Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Structure and Properties of Nanoalloys
Structure and Properties of Nanoalloys
Structure and Properties of Nanoalloys
Ebook725 pages7 hours

Structure and Properties of Nanoalloys

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Structure and Properties of Nanoalloys is devoted to the topic of alloy nanoparticles, the bi-or multicomponent metallic nanoparticles that are often called nanoalloys. The interest in nanoalloys stems from the wide spectrum of their possible applications in the fields of catalysis, magnetism, and optics.

Nanoalloys are also interesting from a basic science point-of-view due to the complexity of their structures and properties. Nanoalloys are presently a very lively research area, with impressive developments in the last ten years. This book meets the need to systematize the wealth of experimental and computational results generated over the last decade.

  • Provides a well-organized, coherent overall structure, with a tutorial style format ideal for teaching and self-study
  • In-depth and fluent descriptions by a single leading academic
  • Presents a wealth of experimental and computational results generated over the last decade
LanguageEnglish
Release dateSep 3, 2016
ISBN9780081002476
Structure and Properties of Nanoalloys
Author

Riccardo Ferrando

Riccardo Ferrando is Professor of Condensed Matter Physics at the University of Genoa, Italy. His research interests include: Theory of stochastic processes. Theory of diffusion in periodic systems. Simulation of surface diffusion and crystal growth. Modelling of metal nanoparticles and nanoalloys. Modelling of aggregation processes in colloidal suspensions. He is author of more than 200 publications in peer reviewed journals.

Related to Structure and Properties of Nanoalloys

Titles in the series (13)

View More

Related ebooks

Technology & Engineering For You

View More

Related articles

Related categories

Reviews for Structure and Properties of Nanoalloys

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Structure and Properties of Nanoalloys - Riccardo Ferrando

    Chapter 1

    Introduction

    Riccardo Ferrando

    Abstract

    Nanoalloys are bi- or multicomponent aggregates of metal atoms in the size range from 1 to 10² nm. Nanoalloys display properties that depend on both size and composition, sharing the features of pure nanoparticles and of bulk alloys. Nanoalloys are thus well suited for a wide range of possible applications, in optics, catalysis and magnetism.

    Keywords

    Metals; Alloys; Nanoparticles; Optics; Catalysis; Magnetism

    This book is devoted to alloy nanoparticles. These are bi- or multicomponent metallic nanoparticles that are often called nanoalloys (indeed the term nanoalloy will be the most frequently used in the following). Typical sizes of nanoalloys range from a few atoms to several million atoms, with diameters from a few Å to ∼10² nm.¹ Nanoalloys, and nanoparticles in general, can be produced and studied in a variety of environments, such as in vapor phase, in colloidal suspensions, embedded in inert matrices or adsorbed on surfaces [1–7].

    Interest in nanoalloys has steadily increased because of the wide spectrum of their possible applications, mainly in the fields of catalysis, magnetism, and optics. As we will see in the following chapters, nanoalloys are also an interesting subject for basic research due to the variety and complexity of their structures and properties. Nanoalloys are thus a very lively research subject, with impressive developments, especially during the last 10 years, and an explosive growth of published material, which continues. This book aims both at summarizing basic knowledge about nanoalloys and at highlighting some of the most exciting recent developments.

    The basic feature of nanoalloys is that their properties can depend both on size (as in pure metal clusters and nanoparticles) and on composition, as in bulk alloys. This surely allows a wider range of tailored applications, but at the same time poses significant experimental challenges. In the following section we give a few examples about the size-dependent properties of nanoparticles and the composition-dependent properties of bulk alloys, ending with some brief historical notes about the use of nanoalloys.

    1.1 Scalable and Nonscalable Regimes in Metal Clusters and Nanoparticles

    Single-element (pure) metallic clusters and nanoparticles present behaviors that are distinct from those of individual atoms and of bulk crystals. The specificity of clusters and nanoparticles shows up in the peculiar size-dependence of their properties, for which two different regimes can be distinguished [5, 8, 9].

    For sufficiently large aggregates, it usually happens that deviation of the properties from the macroscopic bulk limit scales in a smooth way with the size of the aggregate itself. An example of such smooth scaling of the properties is given in Fig. 1.1. There, the size-dependence of the melting temperature Tm of pure (elemental) Au nanoparticles is reported. The melting temperature in nanoparticles is decreased with respect to bulk samples because the large fraction of surface atoms in nanoparticles facilitates the transition to the liquid state. The experimental data of Fig. 1.1 [10] show a decrease of the melting temperature and this behavior is roughly reproduced by a simple law due to Pawlow [6, 11].

