Hostname: page-component-8448b6f56d-qsmjn Total loading time: 0 Render date: 2024-04-18T23:09:44.797Z Has data issue: false hasContentIssue false

Rational design of nanomaterials from assembly and reconfigurability of polymer-tethered nanoparticles

Published online by Cambridge University Press:  23 July 2015

Ryan L. Marson
Affiliation:
Department of Materials Science and Engineering, University of Michigan, Ann Arbor, Michigan 48109, USA
Trung Dac Nguyen
Affiliation:
Department of Chemical Engineering, University of Michigan, Ann Arbor, Michigan 48109, USA
Sharon C. Glotzer*
Affiliation:
Department of Materials Science and Engineering, University of Michigan, Ann Arbor, Michigan 48109, USA Department of Chemical Engineering, University of Michigan, Ann Arbor, Michigan 48109, USA
*
Address all correspondence to Sharon C. Glotzer atsglotzer@umich.edu

Abstract

Polymer-based nanomaterials have captured increasing interest over the past decades for their promising use in a wide variety of applications including photovoltaics, catalysis, optics, and energy storage. Bottom-up assembly engineering based on the self- and directed-assembly of polymer-based building blocks has been considered a powerful means to robustly fabricate and efficiently manipulate target nanostructures. Here, we give a brief review of the recent advances in assembly and reconfigurability of polymer-based nanostructures. We also highlight the role of computer simulation in discovering the fundamental principles of assembly science and providing critical design tools for assembly engineering of complex nanostructured materials.

Type
Polymers/Soft Matter Prospective Articles
Copyright
Copyright © Materials Research Society 2015 

Introduction

We are in the midst of a materials revolution—a revolution in which materials will be designed, optimized, and engineered, rather than merely selected, for targeted properties, behavior, and function. Twenty-first century materials and devices will be made and tailored to target specifications, combining disparate and even competing attributes of multiple materials classes to achieve new functionality. They will be dynamic and responsive, able to reconfigure autonomously or on command, changing their appearance, strength, electronic, and other properties. Examples include reconfigurable automotive “skins” to optimize aerodynamics, smart prosthetics and bionics, shape-shifting robots, and camouflaging coatings with adaptive optical properties that, through biomimicry, match surroundings to avoid detection. This intrinsic tailorability and dynamism will contrast starkly with today's relatively static matter that is largely chosen, rather than designed, for the task at hand.[Reference Glotzer1]

Since the first studies of hard disks conducted in the late 1950s, computer simulation has evolved into a powerful, and in many cases, indispensable, tool for investigating atomic, molecular, and mesoscopic systems. Revolutionary advances in computer architectures and simulation algorithms over the past two decades have enabled computational scientists to elucidate problems spanning many orders of magnitude greater in time and length scales and from various angles. In particular, open-source molecular dynamics (MD) packages such as Gromacs,[Reference Berendsen, van der Spoel and van Drunen2] LAMMPS,[Reference Plimpton3] and HOOMD-Blue[Reference Anderson and Glotzer4, Reference Glaser, Nguyen, Anderson, Lui, Spiga, Millan, Morse and Glotzer5] have facilitated the spread of tools and ideas, as well as set a standard for rigorous and reproducible simulations that the scale from laptops to supercomputing clusters. More interesting, perhaps, is the ability of computational studies to offer predictions that are testable, which prove highly valuable for bottom-up engineering of nanomaterials. Finally, the ability to predictively design new materials is at hand, and available to computational experts and non-experts alike.

The rational design of nanomaterials via computer simulation requires identifying the desired target nanostructures, candidate-building blocks, and efficient assembly pathways. In this regard, computational techniques such as Monte Carlo, MD, and self-consistent field theory have played a vital role in predicting assembled structures obtained with almost arbitrary types of building blocks, including block copolymer and nanoparticle (NP) systems.[Reference Bates and Fredrickson6Reference Xu, Jiang, Zhang and Shi8] Coupled with today's increasingly powerful computing capacity, these techniques allow simulators to rapidly predict assembled morphologies, screen candidate-building blocks, and search for efficient assembly pathways over a wide range of parameters.

Next generation materials require new types of building blocks, made in large quantities, that can self-assemble into complex, functional, and reconfigurable structures.[Reference Glotzer and Solomon9] Bottom-up assembly engineering builds upon the thermodynamic foundations of assembly science to engineer these building blocks in order to optimize the yield (quantity as well as quality) of a desired assembly. This paradigm shift is possible today due to the great strides in synthesis and characterization over the past decade. New NP and colloidal building blocks, comprised of organic, inorganic, and/or biological constituents, and with arbitrary shape and interaction patchiness, can now be made nearly to specification.

One interesting class of nanoscale building block is the tethered nanoparticle (TNP):[Reference Zhang, Horsch, Lamm and Glotzer10] a NP to which a polymer tether (T) is permanently attached, creating a surfactant-like, amphiphilic object with a large (NP) head group and a tail (T). In more complex architectures, multiple tethers, of similar, or different types, attached at different places on the NP, are also possible. The TNP conceptual framework allows for tuning domain properties on nearly every scale.[Reference Zhang, Horsch, Lamm and Glotzer10] When assembled, NPs make up the functional domains of the material, while polymer tethers form the continuous space between these domains; alternatively, the polymer domains can serve as the functional material, while NP domains can form surfaces or other structural elements. Moreover, TNPs create a spectrum of mesoscopic assemblies, including: (1) pseudo-two-dimensional (2D) assemblies (e.g. lamellae),[Reference Iacovella, Horsch, Zhang and Glotzer11Reference Nguyen, Zhang and Glotzer13] (2) micellar structures adopting body-centered cubic (BCC), Frank–Kasper, quasiscrystalline patterns,[Reference Iacovella, Keys and Glotzer14] and (3) complex gyroid,[Reference Iacovella, Keys, Horsch and Glotzer15] diamond,[Reference Iacovella and Glotzer16] or other networks.[Reference Marson, Phillips, Anderson and Glotzer17] With such a wide range of assembly behavior, an enormous design space is accessible for the design and discovery of new materials.

We begin with a brief overview of the available toolkit for assembly engineering of TNP systems, which includes current advances in the synthesis of TNPs, as well as how they are being utilized in self-assembly. We then turn our attention to how computer simulation can be applied to rapidly screen candidate TNPs for useful structures, or even to predict the type of TNPs that might assemble a desired structure. Finally, we discuss how reconfigurability can be used as a tool to create more versatile and responsive structures from TNPs, reminiscent of biological systems.

