Icosahedral twins

Last updated
FCC icosahedral model projected down the 5-fold on the left and 3-fold zone axis orientation on the right. IcotwinModel.png
FCC icosahedral model projected down the 5-fold on the left and 3-fold zone axis orientation on the right.

An icosahedral twin is a nanostructure found in atomic clusters and also nanoparticles with some thousands of atoms. The simplest form of these clusters is twenty interlinked tetrahedral crystals joined along triangular (e.g. cubic-(111)) faces, although more complex variants of the outer surface also occur. A related structure has five units similarly arranged with twinning, which were known as "fivelings" in the 19th century, [1] [2] [3] more recently as "decahedral multiply twinned particles", "pentagonal particles" or "star particles". A variety of different methods (e.g. condensing metal nanoparticles in argon, deposition on a substrate, chemical synthesis) lead to the icosahedral form, and they also occur in virus capsids.

Contents

They occur at small sizes where they have lower surface energies than other configurations. This is balanced by a strain energy, which dominates at larger sizes. This leads to a competition between different forms as a function of size, and often there is a population of different shapes.

Energetics

In a large particle the form that it takes is dominated by the bulk bonding, leading to a Wulff construction shape. However, when the size is reduced there are a significant number of atoms at the surface, and hence the surface energy starts to become important. Icosahedral arrangements, typically because of their smaller surface energy, [4] may be preferred for small clusters. For face centered cubic materials such as gold or silver these structures can be considered as being built from twenty different single crystal units all with three twin facets arranged in icosahedral symmetry, and mainly the low energy {111} external facets. The external surface shape can be generated from a modified Wulff construction. [3] and is also not always that of a simple icosahedron; there can be additional facets at the twin boundaries leading to a more spherical shape. [3] There are several software codes that can be used to calculate the shape as a function if the energy of different surface facets. [5] [6]

With just tetrahedra these structure cannot fill space and there would be gaps, so there is some distortions of the atomic positions, that is elastic deformation to close these gaps. [4] Roland De Wit pointed out that these can be thought of in terms of disclinations, [7] an approach later extended to three dimensions by Elisabeth Yoffe. [8] This leads to a compression in the center of the particles, and an expansion at the surface. [8]

At larger sizes the energy to distort becomes larger than the gain in surface energy, and bulk materials (i.e. sufficiently large clusters) generally revert to one of the crystalline close-packing configurations. In principle they will convert to a simple single crystal with a Wulff construction [9] shape. The size when they become less energetically stable is typically in the range of 10-30 nanometers in diameter, [10] but it does not always happen that the shape changes and the particles can grow to micron sizes. [11]

The most common approach to understand the formation of these particles, first used by Shozo Ino in 1969, [4] is to look at the energy as a function of size comparing these icosahedral twins, decahedral nanoparticles and single crystals. The total energy for each type of particle can be written as the sum of three terms:

Energy landscape for a 75 atom Leonard-Jones cluster for temperature and an order parameter. Energy landscape for 75 atom Leonard-Jones cluster.tif
Energy landscape for a 75 atom Leonard-Jones cluster for temperature and an order parameter.

for a volume , where is the surface energy, is the disclination strain energy to close the gap , and is a coupling term for the effect of the strain on the surface energy via the surface stress, [13] [14] [15] which can be a significant contribution. [16] The sum of these three terms is compared to the total surface energy of a single crystal (which has no strain), and to similar terms for a decahedral particle. Of the three the icosahedral particles have both the lowest total surface energy and the largest strain energy for a given volume. Hence the icosahedral particles are more stable at very small sizes. At large sizes the strain energy can become very large, so it is energetically favorable to have dislocations and/or a grain boundary instead of a distributed strain. [17]

There is no general consensus on the exact sizes when there is a transition in which type of particle is lowest in energy, as these vary with material and also the environment such as gas and temperature; the coupling surface stress term and also the surface energies of the facets are very sensitive to these. [18] [19] [20] In addition, as first described by Michael Hoare and P Pal [21] and R. Stephen Berry [22] [23] and analyzed for these particles by Pulickel Ajayan and Laurence Marks [24] as well as discussed by others such as Amanda Barnard, [25] David J. Wales, [26] [27] [28] Kristen Fichthorn [29] and Francesca Baletto and Riccardo Ferrando, [30] at very small sizes there will be a statistical population of different structures so many different ones will exist at the same time. In many cases nanoparticles are believed to grow from a very small seed without changing shape, and hence what is found reflects the distribution of coexisting structures. [3]

