Hierarchical Assembly of Plasmonic Nanoparticles

5 downloads 1018 Views 3MB Size Report
E-mail : [email protected]. [b] Prof. ... E-mail : [email protected]. Chem. .... to design templates, such as polymethylmethacrylate (PMMA),.
DOI: 10.1002/chem.201500149

Minireview

& Plasmonic Nanoparticles

Hierarchical Assembly of Plasmonic Nanoparticles Cyrille Hamon*[a] and Luis M. Liz-Marzn*[a, b]

Chem. Eur. J. 2015, 21, 1 – 9

These are not the final page numbers! ÞÞ

1

 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

&

&

Minireview blies, with a particular attention on assemblies displaying hierarchical order. Throughout the manuscript we identify the main parameters and methods that can be tuned to achieve long range organization.

Abstract: In this Minireview, we summarize recent advances in the ordering of plasmonic nanoparticles over extended areas on solid substrates. The spotlight is thus focused on examples showing one-, two-, and three-dimensional assem-

1. Introduction

“static” self-assembly as in most of the following cited examples nanoparticles are fixed on a substrate. Thus, “dynamic” self-assembly processes occurring in solution will not be discussed and we direct the readers to recent reviews.[6] In comparison with colloidal systems, assemblies on surfaces can be easily functionalized and cleaned under aggressive conditions, which would not be possible in a colloidal suspension. Furthermore, we only consider approaches yielding assemblies that span over several micrometers (i.e., device scale) and eventually up to the millimeter scale. In the latter, particular attention is paid to hierarchical materials that show particular structures or architectures at different length scales. Solid superstructures made of nanoparticles showing different levels of organization originate from an interplay of forces that are relevant at different length scales.[7] At dimensions smaller than the size of the nanoparticles various nanoscale forces affect the final local packing and can be controlled by tuning the organic shell around the nanoparticles, as well as the surrounding medium, which has been described using theoretical models.[4b, 8] At dimensions larger than the nanoparticles, macroscale forces induced by dragging flows during drying processes affect long-range order. In the following, the spotlight is focused on techniques based on using templates that control liquid flows or direct chemical interactions. Processes involving gravitational, magnetic, or electric fields have also been reported but are beyond the focus of this review.

Plasmonic nanoparticles (NPs) made of gold and silver display exciting optical properties in the visible region due to localized surface plasmon resonances that can be precisely tuned through variations in particle shape and size.[1] Taking into account developments made over the past decade, an extensive library of particle morphologies can be produced, even at the gram scale. Moreover, when such nanoparticles are organized in ensembles, collective properties are obtained that differ from those of individual particles and the resulting optical properties can be further tuned and even amplified.[2] In particular, plasmon coupling in small gaps (1–10 nm) between plasmonic nanoparticles results in intense electric fields (i.e., hotspots) that can be exploited for many purposes, such as sensing, biomaterials, metamaterials design, switching devices, and so forth. However, the implementation of metallic NPs in devices requires their precise placement over large areas and this is the focus of this Minireview, rather than describing the properties and applications of such plasmonic assemblies. Therefore, the challenge relies in assembling individual particles that have an initial random orientation in solution. This field, often referred to as nanotectonics,[3] has rapidly evolved with the development of colloidal synthesis and characterization techniques during the past decade and numerous strategies are currently emerging. For example, interfacial assemblies or electrostatic interactions have shown promise in large area substrate coverage and one-, two-, and three-dimensional assemblies can be achieved.[4] Even if those benchtop techniques permit to scale up the organization of nanoparticles up to several micrometers, there is still a lack of sufficient reproducibility and control over the position of the particles on the substrate. Therefore, benchtop techniques cannot compete at present with lithography or physical vapor deposition techniques. However, the development of more efficient benchtop methods would permit to decrease the overall cost of fabrication and to design even three-dimensional materials in few steps, which would be difficult to manufacture by top-down techniques. Research in this area will have an enormous potential impact and we therefore tried to summarize in this Minireview the recent advances in the field. Since the field of nanoparticles self-assembly is very broad,[5] the discussion is restricted to

1.1. Identification of template features to achieve longrange order We define the term template in this work as a surface of any dimensionality or a matrix containing nano/macro-structured features. This definition encompasses both naturally occurring and synthetically produced templates, by either top-down or bottom-up approaches. In order to overview all self-assembly methods mediated by templates, Yin and co-workers suggested that these strategies could be classified according to the type of template (naturally occurring or synthetically made, etc.) or on the methods employed to fabricate them.[9] We have chosen to restrict the discussion to approaches that meet certain quality criteria, namely reusability, low cost, precise definition of pattern units, high throughput, simplicity of the assembly procedure, and possibility to scale up to macroscale assemblies. Thus, most of the examples highlighted in this Minireview deal with lithographically made templates. Moreover, we chose to cover the field by classifying assemblies according to their dimensions as this is an important parameter to control the plasmonic properties. Indeed, a perturbation of the dielectric constant of the surrounding medium can induce an

[a] Dr. C. Hamon, Prof. Dr. L. M. Liz-Marzn Bionanoplasmonics Laboratory, CIC biomaGUNE Paseo de Miramn 182, 20009 Donostia—San Sebastin (Spain) E-mail: [email protected] [b] Prof. Dr. L. M. Liz-Marzn Ikerbasque, Basque Foundation for Science, 48013 Bilbao (Spain) E-mail: [email protected]

&

&

Chem. Eur. J. 2015, 21, 1 – 9

www.chemeurj.org

2

 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!

