This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Distribution Function for ngn\geq g Quantum Particles

Shimul Akhanjee [email protected]
Abstract

A new quantum mechanical distribution function nI(ε)n^{I}(\varepsilon), is derived for the condition ngn\geq g, where in contrast to the exclusion principle ngn\leq g for fermions, each energy state must be populated by at least one particle. Although the particles share many features with bosons, the anomalous behavior of nI(ε)n^{I}(\varepsilon) precludes Bose-Einstein condensation (BEC) due to the required occupancy of the excited states, which creates a permanently pressurized background at T=0T=0, similar to the degeneracy pressure of fermions. An exhaustive classification scheme is presented for both distinguishable and indistinguishable, particles and energy levels based on Richard Stanley’s twelvefold way in combinatorics.

I Introduction

The statistical distribution function, n(ε)n(\varepsilon) for identical particles has been an essential component of quantum mechanics. Historically, the behavior of n(ε)n(\varepsilon) has been over-determined by key experimental facts in a wide variety of physical systems such as the blackbody spectrum, semiconductor heterostructures, astrophysical spectroscopic measurements, low temperature TT, and condensed matter systemsKittel (2005); Pradhan and Nahar (2011); Leggett (2001, 2006). Theoretical approaches converge, from the grand-canonical ensemble to the micro-canonical ensemble, as well as the more mathematically rigorous Darwin-Fowler method of mean valuesBose (1924); Einstein (1924); Fermi (1926); Dirac (1926); Arnaud et al. (1999); Darwin and Fowler (1922).

Consider a system within the microcanonical ensemble having a fixed number of particles N=jnjN=\sum_{j}{n_{j}}, total energy U=jεjnjU=\sum_{j}{\varepsilon_{j}n_{j}}, and volume VVSchwabl (2002). One can make use of the mathematical structures found in combinatorial counting problems, particularly the number of ways that one can distribute a specified number of balls into a fixed number of boxes as shown in Table 1, known as Richard Stanley’s twelvefold wayStanley (2012). For quantum systems in particular there are only three possible arrangements of the identical balls, of which represent indistinguishable particles, into the labeled boxes that play the role of distinguishable energy states. I will introduce the new case of the second row, third column of Table 1.

Table 1: The Twelvefold Way - How many ways can 𝐧\bf{n} balls be sorted into 𝐠\bf{g} boxes?
{ng}{n\brace g} - Stirling numbers of the 2nd kind
pg(n)p_{\leq g}(n) - integer partitions of nn into at most gg parts
pg(n)p_{g}(n) - integer partitions of nn into exactly gg parts
Ball and Box Set Arbitrary (Any Sorting) Injective (Maximum 1 ball per box) Surjective (Minimum 1 ball per box)
Distinct Balls
Distinct Boxes gng^{n} g!(gn)!\frac{g!}{(g-n)!} g!{ng}g!{n\brace g}
Identical Balls
Distinct Boxes (g+n1n){g+n-1}\choose{n} (gn){g}\choose{n} (n1g1){n-1}\choose{g-1}
Distinct Balls
Identical Boxes j=0g{nj}{\sum}_{j=0}^{g}{n\brace j} 1 if ngn\leq g {ng}n\brace g
Identical Balls
Identical Boxes pg(n)p_{\leq g}(n) 1 if ngn\leq g pg(n)p_{g}(n)

II Unrestricted sorting of nn and gg

Starting with second row, first column of Table 1, the microstate configuration of bosons can be constructed from the distinct orderings of gj1g_{j}-1 lines and njn_{j} circles as shown in Fig.1(a)

tjB=(gj+nj1nj)=(gj+nj1)!nj!(gj1)!t^{B}_{j}={{g_{j}+n_{j}-1}\choose{n_{j}}}=\frac{(g_{j}+n_{j}-1)!}{n_{j}!(g_{j}-1)!} (1)

where the standard manipulations lead to the Bose-Einstein distribution,

njB(ε)=gje(εjμ)/(kBT)1n^{B}_{j}(\varepsilon)=\frac{g_{j}}{\mathrm{e}^{(\varepsilon_{j}-\mu)/(k_{B}T)}-1} (2)

