Next Article in Journal
Information Technology Project Portfolio Implementation Process Optimization Based on Complex Network Theory and Entropy
Next Article in Special Issue
Initial Results of Testing Some Statistical Properties of Hard Disks Workload in Personal Computers in Terms of Non-Extensive Entropy and Long-Range Dependencies
Previous Article in Journal
Multiscale Entropy Analysis of Unattended Oximetric Recordings to Assist in the Screening of Paediatric Sleep Apnoea at Home
Previous Article in Special Issue
Peierls–Bogolyubov’s Inequality for Deformed Exponentials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On the Simplification of Statistical Mechanics for Space Plasmas

Division of Space Science and Engineering, Southwest Research Institute, San Antonio, TX 78238, USA
Entropy 2017, 19(6), 285; https://doi.org/10.3390/e19060285
Submission received: 17 May 2017 / Revised: 15 June 2017 / Accepted: 16 June 2017 / Published: 18 June 2017
(This article belongs to the Collection Advances in Applied Statistical Mechanics)

Abstract

:
Space plasmas are frequently described by kappa distributions. Non-extensive statistical mechanics involves the maximization of the Tsallis entropic form under the constraints of canonical ensemble, considering also a dyadic formalism between the ordinary and escort probability distributions. This paper addresses the statistical origin of kappa distributions, and shows that they can be connected with non-extensive statistical mechanics without considering the dyadic formalism of ordinary/escort distributions. While this concept does significantly simplify the usage of the theory, it costs the definition of a dyadic entropic formulation, in order to preserve the consistency between statistical mechanics and thermodynamics. Therefore, the simplification of the theory by means of avoiding dyadic formalism is impossible within the framework of non-extensive statistical mechanics.

1. Introduction

Kappa distributions were employed to describe numerous space plasma populations in: (i) the inner heliosphere, including solar wind [1,2,3,4,5,6,7,8,9,10,11,12,13], solar spectra [14,15], solar corona [16,17,18,19], solar energetic particles [20,21], corotating interaction regions [22], and related solar flares [23,24,25,26]; (ii) the planetary magnetospheres including the magnetosheath [27,28], near the magnetopause [29], magnetotail [30], ring current [31], plasma sheet [32,33,34], ionosphere [35], magnetospheric substorms [36], magnetospheres of giant planets, such as Jovian [37,38], Saturnian [39,40,41], Uranian [42], Neptunian [43], magnetospheres of planetary moons, such as Io [44] and Enceladus [45], or cometary magnetospheres [46,47]; (iii) the outer heliosphere and the inner heliosheath [7,8,9,10,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62]; (iv) beyond the heliosphere, including HII regions [63], planetary nebula [64,65], and supernova magnetospheres [66]; or other space plasma-related analyses [53,54,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90]. Maxwell distributions have also been used in space science [91], especially due to their simplicity; for example, they are often used to fit the “core” of the observed distributions, that is, the part of the distribution around its maximum.
A breakthrough in the field came with the connection of kappa distributions with statistical mechanics, and specifically with the concept of non-extensive statistical mechanics. This topic has been examined by several authors [67,92,93,94,95,96]. The empirical kappa distribution and the theoretical q-exponential distribution (that results from non-extensive statistical mechanics) are accidentally of the same form under the transformation of indices: q = 1 + 1 / κ . Livadiotis & McComas [94] showed that the consistent connection of the theory and formalism of kappa distributions with non-extensive statistical mechanics is based on four fundamental physical notions and concepts: (i) q-deformed exponential and logarithm functions [97,98]; (ii) escort probability distribution [99]; (iii) Tsallis entropy [100]; and (iv) physical temperature [101,102,103] (that is, the actual temperature of the system). Understanding the statistical origin of kappa distributions led to further theoretical developments and applications, including the (a) foundation theory; (b) plasma formalism; and (c) space plasma applications (for more details, see: [96]).
Non-extensive statistical mechanics was introduced by Tsallis [100] as a generalization of the classical Boltzmann Gibbs (BG) statistical mechanics. The well-known q-exponential function was deduced as a result of the maximization of Tsallis entropy under the constraints of the canonical ensemble. However, this result was facing some fundamental problems: In particular, the canonical distribution of energy ε was not invariant for an arbitrary selection of the ground level of energy, and the internal energy (mean energy), U = ε , was not extensive as it should be for uncorrelated distributions. These inconsistencies as well as other problems were corrected by the later work of Tsallis et al. [104] that considers the escort-expectation values. The modern theory of non-extensive statistics is based on the generalization of two elements, fundamental for statistical mechanics, (i) the classical formulation of Boltzmann-Gibbs entropy, and (ii) the notion of canonical distribution via the formalism of ordinary/escort probabilities; (see: [99,105,106]; see also: [94,95,96]).
The purpose of this paper is to show that the statistical framework of non-extensive statistical mechanics can be deduced as it is without the consideration of the dyadic formalism of ordinary/escort distributions. This concept can significantly simplify the usage of the theory, and make it accessible by the new generation of researchers that struggle to understand and apply it in exotic particle systems out of thermal equilibrium, such as the space plasmas. However, the cost of this simplification is the necessity of having two types of entropic functions. This dyadic entropic formulation preserves the basic fundamental thermodynamic formulae, which is necessary for the consistent connection of statistical mechanics with thermodynamics.
Next, we first expose the existing formalism of non-extensive statistical mechanics, then, in Section 2, we briefly present the above two elements (i) and (ii), and in Section 3, we present the maximization of the entropy under the constraint of canonical ensemble. Then, in Section 4, we present an alternative formalism that skips the consideration of ordinary/escort formalism. In Section 5, we show the consistency of the introduced formalism with thermodynamics, but with a dyadic formalism of entropy that is needed. Finally, in Section 6, we briefly summarize the conclusions.

2. Tsallis Entropy and Ordinary/Escort Formalism

Consider a system described by a discrete energy spectrum { ε k } k = 1 W , which is associated with a discrete probability distribution { p k } k = 1 W . Non-extensive statistical mechanics generalizes the classical formulation of BG entropy and uses the dyadic formalism of ordinary/escort probabilities. For the discrete energy spectrum { ε k } k = 1 W and the associated distribution { p k } k = 1 W , the non-extensive entropy is given by:
S ( q ) = 1 q 1 ( 1 k = 1 W p k q )   or   S ( q ) = 1 q 1 [ 1 ( k = 1 W P k 1 / q ) q ] ,
or in terms of the kappa index:
S ( κ ) = κ ( 1 k = 1 W p k κ + 1 κ )   or   S ( κ ) = κ [ 1 ( k = 1 W P k κ κ + 1 ) κ + 1 κ ] ,
leading to the BG formulation for κ→∞ (or q→1):
S ( q 1 ) = S ( κ ) = k = 1 W p k ln ( p k ) .
Note that all the entropic formulations are given in units of the Boltzmann’s constant, that is, setting k B = 1 for simplicity; this will be restored further below.
On the other hand, the formalism of non-extensive statistical mechanics is built together with the concept of escort probabilities [99]. The escort probability distribution { P k } k = 1 W is constructed from the ordinary probability distribution, { p k } k = 1 W , as P k p k q , k = 1 , , W , coinciding thus, with { p k } k = 1 W for q 1 . In fact, there is a duality between the ordinary { p k } k = 1 W and escort probabilities { P k } k = 1 W , such as P k = P k ( { p k } k = 1 W ; q ) and p k = p k ( { P k } k = 1 W ; 1 q ) , k = 1 , , W , given by:
P k = p k q k = 1 W p k q p k = P k 1 / q k = 1 W P k 1 / q , k = 1 , , W
According to the Tsallis statistical interpretation of the internal energy U , this is given by the escort expectation value of energy ε e s c , that is:
U = ε e s c = k = 1 W P k ε k = k = 1 W p k q ε k k = 1 W p k q ,
where the symbol e s c denotes the escort expectation value. (Hereafter, the subscript “esc” will be ignored, since the expectation value will be taken associated with the escort distributions by default, unless it is stated otherwise.)
It is noted that the ordinary and escort distributions have differences in their meaning and usage. The ordinary distribution is rather an auxiliary mathematical tool related to the information arrangement and its dynamics, while the escort distribution is related to the observed stationary states and measurements (statistical moments).