Minireview optical shift of the localized surface plasmon resonance (LSPR) at the single-particle level, which has been used to build sensing devices. On the other hand, in 2D and 3D assemblies, due to plasmon coupling, the LSPR bands are shifted and broadened because new plasmon modes originate through optical coupling, which broadens the range of applications, from light harvesting to optical waveguide design, among others. Interestingly, 2D assemblies limited to one or few nanoparticles are more sensitive toward surface changes than their 3D counterparts, which have additional topographical features and collective optical effects. After a short description of template-free techniques, we describe methods to achieve 1D and 2D assemblies, and then recent advances in shaping 3D plasmonic materials are presented.

rendering each sample unique and thereby showing promise in anti-counterfeiting device fabrication.[12] However, for numerous other applications it is necessary to improve the reproducibility of the structures, therefore the coffee ring effect should be hindered by increasing the viscosity of the suspension or manipulating the substrate temperature and wettability.[11] Other strategies rely on controlling the deposition of the solution. For example, small drops of a plasmonic ink have been printed on desired locations of a substrate to create single or connected rings (Figure 1 B), thereby manufacturing transparent conductive surfaces by simple inkjet printing.[13] Controlled deposition of plasmonic inks of silver and gold nanoparticles has been achieved by means of a socalled pen on paper approach (Figure 1 C).[14] In this case, the coffee ring effect was suppressed by the use of an absorbent paper as substrate and nanoparticles were rapidly adsorbed on the cellulose fibers by capillary forces. This approach has been used for the fabrication of ultrasensitive surface enhanced Raman scattering (SERS) sensors. Interestingly, the edge of the drop can be mechanically controlled by pinning the colloidal suspension between a translating substrate and a fixed blade leading to local concentration effects due to convective flows.[15–16] In this case the assembly process starts when the thickness of the solvent layer becomes equal to the particle diameter.[17] Recently, Farcau et al. showed that the orientation of plasmonic wires could be further tuned between parallel and perpendicular to the substrate–liquid–air contact line, by controlling substrate temperature (Figure 1 D).[15] They showed that resulting assemblies are conductive and could be implemented as a stress gauge sensor.[16a]

2. Template-Free Methods: Controlled Deposition of Colloidal Dispersions Many works report the assembly of nanoparticles at the pinned edge of a drying drop, resulting from the well-known coffee-ring effect and leading to a higher local concentration of nanoparticles.[10] The coffee-ring structure occurs from capillary flow during an evaporation process.[11] In a remarkable example, under certain conditions of particle coating and surfactant concentration that allow the system to reach an equilibrium between repulsive and attractive forces, two-dimensional close-packed arrays of gold nanotriangles were observed within the coffee ring (Figure 1 A).[10d] In general, it has been observed that supercrystal size can be roughly tuned by the colloid concentration. Our group reported the formation of Au@Ag nanorod supercrystals with a variety of packing densities and orientations at different regions of the coffee ring.[10a] Unfortunately, such assemblies are difficult to reproduce, thus

Cyrille Hamon obtained his Ph.D. from the University of Rennes 1 (France) under the supervision of Pascale Even-Hernandez and Valrie Marchi in 2013. He is currently a postdoctoral fellow in Luis Liz-Marzn laboratory. His current interest focuses on self-assembly of metallic nanoparticles and the optimization of supercrystalline architecture for sensing applications.

Luis M. Liz-Marzn has a Ph.D. from the University of Santiago de Compostela (Spain) and was a postdoc at Utrecht University (The Netherlands) and visiting professor at various universities and research centers. After holding a chair in Physical Chemistry at the University of Vigo (Spain) from 1995–2012, he is currently an Ikerbasque Research Professor and Scientific Director of CIC biomaGUNE in San Sebastian (Spain). His current interests include nanoparticle synthesis and assembly, nanoplasmonics, and nanoparticle-based sensing and diagnostic tools.

Figure 1. Examples of plasmonic supercrystalline structures obtained by controlling deposition on various substrates. A) Gold nanotriangle assemblies obtained by drop casting.[10d] Scale bar is 1 mm. B) Connected rings of silver nanoparticles obtained by inkjet printing.[13] C) Macroscale arrays of various plasmonic nanoparticles obtained by a pen on paper approach.[14] D) Wires of gold colloids obtained by convective self-assembly.[15] B) and D) were adapted with permission from references [13, 15], respectively. Copyright American Chemical Society. Chem. Eur. J. 2015, 21, 1 – 9