Assembly toolkit for TNPs

Tethered NP building blocks

Tremendous effort in synthesis techniques has been devoted to controlling the shape, size, and composition of nanobuilding blocks for high yield and uniformity. Methods such as seed-mediated growth and redox transmetallation have been widely applied, allowing for a zoo of building blocks in various geometries (e.g. spheres, rods, cubes, plates, and stars) and materials (e.g. semi-conductor, metal and metallic oxides).[Reference Kotov18] Other anisotropy dimensions have also been realized via selective surface modification, functionalization, and compartmentalization.[Reference Lahann19] For instance, NPs and colloids can be functionalized with different chemical moieties through the use of appropriate stabilizers and linkers [Fig. 1(a)]. To precisely control the number of the attached functional groups, one can take advantage of techniques based on “click chemistry”[Reference Yue, Liu, Guo, Wu, Dong, Liu, Huang, Wes- demiotis, Cheng and Zhang21, Reference Kolb, Finn and Sharpless27] and DNA conjugation.[Reference Zhang, Lu, Yager, van der Lelie and Gang22, Reference Macfarlane, Lee, Jones, Harris, Schatz and Mirkin23, Reference Lee, Prytkova and Schatz28, Reference Zhang, Macfarlane, Young, Choi, Hao, Auyeung, Liu, Zhou and Mirkin29]

Figure 1. Building blocks synthesized from various techniques: (a) Functionalization (adopted from Ref. Reference Subbiah, Veerapandian and Yun20); (b) click reactions (Ref. Reference Yue, Liu, Guo, Wu, Dong, Liu, Huang, Wes- demiotis, Cheng and Zhang21); DNA conjugation with (c) multiple DNA strands[Reference Zhang, Lu, Yager, van der Lelie and Gang22] and (d) single DNA strand[Reference Macfarlane, Lee, Jones, Harris, Schatz and Mirkin23] per building block; (e) coordination with metal ions,[Reference Li, Kim, Li and Li24] and (f) electrohydrodynamic co-jetting methods that are used to produce multicompartmentalized building blocks[Reference Lahann19, Reference Lee, Yoon and Lahann25, Reference Lee, Yoon, Rahmanic, Hwang, Bhaskar, Mitragotri and Lahann26]. Panel (c) reprinted by permission from Macmillan Publishers Ltd.: Nature Nanotechnology [Reference Zhang, Lu, Yager, van der Lelie and Gang22], copyright 2013. Panel (d) from Ref. Reference Macfarlane, Lee, Jones, Harris, Schatz and Mirkin23. Reprinted with permission from AAAS. Panel (e) adapted from Ref. Reference Li, Kim, Li and Li24 with permission of The Royal Society of Chemistry.

In the case of “click chemistry”[Reference Yue, Liu, Guo, Wu, Dong, Liu, Huang, Wes- demiotis, Cheng and Zhang21, Reference Kolb, Finn and Sharpless27] [Fig. 1(b)], multicomponent building blocks, e.g. surfactants and shape amphiphiles, are created through a series of sequential “click reactions”, namely the copper-catalyzed azide-alkyne cycloaddition reaction and the thiolene reaction.[Reference Yue, Liu, Guo, Wu, Dong, Liu, Huang, Wes- demiotis, Cheng and Zhang21, Reference Kolb, Finn and Sharpless27] This technique has recently been used to interchange nearly every portion of the building block.[Reference Yue, Liu, Guo, Wu, Dong, Liu, Huang, Wes- demiotis, Cheng and Zhang21, Reference Zhang, Yu, Wang, Sun, Hsieh, Li, Dong, Yue, Van Horn and Cheng30Reference Yu, Li, Dong, Yue, Lin, Feng, Huang, Zhang and Cheng32] Examples include the ability to interchange head groups from polyhedral oligomeric silsesquioxane (POSS) cages to gold NPs, as well as to exchange polymer types such as polystyrene or poly(methyl methacrylate) (PMMA); additionally, the size, number, and placement of these chemically distinct domains can be controlled, such as tethering a single corner of a POSS molecule with one or more polymer tethers.[Reference Yue, Liu, Guo, Wu, Dong, Liu, Huang, Wes- demiotis, Cheng and Zhang21, Reference Kolb, Finn and Sharpless27, Reference Zhang, Yu, Wang, Sun, Hsieh, Li, Dong, Yue, Van Horn and Cheng30Reference Yu, Li, Dong, Yue, Lin, Feng, Huang, Zhang and Cheng32]

Related methods have been used to create additional types of polymer–NP composite blocks. Protein–polymer conjugates, for example, were created using a covalent binding maleimide–thiol coupling reaction to create bioconjugates.[Reference Thomas, Glassman and Olsen33Reference Lam and Olsen35] These bioconjugates were also been used in self-assembly experiments, demonstrating the same phase diversity as related building blocks.[Reference Lam and Olsen35] Other materials, such as polyoxometalate anionic metal–oxygen nanocages, have been successfully attached to polymer tethers via covalent bonds.[Reference Han, Xiao, Zhang, Liu, Zheng, He and Wang36, Reference Rieger, Antoun, Lee, Chenal, Pembouong, Haye, Azcarate, Hasenknopf and Lacote37] As these studies demonstrate, a wide range of materials are amenable to this type of modification and are experimentally accessible, opening up an enormous design space for new functional materials.

An alternative strategy that has also produced numerous candidate TNPs is to functionalize NPs or colloids with a finite number of DNA oligomers [Figs 1(c) and 1(d)]. The highly specific interactions between complementary strands of DNA induce attraction between specific NP-type pairs. This suggests a robust framework that has already enabled the design of numerous NP superlattices upon tuning the NP size and bond distance using DNA linkers.[Reference Zhang, Lu, Yager, van der Lelie and Gang22, Reference Macfarlane, Lee, Jones, Harris, Schatz and Mirkin23, Reference Zhang, Macfarlane, Young, Choi, Hao, Auyeung, Liu, Zhou and Mirkin29]

TNP colloidal “molecules” are indeed reminiscent of their molecular counterparts, as their interaction can be made directional and highly specific. For example, coordination polymeric structures exist that link polymeric or molecular subunits using non-covalent interactions, creating adaptive, and responsive polymer-like superstructures. In these building blocks interactions are made highly specific at certain sites via the use of suitable metallic ions and ligands [Fig. 1(e)].[Reference Li, Kim, Li and Li24] Patchy particles and multicompartmentalized particles are realized via selective surface treatment methods[Reference Pawar and Kretzschmar38] and electrohydrodynamic co-jetting [Fig. 1(e)], respectively. Recently, Sacanna and Pine[Reference Sacanna and Pine39] reported a simple, yet generic, route to fabricating colloids with tunable cavity, which enable a host of complex assemblies.

It is interesting to note that techniques now are available to consider the synthesis and fabrication of building blocks whose size and shape can change, in situ, in a controlled manner. For example, colloidal gold nanorods can be shortened or shifted irreversibly to other shapes such as spheres, bent, twisted, or φ-shaped using laser pulses with different wavelengths and widths.[Reference Chang, Shih, Chen, Lai and Wang40Reference Link, Wang and El-sayed42] Kim et al.[Reference Kim, Lee, Lim and Lee43] reported thermally responsive capsule structures with 25 nm diameter pores on shells formed by hierarchical self-assembly of double-tethered rod amphiphiles. Upon heating or cooling, the hydrophilic oligo-(ethylene oxide) coils at one end of the rods shrink or expand, respectively, resulting in a reversible closed/open gating motion of the nanopores. They also demonstrated a reversible transformation between 2D sheets and tubular structures assembled by laterally grafted rod amphiphiles upon heating via a similar mechanism.[Reference Lee, Kim and Lee44] Alternatively, polypeptide-based block copolymers can be used as stimuli-responsive building blocks due to the ability of the polypeptide segments to adapt various conformations.[Reference Chockalingam, Blenner and Banta45Reference Gebhardt, Ahn, Venkatachalam and Savin47] For instance, Gebhardt et al.[Reference Gebhardt, Ahn, Venkatachalam and Savin47] demonstrated that the polypeptide rod segment in the poly(butadiene)–poly(L-lysine) block copolymers undergoes an R-helix to coil transition in response to a change in pH and temperature. Yoo et al.[Reference Yoo and Mitragotri48] synthesized polymeric particles that are able to switch shape in response to changes in temperature, pH, and chemical additives. Multicompartmental particles based on poly-(lactic-co-glycolic acid) are able to morph into various shapes upon heating one of the polymeric compartments above its glass transition temperature.[Reference Lee, Yoon, Rahmanic, Hwang, Bhaskar, Mitragotri and Lahann26]