3. The Entropy Maximization in the Framework of Non-Extensive Statistical Mechanics

In order to obtain the probability distribution { p k } k = 1 W , associated with the conservative physical system of energy spectrum { ε k } k = 1 W , we follow along the famous Gibbs’ path [107], where the entropy S = S ( { p k } k = 1 W ) is maximized under the constraints of canonical ensemble.
The maximum of entropy is derived from:
p j S ( p 1 , p 2 , , p W ) = 0 , j = 1 , , W
On the other hand, the “variables” { p k } k = 1 W are not independent, because of the two constraints: (i) Normalization of the probability distribution; (ii) Fixed value of internal energy; i.e.,
( i ) k = 1 W p k = 1 , ( ii ) k = 1 W p k ε k = U .
Then, the Lagrange method applies, which involves maximizing a functional form where unknown Lagrange multipliers λ 1 , λ 2 , are linearly expressed in terms of the constraints:
G ( { p k } k = 1 W ) = S ( { p k } k = 1 W ) + λ 1 k = 1 W p k + λ 2 k = 1 W p k ε k ,
where the maximization follows by setting ( / p j ) G ( { p k } k = 1 W ) = 0 , j = 1 , , W .
In the case of non-extensive statistical mechanics, the interpretation for the internal energy U is given by the escort expectation value as shown in Equation (3), thus, we maximize the functional form:
G ( p 1 , p 2 , , p W ) = 1 q 1 ( 1 k = 1 W p k q ) + λ 1 k = 1 W p k + λ 2 k = 1 W P k ( { p k } k = 1 W ) ε k .
The maximization of this functional gives:
q q 1 p j q 1 + λ 1 + λ 2 k = 1 W ε k p j P k ( { p k } k = 1 W ) ,
leading to:
p j = 1 Z ¯ [ 1 ( 1 q ) β ( ε j U ) ] 1 1 q , Z ¯ = j = 1 W [ 1 ( 1 q ) β ( ε j U ) ] 1 1 q ,
where we set the auxiliary partition function Z ¯ ( λ 1 q 1 q ) 1 1 q ; the inverse-temperature β = ( k B T ) 1 relates the Lagrangian one, β L = ( k B T L ) 1 , defined by the negative value of the second Lagrange multiplier, β L λ 2 , that is:
β β L / φ q T T L φ q ,   with   φ q k = 1 W p k q .
Note that T (also known as physical temperature, e.g., [102]) represents the actual temperature of the system, thus it does not depend on the q-index. On the other hand, the Lagrangian “temperature” T L is a thermal parameter that depends on the actual temperature T and the q-index, T L = T L ( T ; q ) . At the classical limit of q→1, T L coincides with the system’s temperature [96].
The system is described by the escort distribution, defined by P j ( ε j ) p j ( ε j ) q , i.e.,
P j ( ε j ) = 1 Z ¯ [ 1 + ( q 1 ) β ( ε j U ) ] 1 q 1 1 .
We can rewrite this, using the Q-deformed exponential function [97,98] and its inverse, the Q-logarithm function, defined by:
exp Q ( x ) = [ 1 + ( 1 Q ) x ] + 1 Q 1 , ln Q ( x ) = 1 x 1 Q Q 1 .
We also use the q-deformed “unity function” [94], defined by:
1 Q ( x ) = [ 1 + ( 1 Q ) x ] +
The subscript “+” denotes the cut-off condition, where exp Q ( x ) becomes zero if its base is non-positive.
Setting Q = q, the ordinary/escort distributions are written as:
p j ( ε j ) exp q [ β ( ε j U ) ] P j ( ε j ) exp q q [ β ( ε j U ) ] .
In the above, the escort distribution P j ( ε j ) is expressed by the q-exponential function, in terms of the q-index, but it can be transformed to its equivalent, the kappa distribution, under the transformation of the kappa and q indices:
κ = 1 q 1 q = 1 + 1 κ .
Namely, we derive the discrete kappa distribution:
P j ( ε j ) [ 1 + 1 κ β ( ε j U ) ] κ 1 .
In the continuous description, the kappa distribution attains its well-known form:
P ( ε ) [ 1 + 1 κ β ( ε U ) ] κ 1 ,
(e.g., see [96], Chapter 1, and reference therein).