www.chemeurj.org

These are not the final page numbers! ÞÞ

3

 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

&

&

Minireview 3. One- and Two-Dimensional Assemblies Mediated by Patterned Substrates In this section we focus on methods used to create plasmonic structures from patterned substrates, rather than on the fabrication of the substrate itself. Among the various nanofabrication methods, soft lithography comprises a family of techniques that allow the replication of a predefined structure by using an inexpensive elastomeric material.[18] Polydimethylsiloxane (PDMS) is probably the most widely used synthetic material toward the manufacture of soft lithography templates. PDMS has a number of advantages such as low cost, flexibility, the ability to accurately reproduce tiny features with a high resolution, and simple transfer of assemblies onto another substrate of interest. Other polymeric materials are currently used to design templates, such as polymethylmethacrylate (PMMA), which present the advantage to be easily lifted off after the assembly is completed.[19] 3.1 Topographically and chemically patterned substrates Precise placement of colloids on surfaces can be performed by confining single particles or clusters within topographical features where convective and capillary forces are the main driving forces in the self-assembly process.[16c,d, 17a, 20] In general, the separation between the particles or clusters and the topographical features are of the same order of magnitude: if the features are smaller, no patterning is obtained and conversely bigger features lead to a random organization.[19a, 21] This effect has been exploited by Chen and co-workers using a substrate modified with PMMA trenches of different width, so that an iterative process leads to the selective assembly of gold nanoparticles having different hydrodynamic diameters (Figure 2 A),[22] the spacing between the particles in the 1D array being controlled by adjusting the ionic strength of the colloidal ink. Mirkin’s group demonstrated an alternative way to control interparticle distance using multi-segmented nanowires with one of the components acting as a spacer that can be lifted off after assembly in PDMS trenches.[23] In another work they assembled pairs of gold nanoprisms and concave nanocubes using a PDMS stamp and demonstrated that the depth of the trenches has to be commensurate or larger than the metallic nanoparticles so as to immobilize them, with a yield of 80 % (Figure 2 B).[24] Other strategies based on non-specific (i.e., wettability contrast or electrostatic) interactions are often used to immobilize nanoparticles on patterned surfaces, to increase the adhesion yield and to avoid deposition at non-desired places.[6b, 25] An example is the formation of gold nanorod supercrystals onto a patterned silicon substrate that had been selectively modified with a gold layer to increase the wettability affinity of the particles toward the topographical features.[26] In another example, defined plasmonic nanoparticle cluster arrays were fabricated using template-guided self-assembly, in which the substrate was patterned with ligands carrying opposite charges to those on the nanoparticles (Figure 2 C).[27] In this case, the pattern was engineered by depositing a PMMA layer on a lithographically etched flat substrate that acts as a mask &

&

Chem. Eur. J. 2015, 21, 1 – 9

www.chemeurj.org

Figure 2. Examples of 1D/2D plasmonic NP assemblies obtained using patterned substrates. A) Schematic description of 1D NP assemblies of different sizes with various 1D packing obtained on the same substrate by varying the width of PMMA trenches.[22] B) Selective deposition of gold nanoprisms in template features that fit nanoparticle geometry whereas other particle shapes do not interact.[24] Scale bar: 600 nm. C) Arrays of gold nanoparticles clusters obtained by electrostatic interaction at defined binding sites.[27] Scale bars: left 1 mm; right 200 nm. D) 1D arrays of gold nanorods selectively deposited on a pattern with chemical contrast.[28] A), B), C) and D) were adapted with permission from references [22, 24, 27, 28], respectively. Copyright 2015 American Chemical Society.

for the functionalization of the appropriate ligand. Alternatively, a chemical pattern can be produced via micro-contact printing of patches containing functional groups (electrostatic or wettability contrast) on a flat substrate, thereby allowing to directly create long-range nanoparticle assemblies. Similarly to topographically modified substrates, assembly occurs if the dimensions of the functional patches are commensurate with particle size. In this respect, Vaia and co-workers tuned the packing of gold nanorods by modulating electrostatic interactions between the particles and poly(2-vinylpyridine) chemical patches (Figure 2 D).[28]

4

 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!

Minireview 3.2. Towards three dimensions: combining patterned substrates and supramolecular approaches

terned substrates. Recent work demonstrated the precise integration of DNA origami scaffolds into macroscale surfaces, by using patterned substrates.[30] Cha and co-workers demonstrated macroscale arrangements of 5 nm gold nanoparticles immobilized on triangular DNA origami scaffolds, selectively deposited on the hydrophilic regions of a patterned substrate (Figure 3 A).[30c] In the direction toward three-dimensional assemblies, Mirkin and colleagues demonstrated an epitaxial overgrowth process using DNA-coated gold NPs to cover patterned substrates with a defined number of nanoparticle layers.[31] Overall, the ability of DNA architectures to precisely bind nanoparticles of various composition and morphology, eventually leading to 3D assemblies, is of high significance in the design of plasmonic devices. One of the major limitations of DNA toward directing the assembly of nanoparticles is the high production cost, which has so far hindered its implementation at the industrial level. In this respect, finding complementary and alternative approaches seems a good alternative. Block copolymers (BCPs) have been studied for decades as templates to direct particle organization and to build cost effective nanocomposites.[25c] Understanding the thermodynamics behind nanoparticle organization in the polymer blend is key toward controlling interparticle distances and the overall shape of 3D structures in BCP microdomains that can be extended into larger length scales using external fields or patterned substrates. In the latter, epitaxial assembly, graphoepitaxy, and other methods have been used to control BCP domains at the macroscale.[33] However, only few examples report the organization of metallic nanoparticles in BCPs using patterned substrates. Recently, Xu and co-workers reported the three-dimensional organization of gold nanoparticles included in BCP domains, resulting in large assemblies (up to 3 cm) using a saw-toothed faceted substrate.[32] Remarkably, the nanoparticles followed the alignment of the BCP and formed unidirectional chains (Figure 3 B). In this work, a maximum of ten stacked layers of NPs was obtained because confinement effects imposed by the substrate decrease as the thickness of the nanocomposite increases.

