emergence of hierarchical structural complexities in ......2016/09/13  · emergence of hierarchical...

6
dissipative nonreciprocal devices, a circulator that is controlled by a single quantum system also enables operation in coherent superposi- tion states of routing light in one and the other direction, providing a route toward its applica- tion in future photonic quantum protocols. The demonstrated operation principle is universal in the sense that it can straightforwardly be im- plemented with a large variety of different quan- tum emitters provided that they exhibit circularly polarized optical transitions and that they can be spin-polarized. Using state-of-the-art WGM microresonators (29), one could realize a circula- tor with optical losses below 7% and close-to-unit operation fidelity (26). This would then allow one to almost deterministically process and control photons in an integrated optical environment. Arranging N circulators so that they form a linear array allows one to realize a (2N + 2)-port optical circulator. Moreover, two- and three-dimensional networks of quantum circulators are potential candidates for implementing lattice-based quan- tum computation (30). Such networks would enable the implementation of artificial gauge fields for photons (3133), in which a nonlinearity at the level of single quanta allows for the flux to become a dynamical degree of freedom that interacts with the particles themselves (34). REFERENCES AND NOTES 1. B. J. H. Stadler, T. Mizumoto, IEEE Photonics J. 6,115 (2014). 2. K. Gallo, G. Assanto, K. R. Parameswaran, M. M. Fejer, Appl. Phys. Lett. 79, 314 (2001). 3. L. Fan et al., Science 335, 447450 (2012). 4. B. Peng et al., Nat. Phys. 10, 394398 (2014). 5. H. Lira, Z. Yu, S. Fan, M. Lipson, Phys. Rev. Lett. 109, 033901 (2012). 6. L. D. Tzuang, K. Fang, P. Nussenzveig, S. Fan, M. Lipson, Nat. Photonics 8, 701705 (2014). 7. J. Kim, M. C. Kuzyk, K. Han, H. Wang, G. Bahl, Nat. Phys. 11, 275280 (2015). 8. Y. Shoji, T. Mizumoto, Sci. Technol. Adv. Mater. 15, 014602 (2014). 9. D. Huang et al., Conference on Lasers and Electro-Optics (2016), p. SM3E.1. 10. N. Gisin, R. Thew, Nat. Photonics 1, 165171 (2007). 11. M. A. Nielsen, I. L. Chuang, Quantum Computation and Quantum Information (Cambridge Univ. Press, 2011). 12. R. P. Feynman, Int. J. Theor. Phys. 21, 467488 (1982). 13. M. Pöllinger, A. Rauschenbeutel, Opt. Express 18, 1776417775 (2010). 14. D. OShea, C. Junge, J. Volz, A. Rauschenbeutel, Phys. Rev. Lett. 111, 193601 (2013). 15. C. Junge, D. OShea, J. Volz, A. Rauschenbeutel, Phys. Rev. Lett. 110, 213604 (2013). 16. P. Lodahl et al., https://arxiv.org/abs/1608.00446 (2016). 17. I. Shomroni et al., Science 345, 903906 (2014). 18. S. Rosenblum et al., Nat. Photonics 10, 1922 (2016). 19. C. Sayrin et al., Phys. Rev. X 5, 041036 (2015). 20. H. J. Metcalf, P. van der Straten, Laser Cooling and Trapping (Springer, 1999). 21. D. Jalas et al., Nat. Photonics 7, 579582 (2013). 22. K. Xia et al., Phys. Rev. A 90, 043802 (2014). 23. E. J. Lenferink, G. Wei, N. P. Stern, Opt. Express 22, 1609916111 (2014). 24. I. Söllner et al., Nat. Nanotechnol. 10, 775778 (2015). 25. C. Gonzalez-Ballestero, E. Moreno, F. J. Garcia-Vidal, A. Gonzalez-Tudela, https://arxiv.org/abs/1608.04928 (2016). 26. Materials and methods are available as supplementary materials on Science Online. 27. A. G. White et al., J. Opt. Soc. Am. B 24, 172 (2007). 28. J. Volz, M. Scheucher, C. Junge, A. Rauschenbeutel, Nat. Photonics 8, 965970 (2014). 29. M. Pöllinger, D. OShea, F. Warken, A. Rauschenbeutel, Phys. Rev. Lett. 103, 053901 (2009). 30. R. Raussendorf, H. J. Briegel, Phys. Rev. Lett. 86, 51885191 (2001). 31. J. Koch, A. A. Houck, K. L. Hur, S. M. Girvin, Phys. Rev. A 82, 043811 (2010). 32. M. Hafezi, P. Rabl, Opt. Express 20, 76727684 (2012). 33. M. Schmidt, S. Kessler, V. Peano, O. Painter, F. Marquardt, Optica 2, 635 (2015). 