Classical XY model

Last updated

The classical XY model (sometimes also called classical rotor (rotator) model or O(2) 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 [1] for n = 2.

Contents

Definition

Given a D-dimensional lattice Λ, per each lattice site j ∈ Λ there is a two-dimensional, unit-length vector sj = (cos θj, sin θj)

The spin configuration, s = (sj)j ∈ Λ is an assignment of the angle π < θjπ for each j ∈ Λ.

Given a translation-invariant interaction Jij = J(ij) and a point dependent external field , the configuration energy is

The case in which Jij = 0 except for ij nearest neighbor is called nearest neighbor case.

The configuration probability is given by the Boltzmann distribution with inverse temperature β ≥ 0:

where Z is the normalization, or partition function. [2] The notation indicates the expectation of the random variable A(s) in the infinite volume limit, after periodic boundary conditions have been imposed.

Rigorous results

One dimension

As in any 'nearest-neighbor' n-vector model with free (non-periodic) boundary conditions, if the external field is zero, there exists a simple exact solution. In the free boundary conditions case, the Hamiltonian is

therefore the partition function factorizes under the change of coordinates

This gives

where is the modified Bessel function of the first kind. The partition function can be used to find several important thermodynamic quantities. For example, in the thermodynamic limit (), the free energy per spin is

Using the properties of the modified Bessel functions, the specific heat (per spin) can be expressed as [5]

where , and is the short-range correlation function,

Exact specific heat per spin in the one-dimensional XY model 1D XY Specific Heat.svg
Exact specific heat per spin in the one-dimensional XY model

Even in the thermodynamic limit, there is no divergence in the specific heat. Indeed, like the one-dimensional Ising model, the one-dimensional XY model has no phase transitions at finite temperature.

The same computation for periodic boundary condition (and still h = 0) requires the transfer matrix formalism, though the result is the same. [6]

(Click "show" at right to see the details of the transfer matrix formalism.)

The partition function can be evaluated as

which can be treated as the trace of a matrix, namely a product of matrices (scalars, in this case). The trace of a matrix is simply the sum of its eigenvalues, and in the thermodynamic limit only the largest eigenvalue will survive, so the partition function can be written as a repeated product of this maximal eigenvalue. This requires solving the eigenvalue problem

Note the expansion

which represents a diagonal matrix representation in the basis of its plane-wave eigenfunctions . The eigenvalues of the matrix is simply are modified Bessel functions evaluated at , namely . For any particular value of , these modified Bessel functions satisfy and . Therefore in the thermodynamic limit the eigenvalue will dominate the trace, and so .

This transfer matrix approach is also required when using free boundary conditions, but with an applied field . If the applied field is small enough that it can be treated as a perturbation to the system in zero-field, then the magnetic susceptibility can be estimated. This is done by using the eigenstates computed by the transfer matrix approach and computing the energy shift with second-order perturbation theory, then comparing with the free-energy expansion . One finds [7]

where is the Curie constant (a value typically associated with the susceptibility in magnetic materials). This expression is also true for the one-dimensional Ising model, with the replacement .

Two dimensions

Mean-square mganetization for a 25x25 lattice (Ota: 30x30), suggesting an increase in magnetic moment that is not present in the thermodynamic limit XY Magnetisation.svg
Mean-square mganetization for a 25x25 lattice (Ota: 30x30), suggesting an increase in magnetic moment that is not present in the thermodynamic limit

The two-dimensional XY model with nearest-neighbor interactions is an example of a two-dimensional system with continuous symmetry that does not have long-range order as required by the Mermin–Wagner theorem. Likewise, there is not a conventional phase transition present that would be associated with symmetry breaking. However, as will be discussed later, the system does show signs of a transition from a disordered high-temperature state to a quasi-ordered state below some critical temperature, called the Kosterlitz-Thouless transition. In the case of a discrete lattice of spins, the two-dimensional XY model can be evaluated using the transfer matrix approach, reducing the model to an eigenvalue problem and utilizing the largest eigenvalue from the transfer matrix. Though the exact solution is intractable, it is possible to use certain approximations to get estimates for the critical temperature which occurs at low temperatures. For example, Mattis (1984) used an approximation to this model to estimate a critical temperature of the system as