III njgjn_{j}\leq g_{j} - The exclusion principle

. Next, by examining the second row, second column of Table 1 the resulting distribution represents fermions. This particular occupancy of the energy levels is depicted in Fig.1(b), where this scenario implies that only one particle can occupy a sub-state of gg,

tjF=(gjnj)=gj!nj!(gjnj)!t^{F}_{j}={{g_{j}}\choose{n_{j}}}=\frac{g_{j}!}{n_{j}!(g_{j}-n_{j})!} (3)

This yields the Fermi-Dirac distribution,

njF(ε)=gje(εjμ)/(kBT)+1n^{F}_{j}(\varepsilon)=\frac{g_{j}}{\mathrm{e}^{(\varepsilon_{j}-\mu)/(k_{B}T)}+1} (4)

Refer to caption

Figure 1: Typical configurations of the three combinatorially distinct possibilities for identical particles distributed into distinguishable states. (a) Unrestricted sorting, allowing for more than one particle in a state, in addition to empty states. (b) The exclusion principle, njgjn_{j}\leq g_{j}, with no more than one particle per state, and allowing empty states. (c) The new case introduced here: njgjn_{j}\geq g_{j}, where all sub-states must be occupied by at least one particle, while no upper bound is imposed.

IV njgjn_{j}\geq g_{j} - The inclusion constraint

.

The two preceding scenarios are not exhaustive. The primary purpose of this paper is to demonstrate the combinatorially distinct possibility of the second row, third column of Table 1. The inclusion principle introduced here, is a new case of quantum statistics. A single particle is attached to every positive energy level, requiring njgjn_{j}\geq g_{j}, such that no positive energy level is vacant as shown in Fig.1 (c). Although, a similar occupation of the excited states might be possible in the classical limit kBTεjk_{B}T\gg\varepsilon_{j}, where the phase-space density (number of particles per quantum state) is very high, here it is not assumed that this condition is generated from external factors. Rather, the level occupancy is presupposed as an intrinsic property of the particles. Therefore, the microstate configuration of interest can be adapted from the surjective case for identical balls in distinct boxesStanley (2012),

tjI=(nj1gj1)=(nj1)!(gj1)!(njgj)!t^{I}_{j}={{n_{j}-1}\choose{g_{j}-1}}=\frac{(n_{j}-1)!}{(g_{j}-1)!(n_{j}-g_{j})!} (5)

For large values of nn and gg, the total number of microstates becomes a product, tTjnj!gj!(njgj)!t_{T}\approx\prod_{j}{\frac{n_{j}!}{g_{j}!(n_{j}-g_{j})!}} and after the use of Stirling’s approximation: lnN!NlnNN\ln N!\approx N\ln N-N, the entropy S=kBlntTS=k_{B}\ln{t_{T}} can be expressed as.

S=kBjnjlnnjgjlngj(njgj)ln(njgj)S=k_{B}\sum_{j}{}n_{j}\ln{n_{j}}-g_{j}\ln{g_{j}}-(n_{j}-g_{j})\ln{(n_{j}-g_{j})} (6)

Next, we develop the condition for an entropy maximum. Derivatives are taken with respect to njn_{j}. The macrostate conditions dN=jdnj=0\mathrm{d}N=\sum_{j}{\mathrm{d}n_{j}}=0 and dU=jεjdnj=0\mathrm{d}U=\sum_{j}{\varepsilon_{j}\mathrm{d}n_{j}}=0 are enforced with Lagrange multipliers α\alpha and β\beta,

dSdnj=jln(njnjgj)αβεj=0\frac{\mathrm{d}S}{\mathrm{d}n_{j}}=\sum_{j}\ln\left(\frac{n_{j}}{n_{j}-g_{j}}\right)-\alpha-\beta\varepsilon_{j}=0 (7)