34. S. Walter, F. Marquardt, New J. Phys. 18, 113029 (2016). ACKNOWLEDGMENTS The authors are grateful to J. Simon and M. Levy for helpful discussions. We gratefully acknowledge financial support by the Austrian Science Fund (FWF; SFB FoQuS project no. F 4017 and DK CoQuS project no. W 1210-N16) and the European Commission (IP SIQS, no. 600645). A.H. acknowledges financial support from the Austrian Science Fund (FWF; Meitner Program Project M 1970). SUPPLEMENTARY MATERIALS www.sciencemag.org/content/354/6319/1577/suppl/DC1 Materials and Methods Figs. S1 and S2 Table S1 Reference (35) 13 September 2016; accepted 21 November 2016 Published online 8 December 2016 10.1126/science.aaj2118 NANOMATERIALS Emergence of hierarchical structural complexities in nanoparticles and their assembly Chenjie Zeng, 1 Yuxiang Chen, 1 Kristin Kirschbaum, 2 Kelly J. Lambright, 2 Rongchao Jin 1 * We demonstrate that nanoparticle self-assembly can reach the same level of hierarchy, complexity, and accuracy as biomolecules. The precise assembly structures of gold nanoparticles (246 gold core atoms with 80 p-methylbenzenethiolate surface ligands) at the atomic, molecular, and nanoscale levels were determined from x-raydiffraction studies. We identified the driving forces and rules that guide the multiscale assembly behavior.The protecting ligands self-organize into rotational and parallel patterns on the nanoparticle surface via C-H⋅⋅⋅p interaction, and the symmetry and density of surface patterns dictate directional packing of nanoparticles into crystals with orientational, rotational, and translational orders. Through hierarchical interactions and symmetry matching, the simple building blocks evolve into complex structures, representing an emergent phenomenon in the nanoparticle system. H ierarchical self-assembly of nanoparticles (NPs) into complex architectures across different length scales is an important ca- pability in nanotechnology (14), especially for the bottom-up fabrication for electron- ics, sensors, energy conversion, and storage devices. Such self-assembly can be driven by entropy- dictated maximization of the packing density, as demonstrated in close packing of spheres, binary NPs, rods, and hard polyhedrons (59). Interparticle interactions, such as the electrostatic attraction (10), cDNA binding (11, 12), and patchy NP surfaces (1315), have also been exploited to guide assembly into diverse lattice structures. Despite these advances, NP assembly has not achieved the same level of atomic accuracy as in biological systems. We now demonstrate that NP-assembled struc- tures can reach the same hierarchy and atomic accuracy as biomolecules at the interparticle and intraparticle levels. Through crystalliza- tion of uniform 2.2-nm gold NPs bearing p- methylbenzenethiolate (p-MBT) surface ligands [Au 246 (p-MBT) 80 ], we fully resolve the entire self- assembled structures at atomic (packing of gold atoms), molecular (packing of surface ligands), and nanoscale (packing of NPs) levels by single- crystal x-ray diffraction (SC-XRD) (16). The precise structural information across scales allows an in- depth examination of the forces and the rules that govern the assembly behavior at each level. We reveal that the simple structure of protecting ligands can generate complex patterns on the NP surfaces, and the symmetry and density of the surface patterns further guide the packing of NPs into lattices with orientational, rotational, and translational order. The Au 246 (p-MBT) 80 NPs were synthesized by a two-step size-focusingmethod (16). Briefly, the ligand-coated NPs in a narrow size range from ~10 to ~70 kilodaltons were first made, and then the size focusing process gradually led to the stable Au 246 (p-MBT) 80 NPs (figs. S1 and S2). The as- obtained product (~90% purity) was subject to further solvent fractionation to reach molecu- larly pure Au 246 NPs. Optical absorbance spectra showed a prominent peak at 470 nm and sev- eral weak humps at 400 and 600 nm (fig. S3), indicating the nonplasmonic nature of the Au 246 (p-MBT) 80 NPs. Single crystals were grown by dif- fusion of antisolvent (acetonitrile) into a toluene solution of the pure Au 246 NPs. The structure was determined at the resolution of 0.96 Å by SC-XRD 1580 23 DECEMBER 2016 VOL 354 ISSUE 6319 sciencemag.org SCIENCE 1 Department of Chemistry, Carnegie Mellon University, Pittsburgh, PA 15213, USA. 2 Department of Chemistry and Biochemistry, University of Toledo, Toledo, OH 43606, USA. *Corresponding author. Email: [email protected] RESEARCH | REPORTS on June 9, 2021 http://science.sciencemag.org/ Downloaded from

