Ice-type model

Last updated

In statistical mechanics, the ice-type models or six-vertex models are a family of vertex models for crystal lattices with hydrogen bonds. The first such model was introduced by Linus Pauling in 1935 to account for the residual entropy of water ice. [1] Variants have been proposed as models of certain ferroelectric [2] and antiferroelectric [3] crystals.

Contents

In 1967, Elliott H. Lieb found the exact solution to a two-dimensional ice model known as "square ice". [4] The exact solution in three dimensions is only known for a special "frozen" state. [5]

Description

An ice-type model is a lattice model defined on a lattice of coordination number 4. That is, each vertex of the lattice is connected by an edge to four "nearest neighbours". A state of the model consists of an arrow on each edge of the lattice, such that the number of arrows pointing inwards at each vertex is 2. This restriction on the arrow configurations is known as the ice rule. In graph theoretic terms, the states are Eulerian orientations of an underlying 4-regular undirected graph. The partition function also counts the number of nowhere-zero 3-flows. [6]

For two-dimensional models, the lattice is taken to be the square lattice. For more realistic models, one can use a three-dimensional lattice appropriate to the material being considered; for example, the hexagonal ice lattice is used to analyse ice.

At any vertex, there are six configurations of the arrows which satisfy the ice rule (justifying the name "six-vertex model"). The valid configurations for the (two-dimensional) square lattice are the following:

Sixvertex2.png

The energy of a state is understood to be a function of the configurations at each vertex. For square lattices, one assumes that the total energy is given by

for some constants , where here denotes the number of vertices with the th configuration from the above figure. The value is the energy associated with vertex configuration number .

One aims to calculate the partition function of an ice-type model, which is given by the formula

where the sum is taken over all states of the model, is the energy of the state, is the Boltzmann constant, and is the system's temperature.

Typically, one is interested in the thermodynamic limit in which the number of vertices approaches infinity. In that case, one instead evaluates the free energy per vertex in the limit as , where is given by

Equivalently, one evaluates the partition function per vertex in the thermodynamic limit, where

The values and are related by

Physical justification

Several real crystals with hydrogen bonds satisfy the ice model, including ice [1] and potassium dihydrogen phosphate KH
2
PO
4
[2] (KDP). Indeed, such crystals motivated the study of ice-type models.

In ice, each oxygen atom is connected by a bond to four other oxygens, and each bond contains one hydrogen atom between the terminal oxygens. The hydrogen occupies one of two symmetrically located positions, neither of which is in the middle of the bond. Pauling argued [1] that the allowed configuration of hydrogen atoms is such that there are always exactly two hydrogens close to each oxygen, thus making the local environment imitate that of a water molecule, H
2
O. Thus, if we take the oxygen atoms as the lattice vertices and the hydrogen bonds as the lattice edges, and if we draw an arrow on a bond which points to the side of the bond on which the hydrogen atom sits, then ice satisfies the ice model. Similar reasoning applies to show that KDP also satisfies the ice model.

In recent years, ice-type models have been explored as descriptions of pyrochlore spin ice [7] and artificial spin ice systems, [8] [9] in which geometrical frustration in the interactions between bistable magnetic moments ("spins") leads to "ice-rule" spin configurations being favoured. Recently such analogies have been extended to explore the circumstances under which spin-ice systems may be accurately described by the Rys F-model. [10] [11] [12] [13]

Specific choices of vertex energies

On the square lattice, the energies associated with vertex configurations 1-6 determine the relative probabilities of states, and thus can influence the macroscopic behaviour of the system. The following are common choices for these vertex energies.

The ice model

When modeling ice, one takes , as all permissible vertex configurations are understood to be equally likely. In this case, the partition function equals the total number of valid states. This model is known as the ice model (as opposed to an ice-type model).