    Fig. 1.1 Size dependence of the melting temperature T m of Au nanoparticles. D is the nanoparticle diameter. Source: Reprinted with permission from P. Borel, J.-P. Buffat, Size effect on the melting temperature of gold particles, Phys. Rev. A 13 (1976) 2287–2298. Copyright 1997 American Physical Society.

    This law states that the relative decrease in the melting temperature Tm(N) of a spherical nanoparticle of size N is inversely proportional to its diameter D, which is in turn proportional to N¹/³:

       (1.1)

    However, the size dependence of the aggregate properties may become much less regular and smooth for sufficiently small sizes. This corresponds to the nonscalable regime, in which small is different in an essential way [8], because the behavior of small objects can no longer be deduced from the knowledge of their behavior at larger sizes. Often this is referred to by saying that every atom counts.

    In the case of the melting transition, a very well-known example of nonscalable behavior at small sizes is given by sodium clusters. In Fig. 1.2 the nonmonotonic behavior of the melting temperature of pure Na clusters is reported [12]. Compared to the case of the Au nanoparticles in Fig. 1.1, sizes are smaller. Peaks corresponding to magic melters are evident. Several peaks can be explained by considering the magic sizes of structures in the icosahedral family, as indicated in the figure (see Section 2.1 for geometric magic sizes of different structural families, including the icosahedral family).

    Fig. 1.2 Size dependence of melting temperature (black, top curve), latent heat of fusion per atom q (red, middle curve) and entropy change per atom Δ s (blue, bottom curve) of Na nanoparticles. The proposed structures of the nanoparticles corresponding to the most important peaks are also shown. Source: Reprinted with permission from H. Haberland, T. Hippler, J. Donges, O. Kostko, M. Schmidt, B. von Issendorff, Melting of sodium clusters: Where do the magic numbers come from? Phys. Rev. Lett. 94 (2005) 035701. Copyright 2005 American Physical Society.

    Another example concerning a nonscalable property is reported in Fig. 1.3, in which a series of calculated lowest-energy structures of pure Ag nanoparticles in gas phase is shown [13, 14] for sizes up to about 200 atoms. In this case, the property that is varying in a nonsmooth way with size is the geometric structure itself of the nanoparticle, which changes in a complicated way between icosahedral, decahedral and truncated-octahedral motifs (see Chapter 2). It is expected that this nonscalable regime will then disappear with increasing size, because fcc nanocrystals, which are fragments of the Ag bulk lattice, will finally prevail as the most stable Ag nanoparticle structures. However, the calculations [14] indicate that the scalable regime is attained for sizes of the order of 10⁴ atoms.

    Fig. 1.3 Lowest-energy structures of Ag nanoparticles for different sizes. These structures are calculated by global optimization methods using an atomistic model. More details about the model and global optimization techniques are given in Chapter 4 . The optimal structure changes in a complicated way with size. For sizes 38 and 201, optimal structures are truncated octahedra, while for sizes 75 and 192 they are decahedral, and for sizes 55 and 147 they are icosahedral.

    Further well-known examples of complex size dependence are seen in the electronic and optical properties of metallic nanoparticles [15, 16]. Small metal clusters present a nonsmooth behavior of the ionization energy with size, as shown in Fig. 1.4 [17] for Ag clusters, which reflects the electronic structure of such small clusters in which every (valence) electron counts. Each Ag atom has a single valence electron, so that the number of atoms and the number of quasi-free electrons coincide in a neutral Ag cluster. The odd-even oscillations of the ionization energy of Fig. 1.4 indicate that clusters with unpaired electrons are more easily ionized.

    Fig. 1.4 Odd-even oscillations in the vertical ionization energy of Ag clusters. Since in Ag there is a single valence electron per atom, these oscillations reflect the fact that clusters with unpaired electrons are more easily ionized. The dashed line corresponds to the work function of polycrystalline bulk silver (4.26 eV). Source: Reprinted with permission from C. Jackschath, I. Rabin, W. Schulze, Electron impact ionization of silver clusters Agn, n ≤ 36, Zeit. Phys. D 22 (2) (1992) 517–520. Copyright 1992 Springer.