Upload: others

Post on 29-Jan-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

  • dissipative nonreciprocal devices, a circulatorthat is controlled by a single quantum systemalso enables operation in coherent superposi-tion states of routing light in one and the otherdirection, providing a route toward its applica-tion in future photonic quantum protocols. Thedemonstrated operation principle is universal inthe sense that it can straightforwardly be im-plemented with a large variety of different quan-tum emitters provided that they exhibit circularlypolarized optical transitions and that they canbe spin-polarized. Using state-of-the-art WGMmicroresonators (29), one could realize a circula-tor with optical losses below 7% and close-to-unitoperation fidelity (26). This would then allow oneto almost deterministically process and controlphotons in an integrated optical environment.ArrangingN circulators so that they form a lineararray allows one to realize a (2N + 2)-port opticalcirculator.Moreover, two- and three-dimensionalnetworks of quantum circulators are potentialcandidates for implementing lattice-based quan-tum computation (30). Such networks wouldenable the implementation of artificial gaugefields for photons (31–33), in which a nonlinearityat the level of single quanta allows for the fluxto become a dynamical degree of freedom thatinteracts with the particles themselves (34).

    REFERENCES AND NOTES

    1. B. J. H. Stadler, T. Mizumoto, IEEE Photonics J. 6, 1–15(2014).

    2. K. Gallo, G. Assanto, K. R. Parameswaran, M. M. Fejer, Appl.Phys. Lett. 79, 314 (2001).

    3. L. Fan et al., Science 335, 447–450 (2012).4. B. Peng et al., Nat. Phys. 10, 394–398 (2014).5. H. Lira, Z. Yu, S. Fan, M. Lipson, Phys. Rev. Lett. 109, 033901

    (2012).6. L. D. Tzuang, K. Fang, P. Nussenzveig, S. Fan, M. Lipson,

    Nat. Photonics 8, 701–705 (2014).7. J. Kim, M. C. Kuzyk, K. Han, H. Wang, G. Bahl, Nat. Phys. 11,

    275–280 (2015).8. Y. Shoji, T. Mizumoto, Sci. Technol. Adv. Mater. 15, 014602 (2014).9. D. Huang et al., Conference on Lasers and Electro-Optics

    (2016), p. SM3E.1.10. N. Gisin, R. Thew, Nat. Photonics 1, 165–171 (2007).11. M. A. Nielsen, I. L. Chuang, Quantum Computation and

    Quantum Information (Cambridge Univ. Press, 2011).12. R. P. Feynman, Int. J. Theor. Phys. 21, 467–488 (1982).13. M. Pöllinger, A. Rauschenbeutel, Opt. Express 18, 17764–17775

    (2010).14. D. O’Shea, C. Junge, J. Volz, A. Rauschenbeutel, Phys. Rev.

    Lett. 111, 193601 (2013).15. C. Junge, D. O’Shea, J. Volz, A. Rauschenbeutel, Phys. Rev.

    Lett. 110, 213604 (2013).16. P. Lodahl et al., https://arxiv.org/abs/1608.00446 (2016).17. I. Shomroni et al., Science 345, 903–906 (2014).18. S. Rosenblum et al., Nat. Photonics 10, 19–22 (2016).19. C. Sayrin et al., Phys. Rev. X 5, 041036 (2015).20. H. J. Metcalf, P. van der Straten, Laser Cooling and Trapping

    (Springer, 1999).21. D. Jalas et al., Nat. Photonics 7, 579–582 (2013).22. K. Xia et al., Phys. Rev. A 90, 043802 (2014).23. E. J. Lenferink, G. Wei, N. P. Stern, Opt. Express 22,

    16099–16111 (2014).24. I. Söllner et al., Nat. Nanotechnol. 10, 775–778 (2015).25. C. Gonzalez-Ballestero, E. Moreno, F. J. Garcia-Vidal,

    A. Gonzalez-Tudela, https://arxiv.org/abs/1608.04928 (2016).26. Materials and methods are available as supplementary

    materials on Science Online.27. A. G. White et al., J. Opt. Soc. Am. B 24, 172 (2007).28. J. Volz, M. Scheucher, C. Junge, A. Rauschenbeutel,

    Nat. Photonics 8, 965–970 (2014).29. M. Pöllinger, D. O’Shea, F. Warken, A. Rauschenbeutel, Phys.

    Rev. Lett. 103, 053901 (2009).30. R. Raussendorf, H. J. Briegel, Phys. Rev. Lett. 86, 5188–5191 (2001).

    31. J. Koch, A. A. Houck, K. L. Hur, S. M. Girvin, Phys. Rev. A 82,043811 (2010).

    32. M. Hafezi, P. Rabl, Opt. Express 20, 7672–7684 (2012).33. M. Schmidt, S. Kessler, V. Peano, O. Painter, F. Marquardt,

    Optica 2, 635 (2015).34. S. Walter, F. Marquardt, New J. Phys. 18, 113029 (2016).

    ACKNOWLEDGMENTS

    The authors are grateful to J. Simon and M. Levy for helpfuldiscussions. We gratefully acknowledge financial support by theAustrian Science Fund (FWF; SFB FoQuS project no. F 4017and DK CoQuS project no. W 1210-N16) and the EuropeanCommission (IP SIQS, no. 600645). A.H. acknowledges financial

    support from the Austrian Science Fund (FWF; MeitnerProgram Project M 1970).

    SUPPLEMENTARY MATERIALS

    www.sciencemag.org/content/354/6319/1577/suppl/DC1Materials and MethodsFigs. S1 and S2Table S1Reference (35)

    13 September 2016; accepted 21 November 2016Published online 8 December 201610.1126/science.aaj2118

    NANOMATERIALS

    Emergence of hierarchical structuralcomplexities in nanoparticles andtheir assemblyChenjie Zeng,1 Yuxiang Chen,1 Kristin Kirschbaum,2

    Kelly J. Lambright,2 Rongchao Jin1*

    We demonstrate that nanoparticle self-assembly can reach the same level of hierarchy,complexity, and accuracy as biomolecules. The precise assembly structures of goldnanoparticles (246 gold core atoms with 80 p-methylbenzenethiolate surface ligands) atthe atomic, molecular, and nanoscale levels were determined from x-ray diffraction studies.We identified the driving forces and rules that guide the multiscale assembly behavior. Theprotecting ligands self-organize into rotational and parallel patterns on the nanoparticle surfacevia C-H⋅⋅⋅p interaction, and the symmetry and density of surface patterns dictate directionalpacking of nanoparticles into crystals with orientational, rotational, and translational orders.Through hierarchical interactions and symmetry matching, the simple building blocks evolveinto complex structures, representing an emergent phenomenon in the nanoparticle system.

