当前位置:高等教育资讯网  >  中国高校课件下载中心  >  大学文库  >  浏览文档

复旦大学:《纳米线材料和功能器件》课外阅读内容_模板法合成和自组装方法_1_Murray_Nature06_Structural diversity in binary NP superlattice

资源类别:文库,文档格式:PDF,文档页数:5,文件大小:600.33KB,团购合买
点击下载完整版文档(PDF)

Vol 439 5 January 2006 doi: 10. 1038/nature04414 nature LETTERS Structural diversity in binary nanoparticle superlattices Elena V. Shevchenko 2*+, Dmitri V. Talapin *t, Nicholas A Kotov, Stephen O'Brien& Christopher B. Murray Assembly of small building blocks such as atoms, molecules anoparticles of different materials(Fig. 1). Coherently packed noparticles into macroscopic structures-that is, bottom domains extend up to 10 um in lateral dimensions, and can displa ssembly-is a theme that runs through chemistry, biology and well defined facets(Supplementary Fig. 1). In many cases, several aterial science. Bacteria, macromolecules and nanoparticles' BNSL structures form simultaneously on the same substrate, under can self-assemble, generating ordered structures with a precision identical experimental conditions. The same nanoparticle mixture that challenges current lithographic techniques. The assembly of can assemble into BNSLs with very different stoichiometry and nanoparticles of two different materials into a binary nanoparticle packing symmetry. For example, 11 different BNSL structures were to a large variety of materials(metamaterials) with precisely nanoparticles(Supplementary Fig. 2). We also observe that, in controlled chemical composition and tight placement of the general, BNSLs tolerate much broader y ranges than hard sphere omponents Maximization of the nanoparticle packing density example, AlB2-type BNSLs assembled from different comb has been proposed as the driving force for BNSL formation, nations of PbSe, PbS, Au, Ag, Pd, Fe2O3, CoPt3 and Bi nanoparticles and only a few BNSL structures have been predicted to be n a broad y range( Supplementary Fig. 3). Further, we observe thermodynamically stable. Recently, colloidal crystals with BNSLs that could not be identified as isostructural with specific micrometre-scale lattice spacings have been grown from oppo- intermetallic compounds( Supplementary Fig. 2). This observed itely charged polymethyl methacrylate spheres,. Here we structural diversity of BNSLs defies traditional expectations, and demonstrate formation of more than 15 different BNSL struc- shows the great potential of modular self-assembly at the nanoscale tures, using combinations of semiconducting, metallic and mag. The formation of binary structures with packing density signifi netic nanoparticle building blocks. At least ten of these colloidal cantly lower than the density of single-phase f.c.c. close packing particles determine BNSLstod: ses on sterically stabilized nano- ordering. Moreover, van der Waals, steric or de e for nanoparticle crystalline structures have not been reported previously. We (0.7405)rules out entropy as the main driving force for nanopartic demonstrate that electrical char rticle iometry; additional contributions interactions are not sufficient to explain why these low density BNSLs der Waals, steric and dipolar forces stabilize form, instead of their constituents separating into single-compo the variety of BNSL structures. uperlattices. Opposite electrical charges on nanoparticles could Face-centred-cubic(fcc )ordering of monodisperse hard spheres impart a specific affinity of one type of particle(for example, dispersed in a liquid permits larger local free space available for each dodecanethiol-capped Au, Ag, Pd)for another (typically PbSe sphere compared to the unstructured phase, resulting in higher PbS, Fe2O3, CoPt, and so on, capped with long chain carboxylic translational entropy of the spheres. When the volume fraction of acids). If nanoparticles are oppositely charged, the Coulomb hard spheres approaches -55%, this ordering enhances the total potential would stabilize the BNSl while destabilizing the single- entropy of the system and drives the ordering phase transition. component superlattices. The electrical charges might be Entropy-driven crystallization has been studied in great detail both present on sterically stabilized nanoparticles even in non-polar theoretically and experimentally on monodisperse latex particles, solvents whose behaviour can be approximated by hard spheres 4. In a charges on the nanoparticles that form our BNSI mixture containing spheres of two different sizes(radi Rsmall and studied the electrophoretic mobility of PbSe and Au nanocrystals Rlarge ) the packing symmetry depends on the size ratio of the small Laser Doppler velocimetry allows the distribution of electrophoretic assembly of hard spheres into binary superlattices isostructural electrical charge(Z, in units of e)of a spherical particle in a low ith NaCl, AlB 2 and NaZn13 can be driven by entropy alone without dielectric solvent in absence of electrolyte can be calculated from the any specific energetic interactions between the spheres..Indeed, electrophoretic mobility (ue) where ue= Ze/(rna), n is the viscosity NaZn13- and AlB2-type assemblies of silica particles were found in of the solvent and a is the hydrodynamic diameter of a particle". with natural Brazilian opals and can be grown from latex spheres"?. In aa=10 nm, we obtain He 0.27X10-4Zcm2V-Is-1. These calcu certain y range, the packing density of these structures either exceeds lated values agree well with the peaks in the experimental mobili or is very close to the density of the close-packed f.c. c. lattice distribution for 7.2-nm-diameter PbSe nanocrystals in chloroform (0.7405), while structures with lower packing densities are predicted ( Fig. 2a). Owing to the organic coat (oleic acid), the effective to be unstable, Is hydrodynamic radius of PbSe nanocrystals extends beyond the Despite these predictions, we observed an amazing variety of crystalline core by 1-2 nm, depending on the density of surface BNSLs that self-assemble from colloidal solutions of nearly spherical coverage. The peaks in the mobility distribution curve indicate the arbor, Michigan 48109, USA +Present ading. 500 West 120th Street, New York, New York 10027, USA. Department of Chemical University of Michigan, Ann dress: The Molecular Foundry, Lawrence Berkeley National Laboratory, Berkeley, California 2006 Nature Publishing Group