The KDP model of a ferroelectric

Slater [2] argued that KDP could be represented by an ice-type model with energies

For this model (called the KDP model), the most likely state (the least-energy state) has all horizontal arrows pointing in the same direction, and likewise for all vertical arrows. Such a state is a ferroelectric state, in which all hydrogen atoms have a preference for one fixed side of their bonds.

Rys F model of an antiferroelectric

The Rys model [3] is obtained by setting

The least-energy state for this model is dominated by vertex configurations 5 and 6. For such a state, adjacent horizontal bonds necessarily have arrows in opposite directions and similarly for vertical bonds, so this state is an antiferroelectric state.

The zero field assumption

If there is no ambient electric field, then the total energy of a state should remain unchanged under a charge reversal, i.e. under flipping all arrows. Thus one may assume without loss of generality that

This assumption is known as the zero field assumption, and holds for the ice model, the KDP model, and the Rys F model.

History

The ice rule was introduced by Linus Pauling in 1935 to account for the residual entropy of ice that had been measured by William F. Giauque and J. W. Stout. [14] The residual entropy, , of ice is given by the formula

where is the Boltzmann constant, is the number of oxygen atoms in the piece of ice, which is always taken to be large (the thermodynamic limit) and is the number of configurations of the hydrogen atoms according to Pauling's ice rule. Without the ice rule we would have since the number of hydrogen atoms is and each hydrogen has two possible locations. Pauling estimated that the ice rule reduces this to , a number that would agree extremely well with the Giauque-Stout measurement of . It can be said that Pauling's calculation of for ice is one of the simplest, yet most accurate applications of statistical mechanics to real substances ever made. The question that remained was whether, given the model, Pauling's calculation of , which was very approximate, would be sustained by a rigorous calculation. This became a significant problem in combinatorics.

Both the three-dimensional and two-dimensional models were computed numerically by John F. Nagle in 1966 [15] who found that in three-dimensions and in two-dimensions. Both are amazingly close to Pauling's rough calculation, 1.5.

In 1967, Lieb found the exact solution of three two-dimensional ice-type models: the ice model, [4] the Rys model, [16] and the KDP model. [17] The solution for the ice model gave the exact value of in two-dimensions as

which is known as Lieb's square ice constant.

Later in 1967, Bill Sutherland generalised Lieb's solution of the three specific ice-type models to a general exact solution for square-lattice ice-type models satisfying the zero field assumption. [18]

Still later in 1967, C. P. Yang [19] generalised Sutherland's solution to an exact solution for square-lattice ice-type models in a horizontal electric field.

In 1969, John Nagle derived the exact solution for a three-dimensional version of the KDP model, for a specific range of temperatures. [5] For such temperatures, the model is "frozen" in the sense that (in the thermodynamic limit) the energy per vertex and entropy per vertex are both zero. This is the only known exact solution for a three-dimensional ice-type model.

Relation to eight-vertex model

The eight-vertex model, which has also been exactly solved, is a generalisation of the (square-lattice) six-vertex model: to recover the six-vertex model from the eight-vertex model, set the energies for vertex configurations 7 and 8 to infinity. Six-vertex models have been solved in some cases for which the eight-vertex model has not; for example, Nagle's solution for the three-dimensional KDP model [5] and Yang's solution of the six-vertex model in a horizontal field. [19]

Boundary conditions

This ice model provide an important 'counterexample' in statistical mechanics: the bulk free energy in the thermodynamic limit depends on boundary conditions. [20] The model was analytically solved for periodic boundary conditions, anti-periodic, ferromagnetic and domain wall boundary conditions. The six vertex model with domain wall boundary conditions on a square lattice has specific significance in combinatorics, it helps to enumerate alternating sign matrices. In this case the partition function can be represented as a determinant of a matrix (whose dimension is equal to the size of the lattice), but in other cases the enumeration of does not come out in such a simple closed form.

Clearly, the largest is given by free boundary conditions (no constraint at all on the configurations on the boundary), but the same occurs, in the thermodynamic limit, for periodic boundary conditions, [21] as used originally to derive .