4. Avoiding Ordinary/Escort Formalism

First, we start with a classical case of BG entropy:
S = S ( p 1 , p 2 , , p W ) = k = 1 W p k ln ( p k ) .
The maximization of the entropy under the constraints of canonical ensemble requires ( / p j ) G ( p 1 , p 2 , , p W ) = 0 , with:
G ( p 1 , p 2 , , p W ) = k = 1 W p k ln ( p k ) + λ 1 k = 1 W p k + λ 2 k = 1 W p k ε k ,
leading to 1 ln ( p j ) + λ 1 + λ 2 ε j = 0 , i.e.,
p j ( ε j ) = exp ( λ 1 1 ) exp ( λ 2 ε j ) .
We consider the “mean-less” energies, ε j U :
p j ( ε j ) = exp ( λ 1 1 + λ 2 U ) exp [ λ 2 ( ε j U ) ] ,   or ,
p j ( ε j ) = 1 Z ¯ exp [ β ( ε j U ) ] ,
Z ¯ [ exp ( λ 1 1 + λ 2 U ) ] 1 = j = 1 W exp [ β ( ε j U ) ] ,
where we used the auxiliary partition function Z ¯ . The inverse-temperature β = ( k B T ) 1 is now given by the Lagrangian one β = β L , defined via the Lagrange multiplier (with opposite sign), β L λ 2 .
Next, we proceed with the Tsallis entropy:
S = S ( p 1 , p 2 , , p W ) = 1 q 1 ( 1 k = 1 W p k q ) = 1 q 1 k = 1 W ( p k p k q ) .
Again, the maximization of the entropy under the constraints of the canonical ensemble involves maximizing the functional:
G ( p 1 , p 2 , , p W ) = 1 q 1 k = 1 W ( p k p k q ) + λ 1 k = 1 W p k + λ 2 k = 1 W p k ε k .
Hence, ( / p j ) G ( p 1 , p 2 , , p W ) = 0 , gives:
1 q 1 q q 1 p j q 1 + λ 1 + λ 2 ε j = 0 ,   hence ,
p j ( ε j ) = [ 1 + ( 1 q 1 ) ( λ 1 1 + λ 2 ε j ) ] q 1 1 q 1 = exp q 1 q 1 ( λ 1 1 + λ 2 ε j ) ,   or :
p j ( ε j ) = [ 1 + ( 1 q 1 ) ( λ 1 1 ) ] q 1 1 q 1 [ 1 + ( 1 q 1 ) λ 2 ε j 1 + ( 1 q 1 ) ( λ 1 1 ) ] q 1 1 q 1 = exp q 1 q 1 ( λ 1 1 ) exp q 1 q 1 [ λ 2 ε j 1 q 1 ( λ 1 1 ) ]
where we used the q-deformed unity and exponential functions for Q = q 1 .
We enter again the mean-less energies, ε j U :
p j ( ε j ) = [ 1 + ( 1 q 1 ) ( λ 1 1 + λ 2 U ) ] q 1 1 q 1 [ 1 + ( 1 q 1 ) λ 2 ( ε j U ) 1 + ( 1 q 1 ) ( λ 1 1 + λ 2 U ) ] q 1 1 q 1 ,
which may be rewritten using the Q-deformed exponential function, Equation (12), by setting Q = q 1 ; hence:
p j ( ε j ) = exp q 1 q 1 ( λ 1 1 + λ 2 U ) exp q 1 q 1 [ λ 2 ( ε j U ) 1 + ( 1 q 1 ) ( λ 1 1 + λ 2 U ) ] .
Also, we use again the notions of the auxiliary partition function Z ¯ and the inverse-temperature β:
p j ( ε j ) = 1 Z ¯ exp q 1 q 1 [ β ( ε j U ) ] ,
Z ¯ exp q 1 q 1 ( λ 1 1 + λ 2 U ) = j = 1 W exp q 1 q 1 [ β ( ε j U ) ] ,
where now the actual (inverse) temperature β does not coincide with the Lagrangian β L :
β β L Z ¯ 1 q , β L λ 2 .
It can be easily shown that for the argument φ q j = 1 W p j q , we have φ q = Z ¯ 1 q (next section), and:
β β L Z ¯ 1 q = β L φ q ,   T T L Z ¯ 1 q = T L φ q .
Finally, the distributions can be expressed in terms of the kappa index:
p j ( ε j ) = 1 Z ¯ [ 1 + ( q 1 1 ) β ( ε j U ) ] q 1 q 1 1 = 1 Z ¯ [ 1 + 1 κ β ( ε j U ) ] κ 1 ,
Z ¯ exp q 1 q 1 ( λ 1 1 + λ 2 U ) = j = 1 W [ 1 + 1 κ β ( ε j U ) ] κ 1 ,
where we set:
κ 1 Q 1 , Q = q 1 or κ q 1 q q = κ κ + 1 .
It must be mentioned that in the previous developments, the entropy and the deduced q-exponential distribution (in the framework of the canonical ensemble) are not characterized by the same q-index, but they are inverse to each other. Namely:
S ( p 1 , p 2 , , p W ) = k B 1 q 1 k = 1 W ( p k p k q ) Q = q 1 p j ( ε j ) = 1 Z ¯ [ 1 + ( Q 1 ) β ( ε j U ) ] Q Q 1