Although the approaches we have discussed so far deal with well-organized macroscale assemblies, they are often limited to one or two dimensions. In this subsection we selected two topics of current interest, namely DNA origami and block copolymer assembly, and show how these supramolecular approaches can be combined with patterned surfaces to build assemblies with new functionalities, as well as their potential toward 3D assembly. DNA constructions were introduced in the early 1990s by Seeman and later developed by Rothemund who coined the term DNA origami for a system based on the use of a DNA viral scaffold and its folding through Watson– Crick base pairing.[29] Since then, DNA has become a cornerstone in bottom-up approaches, as precise distances and interactions can be engineered with high precision and reasonable yields, even in solution. The limitation of such systems toward solid-device applications relies on their controlled immobilization on substrates, which can however be achieved using pat-

4. Three-dimensional Assemblies Mediated by Molds In contrast with the previous section, in which template features were commensurate with particle size, we focus now on the formation of supercrystals in three dimensions by using templates that have significantly higher dimensions than the building blocks. Macroscale assemblies with topographical features in three dimensions are of great interest in various fields such as biomaterials and tissue engineering, as the dimensions of the assemblies approach those of cells.[34] In addition, the optical properties of such architectures are characterized by a dramatic broadening of plasmon bands, which is interesting, for example, in light harvesting devices and in sensor design.[6a, 21, 35] It should be noted that the particle assemblies we deal with here were created by evaporation of the solvent from a colloidal dispersion beneath a mask that features a predesigned pattern. The benefits of such techniques, often re-

Figure 3. Two examples illustrating the combination of patterned substrates and supramolecular approaches using DNA origami and block copolymers. A) Schematic representation of assembling 2D NP arrays on triangular DNA structures arranged on a patterned surface and AFM image of the resulting assembly.[30c] Reprinted by permission from Macmillan Publishers Ltd: [Nat. Nanotech.] (reference [30c]), copyright (2010). B) AFM image of organized gold nanoparticles embedded in a BCP matrix deposited on a faceted substrate. Scale bar: 100 nm. The image in the right panel is a GIGAXS pattern demonstrating long range order, over 3 cm.[32] Chem. Eur. J. 2015, 21, 1 – 9

www.chemeurj.org

These are not the final page numbers! ÞÞ

5

 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

&

&

Minireview ferred to as evaporative lithography, are the low cost, high throughput, and no need of substrate modification for reusability of the mask. Notable differences in evaporative lithography techniques can be found depending on the type of masks used in the self-assembly process and thus we review below some of the, in our opinion, most significant advances. In a pioneering work, Lewis and co-workers used a mask consisting of holes organized in hexagonal arrays to control the accumulation of silica microspheres at the uncovered regions, at which evaporation was faster than at the covered parts.[41] The evaporation rate could also be enhanced at the uncovered regions by illumination with an IR lamp.[42] Recently, a mixture of gold nanoparticles and polymer particles were dried by this method, leading to a periodically organized plasmonic composite.[43] More recently, a pyramidal mold was used

to direct the organization of gold nanospheres into macroscale arrays of plasmonic pyramids.[44] In this work, the mold was first filled with the metallic NPs, and then assemblies were transferred onto a flat substrate by stamping. Vakarelski et al. demonstrated formation of gold wires in a macroscale connected network using a colloidal mask of latex beads.[36–37, 45] Importantly, the addition of surfactant was important to stabilize the liquid film and allow the formation of the plasmonic network upon drying (Figure 4 A, B). In order to improve the reproducibility and versatility, the authors used a photoresist as a mask to direct the drying of the gold colloid (Figure 4 C).[38] An alternative method is screen printing, so that the resulting structures could be tuned via the screen mesh.[46] Similarly, Marchi-Artzner and co-workers used a PDMS mold to control meniscus morphologies at the gaps between cylindrical posts, which were engineered by soft lithography allowing great versatility on features morphology and spacing.[47] The authors demonstrated a hierarchical organization of gold nanorods into smectic phases. More recently, our group proposed the use of a PDMS mold composed of an array of micron-sized cavities in which gold nanorod supercrystals could be confined in a precise manner (Figure 4 E).[40] The resulting gold supercrystals displayed a morphology that was reminiscent of the meniscus formed by the solvent in the cavities and could be tuned through the initial particle concentration. Microfluidics is another appealing technique in this field as it allows control over the addition of reactants, easy monitoring of the self-assembly process and versatility over chip design. To overcome the use of expensive syringe pumps and tubing, solvent flow can be induced by capillarity (MIMIC)[48] or by microevaporation,[49] rendering these approaches attractive in the field of particle crystallization. Correa-Duarte and colleagues combined both approaches by introducing a thin PDMS layer at the end of the channel to induce a local fast evaporation at which the assembly of gold and silver nanoparticles occurred and adopted the overall shape of the channel (Figure 4 D).[39]

