Assembly of Gold Nanoparticles into Chiral Superstructures Driven by

Jul 22, 2019 - Read OnlinePDF (2 MB) ... 3D structures of chiral Au NPs by electron tomography (PPTX) ... Showing 1/2: ja9b00700_si_001.pdf. S1...
1 downloads 0 Views 2MB Size
Communication pubs.acs.org/JACS

Cite This: J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Assembly of Gold Nanoparticles into Chiral Superstructures Driven by Circularly Polarized Light Ji-Young Kim,†,‡,▲ Jihyeon Yeom,†,§,▲ Gongpu Zhao,∥,⊥ Heather Calcaterra,†,# Jiyoun Munn,¶ Peijun Zhang,∥,○,△ and Nicholas Kotov*,†,‡,§,#,□ †

Biointerfaces Institute, ‡Department of Materials Science, §Macromolecular Science and Engineering, #Department of Chemical Engineering, University of Michigan, Ann Arbor, Michigan 48109, United States



Department of Structural Biology, University of Pittsburgh, Pittsburgh, Pennsylvania 15261, United States



Downloaded via BUFFALO STATE on July 22, 2019 at 23:07:56 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

David Van Andel Advanced Cryo-Electron Microscopy Suite, Van Andel Research Institute, Grand Rapids, Michigan 49503, United States



COMSOL, Inc., Burlington, Massachusetts 01803, United States



Division of Structural Biology, Wellcome Trust Centre for Human Genetics, University of Oxford, Oxford, OX3 7BN, U.K.



Electron Bio-Imaging Centre, Diamond Light Source, Harwell Science and Innovation Campus, Didcot OX11 0DE, U.K.



Michigan Institute for Translational Nanotechnology, Ypsilanti, Michigan 48198, United States S Supporting Information *

between molecules.6−9 Chirality transfer from photons to matter10,11 has both fundamental and practical importance, as a process that could be, for example, involved in emergence of biomolecular homochirality,12−17 employed as a preparative method for chiral compounds,18,19 and lead to new photonic devices.20 Some enantiomeric excess of small molecules,21 helical polymers,22 supramolecular compounds,10,23 and liquid crystals11,24,25 after circularly polarized light (CPL) illumination was observed. However, the enantiomeric excess for the majority of organic and inorganic reactions carried out using CPL is typically small, namely 0.1−2%.26 Strong chiroptical activity of inorganic nanostructures associated with inorganic matter4,27−30 offers the possibility of increasing the chiral bias of photoinduced processes. The enantioselective assembly of semiconductor nanoparticles (NPs) by chiral photons was observed for CdTe/CdS nanocolloids and led to the formation of right- and left-handed twisted nanoribbons, respectively.16 The enantiomeric excess for this photoinduced reaction exceeded 30%.16,31 Similar processes can potentially be observed for metallic NPs also. However, such processes are difficult to study at the moment, first, because easily identifiable chiral shapes, such as helices or tetrahedrons familiar from organic chemistry, may not necessarily be dominant for self-assembled nanostructures. In fact, some nanoscale assemblies may be geometrically complex and their ensembles appear achiral while they are not. Second, circular dichroism (CD) peaks for inorganic nanostructures are polysemous due to multiple optical processes contributing to their optical activity.4,28,32,33 Their interpretation must include, therefore, careful analysis of chirality at nanoscale dimensions. The latter necessitates three-dimensional (3D) reconstructions of nanostructures using electron tomography. Although facilitated by their high e-beam contrast, acquisition and analysis of nanoscale tomographs is still far from being routine. The

