Maxwell–Jüttner Distribution
   HOME

TheInfoList



OR:

In
physics Physics is the natural science that studies matter, its fundamental constituents, its motion and behavior through space and time, and the related entities of energy and force. "Physical science is that department of knowledge which r ...
, the Maxwell–Jüttner distribution is the distribution of speeds of particles in a hypothetical gas of relativistic particles. Similar to Maxwell's 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's case is that effects of
special relativity In physics, the special theory of relativity, or special relativity for short, is a scientific theory regarding the relationship between space and time. In Albert Einstein's original treatment, the theory is based on two postulates: # The laws ...
are taken into account. In the limit of low temperatures T much less than mc^2/k (where m is the mass of the kind of particle making up the gas, c is the
speed of light The speed of light in vacuum, commonly denoted , is a universal physical constant that is important in many areas of physics. The speed of light is exactly equal to ). According to the special theory of relativity, is the upper limit ...
and k is
Boltzmann's constant The Boltzmann constant ( or ) is the proportionality factor that relates the average relative kinetic energy of particles in a gas with the thermodynamic temperature of the gas. It occurs in the definitions of the kelvin and the gas constant ...
), this distribution becomes identical to the Maxwell–Boltzmann distribution. The distribution can be attributed to Ferencz Jüttner, who derived it in 1911. It has become known as the Maxwell–Jüttner distribution by analogy to the name Maxwell-Boltzmann distribution that is commonly used to refer to Maxwell's distribution.


Definition

As the gas becomes hotter and kT approaches or exceeds mc^2, the probability distribution for \gamma=1/\sqrt in this relativistic Maxwellian gas is given by the Maxwell–Jüttner distribution: : f(\gamma) = \frac \exp \left( - \frac \right) where \beta = \frac =\sqrt, \theta=\frac, and K_2 is the modified
Bessel function Bessel functions, first defined by the mathematician Daniel Bernoulli and then generalized by Friedrich Bessel, are canonical solutions of Bessel's differential equation x^2 \frac + x \frac + \left(x^2 - \alpha^2 \right)y = 0 for an arbitrar ...
of the second kind. Alternatively, this can be written in terms of the momentum as : f(\mathbf) = \frac \exp\left( -\frac\right) where \gamma(p) = \sqrt. The Maxwell–Jüttner equation is covariant, but not manifestly so, and the temperature of the gas does not vary with the gross speed of the gas.


Jüttner distribution graph

A visual representation of the distribution in particle velocities for plasmas at four different temperatures: Where we have defined the thermal parameter \mu=mc^2/k_bT. The four general limits are: * ultrarelativistic temperatures \mu \ll 1 * relativistic temperatures: \mu < 1, * weakly (or mildly) relativistic temperatures: \mu > 1, * low temperatures: \mu \gg 1,


Limitations

Some limitations of the Maxwell–Jüttner distributions are shared with the classical ideal gas: neglect of interactions, and neglect of quantum effects. An additional limitation (not important in the classical ideal gas) is that the Maxwell–Jüttner distribution neglects antiparticles. If particle-antiparticle creation is allowed, then once the thermal energy kT is a significant fraction of mc^2, particle-antiparticle creation will occur and begin to increase the number of particles while generating antiparticles (the number of particles is not conserved, but instead the conserved quantity is the difference between particle number and antiparticle number). The resulting thermal distribution will depend on the
chemical potential In thermodynamics, the chemical potential of a species is the energy that can be absorbed or released due to a change of the particle number of the given species, e.g. in a chemical reaction or phase transition. The chemical potential of a species ...
relating to the conserved particle-antiparticle number difference. A further consequence of this is that it becomes necessary to incorporate statistical mechanics for indistinguishable particles, because the occupation probabilities for low kinetic energy states becomes of order unity. For fermions it is necessary to use Fermi–Dirac statistics and the result is analogous to the thermal generation of electron-
hole A hole is an opening in or through a particular medium, usually a solid body. Holes occur through natural and artificial processes, and may be useful for various purposes, or may represent a problem needing to be addressed in many fields of en ...
pairs in semiconductors. For
boson In particle physics, a boson ( ) is a subatomic particle whose spin quantum number has an integer value (0,1,2 ...). Bosons form one of the two fundamental classes of subatomic particle, the other being fermions, which have odd half-integer spi ...
ic particles, it is necessary to use the
Bose–Einstein statistics In quantum statistics, Bose–Einstein statistics (B–E statistics) describes one of two possible ways in which a collection of non-interacting, indistinguishable particles may occupy a set of available discrete energy states at thermodynamic ...
.See first few paragraphs i

for extended discussion.
Perhaps most significantly, the basic MB distribution has two main issues: it does not extend to particles moving at relativistic speeds, and  it assumes anisotropic temperature (where each DOF hasn’t the same translational kinetic energy). While the classic Maxwell-Juttner distribution generalizes for the case of special relativity, it fails to consider the anisotropic description.