3-colorings of a lattice

The number of states of an ice type model on the internal edges of a finite simply connected union of squares of a lattice is equal to one third of the number of ways to 3-color the squares, with no two adjacent squares having the same color. This correspondence between states is due to Andrew Lenard and is given as follows. If a square has color i = 0, 1, or 2, then the arrow on the edge to an adjacent square goes left or right (according to an observer in the square) depending on whether the color in the adjacent square is i+1 or i1 mod 3. There are 3 possible ways to color a fixed initial square, and once this initial color is chosen this gives a 1:1 correspondence between colorings and arrangements of arrows satisfying the ice-type condition.

See also

Notes

  1. 1 2 3 Pauling, L. (1935). "The Structure and Entropy of Ice and of Other Crystals with Some Randomness of Atomic Arrangement". Journal of the American Chemical Society . 57 (12): 2680–2684. doi:10.1021/ja01315a102.
  2. 1 2 3 Slater, J. C. (1941). "Theory of the Transition in KH2PO4". Journal of Chemical Physics . 9 (1): 16–33. Bibcode:1941JChPh...9...16S. doi:10.1063/1.1750821.
  3. 1 2 Rys, F. (1963). "Über ein zweidimensionales klassisches Konfigurationsmodell". Helvetica Physica Acta . 36: 537.
  4. 1 2 Lieb, E. H. (1967). "Residual Entropy of Square Ice". Physical Review . 162 (1): 162–172. Bibcode:1967PhRv..162..162L. doi:10.1103/PhysRev.162.162.
  5. 1 2 3 Nagle, J. F. (1969). "Proof of the first order phase transition in the Slater KDP model". Communications in Mathematical Physics . 13 (1): 62–67. Bibcode:1969CMaPh..13...62N. doi:10.1007/BF01645270. S2CID   122432926.
  6. Mihail, M.; Winkler, P. (1992). "On the Number of Eularian Orientations of a Graph". SODA '92 Proceedings of the Third Annual ACM-SIAM Symposium on Discrete Algorithms. Society for Industrial and Applied Mathematics. pp. 138–145. ISBN   978-0-89791-466-6.
  7. Bramwell, Steven T; Harris, Mark J (2020-09-02). "The history of spin ice". Journal of Physics: Condensed Matter. 32 (37): 374010. Bibcode:2020JPCM...32K4010B. doi: 10.1088/1361-648X/ab8423 . ISSN   0953-8984. PMID   32554893.
  8. Wang, R. F.; Nisoli, C.; Freitas, R. S.; Li, J.; McConville, W.; Cooley, B. J.; Lund, M. S.; Samarth, N.; Leighton, C.; Crespi, V. H.; Schiffer, P. (January 2006). "Artificial 'spin ice' in a geometrically frustrated lattice of nanoscale ferromagnetic islands". Nature. 439 (7074): 303–306. arXiv: cond-mat/0601429 . Bibcode:2006Natur.439..303W. doi:10.1038/nature04447. ISSN   1476-4687. PMID   16421565. S2CID   1462022.
  9. Perrin, Yann; Canals, Benjamin; Rougemaille, Nicolas (December 2016). "Extensive degeneracy, Coulomb phase and magnetic monopoles in artificial square ice". Nature. 540 (7633): 410–413. arXiv: 1610.01316 . Bibcode:2016Natur.540..410P. doi:10.1038/nature20155. ISSN   1476-4687. PMID   27894124. S2CID   4409371.
  10. Jaubert, L. D. C.; Lin, T.; Opel, T. S.; Holdsworth, P. C. W.; Gingras, M. J. P. (2017-05-19). "Spin ice Thin Film: Surface Ordering, Emergent Square ice, and Strain Effects". Physical Review Letters. 118 (20): 207206. arXiv: 1608.08635 . Bibcode:2017PhRvL.118t7206J. doi:10.1103/PhysRevLett.118.207206. ISSN   0031-9007. PMID   28581768. S2CID   118688211.
  11. Arroo, Daan M.; Bramwell, Steven T. (2020-12-22). "Experimental measures of topological sector fluctuations in the F-model". Physical Review B. 102 (21): 214427. Bibcode:2020PhRvB.102u4427A. doi:10.1103/PhysRevB.102.214427. ISSN   2469-9950. S2CID   222290448.
  12. Nisoli, Cristiano (2020-11-01). "Topological order of the Rys F-model and its breakdown in realistic square spin ice: Topological sectors of Faraday loops". Europhysics Letters. 132 (4): 47005. arXiv: 2004.02107 . Bibcode:2020EL....13247005N. doi:10.1209/0295-5075/132/47005. ISSN   0295-5075. S2CID   221891692.
  13. Schánilec, V.; Brunn, O.; Horáček, M.; Krátký, S.; Meluzín, P.; Šikola, T.; Canals, B.; Rougemaille, N. (2022-07-07). "Approaching the Topological Low-Energy Physics of the F Model in a Two-Dimensional Magnetic Lattice". Physical Review Letters. 129 (2): 027202. Bibcode:2022PhRvL.129b7202S. doi:10.1103/PhysRevLett.129.027202. ISSN   0031-9007. PMID   35867462. S2CID   250378329.
  14. Giauque, W. F.; Stout, Stout (1936). "The entropy of water and third law of thermodynamics. The heat capacity of ice from 15 to 273K". Journal of the American Chemical Society . 58 (7): 1144–1150. Bibcode:1936JAChS..58.1144G. doi:10.1021/ja01298a023.
  15. Nagle, J. F. (1966). "Lattice Statistics of Hydrogen Bonded Crystals. I. The Residual Entropy of Ice". Journal of Mathematical Physics . 7 (8): 1484–1491. Bibcode:1966JMP.....7.1484N. doi:10.1063/1.1705058.
  16. Lieb, E. H. (1967). "Exact Solution of the Problem of the Entropy of Two-Dimensional Ice". Physical Review Letters . 18 (17): 692–694. Bibcode:1967PhRvL..18..692L. doi:10.1103/PhysRevLett.18.692.
  17. Lieb, E. H. (1967). "Exact Solution of the Two-Dimensional Slater KDP Model of a Ferroelectric". Physical Review Letters . 19 (3): 108–110. Bibcode:1967PhRvL..19..108L. doi:10.1103/PhysRevLett.19.108.
  18. Sutherland, B. (1967). "Exact Solution of a Two-Dimensional Model for Hydrogen-Bonded Crystals". Physical Review Letters . 19 (3): 103–104. Bibcode:1967PhRvL..19..103S. doi:10.1103/PhysRevLett.19.103.
  19. 1 2 Yang, C. P. (1967). "Exact Solution of a Two-Dimensional Model for Hydrogen-Bonded Crystals". Physical Review Letters . 19 (3): 586–588. Bibcode:1967PhRvL..19..586Y. doi:10.1103/PhysRevLett.19.586.
  20. Korepin, V.; Zinn-Justin, P. (2000). "Thermodynamic limit of the Six-Vertex Model with Domain Wall Boundary Conditions". Journal of Physics A . 33 (40): 7053–7066. arXiv: cond-mat/0004250 . Bibcode:2000JPhA...33.7053K. doi:10.1088/0305-4470/33/40/304. S2CID   2143060.
  21. Brascamp, H. J.; Kunz, H.; Wu, F. Y. (1973). "Some rigorous results for the vertex model in statistical mechanics". Journal of Mathematical Physics . 14 (12): 1927–1932. Bibcode:1973JMP....14.1927B. doi:10.1063/1.1666271.

Further reading

Related Research Articles

The Ising model, named after the physicists Ernst Ising and Wilhelm Lenz, is a mathematical model of ferromagnetism in statistical mechanics. The model consists of discrete variables that represent magnetic dipole moments of atomic "spins" that can be in one of two states. The spins are arranged in a graph, usually a lattice, allowing each spin to interact with its neighbors. Neighboring spins that agree have a lower energy than those that disagree; the system tends to the lowest energy but heat disturbs this tendency, thus creating the possibility of different structural phases. The model allows the identification of phase transitions as a simplified model of reality. The two-dimensional square-lattice Ising model is one of the simplest statistical models to show a phase transition.

In physics, mathematics and statistics, scale invariance is a feature of objects or laws that do not change if scales of length, energy, or other variables, are multiplied by a common factor, and thus represent a universality.

<span class="mw-page-title-main">Polaron</span> Quasiparticle in condensed matter physics

A polaron is a quasiparticle used in condensed matter physics to understand the interactions between electrons and atoms in a solid material. The polaron concept was proposed by Lev Landau in 1933 and Solomon Pekar in 1946 to describe an electron moving in a dielectric crystal where the atoms displace from their equilibrium positions to effectively screen the charge of an electron, known as a phonon cloud. This lowers the electron mobility and increases the electron's effective mass.

The classical XY model is a lattice model of statistical mechanics. In general, the XY model can be seen as a specialization of Stanley's n-vector model for n = 2.

Ice I<sub>h</sub> Hexagonal crystal form of ordinary ice or frozen water

Ice Ih is the hexagonal crystal form of ordinary ice, or frozen water. Virtually all ice in the biosphere is ice Ih, with the exception only of a small amount of ice Ic that is occasionally present in the upper atmosphere. Ice Ih exhibits many peculiar properties that are relevant to the existence of life and regulation of global climate. For a description of these properties, see Ice, which deals primarily with ice Ih.

In condensed matter physics, the term geometrical frustration refers to a phenomenon where atoms tend to stick to non-trivial positions or where, on a regular crystal lattice, conflicting inter-atomic forces lead to quite complex structures. As a consequence of the frustration in the geometry or in the forces, a plenitude of distinct ground states may result at zero temperature, and usual thermal ordering may be suppressed at higher temperatures. Much studied examples are amorphous materials, glasses, or dilute magnets.

<span class="mw-page-title-main">Hubbard model</span>

The Hubbard model is an approximate model used to describe the transition between conducting and insulating systems. It is particularly useful in solid-state physics. The model is named for John Hubbard.

In the renormalization group analysis of phase transitions in physics, a critical dimension is the dimensionality of space at which the character of the phase transition changes. Below the lower critical dimension there is no phase transition. Above the upper critical dimension the critical exponents of the theory become the same as that in mean field theory. An elegant criterion to obtain the critical dimension within mean field theory is due to V. Ginzburg.

<span class="mw-page-title-main">Hofstadter's butterfly</span> Fractal describing the theorised behaviour of electrons in a magnetic field

In condensed matter physics, Hofstadter's butterfly is a graph of the spectral properties of non-interacting two-dimensional electrons in a perpendicular magnetic field in a lattice. The fractal, self-similar nature of the spectrum was discovered in the 1976 Ph.D. work of Douglas Hofstadter and is one of the early examples of modern scientific data visualization. The name reflects the fact that, as Hofstadter wrote, "the large gaps [in the graph] form a very striking pattern somewhat resembling a butterfly."

In physics, the Bethe ansatz is an ansatz method for finding the exact wavefunctions of certain one-dimensional quantum many-body models. It was invented by Hans Bethe in 1931 to find the exact eigenvalues and eigenvectors of the one-dimensional antiferromagnetic Heisenberg model Hamiltonian. Since then the method has been extended to other models in one dimension: the (anisotropic) Heisenberg chain, the Lieb-Liniger interacting Bose gas, the Hubbard model, the Kondo model, the Anderson impurity model, the Richardson model etc.

<span class="mw-page-title-main">Spin ice</span>

A spin ice is a magnetic substance that does not have a single minimal-energy state. It has magnetic moments (i.e. "spin") as elementary degrees of freedom which are subject to frustrated interactions. By their nature, these interactions prevent the moments from exhibiting a periodic pattern in their orientation down to a temperature much below the energy scale set by the said interactions. Spin ices show low-temperature properties, residual entropy in particular, closely related to those of common crystalline water ice. The most prominent compounds with such properties are dysprosium titanate (Dy2Ti2O7) and holmium titanate (Ho2Ti2O7). The orientation of the magnetic moments in spin ice resembles the positional organization of hydrogen atoms (more accurately, ionized hydrogen, or protons) in conventional water ice (see figure 1).

<span class="mw-page-title-main">Vertex model</span>

A vertex model is a type of statistical mechanics model in which the Boltzmann weights are associated with a vertex in the model. This contrasts with a nearest-neighbour model, such as the Ising model, in which the energy, and thus the Boltzmann weight of a statistical microstate is attributed to the bonds connecting two neighbouring particles. The energy associated with a vertex in the lattice of particles is thus dependent on the state of the bonds which connect it to adjacent vertices. It turns out that every solution of the Yang–Baxter equation with spectral parameters in a tensor product of vector spaces yields an exactly-solvable vertex model.

In statistical mechanics, the two-dimensional square lattice Ising model is a simple lattice model of interacting magnetic spins. The model is notable for having nontrivial interactions, yet having an analytical solution. The model was solved by Lars Onsager for the special case that the external magnetic field H = 0. An analytical solution for the general case for has yet to be found.

In statistical mechanics, the eight-vertex model is a generalisation of the ice-type (six-vertex) models; it was discussed by Sutherland, and Fan & Wu, and solved by Baxter in the zero-field case.

The Lieb–Liniger model describes a gas of particles moving in one dimension and satisfying Bose–Einstein statistics.

In chemistry, ice rules are basic principles that govern arrangement of atoms in water ice. They are also known as Bernal–Fowler rules, after British physicists John Desmond Bernal and Ralph H. Fowler who first described them in 1933.

Dynamical mean-field theory (DMFT) is a method to determine the electronic structure of strongly correlated materials. In such materials, the approximation of independent electrons, which is used in density functional theory and usual band structure calculations, breaks down. Dynamical mean-field theory, a non-perturbative treatment of local interactions between electrons, bridges the gap between the nearly free electron gas limit and the atomic limit of condensed-matter physics.

The Lieb–Robinson bound is a theoretical upper limit on the speed at which information can propagate in non-relativistic quantum systems. It demonstrates that information cannot travel instantaneously in quantum theory, even when the relativity limits of the speed of light are ignored. The existence of such a finite speed was discovered mathematically by Elliott H. Lieb and Derek W. Robinson in 1972. It turns the locality properties of physical systems into the existence of, and upper bound for this speed. The bound is now known as the Lieb–Robinson bound and the speed is known as the Lieb–Robinson velocity. This velocity is always finite but not universal, depending on the details of the system under consideration. For finite-range, e.g. nearest-neighbor, interactions, this velocity is a constant independent of the distance travelled. In long-range interacting systems, this velocity remains finite, but it can increase with the distance travelled.

The Frenkel–Kontorova model, also known as the FK model, is a fundamental model of low-dimensional nonlinear physics.

Exact diagonalization (ED) is a numerical technique used in physics to determine the eigenstates and energy eigenvalues of a quantum Hamiltonian. In this technique, a Hamiltonian for a discrete, finite system is expressed in matrix form and diagonalized using a computer. Exact diagonalization is only feasible for systems with a few tens of particles, due to the exponential growth of the Hilbert space dimension with the size of the quantum system. It is frequently employed to study lattice models, including the Hubbard model, Ising model, Heisenberg model, t-J model, and SYK model.