Assembled nanostructures

While initial experimental efforts were primarily directed at TNP synthesis, attention has increasingly turned to structure assembly. Tethering short polymer segments to NPs produces an enormous design space, comparable with that of copolymers, but with an expanded spectrum of properties and structures beyond the complex phases exhibited by coordination polymers and block copolymers[Reference Bates and Fredrickson6, Reference Meuler, Hillmyer and Bates7, Reference Li, Kim, Li and Li24, Reference Lee, Prytkova and Schatz28, Reference Bates and Fredrickson49Reference Hayashida, Dotera, Takano and Matsushita53] [Figs 2(a)–2(c)].

Figure 2. Static assemblies: (a) Nanotubular assemblies from coordination polymers upon addition of divalent metal ions (red circle),[Reference Li, Kim, Li and Li24] (b) assembled structures from shape amphiphiles, including lamellar, gyroid, hexagonal cylinders, and BCC micelle phases, as reported in[Reference Yue, Liu, Guo, Wu, Dong, Liu, Huang, Wes- demiotis, Cheng and Zhang21]. (c) Dodecagonal quasicrystal formed in a blend of polyisoprene–polystyrene–poly(2-vinylpyridine) star block copolymers/polysterene homopolymer blend.[Reference Hayashida, Dotera, Takano and Matsushita53] (d) Ordered structures formed by gold NPs functionalized with complementary DNA strands (top): NaCl lattice (middle) and simple cubic lattice (bottom)[Reference Zhang, Lu, Yager, van der Lelie and Gang22]. Panel (a) adapted from Ref. Reference Li, Kim, Li and Li24 with permission of The Royal Society of Chemistry. Panel (c) reprinted with permission from Ref. Reference Hayashida, Dotera, Takano and Matsushita53. Copyright (2007) by the American Physical Society. Panel (d) from Ref. Reference Macfarlane, Lee, Jones, Harris, Schatz and Mirkin23. Reprinted with permission from AAAS.

As mentioned previously, direct realizations of TNP assembly have come from the use of “click chemistry” to attach tethers to NPs directly, creating gram quantities of material.[Reference Yue, Liu, Guo, Wu, Dong, Liu, Huang, Wes- demiotis, Cheng and Zhang21, Reference Kolb, Finn and Sharpless27, Reference Zhang, Yu, Wang, Sun, Hsieh, Li, Dong, Yue, Van Horn and Cheng30] TNPs created with this method have been used to self-assemble many of the same structures observed in block copolymers, but with the additional benefit of controlling the specific material composition of each domain of the material. By tuning various portions of the block geometry, different types of structures are accessible. For example, by increasing the length of a polystyrene tether attached to the corner of a POSS cube Cheng et al.[Reference Yu, Yue, Hsieh and Li31] can tune between lamella, the double-gyroid, hexagonal tubes, and BCC micelles, as seen in Fig. 2(b). Moreover, these structures demonstrated a high degree of sensitivity to the block features; by fixing the volume fraction of each component, but adding additional tethers to one corner of the POSS cube, they could induce a phase inversion in the hexagonal tubes phase, resulting in a matrix of POSS and tubes of polymer. Similarly, by replacing a single tail with two tails of half the length, they could select either the double gyroid for the former or hexagonal tubes for the latter. As we will highlight in the following section, computer simulation has played a large role in predicting the types of morphologies we observe in TNP systems and will continue to help guide work as these materials are integrated into devices. It is important to note that while initial experimental work has demonstrated the feasibility of the synthesis and assembly of these materials, there are still enormous opportunities for the design of new building blocks and structures.

For building blocks with highly specific interactions, it is possible to design nanostructures with a higher degree of complexity. For instance, using self-complementary and non-self-complementary DNA strands, Macfarlane et al.[Reference Macfarlane, Lee, Jones, Harris, Schatz and Mirkin23] have showed that one can program the multivalent interaction between DNA-functionalized NPs. With such encoded interactions, the resulting structure can be a NaCl or simple cubic lattice, depending on the size ratio of the inorganic NPs [Fig. 2(d)].

Shape-changing TNPs provide additional possibilities for structural complexity in resulting assemblies. For example, Lee et al.[Reference Lee, Kim and Lee44] showed that a bilayer sheet formed by rod–coil molecules in aqueous solution roll into a tubular structure upon heating and vice versa [Fig. 3(a)]. The driving force for reconfiguration is attributed to the increase or decrease in the excluded volume interaction of the coil segments in response to the change in temperature. As another example, the lattice spacing of the superlattices formed by DNA-grafted gold NPs was shown to be tunable upon adding suitable set and unset complementary DNA strands[Reference Lee, Kim and Lee44] [Fig. 3(b)]. Consequently, one can imagine the application of such dynamic superlattices to responsive photonic band gap materials or automative “skins”. By tuning the interfacial tension via temperature, Gibaud et al.[Reference Gibaud, Barry, Zakhary, Henglin, Ward, Yang, Berciu, Oldenbourg, Hagan, Nicastro, Meyer and Dogic55] reported a novel way to induce morphological transitions in colloidal membranes through changes occurred at microscopic levels [Fig. 3(c)]. Zhang et al.[Reference Zhang, Lu, van der Lelie and Gang56] demonstrated that a simple cubic lattice of palladium nanocubes coated by dodecanethiol ligands in toluene, a poor solvent for the ligands, transforms into a rhombohedral lattice upon solvent evaporation [Fig. 3(d)]. Shifting of the NP shape from cuboid to ellipsoid was attributed to the swelling ligands as the solvent concentration decreased during evaporation.

Figure 3. Reconfigurable assemblies: (a) Reversible transformation between flat sheet and tubule formed by rod–coil macromolecules upon heating and cooling[Reference Lee, Kim and Lee44], (b) BCC structure with lattice spacings tunable by adding set and unset DNA strands[Reference Maye, Kumara, Nykypanchuk, Sherman and Gang54]; (c) transition from a disk into twisted ribbons upon quenching[Reference Gibaud, Barry, Zakhary, Henglin, Ward, Yang, Berciu, Oldenbourg, Hagan, Nicastro, Meyer and Dogic55]; (d) transition from a simple cubic lattice to face-centered cubic lattice assembled by dodecanethiol ligated palladium nanocubes[Reference Zhang, Lu, van der Lelie and Gang56]. Panel (a) from E. Lee, J. K. Kim, and M. Lee. “Reversible scrolling of 2D sheets from self-assembly of laterally grafted amphiphilic rods.” Angew. Chem. Int. Ed. Engl., John Wiley and Sons Publishing Group. Panels (b) and (c) reprinted with permission from Macmillan Publishers Ltd: Nat. Nanotechnol. [Reference Maye, Kumara, Nykypanchuk, Sherman and Gang54], copyright 2009, and Nature [Reference Gibaud, Barry, Zakhary, Henglin, Ward, Yang, Berciu, Oldenbourg, Hagan, Nicastro, Meyer and Dogic55], copyright 2012. Panel (d) reprinted with permission from Ref. Reference Zhang, Lu, van der Lelie and Gang56. Copyright (2011) by the American Physical Society.

Computational design of functional nanostructures

Given the versatility of the assembly toolkit and the enormity of the design space provided by TNPs, what types of structures are possible? Can we design and assemble domain-specific materials with precisely controlled bulk and nanoscale properties?

This question drove an initial flurry of computational investigation into tethered NPs[Reference Zhang, Horsch, Lamm and Glotzer10, Reference Glotzer, Horsch, Iacovella, Zhang, Chan and Zhang57] a decade ago. Nanoparticles, it was hypothesized, would offer a truly limitless design platform, provided they could successfully be used to create composite polymer–NP blocks. Initial investigations focused on establishing relationships between existing theory for polymers and surfactants, and TNPs via simulation.[Reference Iacovella, Horsch, Zhang and Glotzer11, Reference Tschierske58] Toward this end, the focus has been on the phase behavior[Reference Zhang, Horsch, Lamm and Glotzer10, Reference Iacovella, Horsch, Zhang and Glotzer11, Reference Iacovella, Keys, Horsch and Glotzer15, Reference Glotzer, Horsch, Iacovella, Zhang, Chan and Zhang57, Reference Phillips, Iacovella and Glotzer59] and properties[Reference Jayaraman and Schweizer60Reference Nair, Wentzel and Jayaraman64] of TNPs, as a function of the size of the isotropic NPs. Figure 4 shows examples of interesting assembled structures from our studies, including double gyroids,[Reference Marson, Phillips, Anderson and Glotzer17] cylinders and columnar structures,[Reference Iacovella and Glotzer65] bilayer sheets,[Reference Nguyen and Glotzer66] and dodecagonal quasicrystals.[Reference Iacovella, Keys and Glotzer14] Key amongst these findings was the realization that the phase behavior of tethered NPs, to a great extent, is consistent with that of surfactants and liquid crystals.[Reference Iacovella, Horsch, Zhang and Glotzer11] Recently, many of these initial findings have also been corroborated with experiments.[Reference Yu, Yue, Hsieh and Li31]

Figure 4. Simulation predictions: (a) Double gyroid formed by TNP telechelics[Reference Marson, Phillips, Anderson and Glotzer17]; (b) tetragonally cylinder structure, and [6;6;6] columnar structure assembled by di-tethered nanospheres with different planar angles, θ, between two tethers[Reference Iacovella and Glotzer65]; (c) bilayer sheets and honeycomb grid formed by laterally tethered nanorods[Reference Nguyen and Glotzer66]; (d) dodecagonal quasicrystal formed by mono-tethered NPs[Reference Iacovella, Keys and Glotzer14]. Panel (a) reprinted with permission from Ref. Reference Marson, Phillips, Anderson and Glotzer17. Copyright (2014) American Chemical Society.” Panel (b) reprinted with permission from Ref. Reference Iacovella and Glotzer65. Copyright (2009) American Chemical Society. Panel (b) reprinted with permission from Ref. Reference Nguyen and Glotzer66. Copyright (2010) American Chemical Society.

Given the close relationship between tethered NPs and liquid crystals,[Reference Tschierske58] the roles of the shape of the NP head group have also been extensively investigated.[Reference Horsch, Zhang and Glotzer12, Reference Nguyen, Zhang and Glotzer13, Reference Horsch, Zhang and Glotzer67Reference Horsch, Zhang and Glotzer70] In these cases, it was again found that structures and transitions similar to those in liquid crystals were observed, such as lamella, gyroids, cylinders, and micelles. However, the connectivity and coordination of the micelles and networks were altered by the packing of the head groups. Additional types of liquid-crystalline ordering were found within these domains, such as long micelles and lamella,[Reference Horsch, Zhang and Glotzer69] twisted columns and sheets with liquid crystalline order,[Reference Horsch, Zhang and Glotzer12, Reference Nguyen, Zhang and Glotzer13, Reference Horsch, Zhang and Glotzer67] honeycombs,[Reference Nguyen, Zhang and Glotzer13] and gyroids.[Reference Iacovella, Horsch and Glotzer68]

The focus of recent studies has included increasing the architectural complexity of TNP building-blocks, including interactions between blocks and the solvent, as well as tuning their respective block volume fraction, thereby mirroring trends within the copolymer community.[Reference Bates, Hillmyer, Lodge, Bates, Delaney and Fredrickson50] Asymmetry within these domains, for example, can be tuned between single- and double-gyroid morphologies.[Reference Marson, Phillips, Anderson and Glotzer17] Moreover, controlling their architecture and interaction has led to open, simple cubic, and diamond networks,[Reference Capone, Coluzza, LoVerso, Likos and Blaak71] which are known to be useful candidates for photonic applications.[Reference Urbas, Maldovan, DeRege and Thomas72Reference Maldovan and Thomas75] As the spacing and characteristics of individual domains can be tuned in a reliable manner, it is possible to propose new and previously inaccessible structures, as well as create complex and hierarchical morphologies.[Reference Schroder-Turk, Varslot, de Campo, Kapfer and Mickel76Reference Kirkensgaard, Evans, de Campo and Hyde83] Finally, the solvent selectivity is important to the kinds of morphologies that will assemble, and has been found to dramatically affect the nature of self-assembled structures allowing for a wide variety of tunable micellar structures.[Reference Ma, Hu and Wang84]

As the complexity of static assemblies has grown over the past decade, so too has the desire to incorporate these assemblies into devices and other applications. Recently, particular interest has grown in tunable, responsive assemblies. Drawing inspiration from nature and biology, researchers aim to create building blocks that can dynamically respond to either internal or external stimuli. Enzymes and proteins, for example, change their conformation and interactions to activate processes within the body as necessary.[Reference Lai, King and Yeates85] Similarly, simplistic versions of these building blocks could, for example, suppress the formation of close-packed structures that then lead to other complex structures that would be otherwise inaccessible, such as quasicrystals or Frank–Kasper phases.[Reference Iacovella, Keys and Glotzer14, Reference Damasceno, Engel and Glotzer86, Reference Engel, Damasceno, Phillips and Glotzer87]

Toward this end, the work has begun on dynamic blocks that can reconfigure during assembly. Early investigations were largely centered on simply introducing anisotropy in a single dimension to building blocks during a simulation. Using computer simulation, for example, Batista et al.[Reference Batista and Miller88] studied the crystalline packing of deformable spherical colloids. They modeled the shape change in the colloids by allowing the spherical particles to continuously change shape into prolate or oblate ellipsoids. Interestingly, the system was shown to undergo a second transition to an orientationally ordered crystal upon shape shifting. Nguyen et al.[Reference Nguyen and Glotzer66] predicted that when the rod segments in polymer-tethered nanorod systems are shortened or lengthened continuously, the assembled structures reconfigure between various mesophases such as square grids, honeycomb, and bilayer sheets. This idea was extended to more complex structural transitions, including transitions between 2D tiles such as rhombi and pentagons, as well as kinked rods, zigzags, and cross structures.[Reference Nguyen, Jankowski and Glotzer89] Importantly, these studies showed not only reversible transitions between equilibrium assemblies, but also the ability to obtain configurations that would otherwise be inaccessible to static building blocks. Such transitions are exciting because they open up the possibility for obtaining structures that otherwise do not appear in equilibrium self-assembly.[Reference Eshet, Bruneval and Parrinello90]

Conclusion and outlook

With powerful tools in hand both to synthesize TNP-building blocks of arbitrary complexity,[Reference Zhang, Yu, Wang, Sun, Hsieh, Li, Dong, Yue, Van Horn and Cheng30] and to simulate their assembly behavior,[Reference Marson, Phillips, Anderson and Glotzer17] it is tempting to begin exploring complex assemblies that could lead to new applications and devices.[Reference Saba, Thiel, Turner, Hyde, Gu, Grosse-Brauckmann, Neshev, Mecke and Schroder-Turk91Reference Liu, Hou and Gao94] A pressing challenge is unifying these tools before moving into application. We have highlighted just a few of the conceptual design axes, as we highlight in Fig. 5. Despite having already sketched out a framework that encompasses the vast design space,[Reference Glotzer and Solomon9] linking structural prediction to properties within a simulation framework would expedite application-specific design. As new building blocks and assemblies continue to emerge, the fundamental question is no longer if we can build targeted structures of exquisite complexity, but how we can design a target structure? What specific building blocks or set of interactions may lead to its formation, and how can the assembled structure be utilized from a technology standpoint?

Figure 5. Anisotropy axis for relevant assembly dimensions in TNP systems. By tuning NP shape and interaction, researchers can tune the resultant structure and ordering in TNP systems. Additionally, reconfigurability of these building blocks will allow for tunable and responsive next-generation materials.

Knowing the target structure that will suit a particular application is one of the most crucial aspects of the problem. Diamond structures, for example, are known to be excellent candidates for photonics applications.[Reference Urbas, Maldovan, DeRege and Thomas72Reference Maldovan and Thomas75] The challenge is then to supply a proper set of design conditions to produce the structure. Computation allows us to rapidly screen candidate blocks for targeted assemblies. New data science approaches, coupled with high performance computing-enabled prediction of TNP self-assembly, promise to further expedite material design in this space,[Reference Oganov, Lyakhov and Valle95Reference Miskin and Jaeger97] in the spirit of the Materials Genome Initiative.[98]

Tethered NPs and related structures hold a great deal of promise as candidate-building blocks for next generation materials. Their design flexibility makes them useful for a wide range of potential applications. The theory that drives their assembly is well understood, while computation allows us to quickly make predictions about their assembly behavior. Many realizations of TNP assemblies that were initially predicted via simulation have been realized in the laboratory. Researchers now have the ability to predict, design, and assemble targeted structures for a given application. Importantly, these predictions are not specific to particular material types, and thus can guide the synthesis and assembly of many types of TNPs. Moreover, this design approach provides experimentalists flexibility in their design strategy—the same set of results holds equally for polymer-tethered gold particles, or proteins, as for giant surfactants of ligand functionalized POSS molecules, provided the geometric length scales of particle and tether are commensurate. Any materials that are amenable to surface modification can be incorporated into this scheme, providing a versatile materials design platform.

Acknowledgments

Research supported by the US Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering under Award #DE-FG02-02ER46000.

References

1.Glotzer, S.C.: Assembly engineering: materials design for the 21st centruy. Chem. Eng. Sci. 121, 39 (2014).CrossRefGoogle Scholar
2.Berendsen, H.J.C., van der Spoel, D., and van Drunen, R.: GROMACS: a message-passing parallel molecular dynamics implementation. Comput. Phys. Commun. 91, 4356 (1995).CrossRefGoogle Scholar
3.Plimpton, S.: Fast Parallel Algorithms for Short-Range Molecular Dynamics (1995).Google Scholar
4.Anderson, J.A. and Glotzer, S.C.: the development and expansion of HOOMD-blue through six years of GPU proliferation. arXiv preprint arXiv:1308.5587 (2013).Google Scholar
5.Glaser, J., Nguyen, T.D., Anderson, J.A., Lui, P., Spiga, F., Millan, J.A., Morse, D.C., and Glotzer, S.C.: Strong scaling of general-purpose molecular dynamics simulations on GPUs. Comput. Phys. Commun. 192, 97107 (2015).CrossRefGoogle Scholar
6.Bates, F.S. and Fredrickson, G.H.: Block copolymers—designer soft materials. Phys. Today 52, 32 (1999).CrossRefGoogle Scholar
7.Meuler, A.J., Hillmyer, M.A., and Bates, F.S.: Ordered network mesostructures in block polymer materials. Macromolecules 42, 72217250 (2009).CrossRefGoogle Scholar
8.Xu, W., Jiang, K., Zhang, P., and Shi, A.-C.: A strategy to explore stable and metastable ordered phases of block copolymers. J. Phys. Chem. B 117, 52965305 (2013).Google Scholar
9.Glotzer, S.C. and Solomon, M.J.: Anisotropy of building blocks and their assembly into complex structures. Nat. Mater. 6, 557562 (2007).Google Scholar
10.Zhang, Z., Horsch, M.A., Lamm, M.H., and Glotzer, S.C.: Tethered nano building blocks: toward a conceptual framework for nanoparticle self-assembly. Nano Lett. 3, 13411346 (2003).Google Scholar
11.Iacovella, C.R., Horsch, M.A., Zhang, Z., and Glotzer, S.C.: Phase diagrams of self-assembled mono-tethered nanospheres from molecular simulation and comparison to surfactants. Langmuir 21, 94889494 (2005).Google Scholar
12.Horsch, M.A., Zhang, Z., and Glotzer, S.C.: Simulation studies of self-assembly of end-tethered nanorods in solution and role of rod aspect ratio and tether length. J. Chem. Phys. 125, 184903 (2006).Google Scholar
13.Nguyen, T.D., Zhang, Z., and Glotzer, S.C.: Molecular simulation study of self-assembly of tethered V-shaped nanoparticles. J. Chem. Phys. 129, 244903 (2008).Google Scholar
14.Iacovella, C.R., Keys, A.S., and Glotzer, S.C.: Self-assembly of soft-matter quasicrystals and their approximants. Proc. Natl. Acad. Sci. U.S.A. 108, 2093520940 (2011).Google Scholar
15.Iacovella, C., Keys, A., Horsch, M., and Glotzer, S.: Icosahedral packing of polymer-tethered nanospheres and stabilization of the gyroid phase. Phys. Rev. E 75, 14 (2007).Google Scholar
16.Iacovella, C.R. and Glotzer, S.C.: Phase behavior of ditethered nanospheres. Soft Matter 5, 44924498 (2009).Google Scholar
17.Marson, R.L., Phillips, C.L., Anderson, J.A., and Glotzer, S.C.: Phase behavior and complex crystal structures of self-assembled tethered nanoparticle telechelics. Nano Lett. 14, 20712078 (2014).Google Scholar
18.Kotov, N.A.: Nanoparticle Assemblies and Superstructures (CRC Press, Boca Raton, FL, 2014).Google Scholar
19.Lahann, J.: Recent progress in nano-biotechnology: compartmentalized micro- and nanoparticles via electrohydrodynamic co-jetting. Small 7, 11491156 (2011).CrossRefGoogle ScholarPubMed
20.Subbiah, R., Veerapandian, M., and Yun, K.S.: Nanoparticles: functionalization and multifunctional applications in biomedical sciences. Curr. Med. Chem. 17, 45594577 (2010).CrossRefGoogle ScholarPubMed
21.Yue, K., Liu, C., Guo, K., Wu, K., Dong, X.-H., Liu, H., Huang, M., Wes- demiotis, C., Cheng, S.Z.D., and Zhang, W.-B.: Exploring shape amphiphiles beyond giant surfactants: molecular design and click synthesis. Polym. Chem. 4, 10561067 (2013).Google Scholar
22.Zhang, Y., Lu, F., Yager, K.G., van der Lelie, D., and Gang, O.: A general strategy for the DNA-mediated self-assembly of functional nanoparticles into heterogeneous systems. Nat. Nanotechnol. 8, 865872 (2013).Google Scholar
23.Macfarlane, R.J., Lee, B., Jones, M.R., Harris, N., Schatz, G.C., and Mirkin, C.A.: Nanoparticle Superlattice Engineering with DNA. Science 334, 204208 (2011).CrossRefGoogle ScholarPubMed
24.Li, W., Kim, Y., Li, J., and Li, M.: Dynamic self-assembly of coordination polymers in aqueous solution. Soft Matter 10, 52315242 (2014).Google Scholar
25.Lee, K.J., Yoon, J., and Lahann, J.: Recent advances with anisotropic particles. Curr. Opin. Colloid Interface Sci. 16, 195202 (2011).CrossRefGoogle Scholar
26.Lee, K., Yoon, J., Rahmanic, S., Hwang, S., Bhaskar, S., Mitragotri, S., and Lahann, J.: Spontaneous shape reconfigurations in multicompartmental microcylinders. Proc. Natl. Acad. Sci. U.S.A. 109, 1605716062 (2012).Google Scholar
27.Kolb, H.C., Finn, M.G., and Sharpless, K.B.: Click chemistry: diverse chemical function from a few good reactions. Angew. Chem. Int. Ed. Engl. 40, 20042021 (2001).Google Scholar
28.Lee, O.-S., Prytkova, T.R., and Schatz, G.C.: Using DNA to link gold nanoparticles, polymers and molecules: a theoretical perspective. J. Phys. Chem. Lett. 1, 17811788 (2010).Google Scholar
29.Zhang, C., Macfarlane, R.J., Young, K.L., Choi, C.H.J., Hao, L., Auyeung, E.- l, Liu, G., Zhou, X., and Mirkin, C.A.: A general approach to DNA-programmable atom equivalents. Nat. Mater. 12, 741746 (2013).Google Scholar
30.Zhang, W.B., Yu, X., Wang, C.L., Sun, H.J., Hsieh, I.F., Li, Y., Dong, X.H., Yue, K., Van Horn, R., and Cheng, S.Z.D.: macromolecular science: from “nanoatoms” to giant molecules. Macromolecules 47, 12211239 (2014).Google Scholar
31.Yu, X., Yue, K., Hsieh, I.-F., and Li, Y.: Giantsurfactants provide a versatile platform for sub-10-nm nanostructure engineering. Proc. Natl. Acad. Sci. U.S.A. 110, 1007810083 (2013).Google Scholar
32.Yu, X., Li, Y., Dong, X.-H., Yue, K., Lin, Z., Feng, X., Huang, M., Zhang, W.-B., and Cheng, S.Z.D.: Giant surfactants based on molecular nanoparticles: Precise synthesis and solution self-assembly. J. Polym. Sci. B: Polym. Phys., 13091325 (2014).Google Scholar
33.Thomas, C.S., Glassman, M.J., and Olsen, B.D.: Solid-state nanostructured materials from self-assembly of a globular protein–polymer diblock copolymer. ACS Nano 5, 56975707 (2011).Google Scholar
34.Olsen, B.D.: Self-assembly of globular-protein-containing block copolymers. Macromol. Chem. Phys. 214, 16591668 (2013).Google Scholar
35.Lam, C.N. and Olsen, B.D.: Phase transitions in concentrated solution self-assembly of globular protein–polymer block copolymers. Soft Matter 9, 23932402 (2013).Google Scholar
36.Han, Y., Xiao, Y., Zhang, Z., Liu, B., Zheng, P., He, S., and Wang, W.: Synthesis of polyoxometalate–polymer hybrid polymers and their hybrid vesicular assembly. Macromolecules 42, 65436548 (2009).Google Scholar
37.Rieger, J., Antoun, T., Lee, S.-H., Chenal, M., Pembouong, G., Haye, J.L., Azcarate, I., Hasenknopf, B., and Lacote, E.: Synthesis and characterization of a thermoresponsive polyoxometalate–polymer hybrid. Chemistry 18, 33553361 (2012).Google Scholar
38.Pawar, A.B. and Kretzschmar, I.: Fabrication, assembly, and application of patchy particles. Macro- Mol. Rapid Commun. 31, 150168 (2010).Google Scholar
39.Sacanna, S. and Pine, D.: Shape-anisotropic colloids: building blocks for complex assemblies. Curr. Opin. Colloid Interface Sci. 16, 96105 (2011).CrossRefGoogle Scholar
40.Chang, S.S., Shih, C.W., Chen, C.D., Lai, W.C., and Wang, C.R.C.: The shape transition of gold nanorods. Langmuir 15, 701709 (1999).CrossRefGoogle Scholar
41.Link, S., Burda, C., Nikoobakht, B., and El-sayed, M.A.: Laser-induced shape changes of colloidal gold nanorods using femtosecond and nanosecond laser pulses. J. Phys. Chem. B 104, 61526153 (2000).Google Scholar
42.Link, S., Wang, Z.L., and El-sayed, M.A.: How does a gold nanorod melt? J. Phys. Chem. B 104, 78677870 (2000).Google Scholar
43.Kim, J.K., Lee, E., Lim, Y.B., and Lee, M.: Supramolecular capsules with gated pores from an amphiphilic rod assembly. Angew. Chem. Int. Ed. Engl. 47, 46624666 (2008).Google Scholar
44.Lee, E., Kim, J.K., and Lee, M.: Reversible scrolling of two-dimensional sheets from self-assembly of laterally-grafted amphiphilic rods. Angew. Chem. Int. Ed. Engl. 48, 3657 (2009).Google Scholar
45.Chockalingam, K., Blenner, M., and Banta, S.: Design and application of stimulus-responsive peptide systems. Prot. Eng., Des. Sel. 20, 155161 (2007).Google Scholar
46.Gebhardt, K.E., Ahn, S., Venkatachalam, G., and Savin, D.A.: Rod-sphere transition in polybutadiene-poly(l-lysine) block copolymer assemblies. Langmuir 23, 28512856 (2007).CrossRefGoogle ScholarPubMed
47.Gebhardt, K.E., Ahn, S., Venkatachalam, G., and Savin, D.A.: Role of secondary structure changes on the morphology of polypeptide-based block copolymer vesicles. J. Colloid Interface Sci. 317, 7076 (2008).Google Scholar
48.Yoo, J.W. and Mitragotri, S.: Polymer particles that switch shape in response to a stimulus. Proc. Nat. Acad. Sci. U.S.A. 107, 1120511210 (2010).CrossRefGoogle ScholarPubMed
49.Bates, F.S. and Fredrickson, G.H.: Block copolymer thermodynamics: theory and experiment. Annu. Rev. Phys. Chem. 41, 525557 (1990).Google Scholar
50.Bates, F.S., Hillmyer, M.A., Lodge, T.P., Bates, C.M., Delaney, K.T., and Fredrickson, G.H.: Multiblock polymers: panacea or Pandora's box? Science 336, 434440 (2012).Google Scholar
51.Epps, T.H. III, Cochran, E.W., Hardy, C.M., Bailey, T.S., Waletzko, R.S., and Bates, F.S.: Network phases in ABC triblock copolymers. Macromolecules 37, 70857088 (2004).Google Scholar
52.Qin, J., Bates, F.S., and Morse, D.C.: Phase behavior of nonfrustrated ABC triblock copolymers: weak and intermediate segregation. Macromolecules 43, 51285136 (2010).Google Scholar
53.Hayashida, K., Dotera, T., Takano, A., and Matsushita, Y.: Polymeric quasicrystal: mesoscopic quasicrystalline tiling in ABC star polymers. Phys. Rev. Lett. 98, 14 (2007).Google Scholar
54.Maye, M.M., Kumara, M.T., Nykypanchuk, D., Sherman, W.B., and Gang, O.: Switching binary states of nanoparticle superlattices and dimer clusters by DNA strands. Nat. Nanotechnol. 5, 116120 (2010).Google Scholar
55.Gibaud, T., Barry, E., Zakhary, M.J., Henglin, M., Ward, A., Yang, Y., Berciu, C., Oldenbourg, R., Hagan, M.F., Nicastro, D., Meyer, R.B., and Dogic, Z.: Reconfigurable self-assembly through chiral control of interfacial tension. Nature 481, 348351 (2012).Google Scholar
56.Zhang, Y., Lu, F., van der Lelie, D., and Gang, O.: Continuous phase transformation in nanocube assemblies. Phys. Rev. Lett. 107, 135701 (2011).Google Scholar
57.Glotzer, S.C., Horsch, M.A., Iacovella, C.R., Zhang, X.L., Chan, E., and Zhang, X.: Self-assembly of anisotropic tethered nanoparticle shape amphiphiles. Curr. Opin. Colloid Interface Sci. 10, 287295 (2005).Google Scholar
58.Tschierske, C.: Liquid crystal engineering–new complex mesophase structures and their relations to polymer morphologies, nanoscale patterning and crystal engineering. Chem. Soc. Rev. 36, 19301970 (2007).Google Scholar
59.Phillips, C.L., Iacovella, C.R., and Glotzer, S.C.: Stability of the double gyroid phase to nanoparticle polydispersity in polymer-tethered nanosphere systems. Soft Matter 6, 16931703 (2010).Google Scholar
60.Jayaraman, A. and Schweizer, K.S.: effective interactions, structure, and phase behavior of lightly tethered nanoparticles in polymer melts. Macromolecules 41, 94309438 (2008).CrossRefGoogle Scholar
61.Jayaraman, A. and Schweizer, K.S.: Structure and assembly of dense solutions and melts of single tethered nanoparticles. J. Chem. Phys. 128, 164904 (2008).Google Scholar
62.Jayaraman, A. and Schweizer, K.S.: Effective interactions and self-assembly of hybrid polymer grafted nanoparticles in a homopolymer matrix. Macromolecules 42, 84238434 (2009).CrossRefGoogle Scholar
63.Hall, L.M., Jayaraman, A., and Schweizer, K.S.: Molecular theories of polymer nanocomposites. Curr. Opin. Solid State Mater. Sci. 14, 3848 (2010).Google Scholar
64.Nair, N., Wentzel, N., and Jayaraman, A.: Effect of bidispersity in grafted chain length on grafted chain conformations and potential of mean force between polymer grafted nanoparticles in a homopolymer matrix. J. Chem. Phys. 134, 194906 (2011).CrossRefGoogle Scholar
65.Iacovella, C.R. and Glotzer, S.C.: Complex crystal structures formed by the self-assembly of ditethered nanospheres. Nano Lett. 9, 12061211 (2009).Google Scholar
66.Nguyen, T.D. and Glotzer, S.C.: Reconfigurable assemblies of shape-changing nanorods. ACS nano 4, 25852594 (2010).Google Scholar
67.Horsch, M.A., Zhang, Z., and Glotzer, S.C.: Self-assembly of end-tethered nanorods in a neat system and role of block fractions and aspect ratio. Soft Matter 6, 945954 (2010).Google Scholar
68.Iacovella, C.R., Horsch, M.A., and Glotzer, S.C.: Local ordering of polymer-tethered nanospheres and nanorods and the stabilization of the double gyroid phase. J. Chem. Phys. 129, 044902 (2008).Google Scholar
69.Horsch, M.A., Zhang, Z., and Glotzer, S.C.: Self-assembly of laterally-tethered nanorods. Nano Lett. 6, 24062413 (2006).Google Scholar
70.Horsch, M., Zhang, Z., and Glotzer, S.: Self-assembly of polymer-tethered nanorods. Phys. Rev. Lett. 95, 14 (2005).Google Scholar
71.Capone, B., Coluzza, I., LoVerso, F., Likos, C.N., and Blaak, R.: telechelic star polymers as self-assembling units from the molecular to the macroscopic scale. Phys. Rev. Lett. 109, 238301 (2012).Google Scholar
72.Urbas, A.M., Maldovan, M., DeRege, P., and Thomas, E.L.: Bicontinuous cubic block copolymer photonic crystals. Adv. Mater. 14, 18501853 (2002).CrossRefGoogle Scholar
73.Maldovan, M., Urbas, A., Yufa, N., Carter, W., and Thomas, E.: Photonic properties of bicontinuous cubic microphases. Phys. Rev. B 65, 15 (2002).Google Scholar
74.Maldovan, M., Ullal, C.K., Carter, W.C., and Thomas, E.L.: Exploring for 3D photonic bandgap structures in the 11 f.c.c. space groups. Nat. Mater. 2, 664667 (2003).Google Scholar
75.Maldovan, M. and Thomas, E.L.: Diamond-structured photonic crystals. Nat. Mater. 3, 593600 (2004).Google Scholar
76.Schroder-Turk, G.E., Varslot, T., de Campo, L., Kapfer, S.C., and Mickel, W.: A bicontinuous mesophase geometry with hexagonal symmetry. Langmuir: ACS J. Surfaces Colloids 27, 1047510483 (2011).Google Scholar
77.Matsushita, Y., Hayashida, K., Dotera, T., and Takano, A.: Kaleidoscopic morphologies from ABC star-shaped terpolymers. J. Phys.: Condens. Matter 23, 284111 (2011).Google Scholar
78.Kirkensgaard, J.J.K.: Kaleidoscopic tilings, networks and hierarchical structures in blends of 3-miktoarm star terpolymers. Interface Focus 2, 602607 (2012).Google Scholar
79.Kirkensgaard, J.J.K.: Striped networks and other hierarchical structures in A {m}B {m}C {n} (2 m+n)-miktoarm star terpolymer melts. Phys. Rev. E 85, 031802 (2012).Google Scholar
80.Evans, M.E., Robins, V., and Hyde, S.T.: Periodic entanglement II: weavings from hyperbolic line patterns. Acta Crystallogr. A A69, 262275 (2013).Google Scholar
81.Evans, M.E. and Hyde, S.T.: Periodic entanglement I: networks from hyperbolic reticulations. Acta Crystallogr. A 241261 (2013).Google Scholar
82.Schroder-Turk, G.E., de Campo, L., Evans, M.E., Saba, M., Kapfer, S.C., Varslot, T., Grosse-Brauckmann, K., Ramsden, S., and Hyde, S.T.: Polycontinuous geometries for inverse lipid phases with more than two aqueous network domains. Faraday Discuss. 161, 215247 (2013).Google Scholar
83.Kirkensgaard, J.J., Evans, M.E., de Campo, L., and Hyde, S.T.: Hierarchical self-assembly of a striped gyroid formed by threaded chiral mesoscale networks. Proc. Nat. Acad. Sci. U.S.A. 111, 12711276 (2014).Google Scholar
84.Ma, S., Hu, Y., and Wang, R.. Self-assembly of polymer tethered molecular nanoparticle shape amphiphiles in selective solvents. Macromolecules, 150424124622003 (2015).Google Scholar
85.Lai, Y.T., King, N.P., and Yeates, T.O.: Principles for designing ordered protein assemblies. Trends Cell Biol. 22, 653661 (2012).Google Scholar
86.Damasceno, P.F., Engel, M., and Glotzer, S.C.: predictive self-assembly of polyhedra into complex structures. Science 337, 453457 (2012).Google Scholar
87.Engel, M., Damasceno, P.F., Phillips, C.L., and Glotzer, S.C.: Computational self-assembly of a one-component icosahedral quasicrystal. Nat. Mater. 14, 18 (2014).Google Scholar
88.Batista, V.M.O. and Miller, M.A.: Crystallization of deformable spherical colloids. Phys. Rev. Lett. 105, 088305 (2010).Google Scholar
89.Nguyen, T.D., Jankowski, E., and Glotzer, S.C: Self-assembly and reconfigurability of shape-shifting particles. ACS Nano 5, 88928903 (2011).Google Scholar
90.Eshet, H., Bruneval, F., and Parrinello, M.: New Lennard-Jones metastable phase. J. Chem. Phys. 129, 026101 (2008).Google Scholar
91.Saba, M., Thiel, M., Turner, M., Hyde, S., Gu, M., Grosse-Brauckmann, K., Neshev, D., Mecke, K., and Schroder-Turk, G.: Circular dichroism in biological photonic crystals and cubic chiral nets. Phys. Rev. Lett. 106, 103902 (2011).Google Scholar
92.Turner, M.D., Saba, M., Zhang, Q., Cumming, B.P., Schroder-Turk, G.E., and Gu, M.: Miniature chiral beamsplitter based on gyroid photonic crystals. Nat. Photonics 7, 15 (2013).Google Scholar
93.Saba, M., Turner, M.D., Mecke, K., Gu, M., and Schroder-Turk, G.E.: Group theory of circular-polarization effects in chiral photonic crystals with four-fold rotation axes applied to the eight-fold intergrowth of gyroid nets. Phys. Rev. B: Condens. Matter Mater. Phys. 88, 116 (2013).Google Scholar
94.Liu, F., Hou, Y., and Gao, S.: Well-ordered nanohybrids and nanoporous materials from gyroid block copolymer templates. Chem. Soc. Rev. 44, 19742018 (2014).Google Scholar
95.Oganov, A.R., Lyakhov, A.O., and Valle, M.: How evolutionary crystal structure prediction works–and why. Acc. Chem. Res. 44, 227237 (2011).Google Scholar
96.Hansen, N. and Ostermeier, A.: Completely derandomized self-adaptation in evolution strategies. Evol. Comput. 9, 159195 (2001).Google Scholar
97.Miskin, M.Z. and Jaeger, H.M.: Evolving design rules for the inverse granular packing problem. Soft Matter 10, 3708 (2014).Google Scholar
98.https://www.whitehouse.gov/mgi [accessed April 12, 2015].Google Scholar
Figure 0

Figure 1. Building blocks synthesized from various techniques: (a) Functionalization (adopted from Ref. 20); (b) click reactions (Ref. 21); DNA conjugation with (c) multiple DNA strands[22] and (d) single DNA strand[23] per building block; (e) coordination with metal ions,[24] and (f) electrohydrodynamic co-jetting methods that are used to produce multicompartmentalized building blocks[19,25,26]. Panel (c) reprinted by permission from Macmillan Publishers Ltd.: Nature Nanotechnology[22], copyright 2013. Panel (d) from Ref. 23. Reprinted with permission from AAAS. Panel (e) adapted from Ref. 24 with permission of The Royal Society of Chemistry.

Figure 1

Figure 2. Static assemblies: (a) Nanotubular assemblies from coordination polymers upon addition of divalent metal ions (red circle),[24] (b) assembled structures from shape amphiphiles, including lamellar, gyroid, hexagonal cylinders, and BCC micelle phases, as reported in[21]. (c) Dodecagonal quasicrystal formed in a blend of polyisoprene–polystyrene–poly(2-vinylpyridine) star block copolymers/polysterene homopolymer blend.[53] (d) Ordered structures formed by gold NPs functionalized with complementary DNA strands (top): NaCl lattice (middle) and simple cubic lattice (bottom)[22]. Panel (a) adapted from Ref. 24 with permission of The Royal Society of Chemistry. Panel (c) reprinted with permission from Ref. 53. Copyright (2007) by the American Physical Society. Panel (d) from Ref. 23. Reprinted with permission from AAAS.

Figure 2

Figure 3. Reconfigurable assemblies: (a) Reversible transformation between flat sheet and tubule formed by rod–coil macromolecules upon heating and cooling[44], (b) BCC structure with lattice spacings tunable by adding set and unset DNA strands[54]; (c) transition from a disk into twisted ribbons upon quenching[55]; (d) transition from a simple cubic lattice to face-centered cubic lattice assembled by dodecanethiol ligated palladium nanocubes[56]. Panel (a) from E. Lee, J. K. Kim, and M. Lee. “Reversible scrolling of 2D sheets from self-assembly of laterally grafted amphiphilic rods.” Angew. Chem. Int. Ed. Engl., John Wiley and Sons Publishing Group. Panels (b) and (c) reprinted with permission from Macmillan Publishers Ltd: Nat. Nanotechnol.[54], copyright 2009, and Nature[55], copyright 2012. Panel (d) reprinted with permission from Ref. 56. Copyright (2011) by the American Physical Society.

Figure 3

Figure 4. Simulation predictions: (a) Double gyroid formed by TNP telechelics[17]; (b) tetragonally cylinder structure, and [6;6;6] columnar structure assembled by di-tethered nanospheres with different planar angles, θ, between two tethers[65]; (c) bilayer sheets and honeycomb grid formed by laterally tethered nanorods[66]; (d) dodecagonal quasicrystal formed by mono-tethered NPs[14]. Panel (a) reprinted with permission from Ref. 17. Copyright (2014) American Chemical Society.” Panel (b) reprinted with permission from Ref. 65. Copyright (2009) American Chemical Society. Panel (b) reprinted with permission from Ref. 66. Copyright (2010) American Chemical Society.

Figure 4

Figure 5. Anisotropy axis for relevant assembly dimensions in TNP systems. By tuning NP shape and interaction, researchers can tune the resultant structure and ordering in TNP systems. Additionally, reconfigurability of these building blocks will allow for tunable and responsive next-generation materials.