© 2006 Nature Publishing Group Structural diversity in binary nanoparticle superlattices Elena V. Shevchenko1,2*†, Dmitri V. Talapin1 *†, Nicholas A. Kotov3 , Stephen O’Brien2 & Christopher B. Murray1 Assembly of small building blocks such as atoms, molecules and nanoparticles into macroscopic structures—that is, ‘bottom up’ assembly—is a theme that runs through chemistry, biology and material science. Bacteria1 , macromolecules2 and nanoparticles3 can self-assemble, generating ordered structures with a precision that challenges current lithographic techniques. The assembly of nanoparticles of two different materials into a binary nanoparticle superlattice (BNSL)3–7 can provide a general and inexpensive path to a large variety of materials (metamaterials) with precisely controlled chemical composition and tight placement of the components. Maximization of the nanoparticle packing density has been proposed as the driving force for BNSL formation3,8,9, and only a few BNSL structures have been predicted to be thermodynamically stable. Recently, colloidal crystals with micrometre-scale lattice spacings have been grown from oppo￾sitely charged polymethyl methacrylate spheres10,11. Here we demonstrate formation of more than 15 different BNSL struc￾tures, using combinations of semiconducting, metallic and mag￾netic nanoparticle building blocks. At least ten of these colloidal crystalline structures have not been reported previously. We demonstrate that electrical charges on sterically stabilized nano￾particles determine BNSL stoichiometry; additional contributions from entropic, van der Waals, steric and dipolar forces stabilize the variety of BNSL structures. Face-centred-cubic (f.c.c.) ordering of monodisperse hard spheres dispersed in a liquid permits larger local free space available for each sphere compared to the unstructured phase, resulting in higher translational entropy of the spheres. When the volume fraction of hard spheres approaches ,55%, this ordering enhances the total entropy of the system and drives the ordering phase transition. Entropy-driven crystallization has been studied in great detail both theoretically12 and experimentally on monodisperse latex particles, whose behaviour can be approximated by hard spheres13,14. In a mixture containing spheres of two different sizes (radii Rsmall and Rlarge), the packing symmetry depends on the size ratio of the small and large spheres (g ¼ Rsmall/Rlarge)3,8. Calculations show that assembly of hard spheres into binary superlattices isostructural with NaCl, AlB2 and NaZn13 can be driven by entropy alone without any specific energetic interactions between the spheres9,15. Indeed, NaZn13- and AlB2-type assemblies of silica particles were found in natural Brazilian opals16 and can be grown from latex spheres17. In a certain g range, the packing density of these structures either exceeds or is very close to the density of the close-packed f.c.c. lattice (0.7405), while structures with lower packing densities are predicted to be unstable8,15. Despite these predictions, we observed an amazing variety of BNSLs that self-assemble from colloidal solutions of nearly spherical nanoparticles of different materials (Fig. 1). Coherently packed domains extend up to 10 mm in lateral dimensions, and can display well defined facets (Supplementary Fig. 1). In many cases, several BNSL structures form simultaneously on the same substrate, under identical experimental conditions. The same nanoparticle mixture can assemble into BNSLs with very different stoichiometry and packing symmetry. For example, 11 different BNSL structures were prepared from the same batches of 6.2 nm PbSe and 3.0 nm Pd nanoparticles (Supplementary Fig. 2). We also observe that, in general, BNSLs tolerate much broader g ranges than hard spheres: for example, AlB2-type BNSLs assembled from different combi￾nations of PbSe, PbS, Au, Ag, Pd, Fe2O3, CoPt3 and Bi nanoparticles in a broad g range (Supplementary Fig. 3). Further, we observe BNSLs that could not be identified as isostructural with specific intermetallic compounds (Supplementary Fig. 2). This observed structural diversity of BNSLs defies traditional expectations, and shows the great potential of modular self-assembly at the nanoscale. The formation of binary structures with packing density signifi- cantly lower than the density of single-phase f.c.c. close packing (0.7405) rules out entropy as the main driving force for nanoparticle ordering. Moreover, van der Waals, steric or dipolar interparticle interactions are not sufficient to explain why these low density BNSLs form, instead of their constituents separating into single-component superlattices. Opposite electrical charges on nanoparticles could impart a specific affinity of one type of particle (for example, dodecanethiol-capped Au, Ag, Pd) for another (typically PbSe, PbS, Fe2O3, CoPt3 and so on, capped with long chain carboxylic acids). If nanoparticles are oppositely charged, the Coulomb potential would stabilize the BNSL while destabilizing the single￾component superlattices. The electrical charges might be present on sterically stabilized nanoparticles even in non-polar solvents18–20. To measure charges on the nanoparticles that form our BNSL, we studied the electrophoretic mobility of PbSe and Au nanocrystals. Laser Doppler velocimetry allows the distribution of electrophoretic mobilities within an ensemble of nanoparticles to be measured. The electrical charge (Z, in units of e) of a spherical particle in a low dielectric solvent in absence of electrolyte can be calculated from the electrophoretic mobility (me) where me ¼ Ze=ð3phaÞ; h is the viscosity of the solvent and a is the hydrodynamic diameter of a particle21. With a ¼ 10 nm, we obtain me < 0:27 £ 1024Z cm2 V21 s 21: These calcu￾lated values agree well with the peaks in the experimental mobility distribution for 7.2-nm-diameter PbSe nanocrystals in chloroform (Fig. 2a). Owing to the organic coat (oleic acid), the effective hydrodynamic radius of PbSe nanocrystals extends beyond the crystalline core by 1–2 nm, depending on the density of surface coverage. The peaks in the mobility distribution curve indicate the LETTERS 1 IBM Research Division, T. J. Watson Research Center, 1101 Kitchawan Road, Yorktown Heights, New York 10598, USA. 2 Department of Applied Physics & Applied Mathematics, Columbia University, 200 SW Mudd Building, 500 West 120th Street, New York, New York 10027, USA. 3 Department of Chemical Engineering, University of Michigan, Ann Arbor, Michigan 48109, USA. †Present address: The Molecular Foundry, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA. *These authors contributed equally to this work. Vol 439|5 January 2006|doi:10.1038/nature04414 55

LETTERS NATURETVol 439 5 January 2006 presence of particles with charges -e, 0, e and 2e in a colloidal charged PbSe nanocrystals and reduces the concentration of pos olution of monodisperse PbSe nanocrystals tively charged nanocrystals(Fig. 2d). Surveys of man We found that the charges on PbSe nanocrystals can be altered revealed that the additives reliably shifted the distribution of charge adding surfactant molecules like carboxylic acids and tri-n- states; however, the initial proportion of particles in each charge state alkylphosphine oxides. Addition of oleic acid increases the popu- was dependent somewhat on sample processing. Both neutral and lation of positively charged PbSe nanocrystals at the expense of the negatively charged nanoparticles were detected in chloroform solt negatively charged and neutral nanocrystals. Depending on the tions of 4.8 nm dodecanethiol-capped Au nanocrystals(Supplemen- amount of acid added, the majority of nanocrystals can be adjusted tary Fig. 5). After addition of oleic acid most Au nanoparticles to have either one or two positive charges(Fig 2b and c). Addition of become negatively charged (Fig. 2e), whereas the addition of TOPO oleic acid increases the solutions viscosity, causing the peaks to shift neutralizes the Au nanoparticles(Fig. 2f). The charges on PbSe and towards lower mobility(compare Fig. 2a-c). The addition of tri-n- Au nanoparticles could originate from deviations in nanocrystal octylphosphine oxide(TOPO)increases the population of negatively stoichiometry and adsorption/desorption of charged capping e 00} 100 k Figure 1 I TEM images of the characteristic projections of the binary g, 7.2 nm PbSe and 4.2 nm Ag: h, 6.2 nm PbSe and 3.0nm Pd; i, 7. 2 nm PbSe superlattices, self-assembled from different and 5.0nm Au; 3, 5.8 nm PbSe and 3.0 nm Pd; k, 7.2 bSe and 4.2 nm Ag nit cells of the corresponding three- nd 1, 6.2 nm PbSe and 3.0nm Pd nanoparticles Scale bars: a-c, e, f, perlattices are assembled from a, 13.4 nm y-Fe,O, and 5.0 nm Au; 20 nm; d, g, h, 10 nm. The lattice projection is labelled in each panel abov b, 7.6nm PbSe and 50nm Au; c, 6.2 nm PbSe and 3.0 nm Pd; d, 6.7 nm Pbs the scale bar. The modelled projections of the binary superlattices are shown and 3.0 nm Pd; e, 6.2 nm PbSe and 3.0 nm Pd; f, 5.8 nm PbSe and 3.0 nm Pd; in Supplementary Fig 4 2006 Nature Publishing Group