Derivation

The Maxwell–Boltzmann (MB) distribution PM describes the velocities u or the kinetic energy \epsilon = \fracm\mathbf^ of the particles at thermal equilibrium, far from the limit of the speed of light, i.e: P_\left( \mathbf \right)\mathbfm^\mathbf^\mathbf^\exp\left( - \frac \right).\ \ \ \ (1a) \theta \equiv \sqrt,\ \ u \ll c Or, in terms of the kinetic energy: P_(\epsilon;T)=\fracexp(-\frac)\epsilon^ (1b) \epsilon \ll mc^ where θ is the temperature in speed dimensions, called thermal speed, and d denotes the kinetic degrees of freedom of each particle. (Note that the temperature is defined in the fluid’s rest frame, where the bulk speed ub is zero. In the non-relativistic case, this can be shown by using ε = \fracm(\mathbf\mathbf_)^ . The relativistic generalization of Eq. (1a), that is, the Maxwell–Jüttner (MJ) distribution, is given by: P_(\gamma)\propto\gamma^\beta(\gamma ) \bullet e^,\Theta \equiv \frac,\ E_ = mc^\ \ \ \ (2) where β ≡ \frac and γ (β) ≡ (1-\beta^)^\ . (Note that the inverse of the unitless temperature \Theta is the relativistic coldness ζ , Rezzola and Zanotti, 2013.) This distribution (Eq. 2) can be derived as follows. According to the relativistic formalism for the particle momentum and energy, we have \mathbfmc \bullet \gamma\left( \beta \right) \bullet \mathbf,\ E\left( \beta \right) = \gamma(\beta) \bullet E_\ \ \ \ \ \ \ \ \ (3) \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ While the kinetic energy is given by \epsilon = E - E_ = (\gamma - 1) \bullet E_. The Boltzmann distribution of a Hamiltonian is Pmj(H) \propto \exp In the absence of a potential energy, H is simply given by the particle energy E, thus: P_\left( E \right)\operatorname\left( - \frac \right) \propto \exp\left( - \frac \right)\ \ \ \ \ (4a) \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ (Note that E is the sum of the kinetic ε and inertial energy E_,\frac = \frac). Then, when we include the d-dimensional density of states: P_\left( \gamma \right) \propto p\left( \gamma \right)^\frac \bullet \exp\left( - \frac \right)\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ (4b) \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ So that:
Where d\Omega_\ denotes the d-dimensional solid angle. For isotropic distributions, we have
\int P_(p)dp_1....dp_d\propto \int exp \fractimes \ \ \ \ \ \ (5a) (\gamma)^\frac\Omega_dd\gamma\equiv \int_d\Omega_d\bullet\int P_(\gamma)d\gamma, or P_(\gamma)\propto exp \fracbullet p(\gamma)^\frac \ \ \ \ \ \ (5b) Then, d\left( \text \right) = \gamma\left( \gamma^ - 1 \right)^d\gamma = \beta^\text so that: \ \ \ \ \ \ \ (6) Or: P_(\gamma)\propto \gamma^\beta^,e^\propto\gamma(\gamma^2-1)^\bullet e^ \ \ \ \ \ (7) \ Now, because \frac = \frac. Then, we normalize the distribution Eq. (7). We set P_(p,\Theta)dp_1dp_2....dp_d \ \ \ \ \ \ (8) =N\bullet e^\fracdp_1dp_2....dp_d And the angular integration: Where B_ = 2\pi^/\Gamma\left( \frac \right) is the surface of the unit d-dimensional sphere. Then, using the identity \gamma^ = \left( \frac \right)^ + 1 we have:
P_(\mathbf;\mathbf\ )dp_dp_\ldots dp_\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ (9)\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
= N \bullet \fracB_\left( \text \right)^ \bullet e^\left( \gamma^ - 1 \right)^d\left( \gamma^ - 1 \right). and Where we have defined the integral: I_ \equiv \int_^d\gamma.\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ (11) \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ The Macdonald function (Avramovitz and Stegun, 1972, p.376) is defined by: K_n(z)\equiv \frac\int_1^\infin e^(\gamma^2-1)^d\gamma \ \ \ \ \ \ (12) So that, by setting n = \frac,\ z = \frac we obtain: I_d = \Gamma(\frac+1)\pi^K_(\frac)(2\Theta)^ \ \ \ \ \ (13) Hence, Or N = \ \pi^2^\left( \text \right)^ \bullet \Theta^K_\left( \frac \right)^\ ,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \left( 14b \right) The inverse of the normalization constant gives the partition function Z \equiv \frac: Z = \pi^2^\left( \text \right)^ \bullet \Theta^K_\left( \frac \right),\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \left( 14c \right) Therefore, the normalized distribution is: P_(P;\Theta)dp_1dp_2...dp_d \ \ \ \ \ (15a) =\pi^2^(mc)^\bullet\Theta^K_(\frac)^ \times exp \fracp_1dp_2....dp_d Or we may derive the normalized distribution in terms of: P_(\gamma;\Theta)d\gamma \ \ \ \ \ \ \ (15b) =\fracK_(\frac)^\Theta^e^(\gamma^2-1)^ \gamma d\gamma Note that \Theta can be shown to coincide with the thermodynamic definition of temperature. Also useful is the expression of the distribution in the velocity space (Dunkel et al., 2007). Given that \frac = \gamma^, we have: Hence \ \ \ \ \ (15c) Take d=3 (the “classic case” in our world): P_(p;\Theta)dp_1dp_2.....dp_d \ \ \ \ \ \ (16a) = \frac\left( \text \right)^ \bullet \fracK_\left( \frac \right)^ \bullet e^dp_dp_dp_ And \ \ \ \ \ (16b) \frac \bullet \fracK_\left( \frac \right)^ \bullet \exp\left\lbrack - \frac\ \right\rbrack\gamma\left( \beta \right)^d\beta_d\beta_d\beta_ Note that when the MB distribution clearly deviates from the MJ distribution of the same temperature and dimensionality, one can misinterpret and deduce a different MB distribution that will give a good approximation to the MJ distribution. This new MB distribution can be either (i) a convected MB distribution, that is, an MB distribution with the same dimensionality, but with different temperature Tmb and bulk speed ub (or bulk energyE_b\equiv \fracmu+u_B^2), or (ii) an MB distribution with the same bulk speed, but with different temperature TMB and degrees of freedom dMB. These two types of approximations are illustrated.


References

{{DEFAULTSORT:Maxwell-Juttner distribution Gases Special relativity