    One may also note that in Fig. 1.4, there are some peaks (those at sizes 2, 8, and 20), which are followed by stronger decreases of the ionization energy. The same type of behavior is present also in Au clusters [18]. This result can be interpreted treating Ag (or Au) clusters as superatoms, in which the valence electrons of the cluster play a role that is somewhat analogous to that of electrons in atoms, thus presenting the shell-closure phenomena.

    The simplest treatment of electronic shell closure in clusters is given by the spherical jellium model [5, 19, 20]. This model assumes a uniform background of positive charge, given by the ions, in which the quasi-free valence electrons are subjected to an external potential, and fill discrete energy levels. Simple forms of this potential are the infinitely deep spherical well and the harmonic well. The solution of the single-electron Schrödinger equation for the spherical well gives the following series of magic numbers, for which electrons form closed shells: 2, 8, 18, 20, 34, 40, 58, etc. On the other hand, the harmonic well gives the series: 2, 8, 20, 40, 70, etc. Special electronic stability is expected for clusters with these numbers of valence electrons. Both series include sizes 2, 8, and 20, showing that Ag cluster ionization results can be interpreted by the spherical jellium model, with the indication that the harmonic model might be more appropriate, because it does not present shell closure for size 18.

    Scalable and nonscalable regimes also can be distinguished when considering the interaction of clusters with adsorbed species. An example is given in Fig. 1.5 in which the adsorption of molecular oxygen on free silver clusters has been measured as a function of cluster size N, finding a very irregular regime up to N = 40, followed by a smooth behavior. Above N = 40 the quantity of adsorbed oxygen becomes a scalable property. Note, however, that the geometric structure of Ag clusters is not yet scalable for sizes around 40, because the experiments have shown that the most stable structures are of the icosahedral family [21] up to at least N = 55, and the calculations [14] indicate that, indeed, the structures become scalable to the bulk limit for much larger sizes (see Fig. 1.3) [22].

    Fig. 1.5 Silver clusters pass through a reaction chamber in a flow of helium with a partial oxygen pressure of 6 × 10 ³ mbar. The time of interaction between gas and clusters is long enough to reach an equilibrium between adsorption and evaporation. Differential pumping separates the clusters from the gas and a time-of-flight mass spectrometer detects the mean number of adsorbed O 2 per cluster. This number 〈 #O 2 〉 is used as a measure of adsorption. Source: Reprinted with permission from M. Schmidt, C. Bréchignac, Silver and oxygen: Transition from clusters to nanoparticles, Compt. Rend. Phys. 17 (2016) 481–484. Copyright 2016 Elsevier Masson SAS. All rights reserved.

    This last point allows us to remark that scalability does not depend on the specific series of clusters of interest alone (being, e.g., Au or Ag clusters of increasing size), but also on the specific property being measured. A discussion of scalability and nonscalability issues about the adsorption of atoms and small molecules on transition and noble metal clusters is given in Ref. [23] and references therein.

    1.2 Bulk alloys: Composition-dependent properties

    Let us now consider a completely different class of systems, namely macroscopic metallic alloys. Metallic alloys have played an enormous role in the development of human civilization. This is strikingly evident from the Bronze Age, which dates back from the 3rd millennium BC, whose name itself is that of a metallic alloy.

    The historical importance of alloys stems from the fact that the range of properties of metallic systems can be greatly extended by taking mixtures of elements, which means that the properties of macroscopic alloys are composition-dependent.

    The dependence of alloy properties on composition is evident in their bulk phase diagrams. Consider, for example, the bulk phase diagram of the binary Ag-Cu system, as shown in Fig. 1.6 [24]. This is a very weakly miscible system, with a large miscibility gap in the solid state. In the miscibility gap, the system separates in an Ag-rich α phase and in a Cu-rich β phase, whose relative abundances depend on the overall composition of the alloy.

    Fig. 1.6 Phase diagram of Ag-Cu bulk alloys [24]. Temperature T is in K.

    There are systems in which miscibility is even weaker, like Ag-Co and Ag-Ni, where the miscibility gap extends even to the liquid state (see the Ag-Ni phase diagram of Fig. 5.17 in Section 5.4.1). Other systems presenting very weak miscibility are Au-Co and Au-Ni.

    On the other hand, there are systems presenting a completely different behavior, with the two metals intermixing for all compositions at the solid state in a wide range of temperatures, in which the metals form solid solutions. An example of this type is given by the phase diagram of Ag-Pd in Fig. 1.7 [25]. Other systems presenting solid-state miscibility in a wide portion of the phase diagrams (and, in some cases, together with a low-T miscibility gap) are Pd-Pt, Au-Pt, Cu-Ni, and Ag-Au.