© 2006 Nature Publishing Group presence of particles with charges 2e, 0, e and 2e in a colloidal solution of monodisperse PbSe nanocrystals. We found that the charges on PbSe nanocrystals can be altered by adding surfactant molecules like carboxylic acids and tri-n￾alkylphosphine oxides. Addition of oleic acid increases the popu￾lation of positively charged PbSe nanocrystals at the expense of the negatively charged and neutral nanocrystals. Depending on the amount of acid added, the majority of nanocrystals can be adjusted to have either one or two positive charges (Fig. 2b and c). Addition of oleic acid increases the solutions’ viscosity, causing the peaks to shift towards lower mobility (compare Fig. 2a–c). The addition of tri-n￾octylphosphine oxide (TOPO) increases the population of negatively charged PbSe nanocrystals and reduces the concentration of posi￾tively charged nanocrystals (Fig. 2d). Surveys of many samples revealed that the additives reliably shifted the distribution of charge states; however, the initial proportion of particles in each charge state was dependent somewhat on sample processing. Both neutral and negatively charged nanoparticles were detected in chloroform solu￾tions of 4.8 nm dodecanethiol-capped Au nanocrystals (Supplemen￾tary Fig. 5). After addition of oleic acid most Au nanoparticles become negatively charged (Fig. 2e), whereas the addition of TOPO neutralizes the Au nanoparticles (Fig. 2f). The charges on PbSe and Au nanoparticles could originate from deviations in nanocrystal stoichiometry and adsorption/desorption of charged capping Figure 1 | TEM images of the characteristic projections of the binary superlattices, self-assembled from different nanoparticles, and modelled unit cells of the corresponding three-dimensional structures. The superlattices are assembled from a, 13.4 nm g-Fe2O3 and 5.0 nm Au; b, 7.6 nm PbSe and 5.0 nm Au; c, 6.2 nm PbSe and 3.0 nm Pd; d, 6.7 nm PbS and 3.0 nm Pd; e, 6.2 nm PbSe and 3.0 nm Pd; f, 5.8 nm PbSe and 3.0 nm Pd; g, 7.2 nm PbSe and 4.2 nm Ag; h, 6.2 nm PbSe and 3.0 nm Pd; i, 7.2 nm PbSe and 5.0 nm Au; j, 5.8 nm PbSe and 3.0 nm Pd; k, 7.2 nm PbSe and 4.2 nm Ag; and l, 6.2 nm PbSe and 3.0 nm Pd nanoparticles. Scale bars: a–c, e, f, i–l, 20 nm; d, g, h, 10 nm. The lattice projection is labelled in each panel above the scale bar. The modelled projections of the binary superlattices are shown in Supplementary Fig. 4. LETTERS NATURE|Vol 439|5 January 2006 56

NATUREIVol 439 5 January 2006 LETTERS ligands. Although these additives are effective in adjusting the particle example, ABA, ABS, AB6, AB13)might form when both charged and charge states, the specific interactions by which this charge tuning neutral nanoparticles of type B are incorporated into the structures. occurs will require further study( Supplementary Discussion 1 The presence of differently charged nanoparticles in the colloidal In the presence of oleic acid, PbSe and Au nanoparticles are solutions(Fig. 2a and Supplementary Fig. 5)could also contribute to oppositely charged(Fig. 2e). The Coulomb potential between two the simultaneous formation of different BNSLs Intentional addition oppositely charged nanoparticles(Z==1)separated by 10 nm of a of a large concentration of charged species into a solution of solvent like chloroform is comparable with kTat room temperature, nanoparticles might reduce the Debye screening length down to and solutions of mixed PbSe and metal nanoparticles retain stability Ro, relaxing the strict rules for BNSL charge neutrality and allowing a for several weeks. The relatively small interparticle potential range of new structures to be formedto. favours annealing of the BNSLs as they grow. For a NaCl-type Tuning the charge state of the nanoparticles allows us to direct the BNSL with Z+=1, Z-=-l and the nearest-neighbour distance self-assembly process. Reproducible switching between different Ro=11.5nm(Fig 1a), the Coulomb binding energy per unit cell is BNSL structures has been achieved by adding small amounts of estimated to be UCoul mz 4tEEORo)N-0 1ev (or about carboxylic acids, TOPO or dodecylamine to colloidal solutions of 4kT at the superlattice growth temperature, 50C), where PbSe(PbS, Fe203, and so on) and metal(Au, Ag, Pd)nanocrystal M=-17476 is the Madelung constant. The Coulomb binding Figure 3 demonstrates how these additives direct the formation of energy is comparable to the van der Waals attractive energy expected specific BNSL structures Combining native solutions of 6.2 nm PbSe forces(1/R)can rival long-range Coulomb energy(1/R)only at results in the formation of several BNSL structures with MgZn2 and the nanometre scale. In BNSLs, we can neglect screening of the cuboctahedral ABi3 lattices dominating. However, the same nano- Coulomb potential by charged species in solution because the Debye particles assemble into orthorhombic AB- and AlBz-type super screening length(-10"cm)is much larger than Ro(refs 10, 11). In lattices after adding oleic acid(Fig. 3a),and into NaZn13-or an ABx BNSL where A and B hold opposite charges, the Coulom potential per ABx molecule' is UCoul = -a+B(xZ-+Z+N2/3 here a and B are positive constants and N is the number of TOPO assembled nanoparticles( Supplementary Discussion 2) Coulomb energy determines the stoichiometry of the growing BNSL.An extended three-dimensional BNSL can form only if the positive and negative charges compensate each other. If during E growth the BNSL accumulates non-compensated charge, eventually m UCoul changes sign from negative to positive and the growth is self limiting. The superlattice nucleation stage should be less sensitive to 8 the Coulomb interactions. Indeed, we observed that many small domains with different BNSL structures can simultaneously nucleate E one or two structures grow to larger length scales 6 on the same substrate, but their size does not exceed -10- nano- (-106-10 particles). BNSLs with many particles per unit cell (for o1 dig nm d 0.8-0.40 0.4 Figure 2 I Electrophoretic mobility of PbSe and Au nanocrystals in 20m chloroform. a-d, Distribution of electrophoretic mobility for 7. 2nm PbSe tals a, PbSe nanocrystals washed to remove excess of cappin Figure 3 TEM images of binary superlattices self-assembled in the inds. The grey bars show mobilities predicted for nanocrystals with ence of 4 mM oleic acid (left column) and 6 mM tri-n-octylphosphi arges of-1, 0, I and 2(in units of e). b-d, Electrophoretic mobility of oxide, TOPo (right column). a, 6.2 nm PbSe and 3.0 nm Pd PbSe nanocrystals in the presence of b, 0.02 M oleic acid, c, 0.06 M oleic acid anopartides self-assembled into orthorhombic AB- and AlB,type BNSls, and d, 0.05 M tri-n-octylphosphine oxide. e, f, Comparison of nd b, into NaZn3-type BNSL. c, d, 7.2 nm PbSe and 4.2 nm Ag electrophoretic mobilities of 7. 2 nm PbSe and 4.8 nm Au nanocrystals in the nanoparticles self-assembled into orthorhombic AB and cuboctahedral presence of e, 0.02 M oleic acid and f, 0.05 M tri-n-octylphosphine oxide, AB13 BNSLs, respectively. e, f, 6.2 nm PbSe and 5.0 nm Au nanoparticles self respectively. a u, arbitrary units assembled into CuAu-type and Ca Cus-type BNSLs, respectivel 2006 Nature Publishing Group