The 2D XY model has also been studied in great detail using Monte Carlo simulations, for example with the Metropolis algorithm. These can be used to compute thermodynamic quantities like the system energy, specific heat, magnetization, etc., over a range of temperatures and time-scales. In the Monte Carlo simulation, each spin is associated to a continuously-varying angle (often, it can be discretized into finitely-many angles, like in the related Potts model, for ease of computation. However, this is not a requirement.) At each time step the Metropolis algorithm chooses one spin at random and rotates its angle by some random increment . This change in angle causes a change in the energy of the system, which can be positive or negative. If negative, the algorithm accepts the change in angle; if positive, the configuration is accepted with probability , the Boltzmann factor for the energy change. The Monte Carlo method has been used to verify, with various methods, the critical temperature of the system, and is estimated to be [9] . The Monte Carlo method can also compute average values that are used to compute thermodynamic quantities like magnetization, spin-spin correlation, correlation lengths, and specific heat. These are important ways to characterize the behavior of the system near the critical temperature. The magnetization and squared magnetization, for example, can be computed as

Specific heat of the two-dimensional XY model, computed by Monte Carlo simulation on square lattices of size up to 4096 x 4096 (Nguyen: ), showing a feature at
k
B
T
/
J
[?]
1.167
{\displaystyle k_{\rm {B}}T/J\approx 1.167}
, above the K-T transition. Inset shows the position of the peak as a function of lattice size. XY Specifc Heat 2D.png
Specific heat of the two-dimensional XY model, computed by Monte Carlo simulation on square lattices of size up to 4096 x 4096 (Nguyen: ), showing a feature at , above the K-T transition. Inset shows the position of the peak as a function of lattice size.

where are the number of spins. The mean magnetization characterizes the magnitude of the net magnetic moment of the system; in many magnetic systems this is zero above a critical temperature and becomes non-zero spontaneously at low temperatures. Similarly the mean-squared magnetization characterizes the average of the square of net components of the spins across the lattice. Either of these are commonly used to characterize the order parameter of a system. Rigorous analysis of the XY model shows the magnetization in the thermodynamic limit is zero, and that the square magnetization approximately follows [11] , which vanishes in the thermodynamic limit. Indeed, at high temperatures this quantity approaches zero since the components of the spins will tend to be randomized and thus sum to zero. However at low temperatures for a finite system, the mean-square magnetization increases, suggesting there are regions of the spin space that are aligned to contribute to a non-zero contribution. The magnetization shown (for a 25x25 lattice) is one example of this, that appears to suggest a phase transition, while no such transition exists in the thermodynamic limit.

Furthermore, using statistical mechanics one can relate thermodynamic averages to quantities like specific heat by calculating

The specific heat is shown at low temperatures near the critical temperature . There is no feature in the specific heat consistent with critical behavior (like a divergence) at this predicted temperature. Indeed, estimating the critical temperature comes from other methods, like from the helicity modulus, or the temperature dependence of the divergence of susceptibility. [12] However, there is a feature in the specific heat in the form of a peak at . This peak position and height have been shown not to depend on system size, for lattices of linear size greater than 256; indeed, the specific heat anomaly remains rounded and finite for increasing lattice size, with no divergent peak.

The nature of the critical transitions and vortex formation can be elucidated by considering a continuous version of the XY model. Here, the discrete spins are replaced by a field representing the spin's angle at any point in space. In this case the angle of the spins must vary smoothly over changes in position. Expanding the original cosine as a Taylor series, the Hamiltonian can be expressed in the continuum approximation as

Color map of the (discrete) two-dimensional XY model in a 250x250 lattice at
k
B
T
/
J
=
0.4
{\displaystyle k_{\rm {B}}T/J=0.4}
. Each spin is represented by a color that corresponds to an angle between
(
-
p
,
p
]
{\displaystyle (-\pi ,\pi ]} XY Color Map.svg
Color map of the (discrete) two-dimensional XY model in a 250x250 lattice at . Each spin is represented by a color that corresponds to an angle between