Evidently, a dimensional analysis of the thermodynamic potential dU=1βdSαβdN\mathrm{d}U=\frac{1}{\beta}\mathrm{d}S-\frac{\alpha}{\beta}\mathrm{d}N, reveals the correspondence with kBT=1/βk_{B}T=1/\beta and the chemical potential μ=α/β\mu=-\alpha/\beta. After solving for njn_{j}, the final expression becomes,

njI(εj)=gjeβ(εjμ)eβ(εjμ)1\boxed{n^{I}_{j}(\varepsilon_{j})=\frac{g_{j}\mathrm{e}^{\beta(\varepsilon_{j}-\mu)}}{\mathrm{e}^{\beta(\varepsilon_{j}-\mu)}-1}} (8)

Consider a three dimensional, non-interacting gas of these particles with energy εp=p22m\varepsilon_{p}=\frac{p^{2}}{2m}. Apparently, the fixed background of excited states will have important thermodynamic consequences. It may be tempting to assign a spin ss to such quantum particles, where in the absence of a magnetic field, the degeneracy factor is g=2s+1g=2s+1. However, no assumptions about the permutation symmetry under the exchange of two particles should be made without a more rigorous development of the Fock space. Naively, in the number occupancy basis the eigenstates are given by,

|np1,np2,npN(|ψp1)np1(|ψp2)np2,(|ψpN)npN\left|n_{p_{1}},n_{p_{2}},\cdots n_{p_{N}}\right>\propto\left(\left|\psi_{p_{1}}\right>\right)^{n_{p_{1}}}\left(\left|\psi_{p_{2}}\right>\right)^{n_{p_{2}}},\cdots\left(\left|\psi_{p_{N}}\right>\right)^{n_{p_{N}}} (9)

In order to enforce the inclusion principle mathematically, the state |Γ\left|\Gamma\right\rangle defined below,

|Γ=|1,1,1,\left|\Gamma\right\rangle=\left|1,1,1,\cdots\right> (10)

must vanish after applying annihilation operator ap|Γ=0a_{p}\left|\Gamma\right\rangle=0, for all values of pp. This condition will definitely have consequences for the structure of the wavefunctions. On the other hand, for the usual bosonic case, a “simple” BEC transition occurs where the p=0\vec{p}=0 state is macroscopically occupied at T=0T=0Einstein (1924); Leggett (2006),

|BEC=|N,0,0,(ap=0)N1p0ap|Γ=0\left|BEC\right>=\left|N,0,0,\cdots\right>\propto\left(a_{p=0}^{{\dagger}}\right)^{N-1}\prod_{p\neq 0}a_{p}\left|\Gamma\right>=0 (11)

Thus, the |BEC\left|BEC\right> state is forbidden because of the condition ap|Γ=0a_{p}\left|\Gamma\right\rangle=0.

Alternatively, one can derive Eq.(8) exactly, with no approximations by applying the grand canonical ensemble(Kardar, 2007). The conventional approach takes on summations over each occupation number npn_{p}, with allowed values: [0,1][0,1] for fermions and [0,1,2,][0,1,2,\dots] for bosons. For the new case considered here: [1,2,][1,2,\dots] or np0n_{p}\neq 0. The grand partition function for that system becomes,

𝒵G=𝑝np0eβ(εpμ)np=𝑝eβ(εpμ)1eβ(εpμ)\mathcal{Z}_{G}=\underset{p}{\prod}\underset{n_{p}\neq 0}{\sum}e^{-\beta(\varepsilon_{p}-\mu)n_{p}}=\underset{p}{\prod}\frac{e^{-\beta(\varepsilon_{p}-\mu)}}{1-e^{-\beta(\varepsilon_{p}-\mu)}} (12)

where the geometric series converges only if eβ(εμ)<1e^{-\beta\left(\varepsilon-\mu\right)}<1, which is true for μ<0\mu<0. Therefore, the average particle number for a single particle sub-state: 𝒵G=𝑝𝒵p\mathcal{Z}_{G}=\underset{p}{\prod}\mathcal{Z}_{p} becomes,

np=kBT1𝒵p(𝒵pμ)V,T=eβ(εpμ)eβ(εpμ)1\left\langle n_{p}\right\rangle=k_{B}T\frac{1}{\mathcal{Z}_{p}}\left(\frac{\partial\mathcal{Z}_{p}}{\partial\mu}\right)_{V,T}=\frac{\mathrm{e}^{\beta(\varepsilon_{p}-\mu)}}{\mathrm{e}^{\beta(\varepsilon_{p}-\mu)}-1} (13)

Hence, Eq.(8) that was derived earlier within the microcanonical ensemble is also confirmed by the grand canonical ensemble approach, given that for a single particle level, gnp=nI(εp)g\left\langle n_{p}\right\rangle=n^{I}(\varepsilon_{p}). As expected, the particle number variance σN2=kBT(npμ)V,T(kBT)2(εμ)2\sigma_{N}^{2}=k_{B}T\left(\frac{\partial\left\langle n_{p}\right\rangle}{\partial\mu}\right)_{V,T}\sim\frac{(k_{B}T)^{2}}{(\varepsilon-\mu)^{2}}, retains the bosonic form at kBTεjk_{B}T\gg\varepsilon_{j} since the fixed occupancy of excited states should not significantly contribute to σN\sigma_{N}.

Refer to caption

Figure 2: Comparison of the 3 quantum distributions. Unlike (nB)/g(n^{B})/g and (nF)/g(n^{F})/g, which decay to zero at (εμ)kBT(\varepsilon-\mu)\gg k_{B}T, (nB)/g(n^{B})/g saturates at unity.

V Thermodynamics

. It is useful to express Eq.(8) as,

njI(εj)=njB(εj)+gn_{j}^{I}(\varepsilon_{j})=n_{j}^{B}(\varepsilon_{j})+g (14)

An analysis of various thermodynamic quantities in terms of different components will help to elucidate the physical properties of the system. The p=0\vec{p}=0 contribution should be separated and treated carefully, especially as T0T\rightarrow 0, or where the fugacity z=eμβ1z=e^{\mu\beta}\rightarrow 1. In the limit of large VV, and constant specific volume v=V/Nv=V/N, the sums over discrete energy levels for the excited states can be replaced by integrals over gp0=gV(2π)3d3pg{\sum}_{\vec{p}\neq 0}{\cdots}=g\frac{V}{\left(2\pi\hbar\right)^{3}}\int d^{3}p. The average particle number becomes the sum of three terms,

N=pnI(εp)=N0(z)+N1(z)+N2N=\sum_{p}{n^{I}(\varepsilon_{p})}=N_{0}(z)+N_{1}(z)+N_{2} (15)

The first term describes the number of particles occupying the ground state energy,

N0(z)=nI(εp=0)=g1zN_{0}(z)=n^{I}(\varepsilon_{p=0})=\frac{g}{1-z} (16)

of which, diverges at z=1z=1. Moreover, N1(z)N_{1}(z) and N2N_{2} account for the excited states of the nBn^{B} term and occupied background respectively,

N1(z)=V(2π)3nB(εp)d3p=gNvλ3f3/2+(z)N_{1}(z)=\frac{V}{\left(2\pi\hbar\right)^{3}}\int{n^{B}\left(\varepsilon_{p}\right)d^{3}p}=\frac{gNv}{\lambda^{3}}f_{3/2}^{+}(z) (17)
N2=gVm3/22π230Ωε1/2𝑑ε=43πgNvλ3(ΩkBT)3/2N_{2}=\frac{gVm^{3/2}}{\sqrt{2}\pi^{2}\hbar^{3}}\int_{0}^{\Omega}\varepsilon^{1/2}d\varepsilon=\frac{4}{3\sqrt{\pi}}\frac{gNv}{\lambda^{3}}\left(\frac{\Omega}{k_{B}T}\right)^{3/2} (18)

where the thermal wavelength is defined by λ=(2π)/(mkBT)\lambda=\hbar\sqrt{(2\pi)/(mk_{B}T)} and the generalized ζ\zeta function is defined by fν+(z)=1(ν1)!0xν1z1ex1𝑑xf_{\nu}^{+}(z)=\frac{1}{\left(\nu-1\right)!}\int_{0}^{\infty}\frac{x^{\nu-1}}{z^{-1}e^{x}-1}dx{\frac{}{}}Zwillinger (1996). Upon inspection, it is clear that Eq.(18) would seem problematic since it diverges at the upper limit of integration. Consequently, a high energy cutoff Ω\Omega should ensure finite results.

In order to determine whether the system transitions into a “simple” BEC state, it is necessary to study the behavior of the condensate fraction,

ν0=limNN0(z)/(N(z))\nu_{0}=\underset{{N}\rightarrow\infty}{\mathrm{\lim}}N_{0}(z)/(N(z)) (19)

Since N1(z=1)N_{1}(z=1) has a limiting value in the conventional BEC transition, N1(z=1)T3/2N_{1}(z=1)\propto T^{3/2}, and therefore the number of excited states arising from N1(z=1)N_{1}(z=1) vanishes at low temperatures. However, since N2N_{2} is independent of both TT and zz, there cannot be a macroscopic occupation of the ground state unless Ω\Omega is sufficiently small. Thus, ν0\nu_{0} can never reach the value of 11 and a complete BEC transition is not possible. This should be apparent from the outset since a significant fraction of the excited states are permanently occupied and can never move into the ground state energy.

The pressure PP can be determined from the grand potential Φ\Phi, starting with Eq.(12):

P\displaystyle P =ΦV=1Vβln(𝒵G)=1Vβpln(eβ(εpμ)1eβ(εpμ))\displaystyle=-\frac{\Phi}{V}=\frac{1}{V\beta}\mathrm{\ln}\left(\mathcal{Z}_{G}\right)=\frac{1}{V\beta}{\sum_{p}\mathrm{\ln}\left(\frac{e^{-\beta(\varepsilon_{p}-\mu)}}{1-e^{-\beta(\varepsilon_{p}-\mu)}}\right)} (20)
=gkBTλ3f5/2+(z)+45πgλ3Ω5/2(kBT)3/2μN2V\displaystyle=\frac{gk_{B}T}{\lambda^{3}}f^{+}_{5/2}(z)+\frac{4}{5\sqrt{\pi}}\frac{g}{\lambda^{3}}\frac{\Omega^{5/2}}{(k_{B}T)^{3/2}}-\frac{\mu N_{2}}{V}

The common wisdom suggests that as the distribution function of a quantum gas flattens, which generally occurs as TT increases, then PP increases, reflecting a shift away from quantum effects and toward classical ideal gas behavior. Furthermore, higher excited state occupation results in a higher average kinetic energy of the gas, which translates directly to higher PP, since it is associated with particle collisions and their average momentum transfer. When approaching the T0T\to 0 limit of Eq.(20), PΩ5/2P\sim\Omega^{5/2}, which is similar to the degeneracy pressure of fermions PF=25εFNVP_{F}=\frac{2}{5}\varepsilon_{F}\frac{N}{V}, where the cutoff Ω\Omega is analogous to the Fermi level, εF\varepsilon_{F}.

VI Conclusion

After examining the second row of Table. 1, an exhaustive analysis of the possible distribution functions for identical particles populating distinguishable energy levels has been undertaken. However, the fourth row suggests the possibility of indistinguishable energy levels. The Gibbs paradox points out that from from a classical standpoint, the non-extensivity of the entropy arises due to the neglect of the factor 1/N!1/N! when over-counting configurations of the partition function for identical particlesSchwabl (2002). It would be an interesting endeavor to study the consequences of the microstates being constructed from different integer partitions of nn into gg parts. New paradoxical inconsistencies could arise from enforcing the mathematical conditions that account for identical energy states.

To conclude, a classification scheme for identical particles has been developed by applying important results from enumerative combinatorics, namely the twelvefold way of nn balls sorted into gg boxes. The distribution function, nI(ε)n^{I}(\varepsilon) for ngn\geq g quantum particles has, for the first time, been derived exactly from within the microcanonical and the grand canonical ensembles. At first glance, many of its features are similar to bosons however a “simple”, non-fragmented BEC state is prevented and the system exhibits a T=0T=0 pressure that is energy cutoff dependent. In other words, the system shares features of both fermions and bosons. Such particles could have tremendous implications in various high energy, astrophysical and cosmological theories, specifically dark matter candidates. Unlike ordinary matter, dark matter is “collisionless” under normal conditions, meaning dark matter particles rarely interact with each other or with regular matter in a way that would create traditional pressureBertone and Hooper (2018). Furthermore, the inclusion principle could possibly explain some anomalies in astrophysical observations. For example, in galactic halos, a non-fermionic T=0T=0 pressure present in dark matter could act as a stabilizing factor against gravitational collapse. In galaxy clusters, this helps the dark matter halo retain its shape, size, and density profileGarrett and Dūda (2011).

VII Conflicts of Interest

The author has no conflicts of interest to declare that are relevant to the contents of this article.

References

  • Kittel (2005) C. Kittel, Introduction to Solid State Physics (8th ed.) (Wiley, New York, 2005).
  • Pradhan and Nahar (2011) A. K. Pradhan and S. N. Nahar, Atomic Astrophysics and Spectroscopy (Cambridge University Press, New York, 2011).
  • Leggett (2001) A. J. Leggett, Rev. Mod. Phys. 73, 307 (2001).
  • Leggett (2006) A. J. Leggett, Quantum Liquids Bose Condensation and Pairing in Condensed Matter Systems (Oxford, New York, 2006).
  • Bose (1924) S. Bose, Z.Physik 26, 178 (1924).
  • Einstein (1924) A. Einstein, Sitzber. Kgl. Preuss. Akad. Wiss 1924, 261 (1924).
  • Fermi (1926) E. Fermi, Rend. Lincei 3 (1926).
  • Dirac (1926) P. Dirac, Rend. Lincei 112 (1926).
  • Arnaud et al. (1999) J. Arnaud, J. M. Boé, L. Chusseau,  and F. Philippe, American Journal of Physics 67, 215 (1999).
  • Darwin and Fowler (1922) C. Darwin and R. Fowler, Lond.Edinb.Dubl.Phil.Mag 44, 450 (1922).
  • Schwabl (2002) F. Schwabl, Statistical Mechanics (Springer, Berlin, 2002).
  • Stanley (2012) R. P. Stanley, Enumerative Combinatorics: Volume 1, 2nd Edition (Cambridge University Press, Cambridge, 2012).
  • Kardar (2007) M. Kardar, Statistical Physics of Particles (Cambridge University Press, Cambridge, 2007).
  • Zwillinger (1996) D. Zwillinger, CRC Standard Mathematical Tables and Formulae, 30th edition (CRC Press, Boca Raton, 1996).
  • Bertone and Hooper (2018) G. Bertone and D. Hooper, Rev. Mod. Phys. 90, 045002 (2018).
  • Garrett and Dūda (2011) K. Garrett and G. Dūda, Advances in Astronomy 2011, 968283 (2011).