    Hierarchical self-assembly of nanoparticles(NPs) into complex architectures acrossdifferent length scales is an important ca-pability in nanotechnology (1–4), especiallyfor the bottom-up fabrication for electron-

    ics, sensors, energy conversion, and storage devices.Such self-assembly can be driven by entropy-dictated maximization of the packing density,as demonstrated in close packing of spheres,binary NPs, rods, and hard polyhedrons (5–9).Interparticle interactions, such as the electrostaticattraction (10), cDNA binding (11, 12), and patchyNP surfaces (13–15), have also been exploited toguide assembly into diverse lattice structures.Despite these advances, NP assembly has notachieved the same level of atomic accuracyas in biological systems.Wenowdemonstrate thatNP-assembled struc-

    tures can reach the same hierarchy and atomicaccuracy as biomolecules at the interparticleand intraparticle levels. Through crystalliza-tion of uniform 2.2-nm gold NPs bearing p-methylbenzenethiolate (p-MBT) surface ligands[Au246(p-MBT)80], we fully resolve the entire self-

    assembled structures at atomic (packing of goldatoms), molecular (packing of surface ligands),and nanoscale (packing of NPs) levels by single-crystal x-ray diffraction (SC-XRD) (16). The precisestructural information across scales allows an in-depth examination of the forces and the rules thatgovern the assembly behavior at each level. Wereveal that the simple structure of protectingligands can generate complex patterns on theNP surfaces, and the symmetry and density ofthe surface patterns further guide the packingof NPs into lattices with orientational, rotational,and translational order.The Au246(p-MBT)80 NPs were synthesized by

    a two-step “size-focusing” method (16). Briefly,the ligand-coatedNPs in a narrow size range from~10 to ~70 kilodaltons were first made, and thenthe size focusing process gradually led to the stableAu246(p-MBT)80 NPs (figs. S1 and S2). The as-obtained product (~90% purity) was subject tofurther solvent fractionation to reach molecu-larly pure Au246 NPs. Optical absorbance spectrashowed a prominent peak at 470 nm and sev-eral weak humps at 400 and 600 nm (fig. S3),indicating the nonplasmonic nature of the Au246(p-MBT)80 NPs. Single crystals were grown by dif-fusion of antisolvent (acetonitrile) into a toluenesolution of the pure Au246 NPs. The structure wasdetermined at the resolution of 0.96 Å by SC-XRD

    1580 23 DECEMBER 2016 • VOL 354 ISSUE 6319 sciencemag.org SCIENCE

    1Department of Chemistry, Carnegie Mellon University,Pittsburgh, PA 15213, USA. 2Department of Chemistry andBiochemistry, University of Toledo, Toledo, OH 43606, USA.*Corresponding author. Email: [email protected]

    RESEARCH | REPORTSon June 9, 2021

    http://science.sciencemag.org/

    Dow

    nloaded from

    http://www.sciencemag.org/content/354/6319/1577/suppl/DC1http://science.sciencemag.org/

  • (tables S1 and S2). The individual NPs have anearly spherical shape, with a metal core diam-eter of 2.2 nm and an overall diameter (includingthe ligand shell) of 3.3 nm (fig. S4).In contrast to conventional spherical NPs,

    which are typically packed into superlatticeswith simple translational symmetry, such as face-centered cubic (fcc) or body-centered cubic (bcc)(11, 17), the Au246(p-MBT)80 NPs packed into amore complex monoclinic lattice (Fig. 1). In the(001) plane of the crystal lattice, the NPs or-ganized into a square lattice (Fig. 1A), akin to thepacking mode in the {100} plane of the fcc lattice.The interparticle distance of 3.1 nm (less than the3.3 nm overall size of the NP) arose from ligandinterlocking. These two-dimensional (2D) latticesstacked up obliquely, instead of perpendicularlyalong the z direction (Fig. 1B), so the overall3D lattice deviated from the fcc lattice. The NPpacking density of ~60% is less than the 74% ofthe fcc lattice. Such monoclinic packing shouldbe driven by specific interparticle interactions,because entropy alone would favor close packing(5–9). Indeed, when zooming into the surfacesof the NPs, the monoclinic packing was correlatedto the alignment of surface ligands among NPs(Fig. 1, D to F, green, blue, red).The p-MBT ligands were highly ordered and

    self-organized into two different patterns on the

    gold sphere (Fig. 2A). At the pole site of thegold sphere, 25 of the p-MBTs were rotationallyarranged into four pentagonal circles (Fig. 2B,highlighted in red, blue, blue, green). Each circlehad the same “latitude” and rotational direction.Instead of “on-top” adsorption, the thiolates (char-acterized by S-C bond vectors) “tilted” away fromthe radial direction of the gold sphere becauseof the gold-thiolate binding geometry (18). Thispattern we call a-rotation created a “singularity”(19) at the pole (Fig. 2B). At the waist site, six ofthe p-MBT ligands were aligned into three al-ternating parallel pairs to form a pattern wecall b-parallel (Fig. 2C). Five of these b-parallelpatches circled up and covered the waist of theNP. The packing densities of ligands are ~14ligands nm–2 for a-rotation and ~6 ligands nm–2

    for b-parallel as measured based on the surfacearea of the inner gold sphere (Fig. 2A, magentapolyhedron). The clockwise and counterclockwiserotational arrangement ofp-MBT ligands induceschirality in the NP, and both chiral isomers (de-notedR/L) participate in the crystal packing (Fig.1, C and F). The NPs with the same chirality arepacked in the same square layer, and the neigh-boring square layers are composed of NPs withopposite chirality (Fig. 1C).Such rotational and parallel self-assembled