For systems where icosahedral and decahedral morphologies are both relatively low in energy, the competition between these structures has implications for structure prediction and for the global thermodynamic and kinetic properties. These result from a double funnel energy landscape [31] [32] where the two families of structures are separated by a relatively high energy barrier at the temperature where they are in thermodynamic equilibrium. This arises for a cluster of 75 atoms with the Lennard-Jones potential, where the global potential energy minimum is decahedral, and structures based upon incomplete Mackay icosahedra [33] are also low in potential energy, but higher in entropy. The free energy barrier between these families is large compared to the available thermal energy at the temperature where they are in equilibrium. An example is shown in the figure, with probability in the lower part and energy above with axes of an order parameter and temperature . At low temperature the 75 atom decahedral cluster (Dh) is the global free energy minimum, but as the temperature increases the higher entropy of the competing structures based on incomplete icosahedra (Ic) causes the finite system analogue of a first-order phase transition; at even higher temperatures a liquid-like state is favored. [12]

Ubiquity

Electron micrograph of two Icosahedral adenoviruses, with an illustration to show the shape. Icosahedral Adenoviruses.jpg
Electron micrograph of two Icosahedral adenoviruses, with an illustration to show the shape.

Most modern analysis of these shapes in nanoparticles started with the observation of icosahedral and decahedral particles by Shozo Ino and Shiro Ogawa in 1966-67, and independently but slightly later (which they acknowledged) in work by John Allpress and John Veysey Sanders. In both cases these were for vacuum deposition of metal onto substrates in very clean (ultra-high vacuum) conditions, where nanoparticle islands of size 10-50 nm were formed during thin film growth. Using transmission electron microscopy and diffraction these authors demonstrated the presence of the units in the particles, and also the twin relationships. They called the five-fold and icosahedral crystals multiply twinned particles (MTPs). In the early work near perfect icosahedron shapes were formed, so they were called icosahedral MTPs, the names connecting to the icosahedral () point group symmetry.These forms occur for both elemental nanoparticles [34] [35] as well as alloys [36] [37] and colloidal crystals. [38] A related form also exists in icosahedral viruses. [39] [40]

Quasicrystals are un-twinned structures with long range rotational but not translational periodicity, that some initially tried to explain away as icosahedral twinning. [41]

See also

Related Research Articles

<span class="mw-page-title-main">Quasicrystal</span> Ordered chemical structure with no repeating pattern

A quasiperiodic crystal, or quasicrystal, is a structure that is ordered but not periodic. A quasicrystalline pattern can continuously fill all available space, but it lacks translational symmetry. While crystals, according to the classical crystallographic restriction theorem, can possess only two-, three-, four-, and six-fold rotational symmetries, the Bragg diffraction pattern of quasicrystals shows sharp peaks with other symmetry orders—for instance, five-fold.

<span class="mw-page-title-main">Quantum dot</span> Nano-scale semiconductor particles

Quantum dots (QDs) or semiconductor nanocrystals are semiconductor particles a few nanometres in size with optical and electronic properties that differ from those of larger particles via quantum mechanical effects. They are a central topic in nanotechnology and materials science. When a quantum dot is illuminated by UV light, an electron in the quantum dot can be excited to a state of higher energy. In the case of a semiconducting quantum dot, this process corresponds to the transition of an electron from the valence band to the conduction band. The excited electron can drop back into the valence band releasing its energy as light. This light emission (photoluminescence) is illustrated in the figure on the right. The color of that light depends on the energy difference between the discrete energy levels of the quantum dot in the conduction band and the valence band.

A thin film is a layer of materials ranging from fractions of a nanometer (monolayer) to several micrometers in thickness. The controlled synthesis of materials as thin films is a fundamental step in many applications. A familiar example is the household mirror, which typically has a thin metal coating on the back of a sheet of glass to form a reflective interface. The process of silvering was once commonly used to produce mirrors, while more recently the metal layer is deposited using techniques such as sputtering. Advances in thin film deposition techniques during the 20th century have enabled a wide range of technological breakthroughs in areas such as magnetic recording media, electronic semiconductor devices, integrated passive devices, light-emitting diodes, optical coatings, hard coatings on cutting tools, and for both energy generation and storage. It is also being applied to pharmaceuticals, via thin-film drug delivery. A stack of thin films is called a multilayer.