5. Connection to Thermodynamics

First, we show the following relations:
φ q = Z ¯ 1 q ,
β L Z ¯ = Z ¯ β U β L .
The identity (32a) is derived from:
φ q = j = 1 W p j q = j = 1 W p j p j q 1 = Z ¯ 1 q j = 1 W p j [ 1 + ( q 1 1 ) β ( ε j U ) ] = Z ¯ 1 q j = 1 W p j = Z ¯ 1 q .
Then, we observe that:
Z ¯ = k = 1 W [ 1 + ( q 1 1 ) β ( ε j U ) ] 1 q 1 = k = 1 W [ 1 + ( q 1 1 ) β ( ε j U ) ] 1 q 1 + 1 ,
because:
k = 1 W [ 1 + ( q 1 1 ) β ( ε j U ) ] 1 q 1 ( ε j U ) = 0 .
Hence, we show (32b):
β L Z ¯ = k = 1 W [ 1 + ( q 1 1 ) β ( ε j U ) ] 1 q 1 [ β β L ( ε j U ) + β U β L ] = Z ¯ β U β L .
We define the “actual” partition function, as follows:
Z Z ¯ [ 1 + ( q 1 ) β U ] 1 q 1 = Z ¯ exp q ( β U ) o r ln q Z = ln q Z ¯ β L U
As we will see, Equation (32a, b), together with the definition of the partition function Z in (34), lead to the relations required for the connection of statistical mechanics with thermodynamics. First, the entropy is related with the auxiliary partition function Z ¯ :
S = k B 1 φ q q 1 = k B 1 Z ¯ 1 q q 1 = k B ln q Z ¯ .
Then, from Equation (32b), we obtain:
β L ln q Z ¯ = Z ¯ q β L Z ¯ = Z ¯ 1 q β U β L = β L U β L .
Hence, Equation (35) leads to:
( S / k B ) β L = β L ln q Z ¯ = β L U β L ,   or :
S U = k B β L = 1 T L .
Next, using again Equation (36):
β L ln q Z ¯ = U β L = ( β L U ) β L U or U = β L ( ln q Z ¯ β L U )
and the actual partition function Z defined in Equation (34), we obtain:
U = ln q Z β L .
Equation (34) can be rewritten as U ln q Z ¯ / β L = ln q Z / β L , and taking into account Equation (35), we arrive at:
S = k B ln q Z + U T L ,
and the Helmholtz free energy:
F U T L S = k B T L ln q Z .
(Equations (37a–d) constitute the basic formulae connecting statistical mechanics with thermodynamics; e.g., see other similar cases in [108,109].)
Next, we define an alternative entropic formula. Let the standard Tsallis entropy involved in Equation (22) be denoted by S q . An associated auxiliary formulation of entropy S q 1 is defined by:
exp q 1 ( S q 1 ) exp q ( S q ) = Z ¯   or :
[ 1 + ( 1 q 1 ) S q 1 ] 1 1 q 1 = [ 1 + ( 1 q ) S q ] 1 1 q ,
(where we have again set k B = 1 ). The actual temperature of the system, T, is given by the physical temperature (Equation (10)), and can be related to the entropy via:
[ 1 + ( 1 q ) S q ] 1 S q U = φ q 1 T L 1 = T 1 .
Both entropic formulations, S q and S q 1 , are associated with the same physical temperature T, which means that both quantities are equivalently related to the actual temperature of the system. Indeed, starting from Equation (38a), we obtain:
U exp q ( S q ) = U exp q 1 ( S q 1 ) exp q q ( S q ) S q U = exp q 1 q 1 ( S q 1 ) S q 1 U [ 1 + ( 1 q ) S q ] 1 exp q ( S q ) S q U = [ 1 + ( 1 q 1 ) S q 1 ] 1 exp q 1 ( S q 1 ) S q 1 U ,
hence:
[ 1 + ( 1 q ) S q ] 1 S q U = [ 1 + ( 1 q 1 ) S q 1 ] 1 S q 1 U T 1 .
Then, the entropy must increase when the system transits through the stationary states toward thermal equilibrium (κ0→∞); in other words:
( S κ ) κ 0 > 0 , with   κ 1 q 1 1
As we will see below, in the continuous description, the entropy Sq does not follow the above condition, i.e., ( S q / κ ) κ < 0 . However, the dyadic entropy does, i.e., ( S q 1 / κ ) κ 0 > 0 .
In the continuous description, the kappa distribution of particle velocity is written as:
p ( u ) = 1 Z ¯ [ 1 + 1 κ ε ( u ) U k B T ] κ 1 , Z ¯ = + [ 1 + 1 κ ε ( u ) U k B T ] κ 1 d u 1 d u 2 d u f
(for f kinetic degrees of freedom). The argument φ q , which is related with the entropy as shown in Equation (35), is given by:
φ q + [ p ( u ) σ u f ] q d u 1 d u 2 d u f σ u f = σ u f ( q 1 ) + p ( u ) q d u 1 d u 2 d u f = σ u f ( q 1 ) + p ( u ) q 1 p ( u ) d u 1 d u 2 d u f = σ u f ( q 1 ) Z ¯ 1 q + [ 1 + 1 κ ε ( u ) U k B T ] p ( u ) d u 1 d u 2 d u f = σ u f ( q 1 ) Z ¯ 1 q = ( Z ¯ / σ u f ) 1 q ,
where the speed scale σ u is used, in order to derive the number of the possible states that the system can attain in the velocity space, i.e., d u 1 d u 2 d u f / σ u f (see: [24,53,96], Chapter 2).
We can define the dimensionless probability distribution density and partition function:
p σ ( u ) = 1 Z ¯ [ 1 + 1 κ ε ( u ) U k B T ] κ 1 , Z ¯ σ + [ 1 + 1 κ ε ( u ) U k B T ] κ 1 d u 1 d u 2 d u f σ u f
Hence:
φ q = ( Z ¯ / σ u f ) 1 q = Z ¯ σ 1 q ,
which reads the continuous version of Equation (32a). We calculate φ q as follows:
φ q + [ p ( u ) σ u f ] q d u 1 d u 2 d u f σ u f = σ u f ( q 1 ) + p ( u ) q d u 1 d u 2 d u f = σ u f κ + 1 p ( u ) κ κ + 1 d u ,
where we set σ σ u / θ . Then, we have:
φ q = [ ( σ 2 π e ) 1 2 d N C Γ ( κ 0 + d 2 N C ) ( κ 0 + d 2 N C ) κ 0 + 1 2 d N C κ 0 κ 0 Γ ( κ 0 ) e 1 2 d N C ] 1 κ + 1 = [ ( σ 2 π e ) 1 2 d N C Γ ( κ 0 + d 2 N C ) ( κ 0 + d 2 N C ) κ 0 + 1 2 d N C κ 0 κ 0 Γ ( κ 0 ) e 1 2 d N C ] 1 q ,
where we set κ = κ 0 + 1 2 f and f = d N C ; NC is the number of particles and d denotes the per-particle kinetic and potential degrees of freedom. Hence:
φ q = Z ¯ σ 1 q ,   Z ¯ σ = σ ˜ 1 2 d N C Γ ( κ 0 + d 2 N C ) ( κ 0 + d 2 N C ) κ 0 + 1 2 d N C κ 0 κ 0 Γ ( κ 0 ) e 1 2 d N C ,
where we modified the speed scale parameter to:
σ ˜ σ 2 π e = ( σ u / θ ) 2 π e .
The entropy is associated with Z ¯ σ similar to Equation (35), according to:
S = k B 1 φ q q 1 = k B 1 Z ¯ σ 1 q q 1 = k B ln q ( Z ¯ σ ) ,   or
exp q ( S q / k B ) = Z ¯ σ = exp q 1 ( S q 1 / k B ) .
In terms of the kappa index, Equation (49) is written as:
S q = k B ( κ + 1 ) ( 1 Z ¯ σ 1 κ + 1 ) S q 1 = k B κ ( 1 Z ¯ σ 1 κ )
Figure 1 plots both the entropic formulations S q and S q 1 , as a function of the (invariant) kappa index κ0. We observe that the entropy S q decreases monotonically when the kappa index increases, while S q 1 increases when the kappa index increases for stationary states near thermal equilibrium (κ0→∞). For large values of the kappa index tending to infinity, both entropic formulations tend to be equal to each other, recovering BG entropy. The two functional behaviors are summarized as:
( S q κ ) κ 0 < 0 and ( S q 1 κ ) κ 0 > 0
where κ ( Q 1 ) 1 with Q = q 1 .

6. Conclusions

This paper shows that the maximization of Tsallis entropy can directly lead to the canonical probability distribution expressed by a q-exponential or kappa distribution, without using the dyadic formalism of ordinary/escort probability distributions. In particular, it was shown (using discrete description for simplicity) that the derived distribution is:
S q ( p 1 , p 2 , , p W ) = k B 1 q 1 k = 1 W ( p k p k q ) Q = q 1 p j ( ε j ) = 1 Z [ 1 + ( Q 1 ) β ( ε j U ) ] Q Q 1
The above distribution depends on the Q-index; it leads to the kappa distribution by the standard transformation between kappa and the Q index: κ ( Q 1 ) 1 . The Q index, however, is not exactly the same as the entropic index q, but is its inverse, Q = q 1 . The derived kappa distribution is applicable for positive kappa indices, i.e., κ>0 or Q>1, that is, the entropic index q<1. However, the standard entropic formulation is associated with the kappa distribution is characterized by q>1.
This inconsistency has some unphysical result: The entropy does not increase with the increase of the kappa index for stationary states near thermal equilibrium (large kappa indices). As it was shown for the continuous description, for example, it appears that entropy monotonically decreases with the kappa index. In order to avoid this undesired result, a dyadic entropic formulation with index Q = q 1 instead of q, must be defined:
exp q ( S q / k B ) = Z ¯ = exp q 1 ( S q 1 / k B ) , or S q = k B ln q ( Z ¯ ) S q 1 = k B ln q 1 ( Z ¯ )
In summary, the escort expectation approach is the proper way to formulate non-extensive statistical mechanics. If we erroneously attempt to avoid the use of escort probability, then we will run into a physically inconsistent situation that involves the connection of statistical mechanics with thermodynamics. In such a case, a double entropic formulation must be defined to resolve the inconsistency. More precisely, it is shown that if the known dyadic distribution formalism is used then the entropic formulation is unique; however, if a unique distribution formulation is used then a dyadic entropic formulation is necessary, according to the scheme in Figure 2. Therefore, we conclude that simplification by means of avoiding dyadic formalism (either of distributions or entropic formulations) is impossible within the framework of non-extensive statistical mechanics (in order to be consistent with thermodynamics).