5. Summary and Outlook We have reviewed the recent literature regarding methods to obtain hierarchical assemblies of metallic nanoparticles according to their dimensionality, as it results in different plasmonic properties and can thus be applied in various fields. Most of the highlighted work has been realized during the past five years and there is no doubt that it will continue expanding in the near future, in particular toward methods that can be combined with supramolecular approaches, as only few examples thereof have been reported. Overall, the techniques presented in this Minireview are relevant to bring benchtop preformed plasmonic nanoparticles into solid devices for real-world applications. As a perspective, we can consider that many pioneering works demonstrated precise placement of colloids within templates using micron-sized non-plasmonic particles as building blocks and some of those methods have been implemented with metal nanoparticles, giving rise to plasmonic structures of

Figure 4. Examples of 3D macroscale assemblies using evaporative lithography. A) Gold nanowires made of nanoparticle chains.[36] B) Images before and after drying of a colloidal film containing a fluorophore, depicting meniscus formation.[37] C) Gold wires obtained by evaporative lithography using a photoresist as a mold.[38] D) 3D nanoparticle supercrystals obtained using microevaporation in microfluidic channels.[39] E) Gold nanorod 3D supercrystals obtained using a PDMS mold displaying arrays of micro-cavities.[40] A) is adapted with permission from reference [36]. Copyright American Physical Society. B),D) and E) were adapted with permission from references [37, 39, 40], respectively. Copyright American Chemical Society.

&

&

Chem. Eur. J. 2015, 21, 1 – 9

www.chemeurj.org

6

 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!

Minireview interest to a broad range of applications. However, there is still plenty of room to translate other colloid assembly methods into plasmonic nanoparticle systems, which are expected to lead to important developments. For instance, semiconductor nanoparticles have been assembled using HOPG templates among others,[50] finding differences in the colloidal properties as compared to plasmonic nanoparticles related to their initial solubility in aqueous or organic solvents, respectively. In this particular case, an upstream work to modify the coating ligands on nanoparticles surfaces could be sufficient to translate self-assembly methods from one type of particles to another. Thus, a closer look at findings in other fields should result in the formation of the next generation of plasmonic devices. Conversely, the techniques reviewed in this Minireview can be generalized to other stable colloidal systems such as proteins, DNA, and so forth. We thus envision interesting developments in hybrid devices based on the combination of inorganic and organic colloidal systems using the self-assembly techniques described in this manuscript.

[11] [12] [13] [14] [15] [16]

[17]

[18]

Acknowledgements [19]

Funding from the European Research Council (ERC Advanced Grant #267867 Plasmaquo) is gratefully acknowledged.

[20]

Keywords: hierarchical assemblies · nanoparticles · patterned substrates · soft lithography · supramolecular chemistry

[21] [22]

[1] a) S. Link, M. A. El-Sayed, Annu. Rev. Phys. Chem. 2003, 54, 331 – 366; b) C. Burda, X. Chen, R. Narayanan, M. A. El-Sayed, Chem. Rev. 2005, 105, 1025 – 1102; c) X. Huang, S. Neretina, M. A. El-Sayed, Adv. Mater. 2009, 21, 4880 – 4910; d) R. Alvarez-Puebla, L. M. Liz-Marzn, F. J. Garca de Abajo, J. Phys. Chem. Lett. 2010, 1, 2428 – 2434. [2] a) A. Guerrero-Martnez, M. Grzelczak, L. M. Liz-Marzn, ACS Nano 2012, 6, 3655 – 3662; b) P. K. Jain, M. A. El-Sayed, Chem. Phys. Lett. 2010, 487, 153 – 164. [3] S. A. Davis, M. Breulmann, K. H. Rhodes, B. Zhang, S. Mann, Chem. Mater. 2001, 13, 3218 – 3226. [4] a) L. Scarabelli, M. Coronado-Puchau, J. J. Giner-Casares, J. Langer, L. M. Liz-Marzn, ACS Nano 2014, 8, 5833 – 5842; b) D. A. Walker, B. Kowalczyk, M. O. de La Cruz, B. A. Grzybowski, Nanoscale 2011, 3, 1316 – 1344. [5] G. M. Whitesides, B. Grzybowski, Science 2002, 295, 2418 – 2421. [6] a) A. Klinkova, R. M. Choueiri, E. Kumacheva, Chem. Soc. Rev. 2014, 43, 3976 – 3991; b) L. Xu, W. Ma, L. Wang, C. Xu, H. Kuang, N. A. Kotov, Chem. Soc. Rev. 2013, 42, 3114 – 3126. [7] M. Grzelczak, J. Vermant, E. M. Furst, L. M. Liz-Marzn, ACS Nano 2010, 4, 3591 – 3605. [8] a) E. Rabani, D. R. Reichman, P. L. Geissler, L. E. Brus, Nature 2003, 426, 271 – 274; b) O. Kletenik-Edelman, E. Ploshnik, A. Salant, R. Shenhar, U. Banin, E. Rabani, J. Phys. Chem. C 2008, 112, 4498 – 4506; c) L. Onsager, Ann. N. Y. Acad. Sci. 1949, 51, 627 – 659; d) S.-Y. Zhang, M. D. Regulacio, M.-Y. Han, Chem. Soc. Rev. 2014, 43, 2301 – 2323. [9] Y. Liu, J. Goebl, Y. Yin, Chem. Soc. Rev. 2013, 42, 2610 – 2653. [10] a) S. Gmez-GraÇa, J. Prez-Juste, R. A. Alvarez-Puebla, A. Guerrero-Martnez, L. M. Liz-Marzn, Adv.Opt. Mater. 2013, 1, 477 – 481; b) R. A. Alvarez-Puebla, A. Agarwal, P. Manna, B. P. Khanal, P. Aldeanueva-Potel, E. Carb-Argibay, N. Pazos-Prez, L. Vigderman, E. R. Zubarev, N. A. Kotov, L. M. Liz-Marzn, Proc. Natl. Acad. Sci. USA 2011, 108, 8157 – 8161; c) A. Guerrero-Martnez, J. Prez-Juste, E. Carb-Argibay, G. Tardajos, L. M. Liz-Marzn, Angew. Chem. Int. Ed. 2009, 48, 9484 – 9488; Angew. Chem. 2009, 121, 9648 – 9652; d) D. A. Walker, K. P. Browne, B. Kowalczyk, B. A. Grzybowski, Angew. Chem. Int. Ed. 2010, 49, 6760 – 6763; Angew. Chem. 2010, 122, 6912 – 6915; e) B. Peng, G. Li, D. Li, S. Dodson, Q. Zhang, J. Chem. Eur. J. 2015, 21, 1 – 9

www.chemeurj.org

These are not the final page numbers! ÞÞ

[23] [24] [25]

[26] [27] [28] [29]

[30]

[31] [32] [33]

7

Zhang, Y. H. Lee, H. V. Demir, X. Y. Ling, Q. Xiong, ACS Nano 2013, 7, 5993 – 6000; f) W. D. Wang, Y. G. Yin, Z. Q. Tan, J. F. Liu, Nanoscale 2014, 6, 9588 – 9593; g) J. W. Xu, J. J. Du, C. Y. Jing, Y. L. Zhang, J. L. Cui, ACS Appl. Mater. Interfaces 2014, 6, 6891 – 6897. H. Yildirim Erbil, Adv.Colloid Interface Sci.2015,in press. C. Schopf, A. Martin, M. Burke, D. Jones, A. Pescaglini, A. O’Riordan, A. J. Quinn, D. Iacopino, J. Mater. Chem. C 2014, 2, 3536 – 3541. M. Layani, M. Gruchko, O. Milo, I. Balberg, D. Azulay, S. Magdassi, ACS Nano 2009, 3, 3537 – 3542. L. Polavarapu, A. L. Porta, S. M. Novikov, M. Coronado-Puchau, L. M. LizMarzn, Small 2014, 10, 3065 – 3071. C. Farcau, H. Moreira, B. t. Viallet, J. r. m. Grisolia, L. Ressier, ACS Nano 2010, 4, 7275 – 7282. a) C. Farcau, N. M. Sangeetha, H. Moreira, B. Viallet, J. Grisolia, D. Ciuculescu-Pradines, L. Ressier, ACS Nano 2011, 5, 7137 – 7143; b) M. Kahraman, M. M. Yazıcı, F. Sahin, M. Culha, Langmuir 2008, 24, 894 – 901; c) C. Kuemin, L. Nowack, L. Bozano, N. D. Spencer, H. Wolf, Adv. Funct. Mater. 2012, 22, 702 – 708; d) C. Kuemin, R. Stutz, N. D. Spencer, H. Wolf, Langmuir 2011, 27, 6305 – 6310; e) B. G. Prevo, J. C. Fuller, O. D. Velev, Chem. Mater. 2005, 17, 28 – 35. a) L. Malaquin, T. Kraus, H. Schmid, E. Delamarche, H. Wolf, Langmuir 2007, 23, 11513 – 11521; b) T. Kraus, L. Malaquin, H. Schmid, W. Riess, N. D. Spencer, H. Wolf, Nat. Nanotechnol. 2007, 2, 570 – 576. a) M. E. Stewart, C. R. Anderton, L. B. Thompson, J. Maria, S. K. Gray, J. A. Rogers, R. G. Nuzzo, Chem. Rev. 2008, 108, 494 – 521; b) Y. Xia, D. Qin, G. M. Whitesides, Nat. Protoc. 2010, 5, 491 – 502; c) B. D. Gates, Q. Xu, M. Stewart, D. Ryan, C. G. Willson, G. M. Whitesides, Chem. Rev. 2005, 105, 1171 – 1196. a) L. Jiang, X. Chen, N. Lu, L. Chi, Acc. Chem. Res. 2014, 47, 3009 – 3017; b) Y. Yin, Y. Lu, B. Gates, Y. Xia, J. Am. Chem. Soc. 2001, 123, 8718 – 8729. a) F. Holzner, C. Kuemin, P. Paul, J. L. Hedrick, H. Wolf, N. D. Spencer, U. Duerig, A. W. Knoll, Nano Lett. 2011, 11, 3957 – 3962; b) N. Pazos-Prez, W. Ni, A. Schweikart, R. A. Alvarez-Puebla, A. Fery, L. M. Liz-Marzan, Chem. Sci. 2010, 1, 174 – 178. E. Kumacheva, R. K. Golding, M. Allard, E. H. Sargent, Adv. Mater. 