<span class="mw-page-title-main">Nanoparticle</span> Particle with size less than 100 nm

A nanoparticle or ultrafine particle is a particle of matter 1 to 100 nanometres (nm) in diameter. The term is sometimes used for larger particles, up to 500 nm, or fibers and tubes that are less than 100 nm in only two directions. At the lowest range, metal particles smaller than 1 nm are usually called atom clusters instead.

Precipitation hardening, also called age hardening or particle hardening, is a heat treatment technique used to increase the yield strength of malleable materials, including most structural alloys of aluminium, magnesium, nickel, titanium, and some steels, stainless steels, and duplex stainless steel. In superalloys, it is known to cause yield strength anomaly providing excellent high-temperature strength.

<span class="mw-page-title-main">Crystal twinning</span> Two separate crystals sharing some of the same crystal lattice points in a symmetrical manner

Crystal twinning occurs when two or more adjacent crystals of the same mineral are oriented so that they share some of the same crystal lattice points in a symmetrical manner. The result is an intergrowth of two separate crystals that are tightly bonded to each other. The surface along which the lattice points are shared in twinned crystals is called a composition surface or twin plane.

Gold clusters in cluster chemistry can be either discrete molecules or larger colloidal particles. Both types are described as nanoparticles, with diameters of less than one micrometer. A nanocluster is a collective group made up of a specific number of atoms or molecules held together by some interaction mechanism. Gold nanoclusters have potential applications in optoelectronics and catalysis.

Nanogeoscience is the study of nanoscale phenomena related to geological systems. Predominantly, this is investigated by studying environmental nanoparticles between 1–100 nanometers in size. Other applicable fields of study include studying materials with at least one dimension restricted to the nanoscale and the transfer of energy, electrons, protons, and matter across environmental interfaces.

Melting-point depression is the phenomenon of reduction of the melting point of a material with a reduction of its size. This phenomenon is very prominent in nanoscale materials, which melt at temperatures hundreds of degrees lower than bulk materials.

Stranski–Krastanov growth is one of the three primary modes by which thin films grow epitaxially at a crystal surface or interface. Also known as 'layer-plus-island growth', the SK mode follows a two step process: initially, complete films of adsorbates, up to several monolayers thick, grow in a layer-by-layer fashion on a crystal substrate. Beyond a critical layer thickness, which depends on strain and the chemical potential of the deposited film, growth continues through the nucleation and coalescence of adsorbate 'islands'. This growth mechanism was first noted by Ivan Stranski and Lyubomir Krastanov in 1938. It wasn't until 1958 however, in a seminal work by Ernst Bauer published in Zeitschrift für Kristallographie, that the SK, Volmer–Weber, and Frank–van der Merwe mechanisms were systematically classified as the primary thin-film growth processes. Since then, SK growth has been the subject of intense investigation, not only to better understand the complex thermodynamics and kinetics at the core of thin-film formation, but also as a route to fabricating novel nanostructures for application in the microelectronics industry.

In crystallography, a disclination is a line defect in which there is compensation of an angular gap. They were first discussed by Vito Volterra in 1907, who provided an analysis of the elastic strains of a wedge disclination. By analogy to dislocations in crystals, the term, disinclination, was first used by Frederick Charles Frank and since then has been modified to its current usage, disclination. They have since been analyzed in some detail particularly by Roland deWit.

<span class="mw-page-title-main">Wulff construction</span> Lowest energy shape of a single crystal

The Wulff construction is a method to determine the equilibrium shape of a droplet or crystal of fixed volume inside a separate phase. Energy minimization arguments are used to show that certain crystal planes are preferred over others, giving the crystal its shape. It is of fundamental importance in a number of areas ranging from the shape of nanoparticles and precipitates to nucleation. It also has more applied relevance in areas such as the shapes of active particles in heterogeneous catalysis.

<span class="mw-page-title-main">Surface stress</span> Change of surface energy with strain

Surface stress was first defined by Josiah Willard Gibbs (1839–1903) as the amount of the reversible work per unit area needed to elastically stretch a pre-existing surface. Depending upon the convention used, the area is either the original, unstretched one which represents a constant number of atoms, or sometimes is the final area; these are atomistic versus continuum definitions. Some care is needed to ensure that the definition used is also consistent with the elastic strain energy, and misinterpretations and disagreements have occurred in the literature.

<span class="mw-page-title-main">Silver nanoparticle</span> Ultrafine particles of silver between 1 nm and 100 nm in size

Silver nanoparticles are nanoparticles of silver of between 1 nm and 100 nm in size. While frequently described as being 'silver' some are composed of a large percentage of silver oxide due to their large ratio of surface to bulk silver atoms. Numerous shapes of nanoparticles can be constructed depending on the application at hand. Commonly used silver nanoparticles are spherical, but diamond, octagonal, and thin sheets are also common.

<span class="mw-page-title-main">Self-assembly of nanoparticles</span> Physical phenomenon

Nanoparticles are classified as having at least one of its dimensions in the range of 1-100 nanometers (nm). The small size of nanoparticles allows them to have unique characteristics which may not be possible on the macro-scale. Self-assembly is the spontaneous organization of smaller subunits to form larger, well-organized patterns. For nanoparticles, this spontaneous assembly is a consequence of interactions between the particles aimed at achieving a thermodynamic equilibrium and reducing the system’s free energy. The thermodynamics definition of self-assembly was introduced by Professor Nicholas A. Kotov. He describes self-assembly as a process where components of the system acquire non-random spatial distribution with respect to each other and the boundaries of the system. This definition allows one to account for mass and energy fluxes taking place in the self-assembly processes.

<span class="mw-page-title-main">Heterogeneous gold catalysis</span>

Heterogeneous gold catalysis refers to the use of elemental gold as a heterogeneous catalyst. As in most heterogeneous catalysis, the metal is typically supported on metal oxide. Furthermore, as seen in other heterogeneous catalysts, activity increases with a decreasing diameter of supported gold clusters. Several industrially relevant processes are also observed such as H2 activation, Water-gas shift reaction, and hydrogenation. One or two gold-catalyzed reactions may have been commercialized.

In statistical mechanics, the KTHNY-theory describes the process of melting of crystals in two dimensions (2D). The name is derived from the initials of the surnames of John Michael Kosterlitz, David J. Thouless, Bertrand Halperin, David R. Nelson, and A. Peter Young, who developed the theory in the 1970s. It is, beside the Ising model in 2D and the XY model in 2D, one of the few theories, which can be solved analytically and which predicts a phase transition at a temperature .

<span class="mw-page-title-main">Shape control in nanocrystal growth</span> Influences on the shape of small crystals

Shape control in nanocrystal growth is the control of the shape of nanocrystals formed in their synthesis by means of varying reaction conditions. This is a concept studied in nanosciences, which is a part of both chemistry and condensed matter physics. There are two processes involved in the growth of these nanocrystals. Firstly, volume Gibbs free energy of the system containing the nanocrystal in solution decreases as the nanocrystal size increases. Secondly, each crystal has a surface Gibbs free energy that can be minimized by adopting the shape that is energetically most favorable. Surface energies of crystal planes are related to their Miller indices, which is why these can help predict the equilibrium shape of a certain nanocrystal.

<span class="mw-page-title-main">Fiveling</span> Five crystals arranged round a common axis

A fiveling, also known as a decahedral nanoparticle, a multiply-twinned particle (MTP), a pentagonal nanoparticle, a pentatwin, or a five-fold twin is a type of twinned crystal that can exist at sizes ranging from nanometers to millimetres. It contains five different single crystals arranged around a common axis. In most cases each unit has a face centered cubic (fcc) arrangement of the atoms, although they are also known for other types of crystal structure.

<span class="mw-page-title-main">Extended Wulff constructions</span> Shapes for crystals with interfaces and twins

Extended Wulff constructions refer to a number of variants of the Wulff construction which is used for a solid single crystal in isolation. They include cases for solid particle on substrates, those with twins and also when growth is important. They are important for many applications such as supported metal nanoparticles used in heterogeneous catalysis or for understanding the shape of twinned nanoparticles being explored for other applications such as drug delivery or optical communications. They are also relevant for macroscopic crystals with twins. Depending upon whether there are twins or a substrate there are different cases as indicated in the decision tree figure.

References

  1. Hofmeister, H. (1998). "Forty Years Study of Fivefold Twinned Structures in Small Particles and Thin Films". Crystal Research and Technology. 33 (1): 3–25. Bibcode:1998CryRT..33....3H. doi:10.1002/(sici)1521-4079(1998)33:1<3::aid-crat3>3.0.co;2-3. ISSN   0232-1300.
  2. H. Hofmeister (2004) "Fivefold twinned nanoparticles" in Encyclopedia of Nanoscience and Nanotechnology (ed. H. S. Nalwa, Amer. Sci. Publ., Stevenson Ranch CA) vol. 3, pp. 431-452 ISBN   1-58883-059-4 pdf.
  3. 1 2 3 4 Marks, L D; Peng, L (2016). "Nanoparticle shape, thermodynamics and kinetics". Journal of Physics: Condensed Matter. 28 (5): 053001. Bibcode:2016JPCM...28e3001M. doi:10.1088/0953-8984/28/5/053001. ISSN   0953-8984. PMID   26792459. S2CID   12503859.
  4. 1 2 3 Ino, Shozo (1969). "Stability of Multiply-Twinned Particles". Journal of the Physical Society of Japan. 27 (4): 941–953. Bibcode:1969JPSJ...27..941I. doi:10.1143/jpsj.27.941. ISSN   0031-9015.
  5. Boukouvala, Christina; Daniel, Joshua; Ringe, Emilie (2021). "Approaches to modelling the shape of nanocrystals". Nano Convergence. 8 (1): 26. Bibcode:2021NanoC...8...26B. doi: 10.1186/s40580-021-00275-6 . ISSN   2196-5404. PMC   8429535 . PMID   34499259.
  6. Rahm, J.; Erhart, Paul (2020). "WulffPack: A Python package for Wulff constructions". Journal of Open Source Software. 5 (45): 1944. Bibcode:2020JOSS....5.1944R. doi: 10.21105/joss.01944 . ISSN   2475-9066.
  7. Wit, R de (1972). "Partial disclinations". Journal of Physics C: Solid State Physics. 5 (5): 529–534. Bibcode:1972JPhC....5..529D. doi:10.1088/0022-3719/5/5/004. ISSN   0022-3719.
  8. 1 2 Howie, A.; Marks, L. D. (1984). "Elastic strains and the energy balance for multiply twinned particles". Philosophical Magazine A. 49 (1): 95–109. Bibcode:1984PMagA..49...95H. doi:10.1080/01418618408233432. ISSN   0141-8610.
  9. Pimpinelli, Alberto; Villain, Jacques (1998). Physics of Crystal Growth (1 ed.). Cambridge University Press. doi:10.1017/cbo9780511622526. ISBN   978-0-521-55198-4.
  10. Baletto, Francesca; Ferrando, Riccardo (2005). "Structural properties of nanoclusters: Energetic, thermodynamic, and kinetic effects". Reviews of Modern Physics. 77 (1): 371–423. Bibcode:2005RvMP...77..371B. doi:10.1103/RevModPhys.77.371. ISSN   0034-6861. S2CID   54700637.
  11. Wei; Vajtai, Robert; Jung, Yung Joon; Banhart, Florian; Ramanath, Ganapathiraman; Ajayan, Pulickel M. (2002). "Massive Icosahedral Boron Carbide Crystals". The Journal of Physical Chemistry B. 106 (23): 5807–5809. doi:10.1021/jp014640f. ISSN   1520-6106.
  12. 1 2 Wales, David J. (2013). "Surveying a complex potential energy landscape: Overcoming broken ergodicity using basin-sampling". Chemical Physics Letters. 584: 1–9. Bibcode:2013CPL...584....1W. doi:10.1016/j.cplett.2013.07.066.
  13. Vermaak, J.S.; Mays, C.W.; Kuhlmann-Wilsdorf, D. (1968). "On surface stress and surface tension". Surface Science. 12 (2): 128–133. doi:10.1016/0039-6028(68)90118-0.
  14. Mays, C.W.; Vermaak, J.S.; Kuhlmann-Wilsdorf, D. (1968). "On surface stress and surface tension". Surface Science. 12 (2): 134–140. Bibcode:1968SurSc..12..134M. doi:10.1016/0039-6028(68)90119-2.
  15. Müller, Pierre; Saùl, Andres; Leroy, Frédéric (2013). "Simple views on surface stress and surface energy concepts". Advances in Natural Sciences: Nanoscience and Nanotechnology. 5 (1): 013002. doi: 10.1088/2043-6262/5/1/013002 . ISSN   2043-6262.
  16. Patala, Srikanth; Marks, Laurence D.; Olvera de la Cruz, Monica (2013). "Thermodynamic Analysis of Multiply Twinned Particles: Surface Stress Effects". The Journal of Physical Chemistry Letters. 4 (18): 3089–3094. doi:10.1021/jz401496d. ISSN   1948-7185.
  17. Romanov, Alexey E.; Vikarchuk, Anatoly A.; Kolesnikova, Anna L.; Dorogin, Leonid M.; Kink, Ilmar; Aifantis, Elias C. (2012). "Structural transformations in nano- and microobjects triggered by disclinations". Journal of Materials Research. 27 (3): 545–551. Bibcode:2012JMatR..27..545R. doi:10.1557/jmr.2011.372. ISSN   0884-2914.
  18. Feibelman, Peter J. (1997). "First-principles calculations of stress induced by gas adsorption on Pt(111)". Physical Review B. 56 (4): 2175–2182. Bibcode:1997PhRvB..56.2175F. doi:10.1103/PhysRevB.56.2175. ISSN   0163-1829.
  19. Graoui, H.; Giorgio, S.; Henry, C.R. (1998). "Shape variations of Pd particles under oxygen adsorption". Surface Science. 417 (2–3): 350–360. Bibcode:1998SurSc.417..350G. doi:10.1016/S0039-6028(98)00688-8.
  20. Wynblatt, P.; Chatain, D. (2009). "Surface segregation anisotropy and the equilibrium crystal shape of alloy crystals". Reviews on Advanced Materials Science. 21: 44–56. S2CID   137869647.
  21. Hoare, M.R.; Pal, P. (1971). "Physical cluster mechanics: Statics and energy surfaces for monatomic systems". Advances in Physics. 20 (84): 161–196. Bibcode:1971AdPhy..20..161H. doi:10.1080/00018737100101231. ISSN   0001-8732.
  22. Berry, R. Stephen; Jellinek, Julius; Natanson, Grigory (1984). "Melting of clusters and melting". Physical Review A. 30 (2): 919–931. Bibcode:1984PhRvA..30..919B. doi:10.1103/PhysRevA.30.919. ISSN   0556-2791.
  23. Berry, R. Stephen. (1993). "Potential surfaces and dynamics: what clusters tell us". Chemical Reviews. 93 (7): 2379–2394. doi:10.1021/cr00023a003. ISSN   0009-2665.
  24. Ajayan, P. M.; Marks, L. D. (1988). "Quasimelting and phases of small particles". Physical Review Letters. 60 (7): 585–587. Bibcode:1988PhRvL..60..585A. doi:10.1103/PhysRevLett.60.585. ISSN   0031-9007. PMID   10038590.
  25. Barnard, Amanda S.; Young, Neil P.; Kirkland, Angus I.; van Huis, Marijn A.; Xu, Huifang (2009). "Nanogold: A Quantitative Phase Map". ACS Nano. 3 (6): 1431–1436. doi:10.1021/nn900220k. ISSN   1936-0851. PMID   19489558.
  26. Uppenbrink, Julia; Wales, David J. (1992). "Structure and energetics of model metal clusters". The Journal of Chemical Physics. 96 (11): 8520–8534. Bibcode:1992JChPh..96.8520U. doi:10.1063/1.462305. ISSN   0021-9606.
  27. Wales, David J.; Doye, Jonathan P. K.; Miller, Mark A.; Mortenson, Paul N.; Walsh, Tiffany R. (2000). Prigogine, I.; Rice, Stuart A. (eds.). Energy Landscapes: From Clusters to Biomolecules. Vol. 115 (1 ed.). Wiley. pp. 39–46. doi:10.1002/9780470141748.ch1. ISBN   978-0-471-39331-3 . Retrieved 2024-04-01.
  28. Wales, David J. (2018). "Exploring Energy Landscapes". Annual Review of Physical Chemistry. 69 (1): 401–425. Bibcode:2018ARPC...69..401W. doi:10.1146/annurev-physchem-050317-021219. ISSN   0066-426X. PMID   29677468.
  29. Zhang, Huaizhong; Khan, Mohd Ahmed; Yan, Tianyu; Fichthorn, Kristen A. (2024). "Size and temperature dependent shapes of copper nanocrystals using parallel tempering molecular dynamics". Nanoscale. 16 (23): 11146–11155. doi: 10.1039/D4NR00317A . ISSN   2040-3364. PMID   38506642.
  30. Mottet, C.; Goniakowski, J.; Baletto, F.; Ferrando, R.; Treglia, G. (2004). "Modeling free and supported metallic nanoclusters: structure and dynamics". Phase Transitions. 77 (1–2): 101–113. Bibcode:2004PhaTr..77..101M. doi:10.1080/1411590310001622473. ISSN   0141-1594.
  31. Wales, David (2001). Energy Landscapes: Applications to Clusters, Biomolecules and Glasses (1 ed.). Cambridge University Press. p. 4590479. doi:10.1017/cbo9780511721724. ISBN   978-0-521-81415-7.
  32. Wales, David J.; Miller, Mark A.; Walsh, Tiffany R. (1998). "Archetypal energy landscapes". Nature. 394 (6695): 758–760. Bibcode:1998Natur.394..758W. doi:10.1038/29487. ISSN   0028-0836.
  33. Mackay, A. L. (1962). "A dense non-crystallographic packing of equal spheres". Acta Crystallographica. 15 (9): 916–918. Bibcode:1962AcCry..15..916M. doi:10.1107/S0365110X6200239X. ISSN   0365-110X.
  34. Zhang, Qingbo; Xie, Jianping; Yang, Jinhua; Lee, Jim Yang (2009). "Monodisperse Icosahedral Ag, Au, and Pd Nanoparticles: Size Control Strategy and Superlattice Formation". ACS Nano. 3 (1): 139–148. doi:10.1021/nn800531q. ISSN   1936-0851. PMID   19206260.
  35. Wang, Helan; Zhou, Shan; Gilroy, Kyle D.; Cai, Zaisheng; Xia, Younan (2017). "Icosahedral nanocrystals of noble metals: Synthesis and applications". Nano Today. 15: 121–144. doi:10.1016/j.nantod.2017.06.011. ISSN   1748-0132.
  36. Akbarzadeh, Hamed; Abbaspour, Mohsen; Mehrjouei, Esmat (2017). "Au@Pt and Pt@Au nanoalloys in the icosahedral and cuboctahedral structures: Which is more stable?". Journal of Molecular Liquids. 242: 1002–1017. doi:10.1016/j.molliq.2017.07.096. ISSN   0167-7322.
  37. Wu, Jianbo; Qi, Liang; You, Hongjun; Gross, Adam; Li, Ju; Yang, Hong (2012). "Icosahedral Platinum Alloy Nanocrystals with Enhanced Electrocatalytic Activities". Journal of the American Chemical Society. 134 (29): 11880–11883. Bibcode:2012JAChS.13411880W. doi:10.1021/ja303950v. ISSN   0002-7863. PMID   22738173.
  38. Xu, Meng; Kim, Eun Ji; Lee, Young Jun; Lee, Hyunsoo; Jung, Kyunghyun; Choi, Jaeyoung; Kim, Shin-Hyun; Kim, YongJoo; Yun, Hongseok; Kim, Bumjoon J. (2024). "Icosahedral supracrystal assembly from polymer-grafted nanoparticles via interplay of interfacial energy and confinement effect". Science Advances. 10 (24): eado0745. Bibcode:2024SciA...10O.745X. doi:10.1126/sciadv.ado0745. PMC   11177942 . PMID   38875331.
  39. Speir, J.A.; Johnson, J.E.; Munshi, S.; Wang, G.; Timothy, S.; Baker, T.S. (1995). "Structures of the native and swollen forms of cowpea chlorotic mottle virus determined by x-ray crystallography and cryo-electron microscopy". doi.org. Retrieved 2025-01-20.
  40. Böttcher, B.; Wynne, S. A.; Crowther, R. A. (1997). "Determination of the fold of the core protein of hepatitis B virus by electron cryomicroscopy". Nature. 386 (6620): 88–91. doi:10.1038/386088a0. ISSN   0028-0836.
  41. Pauling, Linus (1987). "So-called icosahedral and decagonal quasicrystals are twins of an 820-atom cubic crystal". Physical Review Letters. 58 (4). American Physical Society (APS): 365–368. Bibcode:1987PhRvL..58..365P. doi:10.1103/physrevlett.58.365. ISSN   0031-9007. PMID   10034915.