© 2006 Nature Publishing Group ligands. Although these additives are effective in adjusting the particle charge states, the specific interactions by which this charge tuning occurs will require further study (Supplementary Discussion 1). In the presence of oleic acid, PbSe and Au nanoparticles are oppositely charged (Fig. 2e). The Coulomb potential between two oppositely charged nanoparticles (Z ¼ ^1) separated by 10 nm of a solvent like chloroform is comparable with kTat room temperature, and solutions of mixed PbSe and metal nanoparticles retain stability for several weeks. The relatively small interparticle potential favours annealing of the BNSLs as they grow. For a NaCl-type BNSL with Zþ ¼ 1, Z 2 ¼ 21 and the nearest-neighbour distance R0 ¼ 11.5 nm (Fig. 1a), the Coulomb binding energy per unit cell is estimated to be UCoul < MZþZ2e2=ð4p110R0Þ < 20:1 eV (or about 24kT at the superlattice growth temperature, 50 8C), where M ¼ 21.7476 is the Madelung constant. The Coulomb binding energy is comparable to the van der Waals attractive energy expected for a NaCl-type BNSL. The energy of short-range van der Waals forces (,1/R6 ) can rival long-range Coulomb energy (,1/R) only at the nanometre scale. In BNSLs, we can neglect screening of the Coulomb potential by charged species in solution because the Debye screening length (,1024 cm) is much larger than R0 (refs 10, 11). In an ABx BNSL where A and B hold opposite charges, the Coulomb potential per ABx ‘molecule’ is UCoul < 2a þ bðxZ2 þ ZþÞ 2 N2=3; where a and b are positive constants and N is the number of assembled nanoparticles (Supplementary Discussion 2). Coulomb energy determines the stoichiometry of the growing BNSL. An extended three-dimensional BNSL can form only if the positive and negative charges compensate each other. If during growth the BNSL accumulates non-compensated charge, eventually UCoul changes sign from negative to positive and the growth is self￾limiting. The superlattice nucleation stage should be less sensitive to the Coulomb interactions. Indeed, we observed that many small domains with different BNSL structures can simultaneously nucleate on the same substrate, but their size does not exceed ,102 nano￾particles. Only one or two structures grow to larger length scales (,106 –108 particles). BNSLs with many particles per unit cell (for example, AB4, AB5, AB6, AB13) might form when both charged and neutral nanoparticles of type B are incorporated into the structures. The presence of differently charged nanoparticles in the colloidal solutions (Fig. 2a and Supplementary Fig. 5) could also contribute to the simultaneous formation of different BNSLs. Intentional addition of a large concentration of charged species into a solution of nanoparticles might reduce the Debye screening length down to R0, relaxing the strict rules for BNSL charge neutrality and allowing a range of new structures to be formed10. Tuning the charge state of the nanoparticles allows us to direct the self-assembly process. Reproducible switching between different BNSL structures has been achieved by adding small amounts of carboxylic acids, TOPO or dodecylamine to colloidal solutions of PbSe (PbS, Fe2O3, and so on) and metal (Au, Ag, Pd) nanocrystals. Figure 3 demonstrates how these additives direct the formation of specific BNSL structures. Combining native solutions of 6.2 nm PbSe and 3.0 nm Pd nanoparticles (particle concentration ratio ,1:5) results in the formation of several BNSL structures with MgZn2 and cuboctahedral AB13 lattices dominating. However, the same nano￾particles assemble into orthorhombic AB- and AlB2-type super￾lattices after adding oleic acid (Fig. 3a), and into NaZn13- or Figure 2 | Electrophoretic mobility of PbSe and Au nanocrystals in chloroform. a–d, Distribution of electrophoretic mobility for 7.2 nm PbSe nanocrystals. a, PbSe nanocrystals washed to remove excess of capping ligands. The grey bars show mobilities predicted for nanocrystals with charges of 21, 0, 1 and 2 (in units of e). b–d, Electrophoretic mobility of PbSe nanocrystals in the presence of b, 0.02 M oleic acid, c, 0.06 M oleic acid and d, 0.05 M tri-n-octylphosphine oxide. e, f, Comparison of electrophoretic mobilities of 7.2 nm PbSe and 4.8 nm Au nanocrystals in the presence of e, 0.02 M oleic acid and f, 0.05 M tri-n-octylphosphine oxide, respectively. a.u., arbitrary units. Figure 3 | TEM images of binary superlattices self-assembled in the presence of 4 mM oleic acid (left column) and 6 mM tri-n-octylphosphine oxide, TOPO (right column). a, 6.2 nm PbSe and 3.0 nm Pd nanoparticles self-assembled into orthorhombic AB- and AlB2-type BNSLs, and b, into NaZn13-type BNSL. c, d, 7.2 nm PbSe and 4.2 nm Ag nanoparticles self-assembled into orthorhombic AB and cuboctahedral AB13 BNSLs, respectively. e,f, 6.2 nm PbSe and 5.0 nm Au nanoparticles self￾assembled into CuAu-type and CaCu5-type BNSLs, respectively. NATURE|Vol 439|5 January 2006 LETTERS 57