The continuous version of the XY model is often used to model systems that possess order parameters with the same kinds of symmetry, e.g. superfluid helium, hexatic liquid crystals. This is what makes them peculiar from other phase transitions which are always accompanied with a symmetry breaking. Topological defects in the XY model lead to a vortex-unbinding transition from the low-temperature phase to the high-temperature disordered phase. Indeed, the fact that at high temperature correlations decay exponentially fast, while at low temperatures decay with power law, even though in both regimes M(β) = 0, is called Kosterlitz–Thouless transition. Kosterlitz and Thouless provided a simple argument of why this would be the case: this considers the ground state consisting of all spins in the same orientation, with the addition then of a single vortex. The presence of these contributes an entropy of roughly , where is an effective length scale (for example, the lattice size for a discrete lattice) Meanwhile, the energy of the system increases due to the vortex, by an amount . Putting these together, the free energy of a system would change due to the spontaneous formation of a vortex by an amount

In the thermodynamic limit, the system does not favor the formation of vortices at low temperatures, but does favor them at high temperatures, above the critical temperature . This indicates that at low temperatures, any vortices that arise will want to annihilate with antivortices to lower the system energy. Indeed, this will be the case qualitatively if one watches 'snapshots' of the spin system at low temperatures, where vortices and antivortices gradually come together to annihilate. Thus, the low-temperature state will consist of bound vortex-antivortex pairs. Meanwhile at high temperatures, there will be a collection of unbound vortices and antivortices that are free to move about the plane.

To visualize the Ising model, one can use an arrow pointing up or down, or represented as a point colored black/white to indicate its state. To visualize the XY spin system, the spins can be represented as an arrow pointing in some direction, or as being represented as a point with some color. Here it is necessary to represent the spin with a spectrum of colors due to each of the possible continuous variables. This can be done using, for example, a continuous and periodic red-green-blue spectrum. As shown on the figure, cyan corresponds to a zero angle (pointing to the right), whereas red corresponds to a 180 degree angle (pointing to the left). One can then study snapshots of the spin configurations at different temperatures to elucidate what happens above and below the critical temperature of the XY model. At high temperatures, the spins will not have a preferred orientation and there will be unpredictable variation of angles between neighboring spins, as there will be no preferred energetically favorable configuration. In this case, the color map will look highly pixellated. Meanwhile at low temperatures, one possible ground-state configuration has all spins pointed in the same orientation (same angle); these would correspond to regions (domains) of the color map where all spins have roughly the same color.

Various forms of vortices and antivortices, shown in a Monte Carlo simulation at
k
B
T
/
J
=
0.4
{\displaystyle k_{\rm {B}}T/J=0.4} XY Vortices.svg
Various forms of vortices and antivortices, shown in a Monte Carlo simulation at

To identify vortices (or antivortices) present as a result of the Kosterlitz–Thouless transition, one can determine the signed change in angle by traversing a circle of lattice points counterclockwise. If the total change in angle is zero, this corresponds to no vortex being present; whereas a total change in angle of corresponds to a vortex (or antivortex). These vortexes are topologically non-trivial objects that come in vortex-antivortex pairs, which can separate or pair-annihilate. In the colormap, these defects can be identified in regions where there is a large color gradient where all colors of the spectrum meet around a point. Qualitatively, these defects can look like inward- or outward-pointing sources of flow, or whirlpools of spins that collectively clockwise or counterclockwise, or hyperbolic-looking features with some spins pointing toward and some spins pointing away from the defect. As the configuration is studied at long time scales and at low temperatures, it is observed that many of these vortex-antivortex pairs get closer together and eventually pair-annihilate. It is only at high temperatures that these vortices and antivortices are liberated and unbind from one another.

In the continuous XY model, the high-temperature spontaneous magnetization vanishes:

Besides, cluster expansion shows that the spin correlations cluster exponentially fast: for instance

At low temperatures, i.e. β ≫ 1, the spontaneous magnetization remains zero (see the Mermin–Wagner theorem),

but the decay of the correlations is only power law: Fröhlich and Spencer [13] found the lower bound

while McBryan and Spencer found the upper bound, for any

Three and higher dimensions

Independently of the range of the interaction, at low enough temperature the magnetization is positive.

Phase transition

As mentioned above in one dimension the XY model does not have a phase transition, while in two dimensions it has the Berezinski-Kosterlitz-Thouless transition between the phases with exponentially and powerlaw decaying correlation functions.

In three and higher dimensions the XY model has a ferromagnet-paramagnet phase transition. At low temperatures the spontaneous magnetization is nonzero: this is the ferromagnetic phase. As the temperature is increased, spontaneous magnetization gradually decreases and vanishes at a critical temperature. It remains zero at all higher temperatures: this is the paramagnetic phase.

In four and higher dimensions the phase transition has mean field theory critical exponents (with logarithmic corrections in four dimensions).

Three dimensional case: the critical exponents

The three dimensional case is interesting because the critical exponents at the phase transition are nontrivial. Many three-dimensional physical systems belong to the same universality class as the three dimensional XY model and share the same critical exponents, most notably easy-plane magnets and liquid Helium-4. The values of these critical exponents are measured by experiments, Monte Carlo simulations, and can also be computed by theoretical methods of quantum field theory, such as the renormalization group and the conformal bootstrap. Renormalization group methods are applicable because the critical point of the XY model is believed to be described by a renormalization group fixed point. Conformal bootstrap methods are applicable because it is also believed to be a unitary three dimensional conformal field theory.

Most important critical exponents of the three dimensional XY model are . All of them can be expressed via just two numbers: the scaling dimensions and of the complex order parameter field and of the leading singlet operator (same as in the Ginzburg–Landau description). Another important field is (same as ), whose dimension determines the correction-to-scaling exponent . According to a conformal bootstrap computation, [14] these three dimensions are given by:

0.519088(22)
1.51136(22)
3.794(8)

This gives the following values of the critical exponents:

general expression ()numerical value
α-0.01526(30)
β0.34869(7)
γ1.3179(2)
δ4.77937(25)
η0.038176(44)
ν0.67175(10)
ω0.794(8)

Monte Carlo methods give compatible determinations: [15] .

See also

Notes

  1. Stanley, H.E. (1968). "Dependence of Critical Properties on Dimensionality of Spins". Physical Review Letters. 20 (12): 589–592. Bibcode:1968PhRvL..20..589S. doi:10.1103/PhysRevLett.20.589.
  2. Chaikin, P.M.; Lubensky, T.C. (2000). Principles of Condensed Matter Physics. Cambridge University Press. ISBN   978-0521794503.
  3. Ginibre, J. (1970). "General formulation of Griffiths' inequalities". Communications in Mathematical Physics. 16 (4): 310–328. Bibcode:1970CMaPh..16..310G. doi:10.1007/BF01646537. S2CID   120649586.
  4. Aizenman, M.; Simon, B. (1980). "A comparison of plane rotor and Ising models". Physics Letters A. 76 (3–4): 281–282. Bibcode:1980PhLA...76..281A. doi:10.1016/0375-9601(80)90493-4.
  5. Badalian, D. (1996). "On the thermodynamics of classical spins with isotrop Heisenberg interaction in one-dimensional quasi-periodic structures". Physica B. 226 (4): 385–390. Bibcode:1996PhyB..226..385B. doi:10.1016/0921-4526(96)00283-9.
  6. Mattis, D.C. (1984). "Transfer matrix in plane-rotator model". Physics Letters A. 104 A (6–7): 357–360. Bibcode:1984PhLA..104..357M. doi:10.1016/0375-9601(84)90816-8.
  7. Mattis, D. C. (1985). The Theory of Magnetism II. Springer Series in Solid-State Physics. ISBN   978-3-642-82405-0.
  8. Ota, S.; Ota, S.B.; Fahnle, M (1992). "Microcanonical Monte Carlo simulations for the two-dimensional XY model". Journal of Physics: Condensed Matter. 4 (24): 5411. Bibcode:1992JPCM....4.5411O. doi:10.1088/0953-8984/4/24/011. S2CID   250920391.
  9. Hsieh, Y.-D.; Kao, Y.-J.; Sandvik, A.W. (2013). "Finite-size scaling method for the Berezinskii-Kosterlitz-Thouless transition". Journal of Statistical Mechanics: Theory and Experiment. 2013 (9): P09001. arXiv: 1302.2900 . Bibcode:2013JSMTE..09..001H. doi:10.1088/1742-5468/2013/09/P09001. S2CID   118609225.
  10. Nguyen, P.H.; Boninsegni, M. (2021). "Superfluid Transition and Specific Heat of the 2D x-y Model: Monte Carlo Simulation". Applied Sciences. 11 (11): 4931. arXiv: 2105.14112 . doi: 10.3390/app11114931 .
  11. Tobochnik, J.; Chester, G.V. (1979). "Monte Carlo study of the planar spin model". Physical Review B. 20 (9): 3761–3769. Bibcode:1979PhRvB..20.3761T. doi:10.1103/PhysRevB.20.3761.
  12. Binder, K. (2013). Applications of the Monte Carlo Method in Statistical Physics. Springer Science & Business Media. ISBN   978-3-642-51703-7.
  13. Fröhlich, J.; Spencer, T. (1981). "The Kosterlitz–Thouless transition in two-dimensional abelian spin systems and the Coulomb gas". Communications in Mathematical Physics. 81 (4): 527–602. Bibcode:1981CMaPh..81..527F. doi:10.1007/bf01208273. S2CID   73555642.
  14. Chester, Shai M.; Landry, Walter; Liu, Junyu; Poland, David; Simmons-Duffin, David; Su, Ning; Vichi, Alessandro (2020). "Carving out OPE space and precise O(2) model critical exponents". Journal of High Energy Physics. 2020 (6): 142. arXiv: 1912.03324 . Bibcode:2020JHEP...06..142C. doi:10.1007/JHEP06(2020)142. ISSN   1029-8479. S2CID   208910721.
  15. Hasenbusch, Martin (2019-12-26). "Monte Carlo study of an improved clock model in three dimensions". Physical Review B. 100 (22): 224517. arXiv: 1910.05916 . Bibcode:2019PhRvB.100v4517H. doi:10.1103/PhysRevB.100.224517. ISSN   2469-9950. S2CID   204509042.

Related Research Articles

In mechanics and geometry, the 3D rotation group, often denoted SO(3), is the group of all rotations about the origin of three-dimensional Euclidean space under the operation of composition.

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.

<span class="mw-page-title-main">Solid of revolution</span> Type of three-dimensional shape

In geometry, a solid of revolution is a solid figure obtained by rotating a plane figure around some straight line, which may not intersect the generatrix. The surface created by this revolution and which bounds the solid is the surface of revolution.

<span class="mw-page-title-main">Halbach array</span> Special arrangement of permanent magnets

A Halbach array is a special arrangement of permanent magnets that augments the magnetic field on one side of the array while cancelling the field to near zero on the other side. This is achieved by having a spatially rotating pattern of magnetisation.

<span class="mw-page-title-main">Bloch sphere</span> Geometrical representation of the pure state space of a two-level quantum mechanical system

In quantum mechanics and computing, the Bloch sphere is a geometrical representation of the pure state space of a two-level quantum mechanical system (qubit), named after the physicist Felix Bloch.

In linear algebra, a rotation matrix is a transformation matrix that is used to perform a rotation in Euclidean space. For example, using the convention below, the matrix

In solid-state physics, the tight-binding model is an approach to the calculation of electronic band structure using an approximate set of wave functions based upon superposition of wave functions for isolated atoms located at each atomic site. The method is closely related to the LCAO method used in chemistry. Tight-binding models are applied to a wide variety of solids. The model gives good qualitative results in many cases and can be combined with other models that give better results where the tight-binding model fails. Though the tight-binding model is a one-electron model, the model also provides a basis for more advanced calculations like the calculation of surface states and application to various kinds of many-body problem and quasiparticle calculations.

In quantum field theory and statistical mechanics, the Hohenberg–Mermin–Wagner theorem or Mermin–Wagner theorem states that continuous symmetries cannot be spontaneously broken at finite temperature in systems with sufficiently short-range interactions in dimensions d ≤ 2. Intuitively, this means that long-range fluctuations can be created with little energy cost, and since they increase the entropy, they are favored.

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.

The Glauber–Sudarshan P representation is a suggested way of writing down the phase space distribution of a quantum system in the phase space formulation of quantum mechanics. The P representation is the quasiprobability distribution in which observables are expressed in normal order. In quantum optics, this representation, formally equivalent to several other representations, is sometimes preferred over such alternative representations to describe light in optical phase space, because typical optical observables, such as the particle number operator, are naturally expressed in normal order. It is named after George Sudarshan and Roy J. Glauber, who worked on the topic in 1963. Despite many useful applications in laser theory and coherence theory, the Sudarshan–Glauber P representation has the peculiarity that it is not always positive, and is not a bona-fide probability function.

Photon polarization is the quantum mechanical description of the classical polarized sinusoidal plane electromagnetic wave. An individual photon can be described as having right or left circular polarization, or a superposition of the two. Equivalently, a photon can be described as having horizontal or vertical linear polarization, or a superposition of the two.

In geometry, various formalisms exist to express a rotation in three dimensions as a mathematical transformation. In physics, this concept is applied to classical mechanics where rotational kinematics is the science of quantitative description of a purely rotational motion. The orientation of an object at a given instant is described with the same tools, as it is defined as an imaginary rotation from a reference placement in space, rather than an actually observed rotation from a previous placement in space.

The Wigner D-matrix is a unitary matrix in an irreducible representation of the groups SU(2) and SO(3). It was introduced in 1927 by Eugene Wigner, and plays a fundamental role in the quantum mechanical theory of angular momentum. The complex conjugate of the D-matrix is an eigenfunction of the Hamiltonian of spherical and symmetric rigid rotors. The letter D stands for Darstellung, which means "representation" in German.

Spin is an intrinsic form of angular momentum carried by elementary particles, and thus by composite particles such as hadrons, atomic nuclei, and atoms. Spin is quantized, and accurate models for the interaction with spin require relativistic quantum mechanics or quantum field theory.

<span class="mw-page-title-main">Kicked rotator</span>

The kicked rotator, also spelled as kicked rotor, is a paradigmatic model for both Hamiltonian chaos and quantum chaos. It describes a free rotating stick in an inhomogeneous "gravitation like" field that is periodically switched on in short pulses. The model is described by the Hamiltonian

In physics, Berry connection and Berry curvature are related concepts which can be viewed, respectively, as a local gauge potential and gauge field associated with the Berry phase or geometric phase. The concept was first introduced by S. Pancharatnam as geometric phase and later elaborately explained and popularized by Michael Berry in a paper published in 1984 emphasizing how geometric phases provide a powerful unifying concept in several branches of classical and quantum physics.

For many paramagnetic materials, the magnetization of the material is directly proportional to an applied magnetic field, for sufficiently high temperatures and small fields. However, if the material is heated, this proportionality is reduced. For a fixed value of the field, the magnetic susceptibility is inversely proportional to temperature, that is

In NMR spectroscopy, the product operator formalism is a method used to determine the outcome of pulse sequences in a rigorous but straightforward way. With this method it is possible to predict how the bulk magnetization evolves with time under the action of pulses applied in different directions. It is a net improvement from the semi-classical vector model which is not able to predict many of the results in NMR spectroscopy and is a simplification of the complete density matrix formalism.

In pure and applied mathematics, quantum mechanics and computer graphics, a tensor operator generalizes the notion of operators which are scalars and vectors. A special class of these are spherical tensor operators which apply the notion of the spherical basis and spherical harmonics. The spherical basis closely relates to the description of angular momentum in quantum mechanics and spherical harmonic functions. The coordinate-free generalization of a tensor operator is known as a representation operator.

In physics, the Maxwell–Jüttner distribution, sometimes called Jüttner–Synge distribution, is the distribution of speeds of particles in a hypothetical gas of relativistic particles. Similar to the Maxwell–Boltzmann distribution, the Maxwell–Jüttner distribution considers a classical ideal gas where the particles are dilute and do not significantly interact with each other. The distinction from Maxwell–Boltzmann's case is that effects of special relativity are taken into account. In the limit of low temperatures much less than , this distribution becomes identical to the Maxwell–Boltzmann distribution.

References

Further reading