ABSTRACT: Photon-to-matter chirality transfer offers both simplicity and universality to chiral synthesis, but its efficiency is typically low for organic compounds. Besides the fundamental importance of this process relevant for understanding the origin of homochirality on Earth, new pathways for imposing chiral bias during chemical process are essential for a variety of technologies from medicine to informatics. The strong optical activity of inorganic nanoparticles (NPs) affords photosynthetic routes to chiral superstructures using circularly polarized photons. Although plasmonic NPs are promising candidates for such synthetic routes due to the strong rotatory power of highly delocalized plasmonic states (Ma et al. Chem. Rev. 2017, 117 (12), 8041), realization of light-driven synthesis of chiral nanostructures has been more challenging for plasmonic NPs than for the semiconductor due to the short lifetime of the plasmonic states. Here we show that illumination of gold salt solutions with circularly polarized light induces the formation of NPs and their subsequent assembly into chiral nanostructures 10−15 nm in diameter. Despite their seemingly irregular shape, the resulting nanocolloids showed circular dichroism (CD) spectra with opposite polarity after exposure to photons with left and right circular polarization. The sign and spectral position of the CD peaks of illuminated dispersions matched those calculated for nanostructures with complex geometry identified from electron tomography images. Quantification of the complex shapes of NP assemblies using chirality measures revealed a direct correlation with the experimental spectra. The light-driven assembly of chiral nanostructures originates from the asymmetric displacement of NPs in dynamic assemblies by plasmonic fields followed by particle-toparticle attachment. The ability of gold NPs to “lock” the chirality of the incident photons in assembled nanostructures can be used to create a variety of chiral nanomaterials with plasmonic resonances.

C

hiral compounds have an overarching importance for physics, chemistry, biology, and medicine.2−5 Despite the large variety of chemical reactions selective for specific enantiomers, the chemical path to products with mirror asymmetry is based on matter-to-matter chiral bias typically via manipulation of covalent, coordination, or hydrogen bonds © XXXX American Chemical Society

Received: January 20, 2019

A

DOI: 10.1021/jacs.9b00700 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Communication

Journal of the American Chemical Society

Au(III) by citrate to Au(0) clusters photoactive at 543 nm and (2) the growth of Au(0) clusters to form about ∼ 3 nm Au NPs via photoinduced reduction(s). The resulting NPs form dynamic assemblies that acquire anisotropic shapes with chiral or other geometrical bias due to the plasmonic forces shaping the NP assemblies.37,38 To examine this mechanism closer, numerical solutions of the 3D vector Maxwell equations based on a finite-element frequency-domain approach (Comsol Multiphysics 5.4) were obtained. The forces acting on five neighboring 3 nm gold NPs by three different polarizations of light (LNP, RCP, and LCP; intensity of 1 MW/m2) were calculated (Figure 2 and Table 1). The integration of the Maxwell surface stress tensor showed that NP assemblies with achiral shapes experience out-of-plane twisting forces induced by CPL that drive them to become chiral. The direction of the forces and, thus, the resulting assemblies change their handedness depending on the polarization rotation of the incident photons. When the incident photons are linearly polarized, the out-of-plane torque twisting of the nanoassemblies is not observed (Figure 2C). The calculated results match well with the experimental TEM observations and spectroscopic data. The resulting assemblies with chiral disposition of NPs can be locked in shape by a NP−NP merger

metallic NPs are also known for the short lifetime of the excited states, which reduces the probability of photoinduced reactions. Regardless of these challenges, several factors make photonto-matter chirality transfer a promising research direction for metallic NPs. First, the plasmonic states are highly polarizable and, therefore, are strongly affected by circular polarization. Second, CPL illumination can potentially influence both the growth and assembly of metallic NPs. Hot−electron processes34 are likely to be both spin-35 and site-selective36 leading to asymmetric particle growth. Third, the assembly processes should also be affected by the coupling of plasmonic and hydrodynamic fields and thus will be CPL-dependent.37,38 Fourth, plasmonic NPs are known to be catalytic,39,40 which makes them suitable for subsequent chirality transfer into other chiral compounds including biomolecules.41 In the framework of the fundamental and technological importance of asymmetric photosynthesis,19,32,42−46 selfassembly of gold NPs, driven by circularly polarized photons, will diversify the spectrum and deepen our understanding of light-driven chirality transfer reactions. Here, we demonstrate that CPL induces the formation of chiral Au nanostructures. They have complex shapes that might seem arbitrary, but they display preferential handedness nevertheless. CD spectroscopy indicates that gold nanostructures illuminated with left- and right-polarized photons show positive and negative optical activity in the 400−700 nm plasmonic band, respectively. Computational models reproduced the spectral position and polarity of their chiroplasmonic bands using 3D geometry obtained by electron tomography. Quantitative analysis revealed correlations between their sign/amplitude and CD spectroscopic features and chirality measures. An aqueous solution (2 mL) of Au(III) chloride hydrate (HAuCl4) and citrate with a 2.5 mM concentration was illuminated by left- (LCP) or right-handed circularly polarized (RCP) lights at a wavelength of 543 nm. After 50 min of illumination, the formation of red dispersions were observed. The dispersions exhibited CD spectra with bands at 550 nm, characteristic of plasmonic resonances in Au NPs (Figure 1A and B). Nanostructures found in dispersions after CPL illumination using transmission electron microscopy (TEM) have complex 3D shapes based on several assembled NPs (Figure 1C and D). To the best of our knowledge, this is the first reported example of using CPL-induced photoreduction of metal ions in solution to synthesize chiral metal nanostructures. A previous study of photoinduced transformation of plasmonic NPs reported transformation of spherical Ag NPs into racemic prism-like shapes under illumination with unpolarized light.47,48 To elucidate the mechanism of chirality transfer from the incident light to the NPs in dispersion, we acquired TEM images and measured CD spectra at various illumination times (Figure S2). After 5 min of illumination of (HAuCl4) solutions, reduction of Au3+ to Au0 takes place49 and ∼ 2 nm NPs were found (Figure S2A). With continued illumination, the NPs grew to become 3−5 nm in diameter and coalesce into structures with complex shapes and dimensions of ∼ 10−15 nm. The intensity of the CD and UV−vis absorbance bands at 550 nm increased with the illumination time (Figure S2B and C). With a linearly polarized light (LNP) illuminating source, no CD bands appeared at any point during this reaction (Figure S3A and B) although the red dispersion was obtained by Au NP formation with elongated shapes (Figure S3C). Based on these data, the first two steps for formation of chiral Au nanostructures were identified as (1) partial reduction of

Figure 1. Optical properties and shape of Au NPs after 50 min CPL illumination. CD (A) and UV−vis absorbance (B) spectra of dispersions formed under LCP (LH NP, black) and RCP (RH NP, red), respectively. (C, D) High resolution TEM images of NPs obtained with LCP (C) and RCP (D) illumination, respectively.

known as oriented attachment and nonclassical crystallization processes,50−52 with or without an epitaxial match at the interfaces. The tomography images revealed complex geometries of the nanoassemblies with nonobvious handedness (Figure 3A and B; Supplementary Video File) as opposed to twisted ribbons or helixes.16,45 To interpret their chiroptical properties, we directly imported tomography coordinates into CD spectra calculation. Considering that the experimental spectra represent NP dispersions containing a range of various shapes and sizes, calculated spectra of two individual NPs are adequately correlated with the experimental results (Figure 3C), with respect to the polarity of the CD peaks and their spectral placement. To the best of our knowledge, this is the first realization of calculations of chiroptical properties performed on B

DOI: 10.1021/jacs.9b00700 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Communication

Journal of the American Chemical Society

Figure 2. Interparticle forces in the achiral plasmonic assemblies dependent on the polarization of incident light. (A) Three different illumination conditions: linear (black, LNP), left-handed circular (red, LCP), and right-handed circular (blue, RCP) polarization. The arrows show the directions of field for the incident photons as they pass through the nanoassembly (light propagates along the − z direction, light intensity: 1 MW/m2). (B) Assembly model with closely packed five 3 nm NPs (P1 to P5). (C) Volume arrow plot for the forces acting on each NP (red, P1; black, P2; green, P3; blue, P4; magenta, P5) under light with different polarization.

the true experimental geometry of the nanostructures. By comparing experimental CD and UV−vis spectra with the computed optical properties of the particles, the percentage of chiral particles among the synthesized Au NPs under LCP and RCP illumination is estimated as 11.9% and 7.10%, respectively (Supporting Information (SI), Figure S1, eq S1). Modification of this synthesis method by adding some chiral small molecules or capping agent can be considered to increase the enantiomeric excess and precision of shape control. In addition to CD spectra, the chirality of the resulting nanostructures can be enumerated using various chirality measures. The latter can be calculated for different molecules including (bio)organic nanoassemblies, such as proteins, based on their atomic formulas, atomic coordinates, and bonding patterns (see SI).53−55 Calculation of chirality measures for inorganic nanostructures presents, however, a challenge because their atomic structure is not consistent from particle to particle. Note that Pelayo et al.56 and Yeom et al.57 carried out calculations of the geometry of chiral clusters and their Hausdorff chirality measures (HCM) but for computational models and not for actual NPs. The accurate experimental geometries of chiral nanostructures are needed, however, in order to establish the relationship between polarity and amplitude of CD spectra and a chirality measure. A chiral nanostructure may be uniquely represented by a 3D density map, p(x,y,z), of the inorganic material that was calculated based on the imported tomography coordinates.