LETTERS NATURETVol 439 5 January 2006 vA 影 Figure 4 I TEM images and proposed unit cells of binary superlattices and 6.2 nm PbSe nanocrystals. The insets show a, a magnified image, and b self-assembled from triangular nanoplates and spherical nanoparticles c, proposed unit cells of the corresponding superlattices. The structure a, b, Self-assembled from LaF, triangular nanoplates(9.0 nm side)and hown in a forms on silicon oxide surfaces while structures shown in b andc 5.0 nm Au nanoparticles; c, self-assembled from LaF, triangular nanoplates form preferentially on amorphous carbon substrates. uboctahedral ABl3-type BNSLs after the addition of dodecylami DDAB). For synthesis of 5.0 nm Au and 4.2 nm or TOPO, respectively(Fig. 3b). In the ABu3-type BNSL, metal 0.034g AuCIy and 0.025 g AgNOs, respectively, and 0.0925g DDAB 3.0 nm Pd particles assemble into icosahedral(NaZn13)or cuboctahedral nanocrystals were synthesized from 0.0237g Pddl with 0. 157g DDAB. Forty microlitres of a 9.4 M aqueous solution of NaBH4 were added drop-wise to the rounded by 24 metal spheres at the vertices of a snub cube?. In the solution of metal salt with vigorous stirring. After 20 min, 0. 8ml l-dodecanethiol presence of TOPO the metal nanoparticles are neutral (Fig. 2f), were precipitated by adding ethanol, and the solid redispersed in 10 ml toluene in favouring formation of the Pd13(Au13, Ag13)clusters. The clusters of presence of 0. 8 ml 1-dodecanethiol and refluxed for 30 min metal nanoparticles in turn provide screening of the charges on PbSe Fe203 nanocrystals were synthesized by methods adapted from ref 28. Briefly, nanocrystals in the ABi3-type BNSL. Surveys of many samples show that the addition of a carboxylic pentacarbonyl into 10 ml trioctylamine in the presence of 0.65g oleic acid at acid to solutions of PbSe-Pd, PbSe-Au, PbSe-Ag and PbSe-Fe203 270C and 250%C, respectively. After heating of the reaction mixtures at 320C for nanoparticle mixtures results in either AB or AB2 superlattices I h, the reaction mixture was cooled to room temp 0. 17 g trimethylamine Fig 3c, e), whereas the addition of toPo to mixtures of the same N-oxide was added to oxidize the iron nanoparticles to y-Fe2Os, and the reaction (if y>-065)BNSLs(Fig. 3d, f). Thus the space-filling principles Pbse, PbS and LaFs nanocrystals can be found in refs7,29 and30,respectively: and particle charging work in combination to determine the struc- silicon oxide-coated transmission electron microscope(TEM)grid, a silicon cure. Adjusting the relative concentrations of A and B particles can be nitride membrane or an alkyl-functionalized silicon chip) was placed in a glass used as an additional tool with which to control the BNSL structure. vial containing a colloidal solution of nanoparticles. The vial was placed tilted by A:B ratio is" l , tesence of ToPO, AB, BNSLs can form when the 60-70 inside a low-pressure chamber. Ordered binary assemblies formed upon large excess B particles. hylene or chloroform were used as solvents(-1: I by volume). The best binary In contrast to particles with amorphous or polycrystalline mor- assemblies(as determined by the length scale of ordering and a low occurrence of phology, nanocrystals allow exploitation of the inherent crystal defec) under reduced pressure(3.2 kl ely concentrated colloidal solutions were obtained by evaporating relati shape can in turn be used as a powerful tool to engineer the structure Structural analysis. A Philips CM12 TEMoperating at 120 kv was used to image self-assembled from LaFs triangular nanoplates and spherical Au or the crystal orientations, and recording a series of two-dimensional projections PbSe nanocrystals. In the Lafs-Au system, the LaF, nanoplates lie down the major symmetry axes. Tilting of the samples allowed observation of flat on silicon oxide surface(Fig. 4a) and stand on edge when additional orientations not expressed in the plan view images. To assign the assembled on amorphous carbon(Fig. 4b and c), demonstrating observed structures to crystallographic space groups, we built three-dimensional how the choice of substrate can be used to control the BNSL structure. lattice models for the 180 most common space groups using Accelrys M It is specifically at the nanoscale that the van der Waals, electro ro- Modelling 3.1 software. The TEM images ompared with simulated steric repulsion and the directional dipolar interactions can projections to match the symmetry of our superlattices. We also performed a ibute to the interparticle potential with comparable comparison of experimental small-angle electron difraction pattens taken 18, 3, 4,25. These, together with the effects of particle substrate larger and the two-dimensional Fourier transformation power spectra of interactions and space-filling(entropic) factors, combine to deter- real space TEM images and the fast Fourier transform power spectra of the mine the BNSL structure. The non-equilibrium nature of our Electrophoretic mobility measurements. These were performed by electro- evaporative self-assembly process adds additional complexity. Pre- phoretic light scattering using a Zetasizer Nano ZS Series (Malvern), allowing cise control of nanoparticle size, shape and composition allows us to measurements in non-polar organic solvents. We used chloroform solutions engineer electronic, optical and magnetic properties of nanoparticle with nanoparticle concentrations-5 times higher than those used for growing building blocks. Assembling these nanoscale building blocks into a binary superlattices. The concentrations of additives(oleic acid and TOPO)wer wide range of BNSL systems provides a powerful modular approach similar to those used for directing BNSL self-assembly. After preparation, the to the design of 'metamaterials' with programmable physical and colloidal solutions were left in the dark for several hours to allow the systems te chemica equilibrate before each measurement. METHODS Nanoparticle synthesis. Au, Ag and Pd nanoparticles were prepared by 1. Shenton, W Pum, D, Sleytr, U& Mann, S. Synthesis of cadmium sulphide modifying the method of ref. 27. Metal salts were dissolved in 10 ml of toluene superlattices using self-assembled bacterial S-layers. Nature 389, 585-58 with ultrasonication in the presence of dodecyldimethylammonium bromide 2006 Nature Publishing Group