    surface patterns of ligands are reminiscent of

    the a helix and b sheet in proteins, which sug-gests that NPs could exhibit a level of structuralcomplexity comparable to that of biomolecules.The secondary structures of proteins are mainlystabilized by the hydrogen bonds. Here, the sur-face patterns on the NPs are stabilized by inter-molecular C-H⋅⋅⋅p interactions, in which the C-Hbonds from the phenyl rings or themethyl groupsinteract with the p electrons (Fig. 2D). Suchintermolecular interactions were observed in thepacking structures of aromatic molecules andsupramolecules, and the strength is about 1.5 to2.5 kcal mol–1 (20, 21). Specifically, the C-H⋅⋅⋅pinteractions in thea-rotation linked the 25p-MBTligands into five spirals, with the H⋅⋅⋅p distancesranging from2.5 to 3.0 Å andC-H-p angle rangingfrom 112° to 147° (fig. S5). For the b-parallels, thealternating pattern among the three pairs wasalso stabilized by the C-H⋅⋅⋅p interactions (fig. S6).Within each parallel pair, the phenyl rings wereoffset to avoid the repulsion between p electrons.We reason that the surface patterns are intrinsic toNPs instead of being induced by crystallization,because the collective C-H⋅⋅⋅p interactions can gen-erate an energy barrier and stabilize the pattern,similar to the case in which the hydrogen bondscan stabilize the DNA double helix in solution.To study the interparticle interactions, we

    isolated the coordination environment of NPs

    SCIENCE sciencemag.org 23 DECEMBER 2016 • VOL 354 ISSUE 6319 1581

    R

    L

    R

    L

    y

    x

    z

    y

    x

    z

    R

    L

    R

    L

    Fig. 1. Packing structure of the derivatized Au NPs in single crystals. (A to C) View from z direction (A), y direction (B), and x direction (C). Magenta,blue: Au NPs with different chirality; yellow: sulfur; gray: carbon. (D to F) Alignment of surface ligands among the NPs. Gray: ligands located at the waistof the NP; red, blue, green: ligands located at the poles.

    RESEARCH | REPORTSon June 9, 2021

    http://science.sciencemag.org/

    Dow

    nloaded from

    http://science.sciencemag.org/

  • 1582 23 DECEMBER 2016 • VOL 354 ISSUE 6319 sciencemag.org SCIENCE

    - parallelwaist

    pole -rotation

    pole waist

    HC

    Fig. 2. Self-assembled surface patterns of the ligands on the Au NPs. (A) Overall structure of ligands on the surface of NPs. (B) Rotational packingof ligands at the pole site of the NP. (C) Parallel packing of ligands at the waist of the NP. (D) The C-H⋅⋅⋅p interactions for stabilizing the large-scalerotational patterns and parallel patterns.

    1

    2

    3

    4

    5

    1

    2 3

    4

    5

    x

    Fig. 3. Interparticle self-assembly dictated by the ligand density and the symmetry of surface patterns. (A and B) Coordination geometry of NPs in thecrystal lattice: side view (A) and top view (B). (C) Contacting environment among the interparticle ligands.The outside pentagons are located at the bottom ofthe top three nanoparticles in (A), and the central pentagon is located at the top of the central nanoparticle. (D) Side-by-side stacking of the ligands in the NPswith the same chirality. (E) Point-to-point stacking of the ligands in the NPswith opposite chirality. (F) Scheme showing the directional packing of NPs achievedthrough matching the symmetry of surface patterns.

    RESEARCH | REPORTSon June 9, 2021

    http://science.sciencemag.org/

    Dow

    nloaded from

    http://science.sciencemag.org/

  • (Fig. 3), which reflects the orientational, rota-tional, and translational symmetry of the crystallattice. Each NP has six nearest neighbors. Fourof them were within the same square layer andhad the same chirality as the central one (Fig. 3A,yellow arrows). The particle-to-particle distancewas 31.0 Å, whereas the other two had oppositechirality (Fig. 3B, purple arrows) and the distancewas 31.7 Å. These NPs interacted with the centralone by aligning the a-rotation ligands at the polesites, whereas there were fewer interactions ofthe b-parallel ligands at the waist site. The pre-ferred alignment of a-rotation ligands was cor-related to the higher packing density of ligandsin a-rotation (~14 nm-2) compared to the densityin b-parallel (~6 nm-2). In this way, the van derWaals interaction among the contacting ligandswas maximized. This surface-density–dictated as-sembly strategy is in accordance with the assemblybehavior observed in DNA-coated NPs (11, 12, 22),oleic acid–protected nanoplates (15), and nano-crystals (17).In addition to the maximization of ligand in-

    teractions,matching of surface symmetrywas alsoimportant for ordered NP assembly. For a moredetailed picture on interparticle interactions, wefurther isolated the contacting region amongthe top three and the central NPs (Fig. 3A, whiteframe), as reflected in the four pentagons shownin Fig. 3C. Notably, the central NP spontaneouslymatched with the same surface region of itsneighbors when forming packing interaction (Fig.3, C to E). Each contacting area was composed ofabout five pairs of symmetrically identical p-MBTligands located at the corners of the contactingpentagons (Fig. 3, D and E). For the neighboringNPs with the same chirality as the central one,