Acknowledgments

The work was supported in part by the project NNX17AB74G of NASA’s HGI Program.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Maksimovic, M.; Pierrard, V.; Lemaire, J. A kinetic model of the solar wind with Kappa distributions in the corona. Astron. Astrophys. 1997, 324, 725–734. [Google Scholar]
  2. Pierrard, V.; Maksimovic, M.; Lemaire, J. Electron velocity distribution function from the solar wind to the corona. J. Geophys. Res. 1999, 104, 17021–17032. [Google Scholar] [CrossRef]
  3. Mann, G.; Classen, H.T.; Keppler, E.; Roelof, E.C. On electron acceleration at CIR related shock waves. Astron. Astrophys. 2002, 391, 749–756. [Google Scholar] [CrossRef]
  4. Marsch, E. Kinetic physics of the solar corona and solar wind. Living Rev. Sol. Phys. 2006, 3, 1. [Google Scholar] [CrossRef]
  5. Zouganelis, I. Measuring suprathermal electron parameters in space plasmas: Implementation of the quasi-thermal noise spectroscopy with kappa distributions using in situ Ulysses/URAP radio measurements in the solar wind. J. Geophys. Res. 2008. [Google Scholar] [CrossRef]
  6. Štverák, S.; Maksimovic, M.; Travnicek, P.M.; Marsch, E.; Fazakerley, A.N.; Scime, E.E. Radial evolution of nonthermal electron populations in the low-latitude solar wind: Helios, Cluster, and Ulysses Observations. J. Geophys. Res. 2009. [Google Scholar] [CrossRef]
  7. Livadiotis, G.; McComas, D.J. Exploring transitions of space plasmas out of equilibrium. Astrophys. J. 2010, 714, 971. [Google Scholar] [CrossRef]
  8. Livadiotis, G.; McComas, D.J. The influence of pick-up ions on space plasma distributions. Astrophys. J. 2011, 738, 64. [Google Scholar] [CrossRef]
  9. Livadiotis, G.; McComas, D.J. Understanding kappa distributions: A toolbox for space science and astrophysics. Space Sci. Rev. 2013, 75, 183–214. [Google Scholar] [CrossRef]
  10. Livadiotis, G.; McComas, D.J. Near-equilibrium heliosphere-Far-equilibrium heliosheath. In Proceedings of the 13th International Solar Wind Conference, Big Island, HI, USA, 17–22 June 2012; pp. 344–347. [Google Scholar]
  11. Yoon, P.H. Electron kappa distribution and quasi-thermal noise. J. Geophys. Res. 2014, 119, 7074. [Google Scholar] [CrossRef]
  12. Pierrard, V.; Pieters, M. Coronal heating and solar wind acceleration for electrons, protons, and minor ions, obtained from kinetic models based on kappa distributions. J. Geophys. Res. 2015, 119, 9441. [Google Scholar] [CrossRef]
  13. Pavlos, G.P.; Malandraki, O.E.; Pavlos, E.G.; Iliopoulos, A.C.; Karakatsanis, L.P. Non-extensive statistical analysis of magnetic field during the March 2012 ICME event using a multi-spacecraft approach. Phys. A Stat. Mech. Appl. 2016, 464, 149–181. [Google Scholar] [CrossRef]
  14. Dzifčáková, E.; Dudík, J. H to Zn ionization equilibrium for the non-Maxwellian electron κ-distributions: Updated calculations. Astrophys. J. Suppl. Ser. 2013, 206, 6. [Google Scholar] [CrossRef]
  15. Dzifčáková, E.; Dudík, J.; Kotrč, P.; Fárník, F.; Zemanová, A. KAPPA: A package for synthesis of optically thin spectra for the non-Maxwellian κ-distributions based on the Chianti database. Astrophys. J. Suppl. Ser. 2015, 217, 14. [Google Scholar] [CrossRef]
  16. Owocki, S.P.; Scudder, J.D. The effect of a non-Maxwellian electron distribution on oxygen and iron ionization balances in the solar corona. Astrophys. J. 1983, 270, 758–768. [Google Scholar] [CrossRef]
  17. Vocks, C.; Mann, G.; Rausche, G. Formation of suprathermal electron distributions in the quiet solar corona. Astron. Astrophys. 2008, 480, 527–536. [Google Scholar] [CrossRef]
  18. Lee, E.; Williams, D.R.; Lapenta, G. Spectroscopic indication of suprathermal ions in the solar corona. arXiv 2013. [Google Scholar]
  19. Cranmer, S.R. Suprathermal electrons in the solar corona: Can nonlocal transport explain heliospheric charge states? Astrophys. J. Lett. 2014. [Google Scholar] [CrossRef]
  20. Xiao, F.; Shen, C.; Wang, Y.; Zheng, H.; Whang, S. Energetic electron distributions fitted with a kappa-type function at geosynchronous orbit. J. Geophys. Res. 2008. [Google Scholar] [CrossRef]
  21. Laming, J.M.; Moses, J.D.; Ko, Y.K.; Ng, C.K.; Rakowski, C.E.; Tylka, A.J. On the remote detection of suprathermal ions in the solar corona and their role as seeds for solar energetic particle production. Astrophys. J. 2013, 770, 73. [Google Scholar] [CrossRef]
  22. Chotoo, K.; Schwadron, N.A.; Mason, G.M.; Zurbuchen, T.H.; Gloeckler, G.; Posner, A.; Fisk, L.A.; Galvin, A.B.; Hamilton, D.C.; Collier, M.R. The suprathermal seed population for corotaing interaction region ions at 1AU deduced from composition and spectra of H+, He++, and He+ observed by Wind. J. Geophys. Res. 2000, 105, 23107–23122. [Google Scholar] [CrossRef]
  23. Mann, G.; Warmuth, A.; Aurass, H. Generation of highly energetic electrons at reconnection outflow shocks during solar flares. Astron. Astrophys. 2009, 494, 669–675. [Google Scholar] [CrossRef]
  24. Livadiotis, G.; McComas, D.J. Evidence of large scale phase space quantization in plasmas. Entropy 2013, 15, 1118–1132. [Google Scholar] [CrossRef]
  25. Bian, N.; Emslie, G.A.; Stackhouse, D.J.; Kontar, E.P. The formation of a kappa-distribution accelerated electron populations in solar flares. Astrophys. J. 2014, 796, 142. [Google Scholar] [CrossRef]
  26. Jeffrey, N.L.S.; Fletcher, L.; Labrosse, N. First evidence of non-Gaussian solar flare EUV spectral line profiles and accelerated non-thermal ion motion. Astron. Astrophys. 2016, 590, 99. [Google Scholar] [CrossRef]
  27. Formisano, V.; Moreno, G.; Palmiotto, F.; Hedgecock, P.C. Solar Wind Interaction with the Earth’s Magnetic Field 1. Magnetosheath. J. Geophys. Res. 1973, 78, 3714–3730. [Google Scholar] [CrossRef]
  28. Ogasawara, K.; Angelopoulos, V.; Dayeh, M.A.; Fuselier, S.A.; Livadiotis, G.; McComas, D.J.; McFadden, J.P. Characterizing the dayside magnetosheath using ENAs: IBEX and THEMIS observations. J. Geophys. Res. 2013, 118, 3126–3137. [Google Scholar] [CrossRef]