© 2006 Nature Publishing Group cuboctahedral AB13-type BNSLs after the addition of dodecylamine or TOPO, respectively (Fig. 3b). In the AB13-type BNSL, metal particles assemble into icosahedral (NaZn13) or cuboctahedral (cuboctahedral AB13) clusters, with each large PbSe particle sur￾rounded by 24 metal spheres at the vertices of a snub cube7 . In the presence of TOPO the metal nanoparticles are neutral (Fig. 2f), favouring formation of the Pd13 (Au13, Ag13) clusters. The clusters of metal nanoparticles in turn provide screening of the charges on PbSe nanocrystals in the AB13-type BNSL. Surveys of many samples show that the addition of a carboxylic acid to solutions of PbSe–Pd, PbSe–Au, PbSe–Ag and PbSe–Fe2O3 nanoparticle mixtures results in either AB or AB2 superlattices (Fig. 3c, e), whereas the addition of TOPO to mixtures of the same nanoparticles favours growth of AB13 (if g , ,0.65) or AB5 (if g . ,0.65) BNSLs (Fig. 3d, f). Thus the space-filling principles and particle charging work in combination to determine the struc￾ture. Adjusting the relative concentrations of A and B particles can be used as an additional tool with which to control the BNSL structure. For example, in presence of TOPO, AB4 BNSLs can form when the A:B ratio is ,1:1, whereas exclusively AB13 forms in the presence of large excess B particles. In contrast to particles with amorphous or polycrystalline mor￾phology, nanocrystals allow exploitation of the inherent crystal anisotropy to precisely engineer nanocrystal shape22. The nanocrystal shape can in turn be used as a powerful tool to engineer the structure of the self-assembled BNSLs. For example, Fig. 4 shows several BNSLs self-assembled from LaF3 triangular nanoplates and spherical Au or PbSe nanocrystals. In the LaF3–Au system, the LaF3 nanoplates lie flat on silicon oxide surface (Fig. 4a) and stand on edge when assembled on amorphous carbon (Fig. 4b and c), demonstrating how the choice of substrate can be used to control the BNSL structure. It is specifically at the nanoscale that the van der Waals, electro￾static, steric repulsion and the directional dipolar interactions can contribute to the interparticle potential with comparable weight18,23,24,25. These, together with the effects of particle substrate interactions and space-filling (entropic) factors, combine to deter￾mine the BNSL structure. The non-equilibrium nature of our evaporative self-assembly process adds additional complexity26. Pre￾cise control of nanoparticle size, shape and composition allows us to engineer electronic, optical and magnetic properties of nanoparticle building blocks. Assembling these nanoscale building blocks into a wide range of BNSL systems provides a powerful modular approach to the design of ‘metamaterials’ with programmable physical and chemical properties. METHODS Nanoparticle synthesis. Au, Ag and Pd nanoparticles were prepared by modifying the method of ref. 27. Metal salts were dissolved in 10 ml of toluene with ultrasonication in the presence of dodecyldimethylammonium bromide (DDAB). For synthesis of 5.0 nm Au and 4.2 nm Ag nanoparticles, we used 0.034 g AuCl3 and 0.025 g AgNO3, respectively, and 0.0925 g DDAB. 3.0 nm Pd nanocrystals were synthesized from 0.0237 g PdCl2 with 0.157 g DDAB. Forty microlitres of a 9.4 M aqueous solution of NaBH4 were added drop-wise to the solution of metal salt with vigorous stirring. After 20 min, 0.8 ml 1-dodecanethiol was added and the stirring was continued for five more minutes. The nanoparticles were precipitated by adding ethanol, and the solid redispersed in 10 ml toluene in the presence of 0.8 ml 1-dodecanethiol and refluxed for 30 min under nitrogen. Fe2O3 nanocrystals were synthesized by methods adapted from ref. 28. Briefly, 11 nm and 13.4 nm Fe2O3 nanocrystals were synthesized by injecting 0.2 ml iron pentacarbonyl into 10 ml trioctylamine in the presence of 0.65 g oleic acid at 270 8C and 250 8C, respectively. After heating of the reaction mixtures at 320 8C for 1 h, the reaction mixture was cooled to room temperature. 0.17 g trimethylamine N-oxide was added to oxidize the iron nanoparticles to g-Fe2O3, and the reaction mixture was heated to 130 8C for 1.5 h and 320 8C for 1 h. Details of the synthesis of PbSe, PbS and LaF3 nanocrystals can be found in refs 7, 29 and 30, respectively. Preparation of binary superlattices. A substrate (for example, a carbon- or silicon oxide-coated transmission electron microscope (TEM) grid, a silicon nitride membrane or an alkyl-functionalized silicon chip) was placed in a glass vial containing a colloidal solution of nanoparticles. The vial was placed tilted by 608–708 inside a low-pressure chamber. Ordered binary assemblies formed upon evaporation of the solvent. Toluene and mixtures of toluene with tetrachlor￾oethylene or chloroform were used as solvents (,1:1 by volume). The best binary assemblies (as determined by the length scale of ordering and a low occurrence of defects) were obtained by evaporating relatively concentrated colloidal solutions at 45 8C under reduced pressure (,3.2 kPa). Structural analysis. A Philips CM12 TEM operating at 120 kV was used to image the structure of the assemblies. Three-dimensional descriptions of the super￾lattices were developed by surveying large regions of the samples, to categorize all the crystal orientations, and recording a series of two-dimensional projections down the major symmetry axes. Tilting of the samples allowed observation of additional orientations not expressed in the plan view images. To assign the observed structures to crystallographic space groups, we built three-dimensional lattice models for the 180 most common space groups using Accelrys MS Modelling 3.1 software. The TEM images were compared with simulated projections to match the symmetry of our superlattices. We also performed a comparison of experimental small-angle electron diffraction patterns taken over larger areas, and the two-dimensional Fourier transformation power spectra of real space TEM images and the fast Fourier transform power spectra of the simulated projections to assure consistency. Electrophoretic mobility measurements. These were performed by electro￾phoretic light scattering using a Zetasizer Nano ZS Series (Malvern), allowing measurements in non-polar organic solvents. We used chloroform solutions with nanoparticle concentrations ,5 times higher than those used for growing binary superlattices. The concentrations of additives (oleic acid and TOPO) were similar to those used for directing BNSL self-assembly. After preparation, the colloidal solutions were left in the dark for several hours to allow the systems to equilibrate before each measurement. Received 20 August; accepted 2 November 2005. 1. Shenton, W., Pum, D., Sleytr, U. & Mann, S. Synthesis of cadmium sulphide superlattices using self-assembled bacterial S-layers. Nature 389, 585–-587 (1997). Figure 4 | TEM images and proposed unit cells of binary superlattices self-assembled from triangular nanoplates and spherical nanoparticles. a, b, Self-assembled from LaF3 triangular nanoplates (9.0 nm side) and 5.0 nm Au nanoparticles; c, self-assembled from LaF3 triangular nanoplates and 6.2 nm PbSe nanocrystals. The insets show a, a magnified image, and b, c, proposed unit cells of the corresponding superlattices. The structure shown in a forms on silicon oxide surfaces, while structures shown in b and c form preferentially on amorphous carbon substrates. LETTERS NATURE|Vol 439|5 January 2006 58

NATUREIVol 439 5 January 2006 LETTERS 2. Guarini, K W, Black, C T. Yeung S H L Optimization of diblock copolymer 22. Yin, Y. Alivisatos, A P Colloidal nanocrystal synthesis and the film self assembly. Adv Mater. 14, 1290-1294 (2002). rganic-inorganic interface Nature 437, 664-670 (200- 3. Red, F.X., Cho, K-S, Murray, C B. O'Brien, S. Three-dimensional binary 23. Korgel, B. A, Fullam, S, Connolly, S& Fitzmaurice, D Assembly and superlattices of magnetic nanocrystals and semiconductor quantum dots. Ntue423.968-971(200 hys. Chem. B1028379-8388(1998 4. Kiely, C.J., Fink, J, Brust, M., Bethel, D. Schiffrin, D J. Spontaneous ordering 24. Cho, K-S, Talapin, D. V Gaschler, W. Murray, C B. Designing PbSe f bimodal ensembles of nanoscopic gold clusters. Nature 396, 444-446 nanowires and nanorings through oriented attachment of nanoparticles. 1. Am Chem. soc.127,7140-7147(2005) 5. Shevchenko, E. V. et al. Colloidal 25. Ohara, P C Leff, D. V. Heath, J.R. Gelbart, W. M. Crystallization of opals anocrystals. J Am. Chem. Soc. 124, 11480-11485(2002) from polydisperse nanoparticles. Phys. Rev. Lett. 75, 3466-3469(1995) anders, A. E& KorgeL, B. A. Observation of an aB phase in bidisperse 26. Rabani E. Reichman. D. R. Geissler, P L Bru cocrystal superlattices. ChemPhys Chem 6, 61-65(2005) elf-assembly of nanoparticles. Nature 426, 271-274 (2003) henko,E V. Talapin, D. V, OBrien, S& Murray, C B Polymorphism in 27. Prasad, B L V, Stoeva, S L, Sorensen, C.M.& Klabunde, K.J.Digestive ripening of thiolated gold nanoparticles: the effect of alkyl chain length. metamaterials. J. Am. Chem. Soc. 127, 8741-8747(2005 Murray, M. J. Sanders, J V Close-packed structures of spheres of two Langmuir18,7515-7520(2002) 28. Hyeon, T, Lee, S.S, Park, J, Chung, Y. Na, H. B. Synthesis of highly 721-740(1980 9. Eldridge, M. D, Madden, P. A& FrenkeL, D Entropy-driven formation of a n process. J. Am. Chem. Soc. 123, 12798-12801(2001) 29. Hines, M. A. Scholes, G D Colloidal PbS nanocrystals with size-tunable lattice in a hard-sphere binary mixture. Nature 365, 35-37( 1993). 10. Leunissen, M. E et al. lonic colloidal crystals of oppositely charged particles. ar-infrared emission: Observation of post-synthesis self-narrowing of the ture437,235-240(2005) 3o article size distribution. Adv. Mater. 15, 1844-1849(2003) ang Y-w, Sun, x, Si, R, You, L-P. Yan, C-H Single-crystalline and oppositely charged colloids. Phys. Rev. Lett. 95, 128302(2005). onodisperse LaF3 triangular nanoplates from a single-source precursor. J Am. 12. Bolhuis, P G, Frenkel, D, Mau, S -C. Huse, D A Entropy difference between hem.Soc127,3260-3261(2005) the face-centred cubic and hexagonal close-packed crystal structures. Nature 88,235-236(1997) Supplementary Information is linked to the online version of the paper at 13. Pusey, P N. van Megen, W. Phase behaviour of concentrated suspensions of arly hard colloidal spheres. Nature 320, 340-342 (1986) Acknowledgements We thank V. Perebeinos, A van blaaderen, v Crespi L. Herman and L E Brus for discussions and r. l sandstrom for technic niversality and perfection. J. Am. Chem. Soc. 125, 15589-15598(2003). upport. This work was partially supported by the MRSEC Program of the 15. Cottin, X& Monson, P. A Substitutionally ordered solid solutions of hard heres. J. Chem. Phys.102,3354-3360(1995) National Science Foundation, and by the New York State Office of Science Technology and Academic Research (NYSTAR). S.o. is grateful for support from 16. Sanders, J. V s Murray, M. J Ordered ar nts of spheres of two different sizes in opal. Nature 275, 201-203(1978) 17. Hachisu, S& Yoshimura, S Optical demonstration of crystalline Author Contributions E V.S. and D V.T. contributed equally to this work. E.V.S. uperstructures in binary mixtures of latex globules. Nature 283, 188-189 and D V.T. carried out syntheses of nanoparticles, and E V.S. investigated 18. Shim, M. Guyot-Sionnest, P Permanent dipole moment and charges modelling and structural assignment of self-assembled binary superlattices. in colloidal semiconductor quantum dots. J. Chem. Phys. 111, 6955-6964 E.V.S., D.V.T. and N A.K. studied electrophoretic mobility of nanoparticles and 19. Krauss, T. D& Brus, L.E. Charge, polarizability, and photoionizat S.O. and C B M. initiated and supervised the work. D V.T. and C.B.M. wrote the paper. All authors discussed the results and commented on the manuscript. 20. Islam, M. A& Herman, I P Electrodeposition of patterned Cdse films using thermally charged nanocrystals. Appl. Phys. Lett. 80, Author Information sandpermissions. The author no competing 21. OBrien, R W.& White, L R Electrophoretic mobility of a spherical colloida financial interests. Correspondence and requests fo should be article. J. Chem. Soc. Farad. Trans. 1 74, 1607-1626(1978). addressedtoDV.T.(dvtalapinalblgov)orC.B.M.(cbmurray@us.ibm.com) 2006 Nature Publishing Group

© 2006 Nature Publishing Group 2. Guarini, K. W., Black, C. T. & Yeung, S. H. I. Optimization of diblock copolymer thin film self assembly. Adv. Mater. 14, 1290–-1294 (2002). 3. Redl, F. X., Cho, K.-S., Murray, C. B. & O’Brien, S. Three-dimensional binary superlattices of magnetic nanocrystals and semiconductor quantum dots. Nature 423, 968–-971 (2003). 4. Kiely, C. J., Fink, J., Brust, M., Bethel, D. & Schiffrin, D. J. Spontaneous ordering of bimodal ensembles of nanoscopic gold clusters. Nature 396, 444–-446 (1998). 5. Shevchenko, E. V. et al. Colloidal synthesis and self-assembly of CoPt3 nanocrystals. J. Am. Chem. Soc. 124, 11480–-11485 (2002). 6. Saunders, A. E. & Korgel, B. A. Observation of an AB phase in bidisperse nanocrystal superlattices. ChemPhysChem 6, 61–-65 (2005). 7. Shevchenko, E. V., Talapin, D. V., O’Brien, S. & Murray, C. B. Polymorphism in AB13 nanoparticle superlattices: An example of semiconductor-metal metamaterials. J. Am. Chem. Soc. 127, 8741–-8747 (2005). 8. Murray, M. J. & Sanders, J. V. Close-packed structures of spheres of two different sizes II. The packing densities of likely arrangements. Phil. Mag. A 42, 721–-740 (1980). 9. Eldridge, M. D., Madden, P. A. & Frenkel, D. Entropy-driven formation of a superlattice in a hard-sphere binary mixture. Nature 365, 35–-37 (1993). 10. Leunissen, M. E. et al. Ionic colloidal crystals of oppositely charged particles. Nature 437, 235–-240 (2005). 11. Bartlett, P. & Campbell, A. I. Three-dimensional binary superlattices of oppositely charged colloids. Phys. Rev. Lett. 95, 128302 (2005). 12. Bolhuis, P. G., Frenkel, D., Mau, S.-C. & Huse, D. A. Entropy difference between the face-centred cubic and hexagonal close-packed crystal structures. Nature 388, 235–-236 (1997). 13. Pusey, P. N. & van Megen, W. Phase behaviour of concentrated suspensions of nearly hard colloidal spheres. Nature 320, 340–-342 (1986). 14. Wong, S., Kitaev, V. & Ozin, G. A. Colloidal crystal films: Advances in universality and perfection. J. Am. Chem. Soc. 125, 15589–-15598 (2003). 15. Cottin, X. & Monson, P. A. Substitutionally ordered solid solutions of hard spheres. J. Chem. Phys. 102, 3354–-3360 (1995). 16. Sanders, J. V. & Murray, M. J. Ordered arrangements of spheres of two different sizes in opal. Nature 275, 201–-203 (1978). 17. Hachisu, S. & Yoshimura, S. Optical demonstration of crystalline superstructures in binary mixtures of latex globules. Nature 283, 188–-189 (1980). 18. Shim, M. & Guyot-Sionnest, P. Permanent dipole moment and charges in colloidal semiconductor quantum dots. J. Chem. Phys. 111, 6955–-6964 (1999). 19. Krauss, T. D. & Brus, L. E. Charge, polarizability, and photoionization of single semiconductor nanocrystals. Phys. Rev. Lett. 83, 4840–-4843 (1999). 20. Islam, M. A. & Herman, I. P. Electrodeposition of patterned CdSe nanocrystal films using thermally charged nanocrystals. Appl. Phys. Lett. 80, 3823–-3825 (2002). 21. O’Brien, R. W. & White, L. R. Electrophoretic mobility of a spherical colloidal particle. J. Chem. Soc. Farad. Trans. II 74, 1607–-1626 (1978). 22. Yin, Y. & Alivisatos, A. P. Colloidal nanocrystal synthesis and the organic–-inorganic interface. Nature 437, 664–-670 (2005). 23. Korgel, B. A., Fullam, S., Connolly, S. & Fitzmaurice, D. Assembly and self-organization of silver nanocrystal superlattices: Ordered “soft spheres”. J. Phys. Chem. B 102, 8379–-8388 (1998). 24. Cho, K.-S., Talapin, D. V., Gaschler, W. & Murray, C. B. Designing PbSe nanowires and nanorings through oriented attachment of nanoparticles. J. Am. Chem. Soc. 127, 7140–-7147 (2005). 25. Ohara, P. C., Leff, D. V., Heath, J. R. & Gelbart, W. M. Crystallization of opals from polydisperse nanoparticles. Phys. Rev. Lett. 75, 3466–-3469 (1995). 26. Rabani, E., Reichman, D. R., Geissler, P. L. & Brus, L. E. Drying-mediated self-assembly of nanoparticles. Nature 426, 271–-274 (2003). 27. Prasad, B. L. V., Stoeva, S. I., Sorensen, C. M. & Klabunde, K. J. Digestive ripening of thiolated gold nanoparticles: The effect of alkyl chain length. Langmuir 18, 7515–-7520 (2002). 28. Hyeon, T., Lee, S. S., Park, J., Chung, Y. & Na, H. B. Synthesis of highly crystalline and monodisperse maghemite nanocrystallites without a size-selection process. J. Am. Chem. Soc. 123, 12798–-12801 (2001). 29. Hines, M. A. & Scholes, G. D. Colloidal PbS nanocrystals with size-tunable near-infrared emission: Observation of post-synthesis self-narrowing of the particle size distribution. Adv. Mater. 15, 1844–-1849 (2003). 30. Zhang, Y.-W., Sun, X., Si, R., You, L.-P. & Yan, C.-H. Single-crystalline and monodisperse LaF3 triangular nanoplates from a single-source precursor. J. Am. Chem. Soc. 127, 3260–-3261 (2005). Supplementary Information is linked to the online version of the paper at www.nature.com/nature. Acknowledgements We thank V. Perebeinos, A. van Blaaderen, V. Crespi, I. Herman and L. E. Brus for discussions and R. L. Sandstrom for technical support. This work was partially supported by the MRSEC Program of the National Science Foundation, and by the New York State Office of Science, Technology and Academic Research (NYSTAR). S.O. is grateful for support from the DOE and an NSF CAREER award. Author Contributions E.V.S. and D.V.T. contributed equally to this work. E.V.S. and D.V.T. carried out syntheses of nanoparticles, and E.V.S. investigated formation of binary nanoparticle superlattices. E.V.S. and D.V.T. performed modelling and structural assignment of self-assembled binary superlattices. E.V.S., D.V.T. and N.A.K. studied electrophoretic mobility of nanoparticles and worked on modelling self-assembly phenomena in binary nanoparticle colloids. S.O. and C.B.M. initiated and supervised the work. D.V.T. and C.B.M. wrote the paper. All authors discussed the results and commented on the manuscript. Author Information Reprints and permissions information is available at npg.nature.com/reprintsandpermissions. The authors declare no competing financial interests. Correspondence and requests for materials should be addressed to D.V.T. (dvtalapin@lbl.gov) or C.B.M. (cbmurray@us.ibm.com). NATURE|Vol 439|5 January 2006 LETTERS 59

点击下载完整版文档(PDF)VIP每日下载上限内不扣除下载券和下载次数;
按次数下载不扣除下载券;
24小时内重复下载只扣除一次;
顺序:VIP每日次数-->可用次数-->下载券;
已到末页,全文结束
相关文档

关于我们|帮助中心|下载说明|相关软件|意见反馈|联系我们

Copyright © 2008-现在 cucdc.com 高等教育资讯网 版权所有