    Fig. 1.7 Phase diagram of Ag-Pd bulk alloys. Temperature T is in °C. Source: Reprinted with permission from I. Karakaya, W.T. Thompson, The Ag-Pd (silver-palladium) system, Bull. Alloy Phase Diagrams 9 (1988) 237–243. Copyright 1988 Springer.

    A further possibility is that the metals not only intermix, but even form ordered phases, that is, arrangements of atoms presenting long-range order. These phases are often referred to as intermetallic compounds. Examples of ordered phases are discussed in Section 2.2.1. Systems of this class are Fe-Pt (see the phase diagram in Fig. 1.8 [26]), Co-Pt, Au-Cu, and Au-Pd.

    Fig. 1.8 Phase diagram of Fe-Pt bulk alloys. Temperature T is in °C. Source: Reprinted with permission from H. Okamoto, Fe-Pt (iron-platinum), J. Phase Equil. Diffusion 25 (2004) 395. Copyright 2004 Springer.

    It is interesting to note that 3d–5d transition and noble metal alloys present a stronger tendency to form intermetallic compounds than 3d–4d alloys. For example, one may compare Au-Cu to Ag-Cu, or Pt-Ni to Pd-Ni. This behavior has been attributed to the strong relativistic effects on the electrons of 5d transition and noble metals [27]. Relativistic effects cause the contraction of s orbitals, whose energy is shifted downwards. Contracted s orbitals better screen the nuclear charge. This enhanced screening in turn causes a looser binding of d electrons, which increases their energy. For this reason, in a 3d–5d system, the d band of the 5d element gets closer to the d band of the 3d element, reinforcing the binding between unlike atoms. At the same time, the contraction of s orbitals effectively reduces the size of 5d atoms, decreasing size mismatch with 3d atoms. These effects, which are much less important in 3d–4d systems, tend to favor the formation of compounds instead of phase separated arrangements.

    1.3 Marrying nano and alloys: The nanoalloys

    Following the words by Julius Jellinek [28], we can undoubtedly state that there has been growing interest in marrying nano and alloys. This is a natural way to obtain objects whose properties can depend on both size and composition, having therefore a broader range of properties than single-component metal nanoparticles or massive alloys.

    Even though research in nanoalloys has witnessed an explosive development in the last two decades, it is worth noting that nanoalloys were already being produced several centuries ago and were used to obtain quite surprising physical effects. A famous example is the Lycurgus cup (British Museum), shown in Fig. 1.9. This is a masterpiece of Roman glass craftsmanship, produced in the first part of the 4th century AD to celebrate the victory of emperor Constantine the Great against Lycinius. The cup shows the mythical King Lycurgus, who (depending on the version of the myth) tried to kill Ambrosia, a follower of the god Dionysus. She was transformed into a vine that twined around the enraged king and restrained him, eventually killing him. Dionysus and two followers are also shown in the cup’s glass taunting the king.²

    Fig. 1.9 The Lycurgus Cup. (Left) In reflected light. (Right) In transmitted light. Source: From Johnbod. Information on Johnbod retrievable at https://en.wikipedia.org/wiki/User:Johnbod.

    The interesting property of the glass of the Lycurgus cup is dichroism. The dichroic glass of the cup appears green (light gray in print versions) in reflected light and red (dark gray in print versions) in transmitted light. Modern studies have shown that the cup’s dichroism is due to nanoalloys. In fact, when examined using the transmission electron microscopy (TEM), this glass reveals very tiny metallic inclusions (see Fig. 1.10), that are identified as being Ag0.7Au0.3 nanoalloys with diameters of ∼50 nm [29]. These nanoalloys present a surface plasmon resonance [16] in the green-blue region, so that the glass absorbs and diffuses green-blue light while allowing red light to be transmitted.

    Fig. 1.10 TEM image of an Ag-Au alloy nanoparticle embedded in the glass of the Lycurgus Cup. The nanoalloy has composition Ag 0.7 Au 0.3 . Source: Reprinted with permission from D.J. Barber, I.C. Freestone, An investigation of the origin of the color of the Lycurgus cup by analytical transmission electron-microscopy, Archaeometry 32 (1990) 33–45. Copyright 1990 by Wiley-Blackwell.