  29. Ogasawara, K.; Dayeh, M.A.; Funsten, H.O.; Fuselier, S.A.; Livadiotis, G.; McComas, D.J. Interplanetary magnetic field dependence of the suprathermal energetic neutral atoms originated in subsolar magnetopause. J. Geophys. Res. 2015, 120, 964–972. [Google Scholar] [CrossRef]
  30. Grabbe, C. Generation of broadband electrostatic waves in Earth’s magnetotail. Phys. Rev. Lett. 2000, 84, 3614. [Google Scholar] [CrossRef] [PubMed]
  31. Pisarenko, N.F.; Budnik, E.Y.; Ermolaev, Y.I.; Kirpichev, I.P.; Lutsenko, V.N.; Morozova, E.I.; Antonova, E.E. The ion differential spectra in outer boundary of the ring current: November 17, 1995 case study. J. Atm. Sol.-Terr. Phys. 2002, 64, 573–583. [Google Scholar] [CrossRef]
  32. Christon, S.P. A comparison of the Mercury and earth magnetospheres: electron measurements and substorm time scales. Icarus 1987, 71, 448–471. [Google Scholar] [CrossRef]
  33. Wang, C.-P.; Lyons, L.R.; Chen, M.W.; Wolf, R.A.; Toffoletto, F.R. Modeling the inner plasma sheet protons and magnetic field under enhanced convection. J. Geophys. Res. 2003, 108, 1074. [Google Scholar] [CrossRef]
  34. Kletzing, C.A.; Scudder, J.D.; Dors, E.E.; Curto, C. Auroral source region: Plasma properties of the high latitude plasma sheet. J. Geophys. Res. 2003, 108, 1360. [Google Scholar] [CrossRef]
  35. Ogasawara, K.; Livadiotis, G.; Grubbs, G.A.; Jahn, J.-M.; Michell, R.; Samara, M.; Sharber, J.R.; Winningham, J.D. Properties of suprathermal electrons associated with discrete auroral arcs. Geophys. Res. Lett. 2017, in press. [Google Scholar]
  36. Hapgood, M.; Perry, C.; Davies, J.; Denton, M. The role of suprathermal particle measurements in CrossScale studies of collisionless plasma processes. Planet. Space Sci. 2011, 59, 618–629. [Google Scholar] [CrossRef]
  37. Collier, M.R.; Hamilton, D.C. The relationship between kappa and temperature in the energetic ion spectra at Jupiter. Geophys. Res. Lett. 1995, 22, 303–306. [Google Scholar] [CrossRef]
  38. Mauk, B.H.; Mitchell, D.G.; McEntire, R.W.; Paranicas, C.P.; Roelof, E.C.; Williams, D.J.; Krimigis, S.M.; Lagg, A. Energetic ion characteristics and neutral gas interactions in Jupiter’s magnetosphere. J. Geophys. Res. 2004. [Google Scholar] [CrossRef]
  39. Dialynas, K.; Krimigis, S.M.; Mitchell, D.G.; Hamilton, D.C.; Krupp, N.; Brandt, P.C. Energetic ion spectral characteristics in the Saturnian magnetosphere using Cassini/MIMI measurements. J. Geophys. Res. 2009. [Google Scholar] [CrossRef]
  40. Livi, R.; Goldstein, J.; Burch, J.L.; Crary, F.; Rymer, A.M.; Mitchell, D.G.; Persoon, A.M. Multi-instrument analysis of plasma parameters in Saturn’s equatorial, inner magnetosphere using corrections for spacecraft potential and penetrating background radiation. J. Geophys. Res. 2014, 119, 3683. [Google Scholar] [CrossRef]
  41. Carbary, J.F.; Kane, M.; Mauk, B.H.; Krimigis, S.M. Using the kappa function to investigate hot plasma in the magnetospheres of the giant planets. J. Geophys. Res. 2014, 119, 8426–8447. [Google Scholar] [CrossRef]
  42. Mauk, B.H.; Krimigis, S.M.; Keath, E.P.; Cheng, A.F.; Armstrong, T.P.; Lanzerotti, L.G.; Gloeckler, G.; Hamilton, D.C. The hot plasma and radiation environment of the Uranian magnetosphere. J. Geophys. Res. 1987, 92, 15283–15308. [Google Scholar] [CrossRef]
  43. Krimigis, S.M.; Armstrong, T.P.; Axford, W.I.; Bostrom, C.O.; Cheng, A.F.; Gloeckler, G.; Hamilton, D.C.; Keath, E.P.; Lanzerotri, L.J.; Mauk, B.H.; et al. Hot plasma and energetic particles in Neptune’s magnetosphere. Science 1989, 246, 1483. [Google Scholar] [CrossRef] [PubMed]
  44. Moncuquet, M.; Bagenal, F.; Meyer-Vernet, N. Latitudinal structure of the outer Io plasma torus. J. Geophys. Res. 2002, 108, 1260. [Google Scholar] [CrossRef]
  45. Jurac, S.; McGrath, M.A.; Johnson, R.E.; Richardson, J.D.; Vasyliunas, V.M.; Eviatar, A. Saturn: Search for a missing water source. Geophys. Res. Lett. 2002, 29, 2172. [Google Scholar]
  46. Broiles, T.W.; Livadiotis, G.; Burch, J.L.; Chae, K.; Clark, G.; Cravens, T.E.; Davidson, R.; Eriksson, A.; Frahm, R.A.; Fuselier, S.A.; et al. Characterizing cometary electrons with kappa distributions. J. Geophys. Res. 2016, 121, 7407–7422. [Google Scholar] [CrossRef]
  47. Broiles, T.W.; Burch, J.L.; Chae, K.; Cravens, T.E.; Eriksson, A.; Fuselier, S.F.; Frahm, R.F.; Gasc, S.; Goldstein, R.; Henri, P.; et al. Statistical analysis of suprathermal electron drivers at 67P/Churyumov-Gerasimenko. Mon. Not. R. Astron. Soc. 2016, 462, 312–322. [Google Scholar] [CrossRef]
  48. Decker, R.B.; Krimigis, S.M. Voyager observations of low-energy ions during solar cycle 23. Adv. Space Res. 2003, 32, 597–602. [Google Scholar] [CrossRef]
  49. Decker, R.B.; Krimigis, S.M.; Roelof, E.C.; Hill, M.E.; Armstrong, T.P.; Gloeckler, G.; Hamilton, D.C.; Lanzerotti, L.J. Voyager 1 in the foreshock, termination shock, and heliosheath. Science 2005, 309, 2020–2024. [Google Scholar] [CrossRef] [PubMed]
  50. Heerikhuisen, J.; Pogorelov, N.V.; Florinski, V.; Zank, G.P.; le Roux, J.A. The effects of a 𝜅-distribution in the heliosheath on the global heliosphere and ENA flux at 1 AU. Astrophys. J. 2008, 682, 679–689. [Google Scholar] [CrossRef]
  51. Heerikhuisen, J.; Zirnstein, E.; Pogorelov, N. 𝜅-distributed protons in the solar wind and their charge-exchange coupling to energetic hydrogen. J. Geophys. Res. 2015, 120, 1516–1525. [Google Scholar] [CrossRef]
  52. Zank, G.P.; Heerikhuisen, J.; Pogorelov, N.V.; Burrows, R.; McComas, D.J. Microstructure of the heliospheric termination shock: implications for energetic neutral atom observations. Astrophys. J. 2010, 708, 1092. [Google Scholar] [CrossRef]
  53. Livadiotis, G. Lagrangian temperature: Derivation and physical meaning for systems described by kappa distributions. Entropy 2014, 16, 4290–4308. [Google Scholar] [CrossRef]
  54. Livadiotis, G. Curie law for systems described by kappa distributions. Europhys. Lett. 2016, 113, 10003. [Google Scholar] [CrossRef]
  55. Livadiotis, G.; McComas, D.J. Invariant kappa distribution in space plasmas out of equilibrium. Astrophys. J. 2011, 741, 88. [Google Scholar] [CrossRef]