2002, 14, 221 – 224. L. Jiang, Y. Sun, C. Nowak, A. Kibrom, C. Zou, J. Ma, H. Fuchs, S. Li, L. Chi, X. Chen, ACS Nano 2011, 5, 8288 – 8294. X. Zhou, Y. Zhou, J. C. Ku, C. Zhang, C. A. Mirkin, ACS Nano 2014, 8, 1511 – 1516. Y. Zhou, X. Zhou, D. J. Park, K. Torabi, K. A. Brown, M. R. Jones, C. Zhang, G. C. Schatz, C. A. Mirkin, Nano Lett. 2014, 14, 2157 – 2161. a) J. Aizenberg, P. V. Braun, P. Wiltzius, Phys. Rev. Lett. 2000, 84, 2997; b) M. R. Jones, K. D. Osberg, R. J. Macfarlane, M. R. Langille, C. A. Mirkin, Chem. Rev. 2011, 111, 3736 – 3827; c) J. Kao, K. Thorkelsson, P. Bai, B. J. Rancatore, T. Xu, Chem. Soc. Rev. 2013, 42, 2654 – 2678. T. Thai, Y. Zheng, S. H. Ng, S. Mudie, M. Altissimo, U. Bach, Angew. Chem. Int. Ed. 2012, 51, 8732 – 8735; Angew. Chem. 2012, 124, 8862 – 8865. B. Yan, A. Thubagere, W. R. Premasiri, L. D. Ziegler, L. Dal Negro, B. M. Reinhard, ACS Nano 2009, 3, 1190 – 1202. D. Nepal, M. S. Onses, K. Park, M. Jespersen, C. J. Thode, P. F. Nealey, R. A. Vaia, ACS Nano 2012, 6, 5693 – 5701. a) N. C. Seeman, Nature 2003, 421, 427 – 431; b) P. W. K. Rothemund, Nature 2006, 440, 297 – 302; c) S. J. Tan, M. J. Campolongo, D. Luo, W. Cheng, Nat. Nanotechnol. 2011, 6, 268 – 276. a) R. J. Kershner, L. D. Bozano, C. M. Micheel, A. M. Hung, A. R. Fornof, J. N. Cha, C. T. Rettner, M. Bersani, J. Frommer, P. W. K. Rothemund, G. M. Wallraff, Nat. Nanotechnol. 2009, 4, 557 – 561; b) A. Gopinath, P. W. K. Rothemund, ACS Nano 2014, 8, 12030 – 12040; c) A. M. Hung, C. M. Micheel, L. D. Bozano, L. W. Osterbur, G. M. Wallraff, J. N. Cha, Nat. Nanotechnol. 2010, 5, 121 – 126. S. L. Hellstrom, Y. Kim, J. S. Fakonas, A. J. Senesi, R. J. Macfarlane, C. A. Mirkin, H. A. Atwater, Nano Lett. 2013, 13, 6084 – 6090. J. Kao, S.-J. Jeong, Z. Jiang, D. H. Lee, K. Aissou, C. A. Ross, T. P. Russell, T. Xu, Adv. Mater. 2014, 26, 2777 – 2781. a) S. Ouk Kim, H. H. Solak, M. P. Stoykovich, N. J. Ferrier, J. J. de Pablo, P. F. Nealey, Nature 2003, 424, 411 – 414; b) I. Bita, J. K. W. Yang, Y. S. Jung, C. A. Ross, E. L. Thomas, K. K. Berggren, Science 2008, 321, 939 – 943; c) A. Tavakkoli, K. W. Gotrik, A. F. Hannon, A. Alexander-Katz, C. A. Ross, K. K. Berggren, Science 2012, 336, 1294 – 1298; d) M. S. Onses, A. Ramrez-Hernndez, S.-M. Hur, E. Sutanto, L. Williamson, A. G. Alleyne,  2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

&

&

Minireview

[34]

[35] [36] [37] [38] [39]

[40] [41] [42] [43]

&

&

P. F. Nealey, J. J. de Pablo, J. A. Rogers, ACS Nano 2014, 8, 6606 – 6613; e) D. Sundrani, S. B. Darling, S. J. Sibener, Nano Lett. 2004, 4, 273 – 276; f) S. W. Hong, J. Huh, X. Gu, D. H. Lee, W. H. Jo, S. Park, T. Xu, T. P. Russell, Proc. Natl. Acad. Sci. USA 2012, 109, 1402 – 1406. a) M. Zhu, G. Baffou, N. Meyerbrçker, J. Polleux, ACS Nano 2012, 6, 7227 – 7233; b) Y. Engel, J. D. Schiffman, J. M. Goddard, V. M. Rotello, Mater. Today 2012, 15, 478 – 485. Z. H. Nie, A. Petukhova, E. Kumacheva, Nat. Nanotechnol. 2010, 5, 15 – 25. I. U. Vakarelski, D. Y. C. Chan, T. Nonoguchi, H. Shinto, K. Higashitani, Phys. Rev. Lett. 2009, 102, 058303. I. U. Vakarelski, J. O. Marston, S. T. Thoroddsen, Langmuir 2013, 29, 4966 – 4973. X. Tang, S. J. O’Shea, I. U. Vakarelski, Adv.Mater. 2010, 22, 5150 – 5153. J. Angly, A. Iazzolino, J.-B. Salmon, J. Leng, S. P. Chandran, V. Ponsinet, A. Dsert, A. Le Beulze, S. Mornet, M. Trguer-Delapierre, M. A. CorreaDuarte, ACS Nano 2013, 7, 6465 – 6477. C. Hamon, S. Novikov, L. Scarabelli, L. Basabe-Desmonts, L. M. LizMarzn, ACS Nano 2014, 8, 10694 – 10703. a) D. J. Harris, H. Hu, J. C. Conrad, J. A. Lewis, Phys. Rev. Lett. 2007, 98, 148301; b) D. J. Harris, J. A. Lewis, Langmuir 2008, 24, 3681 – 3685. A. Georgiadis, A. F. Routh, M. W. Murray, J. L. Keddie, Soft Matter 2011, 7, 11098 – 11102. A. Utgenannt, J. L. Keddie, O. L. Muskens, A. G. Kanaras, Chem. Commun. 2013, 49, 4253 – 4255.

Chem. Eur. J. 2015, 21, 1 – 9

www.chemeurj.org

[44] M. Alba, N. Pazos-Perez, B. Vaz, P. Formentin, M. Tebbe, M. A. CorreaDuarte, P. Granero, J. Ferr-Borrull, R. Alvarez, J. Pallares, A. Fery, A. R. de Lera, L. F. Marsal, R. A. Alvarez-Puebla, Angew. Chem. Int. Ed. 2013, 52, 6459 – 6463; Angew. Chem. 2013, 125, 6587 – 6591. [45] I. U. Vakarelski, J. W. Kwek, X. Tang, S. J. O’Shea, D. Y. C. Chan, Langmuir 2009, 25, 13311 – 13314. [46] K. Higashitani, C. E. McNamee, M. Nakayama, Langmuir 2011, 27, 2080 – 2083. [47] C. Hamon, M. Postic, E. Mazari, T. Bizien, C. Dupuis, P. Even-Hernandez, A. Jimenez, L. Courbin, C. Gosse, F. Artzner, V. Marchi-Artzner, ACS Nano 2012, 6, 4137 – 4146. [48] a) Y. Xia, E. Kim, G. M. Whitesides, Chem. Mater. 1996, 8, 1558 – 1567; b) E. Kim, Y. Xia, G. M. Whitesides, Adv. Mater. 1996, 8, 245 – 247; c) A. Blmel, A. Klug, S. Eder, U. Scherf, E. Moderegger, E. J. W. List, Org. Electron. 2007, 8, 389 – 395. [49] A. Merlin, J.-B. Salmon, J. Leng, Soft Matter 2012, 8, 3526 – 3537. [50] a) S. Ahmed, K. M. Ryan, Nano Lett. 2007, 7, 2480 – 2485; b) T. Bizien, P. Even-Hernandez, M. Postic, E. Mazari, S. Chevance, A. Bondon, C. Hamon, D. Troadec, L. Largeau, C. Dupuis, C. Gosse, F. Artzner, V. Marchi, Small 2014, 10, 3707 – 3716; c) E. Ploshnik, A. Salant, U. Banin, R. Shenhar, Adv. Mater. 2010, 22, 2774 – 2779; d) J. D. Robert Schreiber, E. M. Roller, T. Zhang, V. J. Schller, P. C. Nickels, J. Feldmann, T. Liedl, Nat. Nanotechnol. 2013, 9, 74 – 78. Published online on && &&, 0000

8

 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!

Minireview

MINIREVIEW & Plasmonic Nanoparticles

Held in suspense: The figure illustrates the assembly of metallic nanoparticles upon drying a colloidal suspension. In this Minireview, the main parameters and methods that can be tuned to achieve such organized networks are defined.

C. Hamon,* L. M. Liz-Marzn* && – && Hierarchical Assembly of Plasmonic Nanoparticles

Plasmonic Building Blocks In their Minireview on page && ff. C. Hamon and L. M. LizMarzn take a look at the recent literature regarding methods to obtain hierarchical assemblies of metallic nanoparticles according to their dimensionality, as it results in different plasmonic properties and can thus be applied in various fields. Most of the highlighted work has been realized during the past five years and there is no doubt that it will continue expanding in the near future. Overall, the techniques presented in this Minireview are relevant to bring benchtop preformed plasmonic nanoparticles into solid devices for real-world applications.

Chem. Eur. J. 2015, 21, 1 – 9

www.chemeurj.org

These are not the final page numbers! ÞÞ

9

 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

&

&