Table 1. Calculated Plasmonic Forces on each NP in the Models in Figure 2 (Light Intensity: 1 MW/m2)a

a

The origin of vectors is at the center of each particle.

Various chirality quantification methods could be applicable to nanoscale structures when they are coarse-grained to a set of nanoscale ‘atoms’. To accomplish this, we developed a process to quantitatively convert the p(x,y,z) function into an idealized molecule. The center of mass of an NP with an arbitrary shape, which will be referred to as the primary center of mass, was placed at the (0,0,0) point of the Cartesian coordinate system. The nanostructure was aligned to have its farthest surface point as the z intercept while the x intercept represents the farthest in the xy plane (Figure 4A). The structure was then sectioned into C

DOI: 10.1021/jacs.9b00700 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Communication

Journal of the American Chemical Society

always positive and provide an overall quantification of asymmetry. The values of all three chirality measured correlate with the amplitude of the CD plasmonic peak in the 400−700 nm spectral window (Figure 1A, 3C). Thus, the chirality quantification based on the coarse graining of electron tomography images may be considered as a design strategy for the chiral plasmonic nanostructures. In conclusion, CPL illumination of a dispersion of plasmonic NP transfers the chirality of photons to the chirality of NP assemblies. The nanoscale products showed opposite chiroptical activities, which depends on the handedness of CPL. Transient interparticle forces dependent on the polarization of incident light generate the chiral bias in the intermediate dynamic assemblies that is subsequently locked in shape by the merger of the nanoscale cores. Tomographic reconstructions made possible identification and quantification of chirality even for assemblies of seemingly irregular shape. Considering the ubiquity of plasmonic NPs,47,48,58−60 similar synthetic protocols based on light-induced forces can be applicable to other dispersions capable of spontaneous assembly into a superstructure with lattice-to-lattice connectivity. Preparation of more uniformly shaped nanoscale assemblies is possible by optimizing the power of the incident light and the NP−NP interactions.10

eight pieces according to octants of the Cartesian system (Figure 4B). Coordinates from their center of masses of the eight

Figure 3. Electron tomography of Au nanostructures obtained after CPL illumination and their calculated CD spectra. (A, B) Experimentally obtained tomographic reconstructions of LH (A) and RH (B) gold nanostructures (see also video file in SI). (C) Calculated CD spectra with geometry of the particle models imported directly from tomographic reconstructions (LH, black; RH, red).



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.9b00700. Materials; Methods: synthesis, optical setup, computation for CD, chirality measure, enantiomeric excess, and mechanical force on NPs under light polarization, electron tomography; Supplementary data: TEM images, CD and absorbance spectra of chiral Au NPs at various CPL illumination time points and achiral Au NPs formed by LNP illumination, calculated plasmonic forces and surface charge densities of modeled spheres under CPL (PDF) Reconstructed 3D structures of chiral Au NPs by electron tomography (PPTX)

Figure 4. Schematic description of collecting 3D density map from tomographic data of LH gold nanostructure shown in Figure 3A. (A) Placing NP in a Cartesian coordinate system having an origin at the primary center of total mass. (B) Dividing NP with the octants of coordinate system. (C) Generating the secondary centers of masses from the eight different partitions.



Table 2. Chirality Quantification for LH and RH Au Nanostructures Osipov−Pickup−Dunmor Chirality index (OPD) Hausdorff chirality Measure (HCM) Continuous Chirality Measure (CCM)

LH Au NP

RH Au NP

0.6493 0.1343 1.315

−4.965 0.1861 1.738

AUTHOR INFORMATION

Corresponding Author

*[email protected] ORCID

Peijun Zhang: 0000-0003-1803-691X Nicholas Kotov: 0000-0002-6864-5804 Author Contributions

sections, referred to as secondary centers of masses, have been calculated (Figure 4C, Table S1). Using the coordinates of the eight secondary center of masses and one primary center of mass (Figure 4C, Table S1), we have calculated three chirality measures: HCM,53 Osipov−Pickup− Dunmur chirality indices (OPD),55 and continuous chirality measure (CCM)54 (Table 2). The Broyden−Fletcher−Goldfarb−Shano (BFGS) numerical procedure was used to minimize HCM with a rotation angle of 1°. When calculating CCM, the eight secondary center of masses were assumed to be connected by a “bond” to the primary center of mass. The features of the CD spectra correspond very well with the chirality measures (Figure 3C). The sign of the OPD matches the signs of the experimental CD spectrum. HCM and CCM are



J-Y.K. and J.Y. contributed equally.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors thanks Dr. Didier Law-Hine and Dr. Hee-jeong Jang for discussions to establish the model for chiral quantification. N.A.K. thanks Prof. Juan Jose Saenz from Donostia International Physics Center for insightful discussions of plasmonic forces. This work was supported by the NSF 1463474 project titled “Energy- and Cost-Efficient Manufacturing Employing Nanoparticles” and Vannewar Bush DoD Fellowship to N.A.K. D

DOI: 10.1021/jacs.9b00700 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Communication

Journal of the American Chemical Society



(22) Müller, M.; Zentel, R. Interplay of Chiral Side Chains and Helical Main Chains in Polyisocyanates. Macromolecules 1996, 29 (5), 1609− 1617. (23) Zhang, S.; Zhou, J.; Park, Y. S.; Rho, J.; Singh, R.; Nam, S.; Azad, A. K.; Chen, H. T.; Yin, X.; Taylor, A. J.; et al. Photoinduced Handedness Switching in Terahertz Chiral Metamolecules. Nat. Commun. 2012, 3 (May), 942−947. (24) Nikolova, L.; Todorov, T.; Ivanov, M.; Andruzzi, F.; Hvilsted, S.; Ramanujam, P. S. Photoinduced Circular Anisotropy in Side-Chain Azobenzene Polyesters. Opt. Mater. (Amsterdam, Neth.) 1997, 8 (4), 255−258. (25) Iftime, G.; Labarthet, F. L.; Natansohn, A.; Rochon, P. Control of Chirality of an Azobenzene Liquid Crystalline Polymer with Circularly Polarized Light. J. Am. Chem. Soc. 2000, 122 (51), 12646−12650. (26) Feringa, B. L.; Van Delden, R. A. Absolute Asymmetric Synthesis: The Origin, Control, and Amplification of Chirality. Angew. Chem., Int. Ed. 1999, 38, 3418−3438. (27) Liu, N.; Liedl, T. DNA-Assembled Advanced Plasmonic Architectures. Chem. Rev. 2018, 118 (6), 3032−3053. (28) Govan, J.; Gun’ko, Y. K. Recent Progress in Chiral Inorganic Nanostructures. Nanoscience 2016, 3, 1−30. (29) Auguié, B.; Alonso-Gómez, J. L.; Guerrero-Martínez, A.; LizMarzán, L. M. Fingers Crossed: Optical Activity of a Chiral Dimer of Plasmonic Nanorods. J. Phys. Chem. Lett. 2011, 2 (8), 846−851. (30) Xia, Y.; Zhou, Y.; Tang, Z. Chiral Inorganic Nanoparticles: Origin, Optical Properties and Bioapplications. Nanoscale 2011, 3 (4), 1374. (31) Toyoda, K.; Miyamoto, K.; Aoki, N.; Morita, R.; Omatsu, T. Using Optical Vortex to Control the Chirality of Twisted Metal Nanostructures. Nano Lett. 2012, 12 (7), 3645−3649. (32) Wang, Y.; Xu, J.; Wang, Y.; Chen, H. Emerging Chirality in Nanoscience. Chem. Soc. Rev. 2013, 42 (7), 2930−2962. (33) Guerrero-Martínez, A.; Alonso-Gómez, J. L.; Auguié, B.; Cid, M. M.; Liz-Marzán, L. M. From Individual to Collective Chirality in Metal Nanoparticles. Nano Today 2011, 6 (4), 381−400. (34) Zhou, L.; Swearer, D. F.; Zhang, C.; Robatjazi, H.; Zhao, H.; Henderson, L.; Dong, L.; Christopher, P.; Carter, E. A.; Nordlander, P.; Halas, N. J. Quantifying Hot Carrier and Thermal Contributions in Plasmonic Photocatalysis. Science 2018, 362 (6410), 69−72. (35) Durán Pachón, L.; Yosef, I.; Markus, T. Z.; Naaman, R.; Avnir, D.; Rothenberg, G. Chiral Imprinting of Palladium with Cinchona Alkaloids. Nat. Chem. 2009, 1 (2), 160−164. (36) Saito, K.; Tatsuma, T. Chiral Plasmonic Nanostructures Fabricated by Circularly Polarized Light. Nano Lett. 2018, 18, 3209− 3212. (37) Boriskina, S. V.; Reinhard, B. M. Molding the Flow of Light on the Nanoscale: From Vortex Nanogears to Phase-Operated Plasmonic Machinery. Nanoscale 2012, 4 (1), 76−90. (38) Lucas, A.; Das Sarma, S. Electronic Sound Modes and Plasmons in Hydrodynamic Two-Dimensional Metals. Phys. Rev. B: Condens. Matter Mater. Phys. 2018, 97 (11), 115449. (39) Sawai, K.; Tatumi, R.; Nakahodo, T.; Fujihara, H. Asymmetric Suzuki-Miyaura Coupling Reactions Catalyzed by Chiral Palladium Nanoparticles at Room Temperature. Angew. Chem. 2008, 120, 7023− 7025. (40) Haruta, M. Size- and Support-Dependency in the Catalysis of Gold. Catal. Today 1997, 36, 153−166. (41) Ramezani-Dakhel, H.; Bedford, N. M.; Woehl, T. J.; Knecht, M. R.; Naik, R. R.; Heinz, H. Nature of Peptide Wrapping onto Metal Nanoparticle Catalysts and Driving Forces for Size Control. Nanoscale 2017, 9 (24), 8401−8409. (42) Gansel, J. K.; Gansel, J. K.; Thiel, M.; Rill, M. S.; Decker, M.; Bade, K.; Saile, V.; von Freymann, G.; Linden, S.; Wegener, M. Gold Helix Photonic Metamaterial as Broadband Circular Polarizer. Science 2009, 325, 1513−1515. (43) Wu, X.; Xu, L.; Liu, L.; Ma, W.; Yin, H.; Kuang, H.; Wang, L.; Xu, C.; Kotov, N. A. Unexpected Chirality of Nanoparticle Dimers and Ultrasensitive Chiroplasmonic Bioanalysis. J. Am. Chem. Soc. 2013, 135 (49), 18629−18636.

REFERENCES

(1) Ma, W.; Xu, L.; De Moura, A. F.; Wu, X.; Kuang, H.; Xu, C.; Kotov, N. A. Chiral Inorganic Nanostructures. Chem. Rev. 2017, 117 (12), 8041. (2) Lodahl, P.; Mahmoodian, S.; Stobbe, S.; Rauschenbeutel, A.; Schneeweiss, P.; Volz, J. Review Chiral Quantum Optics. Nature 2017, 541 (7638), 473−480. (3) Pendry, J. B. A. Chiral Route to Negative Refraction. Science 2004, 306 (5700), 1353−1355. (4) Ma, W.; Xu, L.; de Moura, A. F.; Wu, X.; Kuang, H.; Xu, C.; Kotov, N. A. Chiral Inorganic Nanostructures. Chem. Rev. 2017, 117 (12), 8041−8093. (5) Kuroda, R.; Endo, B.; Abe, M.; Shimizu, M. Chiral Blastomere Arrangement Dictates Zygotic Left − Right Asymmetry Pathway in Snails. Nature 2009, 462 (7274), 790−794. (6) Ohta, T.; Takaya, H.; Noyori, R. Binap-Ruthenium(Ii) Dicarboxylate Complexes - New, Highly Efficient Catalysts for Asymmetric Hydrogenations. Inorg. Chem. 1988, 27 (3), 566−569. (7) Ye, Y.; Cook, T. R.; Wang, S.-P.; Wu, J.; Li, S.; Stang, P. J. SelfAssembly of Chiral Metallacycles and Metallacages from a Directionally Adaptable BINOL-Derived Donor. J. Am. Chem. Soc. 2015, 137 (37), 11896−11899. (8) Smulders, M. M. J.; Stals, P. J. M.; Mes, T.; Paffen, T. F. E.; Schenning, A. P. H. J.; Palmans, A. R. A.; Meijer, E. W. Probing the Limits of the Majority-Rules Principle in a Dynamic Supramolecular Polymer. J. Am. Chem. Soc. 2010, 132 (2), 620−626. (9) Avilés-Moreno, J. R.; Quesada-Moreno, M. M.; López-González, J. J.; Martínez-Haya, B. Chiral Recognition of Amino Acid Enantiomers by a Crown Ether: Chiroptical IR-VCD Response and Computational Study. J. Phys. Chem. B 2013, 117 (32), 9362−9370. (10) Silvera Batista, C. A.; Larson, R. G.; Kotov, N. A. Nonadditivity of Nanoparticle Interactions. Science 2015, 350 (6257), 1242477− 1242477a. (11) Huck, N. P. M.; Jager, W. F.; de Lange, B. De; Feringa, B. L. Dynamic Control and Amplification of Molecular Chirality by Circular Polarized Light. Science 1996, 273 (5282), 1686−1688. (12) Jorissen, A.; Cerf, C. Asymmetric Photoreactions as the Origin of Biomolecular Homochirality: A Critical Review. Origins Life Evol. Biospheres 2002, 32 (2), 129−142. (13) Yang, G.; Zhang, S.; Hu, J.; Fujiki, M.; Zou, G.; Yang, G.; Zhang, S.; Hu, J.; Fujiki, M.; Zou, G. The Chirality Induction and Modulation of Polymers by Circularly Polarized Light. Symmetry 2019, 11 (4), 474. (14) Fukue, T. Recent Progress in Chirality Research Using Circularly Polarized Light; astro-G: 2010, Tokyo. (15) Bailey, J. Circular Polarization in Star- Formation Regions: Implications for Biomolecular Homochirality. Science 1998, 281 (5377), 672−674. (16) Yeom, J.; Yeom, B.; Chan, H.; Smith, K. W.; Dominguez-medina, S.; Bahng, J. H.; Zhao, G.; Chang, W.; Chang, S.; Chuvilin, A.; et al. Chiral Templating of Self-Assembling Nanostructures by Circularly Polarized Light. Nat. Mater. 2015, 14, 66−72. (17) Kotov, N. A. Self-Assembly of Inorganic Nanoparticles: Ab Ovo. Europhys. Lett. 2017, 119 (6), 66008. (18) Ayitou, A. J.-L.; Sivaguru, J. Light-Induced Transfer of Molecular Chirality in Solution: Enantiospecific Photocyclization of Molecularly Chiral Acrylanilides. J. Am. Chem. Soc. 2009, 131 (14), 5036−5037. (19) Kim, J.; Lee, J.; Kim, W. Y.; Kim, H.; Lee, S.; Lee, H. C.; Lee, Y. S.; Seo, M.; Kim, S. Y. Induction and Control of Supramolecular Chirality by Light in Self-Assembled Helical Nanostructures. Nat. Commun. 2015, 6, 1−8. (20) Zhong, Y.; Sisto, T. J.; Zhang, B.; Miyata, K.; Zhu, X.-Y.; Steigerwald, M. L.; Ng, F.; Nuckolls, C. Helical Nanoribbons for UltraNarrowband Photodetectors. J. Am. Chem. Soc. 2017, 139 (16), 5644− 5647. (21) Noorduin, W. L.; Bode, A. A. C.; van der Meijden, M.; Meekes, H.; van Etteger, A. F.; van Enckevort, W. J. P.; Christianen, P. C. M.; Kaptein, B.; Kellogg, R. M.; Rasing, T.; Vlieg, E.; et al. Complete Chiral Symmetry Breaking of an Amino Acid Derivative Directed by Circularly Polarized Light. Nat. Chem. 2009, 1 (12), 729−732. E

DOI: 10.1021/jacs.9b00700 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Communication