    The Romans were thus producing nanoalloys and using their surprising properties without even knowing anything about their existence. It is worth noting that the size of the nanoalloys in the Lycurgus cup glass is appropriate for having a sufficiently sharp surface plasmon resonance, and that nanoalloy composition is tuned in such a way that blue-green light is absorbed. Pure Ag or pure Au nanoparticles of the same size and shape would absorb a different part of the spectrum. In fact, photon-emission studies [30] of Ag-Au nanoalloys adsorbed on the surface of a Al2O3 thin film grown on a NiAl(110) crystal have in fact revealed a red shift of the surface plasmon peak with increasing Au content in the nanoalloy, as shown in Fig. 1.11.

    Fig. 1.11 Normalized photon emission spectra of single Ag-Au alloy clusters on Al 2 O 3 /NiAl(110). The Ag content increases from 0% (top) to 100% (bottom). Source: Reprinted with permission from W. Benten, N. Nilius, N. Ernst, H.-J. Freund, Photon emission spectroscopy of single oxide-supported Ag-Au alloy clusters, Phys. Rev. B 72 (2005) 045403. Copyright 2005 American Physical Society.

    The Lycurgus cup is probably the oldest known example in history of the versatility of nanoalloys, a feature that will emerge in many other cases throughout this book. This versatility is at the root of the wealth of applications of nanoalloys in different fields [31–39]. These applications comprise catalysis, data storage, plasmonics, biosensing, and also environmental and medical applications (the latter including diagnostic, therapeutic, and toxicological applications).

    References

    [1] Haberland H. Clusters of Atoms and Molecules. Berlin: Springer; 1994.

    [2] Marks L.D. Experimental studies of small-particle structures. Rep. Prog. Phys. 1994;57:603–649.

    [3] Martin T.P. Shells of atoms. Phys. Rep. 1996;273:199–241.

    [4] Jellinek J. Theory of Atomic and Molecular Clusters. Berlin: Springer; 1999.

    [5] Johnston R.L. Atomic and Molecular Clusters. London: Taylor and Francis; 2002.

    [6] Baletto F., Ferrando R. Structural properties of nanoclusters: Energetic, thermodynamic, and kinetic effects. Rev. Mod. Phys. 2005;77:371–423.

    [7] Ferrando R., Jellinek J., Johnston R.L. Nanoalloys: From theory to applications of alloy clusters and nanoparticles. Chem. Rev. (Wash. DC). 2008;108:845–910.

    [8] Landman U. Materials by numbers: Computations as tools of discovery. Proc. Natl. Acad. Sci. USA. 2005;102:6671–6678.

    [9] Alonso J.A. Structure and Properties of Atomic Nanoclusters. London: Imperial College Press; 2005.

    [10] Borel P., Buffat J.-P. Size effect on the melting temperature of gold particles. Phys. Rev. A. 1976;13:2287–2298.

    [11] Pawlow P.N. Über die abhängigkeit des schmelzpunktes von der oberflächenenergie eines festen korpers. Z. Phys. Chem. 1909;65:1–35.

    [12] Haberland H., Hippler T., Donges J., Kostko O., Schmidt M., von Issendorff B. Melting of sodium clusters: Where do the magic numbers come from? Phys. Rev. Lett. 2005;94:035701.

    [13] Baletto F., Mottet C., Ferrando R. Microscopic mechanisms of the growth of metastable silver icosahedra. Phys. Rev. B. 2001;63:155408.

    [14] Baletto F., Ferrando R., Fortunelli A., Montalenti F., Mottet C. Crossover among structural motifs in transition and noble-metal clusters. J. Chem. Phys. 2002;116:3856.

    [15] de Heer W.A. The physics of simple metal clusters—experimental aspects and simple models. Rev. Mod. Phys. 1993;65:611–676.

    [16] Kreibig U., Vollmer M. Optical Properties of Metal Clusters. Berlin: Springer-Verlag; 1995.

    [17] Jackschath C., Rabin I., Schulze W. Electron impact ionization of silver clusters Agn, n ≤ 36. Zeit. Phys. D. 1992;22(2):517–520.

    [18] Jackschath C., Rabin I., Schultze W. Electron-impact ionization-potentials of gold and silver clusters Men, n ≤ 22. Phys. Chem. Chem. Phys. 1992;96:1200–1204.

    [19] Knight W.D., Clemenger K., de Heer W.A., Saunders W.A., Chou M.Y., Cohen M.L. Electronic shell structure and abundances of sodium clusters. Phys. Rev. Lett. 1984;52:2141–2144.

    [20] Brack M. The physics of simple metal-clusters—Self-consistent jellium model and semiclassical approaches. Rev. Mod. Phys. 1993;65:677–732.

    [21] Xing X., Danell R.M., Garzón I.L., Michaelian K., Blom M.N., Burns M.M., Parks J.H. Size-dependent fivefold and icosahedral symmetry in silver clusters. Phys. Rev. B. 2005;72:081405.

    [22] Schmidt M., Bréchignac C. Silver and oxygen: Transition from clusters to nanoparticles. Compt. Rend. Phys. 2016;17:481–484.

    [23] Kozlov S.M., Neyman K.M. Catalysis from first principles: Towards accounting for the effects of nanostructuring. Topics Catal. 2013;56:867–873.

    [24] Hultgren R., Desai P.D., Hawkins D.T., Gleiser M., Kelley K.K. Values of the Thermodynamic Properties of Binary Alloys. Berkeley: American Society for Metals; 1981.

    [25] Karakaya I., Thompson W.T. The Ag-Pd (silver-palladium) system. Bull. Alloy Phase Diagr. 1988;9:237–243.

    [26] Okamoto H. Fe-Pt (iron-platinum). J. Phase Equil. Diffusion. 2004;25:395.

    [27] Wang L.G., Zunger A. Why are the 3d-5d compounds CuAu and NiPt stable, whereas the 3d-4d compounds CuAg and NiPd are not. Phys. Rev. B. 2003;67:092103.

    [28] Jellinek J. Nanoalloys: Tuning properties and characteristics through size and composition. Fadaray Discuss. 2008;138:11–53.

    [29] Barber D.J., Freestone I.C. An investigation of the origin of the color of the Lycurgus cup by analytical transmission electron-microscopy. Archaeometry. 1990;32:33–45.

    [30] Benten W., Nilius N., Ernst N., Freund H.-J. Photon emission spectroscopy of single oxide-supported Ag-Au alloy clusters. Phys. Rev. B. 2005;72:045403.

    [31] Piccolo L. Surface studies of catalysis by metals: Nanosize and alloying effects. In: Alloyeau D., Mottet C., Ricolleau C., eds. Nanoalloys—Synthesis, Structure and Properties. Berlin: Springer-Verlag; 2012:369–404.

    [32] Bazin D., Fechete I., Garin F., Barcaro G., Negreiros F.R., Sementa L., Fortunelli A. Reactivity and catalysis by nanoalloys. In: Calvo F., ed. Nanoalloys—From Fundamentals to Emergent Applications. Amsterdam: Elsevier; 2013:283–344.

    [33] Dupuis V., Khadra G., Hillion A., Tamion A., Tuaillon-Combes J., Bardotti L., Tournus F. Intrinsic magnetic properties of bimetallic nanoparticles elaborated by cluster beam deposition. Phys. Chem. Chem. Phys. 2015;17:27996–28004.

    [34] Barcaro G., Sementa L., Fortunelli A., Stener M. Optical properties of nanoalloys. Phys. Chem. Chem. Phys. 2015;17:27952–27967.

    [35] Bagga K., Brougham D.F., Keyes T.E., Brabazon D. Magnetic and noble metal nanocomposites for separation and optical detection of biological species. Phys. Chem. Chem. Phys. 2015;17:27968–27980.

    [36] McNamara K., Tofail S.A.M. Nanosystems: the use of nanoalloys, metallic, bimetallic, and magnetic nanoparticles in biomedical applications. Phys. Chem. Chem. Phys. 2015;17:27981–27995.

    [37] Hajipour M.J., Fromm K.M., Ashkarran A.A., de Aberasturi D.J., de Larramendi I.R., Rojo T., Serpooshan V., Parak W.J., Mahmoudi M. Antibacterial properties of nanoparticles. Trends Biotech. 2012;30(10):499–511.

    [38] Ruparelia J.P., Chatterjee A.K., Duttagupta S.P., Mukherji S. Strain specificity in antimicrobial activity of silver and copper nanoparticles. Acta Biomater. 2008;4:707–716.

    [39] Eremenko A.M., Petrik I.S., Smirnova N.P., Rudenko A.V., Marikvas Y.S. Antibacterial and antimycotic activity of cotton fabrics, impregnated with silver and binary silver/copper nanoparticles. Nanoscale Res. Lett. 2016;11:1–9.


    ☆ "To view the full reference list for the book, click here"

    ¹ In the following we use the term nanoalloy for all bi- or multicomponent aggregates whose size is in the specified range, irrespectively of the arrangement of their constituent elements. The arrangement of the elements in a nanoalloy can thus be intermixed, as in solid solutions and in ordered alloys, or phase-separated, as in core-shell and Janus particles. We will also use the terms cluster and nanoparticle rather loosely. For nanoparticle we mean any aggregate of atoms or molecules whose size is between 1 and 10² nm, that is, in the nanometric size range. The term cluster will be used sometimes to denote nanoparticles of small sizes, containing up to ∼10² atoms (i.e., in the range up to ∼2 nm in diameter), but without specific reference to the scalability or nonscalability of their properties. The term nanocrystal is used solely to denote nanoparticles that are fragments of bulk crystals, whatever their size.

    ² see http://en.wikipedia.org/wiki/Lycurgus_Cup.

    Chapter 2

    Geometric structures and chemical ordering in nanoalloys

    Riccardo Ferrando

    Abstract

    Here we introduce the most important geometric structures (also denoted as structural motifs) and chemical ordering patterns, qualitatively discussing the driving forces that determine their stability. Crystalline and noncrystalline structural motifs are considered. Crystalline motifs comprise fragments of the fcc, bcc, and hcp bulk lattices. Noncrystalline motifs comprise decahedra, icosahedra, and polyicosahedra. Different types of mixing and nonmixing chemical ordering patterns are introduced, from random-solution, to ordered-phase, to core-shell, Janus and multi-shell arrangements.

    Keywords

    Crystalline structures; Icosahedra; Decahedra; Polyicosahedra; Mixing patterns; Core-shell structures; Multishell structures; Excess energy; Janus particles

    2.1 Geometric structures

    Single-element (elemental) metallic nanoparticles present a rich variety of geometric structures [2, 3, 5, 6]. As we have shown in Chapter 1, the geometric structure of an elemental metal nanoparticle can change with size in a complicated way, passing through a series of different shapes. The variety of possible geometries in nanoalloys is even more impressive.

    In this section we describe the most important structural families (in the following denoted also as structural motifs) of nanoparticles. Structures are here considered without taking into account the chemical identities of the atoms at different sites. Therefore what follows has a straightforward application to elemental nanoparticles, whereas for nanoalloys some more caveats have to be taken into account, as we will see in Section 2.2.

    Structural motifs are mainly divided into two classes, the crystalline and the noncrystalline motifs:

    • Crystalline motifs are fragments of bulk crystals, for example, fragments of the face-centered cubic (fcc), body-centered cubic (bcc), and hexagonal close-packed (hcp) lattices, which are the most common crystal lattices that are found in pure metals.

    • Noncrystalline motifs are possible because the constraint of translational invariance does not apply to clusters and nanoparticles, which can thus assume shapes that are not fragments of any crystal lattice. The most common noncrystalline structures for pure metal nanoparticles are the icosahedron (Ih) and the decahedron (Dh), and other fcc twinned structures. However, in nanoalloys, other structures, such as polyicosahedra, also are possible for small sizes.

    In the literature one can find many examples of experimental observation of crystalline and noncrystalline structures in metallic nanoparticles. The results are so numerous that it is impossible to mention all of them. Just to recall a few examples, fcc, decahedral, and icosahedral structures have been identified in pure nanoparticles for all noble metals [40–45].

    2.1.1 Crystalline Structures

    Here we consider fcc, bcc, and hcp nanocrystals because most metals of interest in nanoalloys crystallize in one of these lattices.

    Face-centered cubic (fcc) nanoparticles

    Since we aim at constructing nanoparticles of good energetic stability, we tentatively start by cutting crystal fragments exposing only orientations of low surface energy. In most metals, these correspond to the surfaces with the densest packing of atoms.

    In the fcc lattice, the most compact surface corresponds to the (111) orientation. (111) surfaces are close-packed, because their atoms have six nearest neighbors in the surface plane, which is the maximum possible packing. A possible way to obtain a nanoparticle whose surface contains only (111) facets is to cut an octahedral fragment. Another possibility would be to cut a tetrahedral fragment. However, the structures based on the tetrahedron are much less common than those based on the octahedron (a notable exception is the tetrahedral Au20 cluster [46]), so that we will focus on octahedron-based structures.

    The regular octahedron (Oh) is made of two regular square pyramids sharing their bases (see Fig. 2.1). The eight facets of the regular octahedron are equilateral triangles. The regular octahedron is highly symmetric, since its symmetry group Oh contains 48 elements. An Oh can be characterized by a single integer index nl, which is the number of atoms in each of its 12 edges. The total number of atoms of the octahedron is

    Fig. 2.1 From top to bottom, octahedron, regular truncated octahedron, and cuboctahedron. Each structure is shown in two views. (A) Octahedron of 670 atoms. The two pyramids share a square base with edges of n l = 10 atoms. (B) Truncated octahedron of 586 atoms. It is obtained from the previous octahedron by truncating the vertices. A number of atomic planes n cut = 3 is eliminated at each vertex, so that six 4 × 4 square facets appear. This truncation (which satisfies n l = 3 n cut + 1) ensures that the resulting truncated octahedron is regular, that is, its close-packed facets are regular hexagons. (C) Cuboctahedron of 309 atoms. This is obtained from the octahedron with edges of n l = 9 atoms by removing n cut = 4 planes at each vertex, which satisfies n l = 2 n cut + 1. In the cuboctahedron the close-packed facets degenerate into triangles.

       (2.1)

    which gives the series of geometric magic numbers NOh = 1, 6, 19, 44, 85, … . Eq. (2.1) is derived noting that there are two square atomic planes for all sizes between 1 and nl − 1, and a single nl × nl plane.

    The Oh structure has a large surface-to-volume ratio, and therefore it is not expected to be especially favorable from an energetic point of view. Structures with smaller surface-to-volume ratio can be obtained by performing truncations at the vertices of the octahedron. These truncations create six open (100) facets that usually have an energetic cost because of their higher surface energy in most materials. The optimal degree of truncation from the point of view of the energetics will be discussed later below. Now we focus only on the geometric aspects related to the different degrees of truncation. Octahedra can be truncated symmetrically, by making the same truncation at each vertex, or asymmetrically. In the following we focus on symmetrically truncated octahedra.

    In order to characterize a symmetrically truncated octahedron (TO) we need one more integer besides nl. This second index is ncut, which corresponds to the number of layers eliminated at each vertex (see Fig. 2.1B). When cutting ncut layers,¹2ncut³ + 3ncut² + ncut atoms are eliminated in total, so that the magic numbers for symmetric TOs are

      

    (2.2)

    Such truncation creates six square (100) facets whose edges contain ncut + 1 atoms, whereas the eight original triangular close-packed (111) facets become hexagonal. These hexagons are not regular in general, having three (nonconsecutive) edges of length ncut + 1 and three edges of length nl − 2ncut. Truncated octahedra with regular hexagonal facets are thus possible when ncut + 1 = nl − 2ncut, that is, when nl = 3ncut + 1. If this condition is satisfied, the regular TO is obtained (see Fig. 2.1B), whose geometric magic numbers are

       (2.3)

    which gives the series NTOreg = 38, 201, 586, 1289, … for nl = 4, 7, 10, 13, ….

    Finally, when nl = 2ncut + 1 the hexagonal facets degenerate to triangles and the cuboctahedron (COh) is obtained. In the cuboctahedron, the total area of the open (100) facets is larger than the total area of the close-packed facets. The magic numbers of the cuboctahedra are

       (2.4)

    and, in terms of ncut

      

    (2.5)

    which gives NCOh = 1, 13, 55, 147, 309, … for nl = 1, 3, 5, 7, 9, … and ncut = 0, 1, 2, 3, 4, …, respectively. We will see that the series of geometric magic numbers is the same also for regular Ino decahedra and for Mackay icosahedra, indicating that the cuboctahedron is indeed an onionlike structure made of concentric atomic layers, with the number of concentric layers given by k = ncut + 1.²

    Body-centered cubic (bcc) nanoparticles

    In the bcc lattice the surface with the densest atomic packing is of (110) type. It is possible to cut a bcc fragment presenting only (110) surfaces by building an octahedron (see . Both the base and the shorter edges of the triangles contain the same number nl of atoms, but the atoms along the shorter edges are first neighbors, whereas the atoms along the base are second neighbors. The six vertices of this octahedron are not all equivalent. Top and bottom vertices (see Fig. 2.2A, left panel) are different from the remaining four vertices at the corners of the common base of the two

    Enjoying the preview?
    Page 1 of 1