, 325, 1513. (73) Thiel, M.; Rill, M. S.; von Freymann, G.; Wegener, M. Adv. Mater. 2009, 21, 4680. (74) Radke, A.; Gissibl, T.; Klotzbucher, T.; Braun, P. V.; Giessen, H. Adv. Mater. 2011, 23, 3018. 286 dx.doi.org/10.1021/cm4023437 | Chem. Mater. 2014, 26, 277−286