Journal of the American Chemical Society (44) Hu, T.; Isaacoff, B. P.; Bahng, J. H.; Hao, C.; Zhou, Y.; Zhu, J.; Li, X.; Wang, Z.; Liu, S.; Xu, C.; Biteen, J. S.; Kotov, N. A. SelfOrganization of Plasmonic and Excitonic Nanoparticles into Resonant Chiral Supraparticle Assemblies. Nano Lett. 2014, 14 (12), 6799−6810. (45) Kuzyk, A.; Schreiber, R.; Fan, Z.; Pardatscher, G.; Roller, E.-M.; Högele, A.; Simmel, F. C.; Govorov, A. O.; Liedl, T. DNA-Based SelfAssembly of Chiral Plasmonic Nanostructures with Tailored Optical Response. Nature 2012, 483, 311−314. (46) Yan, W.; Xu, L.; Xu, C.; Ma, W.; Kuang, H.; Wang, L.; Kotov, N. A. Self-Assembly of Chiral Nanoparticle Pyramids with Strong R/S Optical Activity. J. Am. Chem. Soc. 2012, 134 (36), 15114−15121. (47) Lee, G. P.; Shi, Y.; Lavoie, E.; Daeneke, T.; Reineck, P.; Cappel, U. B.; Huang, D. M.; Bach, U. Light-Driven Transformation Processes of Anisotropic Silver Nanoparticles. ACS Nano 2013, 7 (7), 5911− 5921. (48) Jin, R.; Cao, Y.; Mirkin, C. A.; Kelly, K. L.; Schatz, G. C.; Zheng, J. G. Photoinduced Conversion of Silver Nanospheres to Nanoprisms. Science 2001, 294, 1901−1904. (49) Kurihara, K.; Kizling, J.; Stenius, P.; Fendler, J. H. Laser and Pulse Radiolytically Induced Colloidal Gold Formation in Water and in Water-in-Oil Microemulsions. J. Am. Chem. Soc. 1983, 105 (9), 2574− 2579. (50) Penn, R. L.; Banfield, J. F. Imperfect Oriented Attachment: Dislocation Generation in Defect-Free Nanocrystals. Science 1998, 281 (5379), 969−971. (51) Tang, Z.; Kotov, N. A.; Giersig, M. Spontaneous Organization of Single CdTe Nanoparticles into Luminescent Nanowires. Science 2002, 297 (5579), 237−240. (52) De Yoreo, J. J.; Gilbert, P. U. P. A.; Sommerdijk, N. A. J. M.; Penn, R. L.; Whitelam, S.; Joester, D.; Zhang, H.; Rimer, J. D.; Navrotsky, A.; Banfield, J. F. Crystallization by Particle Attachment in Synthetic, Biogenic, and Geologic Environments. Science 2015, 349 (6247), aaa6760. (53) Buda, A. B.; Mislow, K. A Hausdorff Chirality Measure. J. Am. Chem. Soc. 1992, 114 (15), 6006−6012. (54) Zabrodsky, H.; Avnir, D. Continuous Symmetry Measures. 4. Chirality. J. Am. Chem. Soc. 1995, 117, 462−473. (55) Osipov, M. A.; Pickup, B. T.; Dunmur, D. A. A New Twist to Molecular Chirality : Intrinsic Chirality Indices. Mol. Phys. 1995, 84 (6), 1193−1206. (56) Pelayo, J. J.; Garzon, I. L.; Whetten, R. L. Geometric Quantification of Chirality in Ligand-Protected Metal Clusters. J. Phys. Chem. C 2015, 119, 28666. (57) Yeom, J.; Santos, U. S.; Chekini, M.; Cha, M.; De Moura, A. F.; Kotov, N. A. Chiromagnetic Nanoparticles and Gels. Science 2018, 359 (6373), 309. (58) Sun, Y.; Xia, Y. Triangular Nanoplates of Silver: Synthesis, Characterization, and Use as Sacrificial Templates For Generating Triangular Nanorings of Gold. Adv. Mater. 2003, 15 (9), 695−699. (59) Maillard, M.; Huang, P.; Brus, L. Silver Nanodisk Growth by Surface Plasmon Enhanced Photoreduction of Adsorbed [Ag+]. Nano Lett. 2003, 3, 1611−1615. (60) Zhang, J.; Langille, M. R.; Mirkin, C. A. Photomediated Synthesis of Silver Triangular Bipyramids and Prisms : The Effect of PH and BSPP. J. Am. Chem. Soc. 2010, 132 (28), 12502−12510.

F

DOI: 10.1021/jacs.9b00700 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX