Rigidity theory (physics)

Last updated

Rigidity theory, or topological constraint theory, is a tool for predicting properties of complex networks (such as glasses) based on their composition. It was introduced by James Charles Phillips in 1979 [1] and 1981, [2] and refined by Michael Thorpe in 1983. [3] Inspired by the study of the stability of mechanical trusses as pioneered by James Clerk Maxwell, [4] and by the seminal work on glass structure done by William Houlder Zachariasen, [5] this theory reduces complex molecular networks to nodes (atoms, molecules, proteins, etc.) constrained by rods (chemical constraints), thus filtering out microscopic details that ultimately don't affect macroscopic properties. An equivalent theory was developed by P.K. Gupta A.R. Cooper in 1990, where rather than nodes representing atoms, they represented unit polytopes. [6] An example of this would be the SiO tetrahedra in pure glassy silica. This style of analysis has applications in biology and chemistry, such as understanding adaptability in protein-protein interaction networks. [7] Rigidity theory applied to the molecular networks arising from phenotypical expression of certain diseases may provide insights regarding their structure and function.

Contents

In molecular networks, atoms can be constrained by radial 2-body bond-stretching constraints, which keep interatomic distances fixed, and angular 3-body bond-bending constraints, which keep angles fixed around their average values. As stated by Maxwell's criterion, a mechanical truss is isostatic when the number of constraints equals the number of degrees of freedom of the nodes. In this case, the truss is optimally constrained, being rigid but free of stress. This criterion has been applied by Phillips to molecular networks, which are called flexible, stressed-rigid or isostatic when the number of constraints per atoms is respectively lower, higher or equal to 3, the number of degrees of freedom per atom in a three-dimensional system. [8] The same condition applies to random packing of spheres, which are isostatic at the jamming point. Typically, the conditions for glass formation will be optimal if the network is isostatic, which is for example the case for pure silica. [9] Flexible systems show internal degrees of freedom, called floppy modes, [3] whereas stressed-rigid ones are complexity locked by the high number of constraints and tend to crystallize instead of forming glass during a quick quenching.

Derivation of isostatic condition

The conditions for isostaticity can be derived by looking at the internal degrees of freedom of a general 3D network. For nodes, constraints, and equations of equilibrium, the number of degrees of freedom is

The node term picks up a factor of 3 due to there being translational degrees of freedom in the x, y, and z directions. By similar reasoning, in 3D, as there is one equation of equilibrium for translational and rotational modes in each dimension. This yields

This can be applied to each node in the system by normalizing by the number of nodes

where , , and the last term has been dropped since for atomistic systems . Isostatic conditions are achieved when , yielding the number of constraints per atom in the isostatic condition of .

An alternative derivation is based on analyzing the shear modulus of the 3D network or solid structure. The isostatic condition, which represents the limit of mechanical stability, is equivalent to setting in a microscopic theory of elasticity that provides as a function of the internal coordination number of nodes and of the number of degrees of freedom. The problem has been solved by Alessio Zaccone and E. Scossa-Romano in 2011, who derived the analytical formula for the shear modulus of a 3D network of central-force springs (bond-stretching constraints): . [10] Here, is the spring constant, is the distance between two nearest-neighbor nodes, the average coordination number of the network (note that here and ), and in 3D. A similar formula has been derived for 2D networks where the prefactor is instead of . Hence, based on the Zaccone–Scossa-Romano expression for , upon setting , one obtains , or equivalently in different notation, , which defines the Maxwell isostatic condition. A similar analysis can be done for 3D networks with bond-bending interactions (on top of bond-stretching), which leads to the isostatic condition , with a lower threshold due to the angular constraints imposed by bond-bending. [11]

Developments in glass science

Rigidity theory allows the prediction of optimal isostatic compositions, as well as the composition dependence of glass properties, by a simple enumeration of constraints. [12] These glass properties include, but are not limited to, elastic modulus, shear modulus, bulk modulus, density, Poisson's ratio, coefficient of thermal expansion, hardness, [13] and toughness. In some systems, due to the difficulty of directly enumerating constraints by hand and knowing all system information a priori, the theory is often employed in conjunction with computational methods in materials science such as molecular dynamics (MD). Notably, the theory played a major role in the development of Gorilla Glass 3. [14] Extended to glasses at finite temperature [15] and finite pressure, [16] rigidity theory has been used to predict glass transition temperature, viscosity and mechanical properties. [8] It was also applied to granular materials [17] and proteins. [18]

In the context of soft glasses, rigidity theory has been used by Alessio Zaccone and Eugene Terentjev to predict the glass transition temperature of polymers and to provide a molecular-level derivation and interpretation of the Flory–Fox equation. [19] The Zaccone–Terentjev theory also provides an expression for the shear modulus of glassy polymers as a function of temperature which is in quantitative agreement with experimental data, and is able to describe the many orders of magnitude drop of the shear modulus upon approaching the glass transition from below. [19]

In 2001, Boolchand and coworkers found that the isostatic compositions in glassy alloys—predicted by rigidity theory—exist not just at a single threshold composition; rather, in many systems it spans a small, well-defined range of compositions intermediate to the flexible (under-constrained) and stressed-rigid (over-constrained) domains. [20] This window of optimally constrained glasses is thus referred to as the intermediate phase or the reversibility window, as the glass formation is supposed to be reversible, with minimal hysteresis, inside the window. [20] Its existence has been attributed to the glassy network consisting almost exclusively of a varying population of isostatic molecular structures. [16] [21] The existence of the intermediate phase remains a controversial, but stimulating topic in glass science.


See also


Related Research Articles

In condensed matter physics and materials science, an amorphous solid is a solid that lacks the long-range order that is characteristic of a crystal. The terms "glass" and "glassy solid" are sometimes used synonymously with amorphous solid; however, these terms refer specifically to amorphous materials that undergo a glass transition. Examples of amorphous solids include glasses, metallic glasses, and certain types of plastics and polymers.

<span class="mw-page-title-main">Melting</span> Material phase change

Melting, or fusion, is a physical process that results in the phase transition of a substance from a solid to a liquid. This occurs when the internal energy of the solid increases, typically by the application of heat or pressure, which increases the substance's temperature to the melting point. At the melting point, the ordering of ions or molecules in the solid breaks down to a less ordered state, and the solid melts to become a liquid.

<span class="mw-page-title-main">Spin glass</span> Disordered magnetic state

In condensed matter physics, a spin glass is a magnetic state characterized by randomness, besides cooperative behavior in freezing of spins at a temperature called 'freezing temperature' Tf. In ferromagnetic solids, component atoms' magnetic spins all align in the same direction. Spin glass when contrasted with a ferromagnet is defined as "disordered" magnetic state in which spins are aligned randomly or without a regular pattern and the couplings too are random.

<span class="mw-page-title-main">Amorphous metal</span> Solid metallic material with disordered atomic-scale structure

An amorphous metal is a solid metallic material, usually an alloy, with disordered atomic-scale structure. Most metals are crystalline in their solid state, which means they have a highly ordered arrangement of atoms. Amorphous metals are non-crystalline, and have a glass-like structure. But unlike common glasses, such as window glass, which are typically electrical insulators, amorphous metals have good electrical conductivity and can show metallic luster.

<span class="mw-page-title-main">Shear modulus</span> Ratio of shear stress to shear strain

In materials science, shear modulus or modulus of rigidity, denoted by G, or sometimes S or μ, is a measure of the elastic shear stiffness of a material and is defined as the ratio of shear stress to the shear strain:

<span class="mw-page-title-main">Boron trioxide</span> Chemical compound

Boron trioxide or diboron trioxide is the oxide of boron with the formula B2O3. It is a colorless transparent solid, almost always glassy (amorphous), which can be crystallized only with great difficulty. It is also called boric oxide or boria. It has many important industrial applications, chiefly in ceramics as a flux for glazes and enamels and in the production of glasses.

Chalcogenide glass is a glass containing one or more chalcogens. Up until recently, chalcogenide glasses (ChGs) were believed to be predominantly covalently bonded materials and classified as covalent network solids. A most recent and extremely comprehensive university study of more than 265 different ChG elemental compositions, representing 40 different elemental families now shows that the vast majority of chalcogenide glasses are more accurately defined as being predominantly bonded by the weaker van der Waals forces of atomic physics and more accurately classified as van der Waals network solids. They are not exclusively bonded by these weaker vdW forces, and do exhibit varying percentages of covalency, based upon their specific chemical makeup. Polonium is also a chalcogen but is not used because of its strong radioactivity. Chalcogenide materials behave rather differently from oxides, in particular their lower band gaps contribute to very dissimilar optical and electrical properties.

First introduced by M. Pollak, the Coulomb gap is a soft gap in the single-particle density of states (DOS) of a system of interacting localized electrons. Due to the long-range Coulomb interactions, the single-particle DOS vanishes at the chemical potential, at low enough temperatures, such that thermal excitations do not wash out the gap.

<span class="mw-page-title-main">Jamming (physics)</span>

Jamming is the physical process by which the viscosity of some mesoscopic materials, such as granular materials, glasses, foams, polymers, emulsions, and other complex fluids, increases with increasing particle density. The jamming transition has been proposed as a new type of phase transition, with similarities to a glass transition but very different from the formation of crystalline solids.

In applied mathematics, the numerical sign problem is the problem of numerically evaluating the integral of a highly oscillatory function of a large number of variables. Numerical methods fail because of the near-cancellation of the positive and negative contributions to the integral. Each has to be integrated to very high precision in order for their difference to be obtained with useful accuracy.

The chameleon is a hypothetical scalar particle that couples to matter more weakly than gravity, postulated as a dark energy candidate. Due to a non-linear self-interaction, it has a variable effective mass which is an increasing function of the ambient energy density—as a result, the range of the force mediated by the particle is predicted to be very small in regions of high density but much larger in low-density intergalactic regions: out in the cosmos chameleon models permit a range of up to several thousand parsecs. As a result of this variable mass, the hypothetical fifth force mediated by the chameleon is able to evade current constraints on equivalence principle violation derived from terrestrial experiments even if it couples to matter with a strength equal or greater than that of gravity. Although this property would allow the chameleon to drive the currently observed acceleration of the universe's expansion, it also makes it very difficult to test for experimentally.

The glass–liquid transition, or glass transition, is the gradual and reversible transition in amorphous materials from a hard and relatively brittle "glassy" state into a viscous or rubbery state as the temperature is increased. An amorphous solid that exhibits a glass transition is called a glass. The reverse transition, achieved by supercooling a viscous liquid into the glass state, is called vitrification.

In polymer chemistry and polymer physics, the Flory–Fox equation is a simple empirical formula that relates molecular weight to the glass transition temperature of a polymer system. The equation was first proposed in 1950 by Paul J. Flory and Thomas G. Fox while at Cornell University. Their work on the subject overturned the previously held theory that the glass transition temperature was the temperature at which viscosity reached a maximum. Instead, they demonstrated that the glass transition temperature is the temperature at which the free space available for molecular motions achieved a minimum value. While its accuracy is usually limited to samples of narrow range molecular weight distributions, it serves as a good starting point for more complex structure-property relationships.

<span class="mw-page-title-main">Structure of liquids and glasses</span> Atomic-scale non-crystalline structure of liquids and glasses

The structure of liquids, glasses and other non-crystalline solids is characterized by the absence of long-range order which defines crystalline materials. Liquids and amorphous solids do, however, possess a rich and varied array of short to medium range order, which originates from chemical bonding and related interactions. Metallic glasses, for example, are typically well described by the dense random packing of hard spheres, whereas covalent systems, such as silicate glasses, have sparsely packed, strongly bound, tetrahedral network structures. These very different structures result in materials with very different physical properties and applications.

<span class="mw-page-title-main">Fragility (glass physics)</span>

In glass physics, fragility characterizes how rapidly the dynamics of a material slows down as it is cooled toward the glass transition: materials with a higher fragility have a relatively narrow glass transition temperature range, while those with low fragility have a relatively broad glass transition temperature range. Physically, fragility may be related to the presence of dynamical heterogeneity in glasses, as well as to the breakdown of the usual Stokes–Einstein relationship between viscosity and diffusion. Fragility has no direct relationship with the colloquial meaning of the word "fragility", which more closely relates to the brittleness of a material.

A fracton is an emergent topological quasiparticle excitation which is immobile when in isolation. Many theoretical systems have been proposed in which fractons exist as elementary excitations. Such systems are known as fracton models. Fractons have been identified in various CSS codes as well as in symmetric tensor gauge theories.

A set of networks that satisfies given structural characteristics can be treated as a network ensemble. Brought up by Ginestra Bianconi in 2007, the entropy of a network ensemble measures the level of the order or uncertainty of a network ensemble.

Alessio Zaccone is an Italian physicist.

A dipole glass is an analog of a glass where the dipoles are frozen below a given freezing temperature Tf introducing randomness thus resulting in a lack of long-range ferroelectric order. A dipole glass is very similar to the concept of a spin glass where the atomic spins don't all align in the same direction and thus result in a net-zero magnetization. The randomness of dipoles in a dipole glass creates local fields resulting in short-range order but no long-range order.

References

  1. Phillips, J. C. (1979). "Topology of covalent non-crystalline solids I: Short-range order in chalcogenide alloys". Journal of Non-Crystalline Solids. 34 (2): 153–181. Bibcode:1979JNCS...34..153P. doi:10.1016/0022-3093(79)90033-4.
  2. Phillips, J. C. (1981-01-01). "Topology of covalent non-crystalline solids II: Medium-range order in chalcogenide alloys and A-Si(Ge)". Journal of Non-Crystalline Solids. 43 (1): 37–77. Bibcode:1981JNCS...43...37P. doi:10.1016/0022-3093(81)90172-1. ISSN   0022-3093.
  3. 1 2 Thorpe, M. F. (1983). "Continuous deformations in random networks". Journal of Non-Crystalline Solids. 57 (3): 355–370. Bibcode:1983JNCS...57..355T. doi:10.1016/0022-3093(83)90424-6.
  4. Maxwell, J. Clerk (April 1864). "XLV. On reciprocal figures and diagrams of forces". The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science. 27 (182): 250–261. doi:10.1080/14786446408643663. ISSN   1941-5982.
  5. Zachariasen, W. H. (October 1932). "The Atomic Arrangement in Glass". Journal of the American Chemical Society. 54 (10): 3841–3851. doi:10.1021/ja01349a006. ISSN   0002-7863.
  6. Gupta, P. K.; Cooper, A. R. (1990-08-02). "Topologically disordered networks of rigid polytopes". Journal of Non-Crystalline Solids. XVth International Congress on Glass. 123 (1): 14–21. Bibcode:1990JNCS..123...14G. doi:10.1016/0022-3093(90)90768-H. ISSN   0022-3093.
  7. Sharma, Ankush; Maria Brigida Ferraro; Maiorano, Francesco; Francesca Del Vecchio Blanco; Mario Rosario Guarracino (2014). "Rigidity and flexibility in protein-protein interaction networks: A case study on neuromuscular disorders". arXiv: 1402.2304 [q-bio.MN].
  8. 1 2 Mauro, J. C. (May 2011). "Topological constraint theory of glass" (PDF). Am. Ceram. Soc. Bull.[ permanent dead link ]
  9. Bauchy, M.; Micoulaut; Celino; Le Roux; Boero; Massobrio (August 2011). "Angular rigidity in tetrahedral network glasses with changing composition". Physical Review B. 84 (5): 054201. Bibcode:2011PhRvB..84e4201B. doi:10.1103/PhysRevB.84.054201.
  10. Zaccone, A.; Scossa-Romano, E. (2011). "Approximate analytical description of the nonaffine response of amorphous solids". Physical Review B. 83 (18): 184205. arXiv: 1102.0162 . Bibcode:2011PhRvB..83r4205Z. doi:10.1103/PhysRevB.83.184205. S2CID   119256092.
  11. Zaccone, A. (2013). "Elastic Deformations in Covalent Amorphous Solids". Modern Physics Letters B. 27 (5): 1330002. Bibcode:2013MPLB...2730002Z. doi:10.1142/S0217984913300020.
  12. Bauchy, Mathieu (2019-03-01). "Deciphering the atomic genome of glasses by topological constraint theory and molecular dynamics: A review". Computational Materials Science. 159: 95–102. doi:10.1016/j.commatsci.2018.12.004. ISSN   0927-0256. S2CID   139431823.
  13. Smedskjaer, Morten M.; Mauro, John C.; Yue, Yuanzheng (2010-09-08). "Prediction of Glass Hardness Using Temperature-Dependent Constraint Theory". Physical Review Letters. 105 (11): 115503. Bibcode:2010PhRvL.105k5503S. doi:10.1103/PhysRevLett.105.115503. PMID   20867584.
  14. Wray, Peter (7 January 2013). "Gorilla Glass 3 explained (and it is a modeling first for Corning!)". Ceramic Tech Today. The American Ceramic Society. Retrieved 24 January 2014.
  15. Smedskjaer, M. M.; Mauro; Sen; Yue (September 2010). "Quantitative Design of Glassy Materials Using Temperature-Dependent Constraint Theory". Chemistry of Materials. 22 (18): 5358–5365. doi:10.1021/cm1016799.
  16. 1 2 Bauchy, M.; Micoulaut (February 2013). "Transport Anomalies and Adaptative Pressure-Dependent Topological Constraints in Tetrahedral Liquids: Evidence for a Reversibility Window Analogue". Phys. Rev. Lett. 110 (9): 095501. Bibcode:2013PhRvL.110i5501B. doi:10.1103/PhysRevLett.110.095501. PMID   23496720.
  17. Moukarzel, Cristian F. (March 1998). "Isostatic Phase Transition and Instability in Stiff Granular Materials". Physical Review Letters. 81 (8): 1634. arXiv: cond-mat/9803120 . Bibcode:1998PhRvL..81.1634M. doi:10.1103/PhysRevLett.81.1634. S2CID   119436288.
  18. Phillips, J. C. (2004). "Constraint theory and hierarchical protein dynamics". J. Phys.: Condens. Matter. 16 (44): S5065–S5072. Bibcode:2004JPCM...16S5065P. doi:10.1088/0953-8984/16/44/004. S2CID   250821575.
  19. 1 2 Zaccone, A.; Terentjev, E. (2013). "Disorder-Assisted Melting and the Glass Transition in Amorphous Solids". Physical Review Letters. 110 (17): 178002. arXiv: 1212.2020 . Bibcode:2013PhRvL.110q8002Z. doi:10.1103/PhysRevLett.110.178002. PMID   23679782. S2CID   15600577.
  20. 1 2 Boolchand, P.; Georgiev, Goodman (2001). "Discovery of the intermediate phase in chalcogenide glasses" (PDF). Journal of Optoelectronics and Advanced Materials. 3 (3): 703–720. Archived from the original on February 3, 2014.
  21. Bauchy, M.; Micoulaut; Boero; Massobrio (April 2013). "Compositional Thresholds and Anomalies in Connection with Stiffness Transitions in Network Glasses". Physical Review Letters. 110 (16): 165501. Bibcode:2013PhRvL.110p5501B. doi:10.1103/PhysRevLett.110.165501. PMID   23679615.