    the pentagons were aligned side by side (Fig. 3D),whereas for the opposite chirality, the pentagonswere aligned point to point (Fig. 3E). The spacingbetween interacting ligands was ~2.5 Å, and theshortest spacing could reach 1.9 Å (fig. S7). Thisshort distance indicates tight packing amongNPs. The symmetry-matching strategy is akin tothe assembly of gears (Fig. 3F), with each con-tacting area resembling a tooth of the gear.Matching the symmetry facilitates the inter-locking of surface ligands. Eachpentagon has fivepotential contacting areas to connect with fiveother NPs, but only three are occupied becauseof the limited space around the NP.The assembly structure provides a precise de-

    piction of the long-standing issue of ligand ef-fects on the packing structures of NPs (23, 24).It implies that the inhomogeneous but symmet-ric distribution of surface ligands can serve as“sticky bonds” for directional NP assembly. Suchspontaneously organized surface patterns havean effect similar to that of artificially decoratedsurface patches in guiding assembly (3, 13, 14).The perfect uniformity of the NPs was criticalfor maintaining the fidelity of surface patterns.The packing behavior also represents an emer-gent phenomenon (25, 26), in which small andsimple entities (the surface ligands), throughmultiscale interactions, generate larger and morecomplex structures with new features that arenot manifested in the simple entities. If the p-MBT ligand is viewed as the primary structure,then through the C-H⋅⋅⋅p interactions among theligands, a more complex secondary structure ofsurface patterns is generated, and these surfacepatterns, via the density- and symmetry-dictatedpacking rules, further guide the packing of NPs

    into the tertiary structure of the crystal lattice. Itis expected that by controlling the symmetriesof surface ligand patterns, diverse packing struc-tures of NPs could be achieved.The assembly behavior within each NP also

    exhibits such order and hierarchy. The individ-ual NP can be divided into four regions from thecore to surface. The innermost part is a three-shellAu116 Ino decahedron (i-Dh) exposing {111} facetsat the poles and {100} facets at the waist (Fig. 4Aand fig. S8). The i-Dh is one of the energy minimafor packing of metal atoms (27–29). The secondpart is a transition layer containing 90 gold atoms(Fig. 4B, magenta), which “sphericizes” the i-Dh,lowers the surface energy, and provides “foot-holds” for anchoring surface protecting motifs.The two parts together give rise to an Au206 core.The average Au-Au bond length in the i-Dh was2.87 ± 0.05 Å (fig. S9), whereas the bonds asso-ciated with the transition layer showed largerdeviations (2.89 ± 1.12 Å) because of the bindingeffect of surface ligands. The third part is the Au-Sinterfacial layer, in which the surface danglingbonds of the Au206 core are anchored by theprotecting motifs assembled from 40 gold and80 sulfur atoms. The protecting motifs are highlydiverse, in accordance with the rich surface fea-tures of the Au206 core (such as facets and grooves,Fig. 4B). At the poles, the exposed {111} facetsare protected by –S–Au–S–Au–S–motifs and the{111}|{111} grooves are linked by simple bridgingthiolates (Fig. 4C, bottom); at the waist, the{100}|{100} and {111}|{100} grooves are all coveredby –S–Au–S– staplemotifs (Fig. 4C, top). The fourthpart is the surface carbon layer as discussedabove (Fig. 4D). Although each part of the NPfollowed different assembling rules, the five-fold symmetry was always maintained (Fig. 4E).An intriguing question is whether the packingsymmetry of the surface patterns emerged fromthe core or the core symmetry emerged from thesurface. Although each part of NP contributes tothe overall energy minimization, we deduce thatthe surface ligands play a pivotal role in guidingthe intraparticle assembly, because the gold atomsare less selective when packing into specific struc-ture types such as fcc, decahedron, or icosahedron(30). In addition, the rotational patterns of ligandsinduce the chirality in the interfacial layer andtransitional layer (figs. S10 and S11), and packing ofgold atoms alone cannot give rise to chirality. Also,the larger sphere of the ligands likely has a greatersurface energy to minimize, and the structures ofAu NPs were highly sensitive to the subtle changesof ligands (29).

    REFERENCE AND NOTES

    1. M. R. Jones, N. C. Seeman, C. A. Mirkin, Science 347, 1260901 (2015).2. C. B. Murray, C. R. Kagan, M. G. Bawendi, Annu. Rev. Mater. Sci.

    30, 545–610 (2000).3. S. C. Glotzer, M. J. Solomon, Nat. Mater. 6, 557–562 (2007).4. B. A. Grzybowski,W. T. S. Huck,Nat. Nanotechnol. 11, 585–592 (2016).5. E. V. Shevchenko, D. V. Talapin, N. A. Kotov, S. O’Brien,

    C. B. Murray, Nature 439, 55–59 (2006).6. M. Eldridge, P. Madden, D. Frenkel, Nature 365, 35–37 (1993).7. T. Wang et al., Science 338, 358–363 (2012).8. J. Henzie, M. Grünwald, A. Widmer-Cooper, P. L. Geissler,

    P. Yang, Nat. Mater. 11, 131–137 (2011).9. P. F. Damasceno, M. Engel, S. C. Glotzer, Science 337, 453–457 (2012).

    SCIENCE sciencemag.org 23 DECEMBER 2016 • VOL 354 ISSUE 6319 1583

    {111}

    {100}

    Top view: Au116

    Side view: Au116

    -S-Au-S-Au-S- & -S-

    -S-Au-S-

    Fig. 4. Intraparticle self-assembly in the Au NP. (A) Au116 i-Dh kernel, top: side view; bottom: top view.(B) Transition layer structure as colored in magenta. (C) Gold-sulfur interfacial structure containing di-verse surface protecting motifs. (D) Surface carbon layer and overall structure. (E) Intraparticle sym-metry matching and emergent behavior.

    RESEARCH | REPORTSon June 9, 2021

    http://science.sciencemag.org/

    Dow

    nloaded from

    http://science.sciencemag.org/

  • 10. A. M. Kalsin et al., Science 312, 420–424 (2006).11. S. Y. Park et al., Nature 451, 553–556 (2008).12. D. Nykypanchuk, M. M. Maye, D. van der Lelie, O. Gang, Nature

    451, 549–552 (2008).13. Q. Chen, S. C. Bae, S. Granick, Nature 469, 381–384 (2011).14. Y. Wang et al., Nature 491, 51–55 (2012).15. X. Ye et al., Nat. Chem. 5, 466–473 (2013).16. Materials and methods are available as supplementary

    materials on Science Online.17. J. J. Choi et al., J. Am. Chem. Soc. 133, 3131–3138 (2011).18. C. Zeng, C. Liu, Y. Chen, N. L. Rosi, R. Jin, J. Am. Chem. Soc.

    138, 8710–8713 (2016).19. G. A. DeVries et al., Science 315, 358–361 (2007).20. C. A. Hunter, J. K. M. Sanders, J. Am. Chem. Soc. 112,

    5525–5534 (1990).

    21. M. Nishio, Phys. Chem. Chem. Phys. 13, 13873–13900 (2011).22. R. J. Macfarlane et al., Science 334, 204–208 (2011).23. W. D. Luedtke, U. Landman, J. Phys. Chem. 100, 13323–13329 (1996).24. R. L. Whetten et al., Acc. Chem. Res. 32, 397–406 (1999).25. P. W. Anderson, Science 177, 393–396 (1972).26. Q.-F. Sun et al., Science 328, 1144–1147 (2010).27. C. L. Cleveland, U. Landman, J. Chem. Phys. 94, 7376 (1991).28. P. D. Jadzinsky, G. Calero, C. J. Ackerson, D. A. Bushnell,

    R. D. Kornberg, Science 318, 430–433 (2007).29. Y. Chen et al., J. Am. Chem. Soc. 137, 10076–10079 (2015).30. C. Zeng et al., Sci. Adv. 1, e1500045 (2015).

    ACKNOWLEDGMENTS

    R.J. thanks the Air Force Office of Scientific Research [awardno. FA9550-15-1-9999 (FA9550-15-1-0154)] and the Camille

    Dreyfus Teacher-Scholar Awards Program for financial support. Alldata are available in the main text and in the supplementarymaterials. The x-ray crystallographic coordinates have beendeposited in Cambridge Crystallographic Data Centre with CCDCnumber 1511348.

    SUPPLEMENTARY MATERIALS

    www.sciencemag.org/content/354/6319/1580/suppl/DC1Materials and MethodsSupplementary TextFigs. S1 to S11Tables S1 and S2

    22 September 2016; accepted 16 November 201610.1126/science.aak9750

    MIGRATION

    Mass seasonal bioflows of high-flyinginsect migrantsGao Hu,1,2,3* Ka S. Lim,2 Nir Horvitz,4 Suzanne J. Clark,2 Don R. Reynolds,5

    Nir Sapir,6 Jason W. Chapman2,3*

    Migrating animals have an impact on ecosystems directly via influxes of predators, prey,and competitors and indirectly by vectoring nutrients, energy, and pathogens. Althoughlinkages between vertebrate movements and ecosystem processes have been established,the effects of mass insect “bioflows” have not been described. We quantified biomassflux over the southern United Kingdom for high-flying (>150 meters) insects and show that~3.5 trillion insects (3200 tons of biomass) migrate above the region annually. These flowsare not randomly directed in insects larger than 10 milligrams, which exploit seasonallybeneficial tailwinds. Large seasonal differences in the southward versus northward transferof biomass occur in some years, although flows were balanced over the 10-year period. Ourlong-term study reveals a major transport process with implications for ecosystemservices, processes, and biogeochemistry.

    Latitudinal migrations of vast numbers offlying insects, birds, and bats (1–7) lead tohuge seasonal exchanges of biomass andnutrients across the Earth’s surface (8–11).Becausemanymigrant species (particularly

    insects) are extremely abundant (1, 5), seasonalmigrations may profoundly affect communitiesthrough predation and competition while trans-ferring enormous quantities of energy, nutrients,propagules, pathogens, and parasites betweenregions, with substantial effects on essential eco-system services, processes, and biogeochemistry(8–11), and, ultimately, ecosystem function.Latitudinal bird migrations are well charac-

    terized; for example, 2.1 billion passerinesmigrateannually between Europe and Africa (2), inte-gratingmultisensorynavigational information (12),exploiting favorable winds and adopting adaptive

    flight behaviors (13). By comparison, even thoughinsect migration surpasses all other aerial migra-tory phenomena in terms of sheer abundance (1),latitudinal insect migration is largely unquan-tified, in particular for themajority of species thatmigrate hundreds of meters above the ground (5).Specialized radar techniques are required to studythese high-flying insect migrants, as they are toosmall to carry transmitters or to be observed byany other means (14). Until now, radar studieshave been aimed almost exclusively at quantifyingmigrations of relatively few nocturnal species ofagricultural pests (3).We quantified annual abundance and biomass

    of three size categories of diurnal and nocturnalinsects migrating above an area of ~70,000 km2

    of the southern United Kingdom (Fig. 1A), be-tween 150 and 1200 m above ground level (agl)(Fig. 1B), from 2000 to 2009 (15). Abundance andbiomass values for medium (10 to 70 mg) andlarge insects (70 to 500 mg) (referred to collect-ively as “larger insects”) were calculated frommeasurements of >1.8 million individuals (tableS1) detected by vertical-looking entomologicalradars (VLRs) located in the southern UnitedKingdom (Fig. 1A). The VLRs provide a range ofinformation—including bodymass, flight altitude,aerial density, displacement speed, displacementdirection, and flight heading—for all individual

    insects of >10-mg body mass that fly through thevertically pointing beam within the altitude rangeof 150 to 1200 m agl (14). Annual abundance andbiomass values for larger insects migrating overthe study area were extrapolated from the aerialdensities and body masses recorded above theVLR locations (15). The third size category, smallinsects (70%of that biomass was from migration that oc-curred during daytime (Fig. 1C and table S2).Numerically, >99% of individuals were small in-sects; although the 15 billionmediumand 1.5 billionlarge insects made up only 0.4% and 0.05% ofthe annual abundance (table S2), they accountedfor a substantial proportion of the biomass: 12%(380 tons) and 7% (225 tons), respectively (table S3).By analyzing 1320 daytime “mass migrations”

    (15) involving 1.25 million VLR-detected insectsand 898 nocturnal mass migrations involving126,000 insects (table S1), we characterized mi-gration directions of the larger insects during“spring” (May to June), “summer” (July) and “fall”(August to September) (fig. S2 and table S4).Although high-altitude winds blew consistentlytoward the northeast or east in all three seasons(Rayleigh tests; daytime: spring, 60°; summer,66°; fall, 84°; nighttime: spring, 69°; summer, 81°;fall, 101°) (Fig. 2A and table S5), mass migrationsof larger insects did not simply move with theprevailing southwesterly winds. During the spring,mass migrations were consistently toward thenorth (Rayleigh tests; daytime: medium, 333°;large, 329°; nighttime:medium, 349°; large, 349°)(Fig. 2A), and this indicates that migration oc-curred on winds with a significantly more south-erly component than prevailing winds (Watson-Wheeler tests; P < 0.0001 in all cases) (table S5).

    1584 23 DECEMBER 2016 • VOL 354 ISSUE 6319 sciencemag.org SCIENCE

    1College of Plant Protection, Nanjing Agricultural University,Nanjing, China. 2Rothamsted Research, Harpenden,Hertfordshire, UK. 3Centre for Ecology and Conservation, andEnvironment and Sustainability Institute, University of Exeter,Penryn, Cornwall, UK. 4Movement Ecology Laboratory,Department of Ecology, Evolution, and Behavior, The HebrewUniversity, Jerusalem, Israel. 5Natural Resources Institute,University of Greenwich, Chatham, Kent, UK. 6Animal FlightLaboratory, Department of Evolutionary and EnvironmentalBiology, University of Haifa, Haifa, Israel.*Corresponding author. Email: [email protected] (G.H.);[email protected] (J.W.C.)

    RESEARCH | REPORTSon June 9, 2021

    http://science.sciencemag.org/

    Dow

    nloaded from

    http://www.sciencemag.org/content/354/6319/1580/suppl/DC1http://science.sciencemag.org/

  • Emergence of hierarchical structural complexities in nanoparticles and their assemblyChenjie Zeng, Yuxiang Chen, Kristin Kirschbaum, Kelly J. Lambright and Rongchao Jin

    DOI: 10.1126/science.aak9750 (6319), 1580-1584.354Science

    , this issue p. 1580Sciencearrangements that reversed between layers.structure. A hierarchy of interparticle ligand interactions controlled the packing, including sets of chiral packingnearly spherical nanoparticles did not pack into a cubic arrangement but instead formed a lower-symmetry monoclinic

    synthesized nanoparticles with a 246-atom gold core surrounded by 80 4-methylbenzenethiol ligands. Theseet al.Zeng The crystals of a well-defined ligand-covered gold nanoparticle can reveal how packing into a lattice happens.

    Probing packing rules

    ARTICLE TOOLS http://science.sciencemag.org/content/354/6319/1580

    MATERIALSSUPPLEMENTARY http://science.sciencemag.org/content/suppl/2016/12/21/354.6319.1580.DC1

    REFERENCES

    http://science.sciencemag.org/content/354/6319/1580#BIBLThis article cites 29 articles, 10 of which you can access for free

    PERMISSIONS http://www.sciencemag.org/help/reprints-and-permissions

    Terms of ServiceUse of this article is subject to the

    is a registered trademark of AAAS.ScienceScience, 1200 New York Avenue NW, Washington, DC 20005. The title (print ISSN 0036-8075; online ISSN 1095-9203) is published by the American Association for the Advancement ofScience

    Copyright © 2016, American Association for the Advancement of Science

    on June 9, 2021

    http://science.sciencemag.org/

    Dow

    nloaded from

    http://science.sciencemag.org/content/354/6319/1580http://science.sciencemag.org/content/suppl/2016/12/21/354.6319.1580.DC1http://science.sciencemag.org/content/354/6319/1580#BIBLhttp://www.sciencemag.org/help/reprints-and-permissionshttp://www.sciencemag.org/about/terms-servicehttp://science.sciencemag.org/