  56. Livadiotis, G.; McComas, D.J. Non-equilibrium thermodynamic processes: Space plasmas and the inner heliosheath. Astrophys. J. 2012, 749, 11. [Google Scholar] [CrossRef]
  57. Livadiotis, G.; McComas, D.J.; Dayeh, M.A.; Funsten, H.O.; Schwadron, N.A. First sky map of the inner heliosheath temperature using IBEX spectra. Astrophys. J. 2011, 734, 1. [Google Scholar] [CrossRef]
  58. Livadiotis, G.; McComas, D.J.; Randol, M.B.; Funsten, H.O.; Möbius, E.S.; Schwadron, N.A.; Dayeh, M.A.; Zank, G.P.; Frisch, P.C. Pick-up ion distributions and their influence on ENA spectral curvature. Astrophys. J. 2012, 751, 64. [Google Scholar] [CrossRef]
  59. Livadiotis, G.; McComas, D.J.; Schwadron, N.A.; Funsten, H.O.; Fuselier, S.A. Pressure of the proton plasma in the inner heliosheath. Astrophys. J. 2013, 762, 134. [Google Scholar] [CrossRef]
  60. Fuselier, S.A.; Allegrini, F.; Bzowski, M.; Dayeh, M.A.; Desai, M.; Funsten, H.O.; Galli, A.; Heirtzler, D.; Janzen, P.; Kubiak, M.A.; et al. Low energy neutral atoms from the heliosheath. Astrophys. J. 2014, 784, 89. [Google Scholar] [CrossRef]
  61. Zirnstein, E.J.; McComas, D.J. Using kappa functions to characterize outer heliosphere proton distributions in the presence of charge-exchange. Astrophys. J. 2015, 815, 31. [Google Scholar] [CrossRef]
  62. Zank, G.P. Faltering steps into the galaxy: The boundary regions of the heliosphere. Ann. Rev. Astron. Astrophys. 2015, 53, 449–500. [Google Scholar] [CrossRef]
  63. Nicholls, D.C.; Dopita, M.A.; Sutherland, R.S. Resolving the electron temperature discrepancies in h ii regions and planetary nebulae: Κ-distributed electrons. Astrophys. J. 2012, 752, 148. [Google Scholar] [CrossRef]
  64. Nicholls, D.C.; Dopita, M.A.; Sutherland, R.S.; Kewley, L.J.; Palay, E. Measuring nebular temperatures: the effect of new collision strengths with equilibrium and κ-distributed electron energies. Astrophys. J. Supp. Ser. 2013, 207, 21. [Google Scholar] [CrossRef]
  65. Zhang, Y.; Liu, X.-W.; Zhang, B. HI free-bound emission of planetary nebulae with large abundance discrepancies: Two-component models versus κ-distributed electrons. Astrophys. J. 2014, 780, 93. [Google Scholar] [CrossRef]
  66. Raymond, J.C.; Winkler, P.F.; Blair, W.P.; Lee, J.-J.; Park, S. Non-Maxwellian Hα profiles in Tycho’s supernova remnant. Astrophys. J. 2010, 712, 901. [Google Scholar] [CrossRef]
  67. Milovanov, A.V.; Zelenyi, L.M. Functional background of the Tsallis entropy: “Coarse-grained” systems and “kappa” distribution functions. Nonlinear Process. Geophys. 2000, 7, 211–221. [Google Scholar] [CrossRef]
  68. Saito, S.; Forme, F.R.E.; Buchert, S.C.; Nozawa, S.; Fujii, R. Effects of a kappa distribution function of electrons on incoherent scatter spectra. Ann. Geophys. 2000, 18, 1216–1223. [Google Scholar] [CrossRef]
  69. Du, J. The nonextensive parameter and Tsallis distribution for self-gravitating systems. Europhys. Lett. 2004, 67, 893–899. [Google Scholar]
  70. Yoon, P.H.; Rhee, T.; Ryu, C.M. Self-consistent formation of electron κ distribution: 1. Theory. J. Geophys. Res. 2006, 111, 09106. [Google Scholar] [CrossRef]
  71. Raadu, M.A.; Shafiq, M. Test charge response for a dusty plasma with both grain size distribution and dynamical charging. Phys. Plasmas 2007, 14, 012105. [Google Scholar] [CrossRef]
  72. Livadiotis, G. Approach on Tsallis statistical interpretation of hydrogen-atom by adopting the generalized radial distribution function. J. Math. Chem. 2009, 45, 930–939. [Google Scholar] [CrossRef]
  73. Livadiotis, G. Kappa distribution in the presence of a potential energy. J. Geophys. Res. 2015, 120, 880–903. [Google Scholar] [CrossRef]
  74. Livadiotis, G. Kappa and q indices: Dependence on the degrees of freedom. Entropy 2015, 17, 2062–2081. [Google Scholar] [CrossRef]
  75. Tribeche, M.; Mayout, S.; Amour, R. Effect of ion suprathermality on arbitrary amplitude dust acoustic waves in a charge varying dusty plasma. Phys. Plasmas 2009, 16, 043706. [Google Scholar] [CrossRef]
  76. Livadiotis, G.; McComas, D.J. Measure of the departure of the q-metastable stationary states from equilibrium. Phys. Scr. 2010, 82, 035003. [Google Scholar] [CrossRef]
  77. Livadiotis, G.; McComas, D.J. Electrostatic shielding in plasmas and the physical meaning of the Debye length. J. Plasma Phys. 2014, 80, 341–378. [Google Scholar] [CrossRef]
  78. Baluku, T.K.; Hellberg, M.A.; Kourakis, I.; Saini, N.S. Dust ion acoustic solitons in a plasma with kappa-distributed electrons. Phys. Plasmas 2010, 17, 053702. [Google Scholar] [CrossRef]
  79. Le Roux, J.A.; Webb, G.M.; Shalchi, A.; Zank, G.P. A generalized nonlinear guiding center theory for the collisionless anomalous perpendicular diffusion of cosmic rays. Astrophys. J. 2010, 716, 671–692. [Google Scholar] [CrossRef]
  80. Eslami, P.; Mottaghizadeh, M.; Pakzad, H.R. Nonplanar dust acoustic solitary waves in dusty plasmas with ions and electrons following a q-nonextensive distribution. Phys. Plasmas 2011, 18, 102303. [Google Scholar] [CrossRef]
  81. Kourakis, I.; Sultana, S.; Hellberg, M.A. Dynamical characteristics of solitary waves, shocks and envelope modes in kappa-distributed non-thermal plasmas: An overview. Plasma Phys. Control. Fusion 2012, 54, 124001. [Google Scholar] [CrossRef]
  82. Randol, B.M.; Christian, E.R. Coupling of charged particles via Coulombic interactions: Numerical simulations and resultant kappa-like velocity space distribution functions. J. Geophys. Res. 2016, 121, 1907–1919. [Google Scholar] [CrossRef]
  83. Varotsos, P.A.; Sarlis, N.V.; Skordas, E.S. Study of the temporal correlations in the magnitude time series before major earthquakes in Japan. J. Geophys. Res. 2014, 119, 9192–9206. [Google Scholar] [CrossRef]
  84. Nicolaou, G.; Livadiotis, G.; Moussas, X. Long term variability of the polytropic Index of solar wind protons at 1AU. Sol. Phys. 2014, 289, 1371–1378. [Google Scholar] [CrossRef]
  85. Fisk, L.A.; Gloeckler, G. The case for a common spectrum of particles accelerated in the heliosphere: Observations and theory. J. Geophys. Res. 2014, 119, 8733–8749. [Google Scholar] [CrossRef]
  86. Viñas, A.F.; Moya, P.S.; Navarro, R.E.; Valdivia, J.A.; Araneda, J.A.; Muñoz, V. Electromagnetic fluctuations of the whistler-cyclotron and firehose instabilities in a Maxwellian and Tsallis-kappa-like plasma. J. Geophys. Res. 2015, 120, 3307–3317. [Google Scholar] [CrossRef]
  87. Livadiotis, G. Shock strength in space and astrophysical plasmas. Astrophys. J. 2015, 809, 111. [Google Scholar] [CrossRef]
  88. Dos Santos, M.S.; Ziebell, L.F.; Gaelzer, R. Ion firehose instability in a dusty plasma considering product-bi-kappa distributions for the plasma particles. Phys. Plasmas 2016, 23, 013705. [Google Scholar] [CrossRef]
  89. Nicolaou, G.; Livadiotis, G. Misestimation of temperature when applying Maxwellian distributions to space plasmas described by kappa distributions. Astrophys. Space Sci. 2016. [Google Scholar] [CrossRef]
  90. Sarri, G.; Dieckmann, M.E.; Brown, C.R.D.; Cecchetti, C.A.; Hoarty, D.J. Observation and characterization of laser-driven phase space electron holes. Phys. Plasmas 2010. [Google Scholar] [CrossRef]
  91. Hammond, C.M.; Feldman, W.C.; Phillips, J.L.; Goldstein, B.E.; Balogh, A. Solar wind double ion beams and the heliospheric current sheet. J. Geophys. Res. 1996, 100, 7881–7889. [Google Scholar] [CrossRef]
  92. Treumann, R.A. Theory of superdiffusion for the magnetopause. Geophys. Res. Lett. 1997, 24, 1727–1730. [Google Scholar] [CrossRef]
  93. Leubner, M.P. A nonextensive entropy approach to kappa distributions. Astrophys. Space Sci. 2002, 282, 573–579. [Google Scholar] [CrossRef]
  94. Livadiotis, G.; McComas, D.J. Beyond kappa distributions: Exploiting Tsallis statistical mechanics in space plasmas. J. Geophys. Res. 2009. [Google Scholar] [CrossRef]
  95. Livadiotis, G. Statistical background and properties of kappa distributions in space plasmas. J. Geophys. Res. 2015, 120, 1607–1619. [Google Scholar] [CrossRef]
  96. Livadiotis, G. Kappa Distribution: Theory and Applications in Plasmas, 1st ed.; Elsevier: New York, NY, USA, 2017. [Google Scholar]
  97. Silva, R.; Plastino, A.R.; Lima, J.A.S. A Maxwellian path to the q-nonextensive velocity distribution function. Phys. Lett. A 1998, 249, 401–408. [Google Scholar] [CrossRef]
  98. Yamano, T. Some properties of q-logarithmic and q-exponential functions in Tsallis statistics. Phys. A Stat. Mech. Appl. 2002, 305, 486–496. [Google Scholar] [CrossRef]
  99. Beck, C.; Schlögl, F. Thermodynamics of Chaotic Systems; Cambridge University Press: Cambridge, UK, 1993. [Google Scholar]
  100. Tsallis, C. Possible generalization of Boltzmann-Gibbs statistics. J. Stat. Phys. 1988, 52, 479–487. [Google Scholar] [CrossRef]
  101. Abe, S. Correlation induced by Tsallis’ nonextensivity. Phys. A Stat. Mech. Appl. 1999, 269, 403–409. [Google Scholar] [CrossRef]
  102. Abe, S. Heat and entropy in nonextensive thermodynamics: Transmutation from Tsallis theory to Renyi-entropy-based theory. Phys. A Stat. Mech. Appl. 2001, 300, 417–423. [Google Scholar] [CrossRef]
  103. Rama, S.K. Tsallis statistics: Averages and a physical interpretation of the Lagrange multiplier β. Phys. Lett. A 2000, 276, 103–108. [Google Scholar] [CrossRef]
  104. Tsallis, C.; Mendes, R.S.; Plastino, A.R. The role of constraints within generalized nonextensive statistics. Phys. A Stat. Mech. Appl. 1998, 261, 534–554. [Google Scholar] [CrossRef]
  105. Gell-Mann, M.; Tsallis, C. Nonextensive Entropy: Interdisciplinary Applications; Oxford University Press: Oxford, UK, 2004. [Google Scholar]
  106. Tsallis, C. Introduction to Nonextensive Statistical Mechanics; Springer: New York, NY, USA, 2009. [Google Scholar]
  107. Gibbs, J.W. Elementary Principles in Statistical Mechanics; Charles Scribner’s Sons: New York, NY, USA, 1902. [Google Scholar]
  108. Livadiotis, G. Non-Euclidean-normed statistical mechanics. Phys. A Stat. Mech. Appl. 2016, 445, 240–255. [Google Scholar] [CrossRef]
  109. Livadiotis, G. Modeling anisotropic Maxwell–Jüttner distributions: Derivation and properties. Ann. Geophys. 2016, 34, 1–14. [Google Scholar] [CrossRef]
Figure 1. The entropic formulations S q (red) and S q 1 (blue) are plotted as a function of the (invariant) kappa index, for number of particles NC and d=3 kinetic degrees of freedom per particle. We set the number of correlated particles to NC = 10 (left) and NC = 100 (right), and the modified speed scale parameter to σ ˜ = 0.9 (up) and σ ˜ = 0.5 (down) (the Boltzmann constant k B is set to 1.) For large kappa indices, both the entropic formulations tend to be equal to each other, recovering the Boltzmann Gibbs (BG) entropy.
Figure 1. The entropic formulations S q (red) and S q 1 (blue) are plotted as a function of the (invariant) kappa index, for number of particles NC and d=3 kinetic degrees of freedom per particle. We set the number of correlated particles to NC = 10 (left) and NC = 100 (right), and the modified speed scale parameter to σ ˜ = 0.9 (up) and σ ˜ = 0.5 (down) (the Boltzmann constant k B is set to 1.) For large kappa indices, both the entropic formulations tend to be equal to each other, recovering the Boltzmann Gibbs (BG) entropy.
Entropy 19 00285 g001
Figure 2. (a) Dyadic distribution and unique entropic formalisms; (b) Unique distribution and dyadic entropic formalisms.
Figure 2. (a) Dyadic distribution and unique entropic formalisms; (b) Unique distribution and dyadic entropic formalisms.
Entropy 19 00285 g002

Share and Cite

MDPI and ACS Style

Livadiotis, G. On the Simplification of Statistical Mechanics for Space Plasmas. Entropy 2017, 19, 285. https://doi.org/10.3390/e19060285

AMA Style

Livadiotis G. On the Simplification of Statistical Mechanics for Space Plasmas. Entropy. 2017; 19(6):285. https://doi.org/10.3390/e19060285

Chicago/Turabian Style

Livadiotis, George. 2017. "On the Simplification of Statistical Mechanics for Space Plasmas" Entropy 19, no. 6: 285. https://doi.org/10.3390/e19060285

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop