Skip to main content
Log in

Derivation of Isothermal Quantum Fluid Equations with Fermi-Dirac and Bose-Einstein Statistics

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

By using the quantum maximum entropy principle we formally derive, from a underlying kinetic description, isothermal (hydrodynamic and diffusive) quantum fluid equations for particles with Fermi-Dirac and Bose-Einstein statistics. A semiclassical expansion of the quantum fluid equations, up to \(\mathcal{O}(\hbar^{2})\)-terms, leads to classical fluid equations with statistics-dependent quantum corrections, including a modified Bohm potential. The Maxwell-Boltzmann limit and the zero temperature limit are eventually discussed.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Similar content being viewed by others

Notes

  1. Reference [11] is a partial exception, because the fully-quantum model is deduced for a generic entropy. However, the semiclassical expansion assumes Boltzmann entropy.

  2. Note that we have not to rescale the Wigner functions w and \(w_{w_{\mathrm{eq}}}\) because we are already using dimensionless Wigner functions (see Remark 2.1).

  3. A rigorous proof of existence and uniqueness of the constrained minimization problem has been recently obtained by Méhats and Pinaud [26] for the one-dimensional case with periodic boundary conditions.

  4. The proof given in Ref. [11], however, is still more general because the density-operator formalism covers the case of a system confined in a domain Ω∈ℝd, while the Wigner formalism is valid only in the whole-space case.

  5. In the Bose-Einstein case, these models are only suited to describe the non-condensate phase; we shall discuss this point later on, in the semiclassical framework (see Proposition 3.2 and Remark 3.2).

  6. Actually, in Ref. [11] the factor 1/T seems to be missing.

References

  1. Arnold, A.: Self-consistent relaxation-time models in quantum mechanics. Commun. Partial Differ. Equ. 21(3–4), 473–506 (1996)

    Article  MATH  Google Scholar 

  2. Barletti, L., Frosali, G.: Diffusive limit of the two-band k⋅p model for semiconductors. J. Stat. Phys. 139(2), 280–306 (2010)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  3. Barletti, L., Méhats, F.: Quantum drift-diffusion modeling of spin transport in nanostructures. J. Math. Phys. 51(5), 053304 (2010), 20 pp.

    Article  MathSciNet  ADS  Google Scholar 

  4. Bohm, D.: A suggested interpretation of the quantum theory in terms of “hidden variables” I. Phys. Rev. 85, 166–179 (1952)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  5. Bohm, D.: A suggested interpretation of the quantum theory in terms of “hidden variables” II. Phys. Rev. 85, 180–193 (1952)

    Article  MathSciNet  ADS  Google Scholar 

  6. Bourgade, J.P., Degond, P., Méhats, F., Ringhofer, C.: On quantum extensions to classical spherical harmonics expansion/Fokker-Planck models. J. Math. Phys. 47(4), 043302 (2006), 26 pp.

    Article  MathSciNet  ADS  Google Scholar 

  7. Brull, S., Méhats, F.: Derivation of viscous correction terms for the isothermal quantum Euler model. Z. Angew. Math. Mech. 90(3), 219–230 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  8. Dalfovo, F., Giorgini, S., Pitaevskii, L.P., Stringari, S.: Theory of Bose-Einstein condensation in trapped gases. Rev. Mod. Phys. 71(3), 463–512 (1999)

    Article  ADS  Google Scholar 

  9. Degond, P., Ringhofer, C.: Quantum moment hydrodynamics and the entropy principle. J. Stat. Phys. 112(3–4), 587–628 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  10. Degond, P., Méhats, F., Ringhofer, C.: Quantum energy-transport and drift-diffusion models. J. Stat. Phys. 118(3–4), 625–667 (2005)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  11. Degond, P., Gallego, S., Méhats, F.: Isothermal quantum hydrodynamics: derivation, asymptotic analysis, and simulation. Multiscale Model. Simul. 6(1), 246–272 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  12. Dürr, D., Teufel, S.: Bohmian Mechanics. The Physics and Mathematics of Quantum Theory. Springer, Berlin (2009)

    MATH  Google Scholar 

  13. Ferry, D.K., Zhou, J.-R.: Form of the quantum potential for use in hydrodynamic equations for semiconductor device modeling. Phys. Rev. B 48(11), 7944–7950 (1993)

    Article  ADS  Google Scholar 

  14. Folland, G.B.: Harmonic Analysis in Phase Space. Princeton University Press, Princeton (1989)

    MATH  Google Scholar 

  15. Gallego, S., Méhats, F.: Numerical approximation of a quantum drift-diffusion model. C. R. Math. Acad. Sci. Paris 339(7), 519–524 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  16. Gallego, S., Méhats, F.: Entropic discretization of a quantum drift-diffusion model. SIAM J. Numer. Anal. 43(5), 1828–1849 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  17. Gasser, I., Markowich, P.A.: Quantum hydrodynamics, Wigner transforms and the classical limit. Asymptot. Anal. 14(2), 97–116 (1997)

    MathSciNet  MATH  Google Scholar 

  18. Jüngel, A.: Transport Equations for Semiconductors. Springer, Berlin (2009)

    Book  Google Scholar 

  19. Jüngel, A.: Dissipative quantum fluid models. Riv. Mat. Univ. Parma 3(2), 217–290 (2012)

    Google Scholar 

  20. Jüngel, A., Matthes, D.: A derivation of the isothermal quantum hydrodynamic equations using entropy minimization. Z. Angew. Math. Mech. 85(11), 806–814 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  21. Jüngel, A., Matthes, D., Milišić, J.P.: Derivation of new quantum hydrodynamic equations using entropy minimization. SIAM J. Appl. Math. 67(1), 46–68 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  22. Jüngel, A., Krause, S., Pietra, P.: Diffusive semiconductor moment equations using Fermi-Dirac statistics. Z. Angew. Math. Phys. 62(4), 623–639 (2011)

    Article  MathSciNet  Google Scholar 

  23. Levermore, C.D.: Moment closure hierarchies for kinetic theories. J. Stat. Phys. 83(5–6), 1021–1065 (1996)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  24. Lewin, L.: Polylogarithms and Associated Functions. North Holland, New York (1981)

    MATH  Google Scholar 

  25. Madelung, E.: Quantentheorie in hydrodynamischer Form. Z. Phys. 40, 322–326 (1926)

    ADS  MATH  Google Scholar 

  26. Méhats, F., Pinaud, O.: An inverse problem in quantum statistical physics. J. Stat. Phys. 140, 565–602 (2010)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  27. Trovato, M., Reggiani, L.: Quantum maximum entropy principle for a system of identical particles. Phys. Rev. E 81, 021119 (2010), 11 pp.

    Article  ADS  Google Scholar 

  28. Trovato, M., Reggiani, L.: Quantum maximum-entropy principle for closed quantum hydrodynamic transport within a Wigner function formalism. Phys. Rev. E 84, 061147 (2011), 29 pp.

    Article  ADS  Google Scholar 

  29. Von Neumann, J.: Mathematical Foundations of Quantum Mechanics. Princeton University Press, Princeton (1955)

    MATH  Google Scholar 

  30. Wigner, E.: On the quantum correction for thermodynamic equilibrium. Phys. Rev. 40, 749–759 (1932)

    Article  ADS  Google Scholar 

  31. Wood, D.C.: The computation of polylogarithms. University of Kent computing laboratory, Technical report 15/92 (1992)

  32. Zachos, C.K., Fairlie, D.B., Curtright, T.L. (eds.): Quantum Mechanics in Phase Space. An Overview with Selected Papers. World Scientific Series in 20th Century Physics, vol. 34. World Scientific, Hackensack (2005)

    MATH  Google Scholar 

  33. Zamponi, N.: Some fluid-dynamic models for quantum electron transport in graphene via entropy minimization. Kinet. Relat. Models 5(1), 203–221 (2012)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

This work was partially supported by the Italian Ministry of University, PRIN “Mathematical problems of kinetic theories and applications”, prot. 2009NAPTJF_003.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Luigi Barletti.

Appendices

Appendix A: Moments of Fermi and Bose Distributions and Related Integrals

We recall that the polylogarithm of order s, with s∈ℝ, is defined in the complex unit disc by the power series

$$\operatorname {\mathrm {Li}}_s(z) = \sum_{k=1}^\infty \frac{z^k}{k^s}, \quad {\vert {z} \vert } < 1, $$

and can be analytically continued to a larger domain (depending on s). To our purposes it will be enough to know that \(\operatorname {\mathrm {Li}}_{s}(z)\) is always well defined, real-valued and regular for z∈(−∞,1), and

$$ \lim_{z \to 1^-} \operatorname {\mathrm {Li}}_s(z) = \left\{ \begin{array}{l@{\quad }l} \zeta(s),& \mbox{if }s>1,\\[5pt] +\infty, &\mbox{if }s \leq 1, \end{array} \right. $$
(A.1)

where ζ is the Riemann zeta function. The polylogarithms are strictly connected with the moments of FD and BE distributions.

Definition A.1

For λez>−1, \(\lambda \not= 0\), and s∈ℝ we define

$$ \phi_s(z) = - \frac{1}{\lambda } \operatorname {\mathrm {Li}}_s \bigl(-\lambda \mathrm {e}^z\bigr), $$
(A.2)

where \(\operatorname {\mathrm {Li}}_{s}\) is the polylogarithm of order s. From known identities [24] we have that, for s>0, the above definition is equivalent to

$$ \phi_s(z) = \frac{1}{\varGamma(s)} \int _0^\infty \frac{t^{s-1}}{e^{t-z} + \lambda }\,dt \quad (s > 0) $$
(A.3)

(known as Fermi integral).

Note that ϕ s (z) is defined for all z∈ℝ if λ>0, and for z<−log|λ| if λ<0. In particular, in the FD case, λ=1, ϕ s (z) is defined on the whole real line while in the BE case, λ=−1, only for z<0.

The following properties of the functions ϕ s can be easily deduced from the properties of polylogarithms (see e.g. Refs. [24, 31]):

(A.4a)
(A.4b)
(A.4c)
(A.4d)
(A.4e)

Here, “f(x)∼g(x) for xy” means lim xy f(x)/g(x)=1.

Starting from the identity (A.3), we shall now compute explicit expressions, in terms of the functions ϕ s , of all the kinds of integrals that have been encountered throughout this paper.

Lemma A.1

For λez>−1, k=1,2,3,… and s>0, let us consider the integrals

$$ I_k^s(z) = \frac{1}{\varGamma(s)} \int _0^\infty \frac{t^{s-1}}{ (\mathrm {e}^{t-z} + \lambda )^k}\,dt. $$
(A.5)

Then, \(I_{1}^{s}\) is given by Eq. (A.3) and higher values of k are recursively obtained by

$$ I_{k+1}^s(z) = \frac{1}{\lambda} \biggl( I_k^s(z) - \frac{1}{k} \frac{dI_k^s}{dz}(z) \biggr). $$
(A.6)

In particular (omitting the argument z),

$$ \everymath{\displaystyle }\begin{array}{l} I_1^s = \phi_s,\\[5pt] I_2^s = \lambda ^{-1} (\phi_s - \phi_{s-1} ),\\[5pt] I_3^s =\lambda ^{-2} \biggl(\phi_s - \frac{3}{2}\phi_{s-1} + \frac{1}{2}\phi_{s-2} \biggr),\\[8pt] I_4^s = \lambda ^{-3} \biggl(\phi_s - \frac{11}{6}\phi_{s-1} + \phi_{s-2} - \frac{1}{6}\phi_{s-3} \biggr). \end{array} $$
(A.7)

Proof

The recursive formula (A.6) follows immediately from a formal derivation of \(I_{k}^{s}(z)\) with respect to z; Eq. (A.7) follows from (A.3) and (A.6) by using the property (A.4e). □

Equation (A.7) suggests that we can look for an expression for \(I_{k}^{s}\) of this kind:

$$ I_k^s = \frac{1}{\lambda^{k-1}} \sum _{j=0}^{k-1} c^k_j \phi_{s-j}, \quad k \geq 1, $$
(A.8)

where \(c^{k}_{j}\) are numerical coefficients (independent on s) to be determined. Since

$$ \frac{d I_k^s}{dz}(z) = I_k^{s-1}(z) $$
(A.9)

(as it is apparent from Lemma A.1), from the recursive relation (A.6) we can write the equivalent relation

$$ I_{k+1}^s = \frac{1}{\lambda} \biggl( I_k^s - \frac{1}{k} I_k^{s-1} \biggr). $$
(A.10)

Inserting (A.8) into (A.10) yields

$$\sum_{j=0}^{k} c^{k+1}_j \phi_{s-j} = \sum_{j=0}^{k-1} c^k_j \phi_{s-j} - \frac{1}{k}\sum_{j=1}^{k} c^k_{j-1} \phi_{s-j}, \quad k \geq 2. $$

Equating the coefficients of ϕ s (j=0) we obtain \(c^{k+1}_{0} = c^{k}_{0}\), and then (since \(c^{1}_{0} = 1\), as follows form (A.8) with k=1)

$$ c^k_0 = 1, \quad k \geq 1; $$
(A.11a)

equating the coefficients of ϕ sk (j=k) we obtain \(c^{k+1}_{k} = - \frac{1}{k} c^{k}_{k-1}\), and then

$$ c^k_{k-1} = \frac{(-1)^k}{(k-1)!}, \quad k \geq 1; $$
(A.11b)

finally, equating the coefficients of ϕ sj , with 1≤jk−1, we obtain

$$ c^{k+1}_j = c^k_j - \frac{1}{k} c^k_{j-1}, \quad k \geq 1,\ 1 \leq j \leq k-1. $$
(A.11c)

By using the recursive relations (A.11a)–(A.11c) one can easily generate all the coefficients of the expansion (A.8).

Proposition A.1

Let \(I_{k}^{s}(z)\) be given as in the previous lemma. Then,

(A.12a)
(A.12b)
(A.12c)

where, as usual, \(n_{d} := (2\pi T)^{\frac{d}{2}}\).

Proof

By using polar coordinates it is easily shown that

$$ \int_{\mathbb {R}^d} \frac{{\vert {p} \vert }^{2m}}{ (\mathrm {e}^{\frac{{\vert {p} \vert }^2}{2T} - z} + \lambda )^k}\,dp = \frac{ n_d\, (2T)^m \varGamma(\frac{d}{2}+m)}{\varGamma(\frac{d}{2})}\, I_k^{\frac{d}{2}+m}(z). $$
(A.13)

This formula immediately yields Eq. (A.12a) and, by obvious symmetry considerations, Eq. (A.12b). The derivation of Eq. (A.12c) requires more explanations. We first show that

$$ J_d(z) := \int_{\mathbb {R}^d} \frac{p_i^4}{ (\mathrm {e}^{\frac{{\vert {p} \vert }^2}{2T} - z} + \lambda )^k}\,dp = 3 n_d T^2 I_k^{\frac{d}{2}+2}(z) $$
(A.14)

(which is of course independent on i). We proceed by double induction on d. The cases d=1,2 can be easily verified by direct computations. Then, assuming (A.14) to be valid for d, for d+2 we can write

$$J_{d+2}(z) = \int_{\mathbb {R}^2} J_d \biggl(z-\frac{{\vert {q} \vert }^2}{2T} \biggr) \,dq = 3 n_d 2\pi T^2 \int_0^\infty I_k^{\frac{d}{2}+2} \biggl(z - \frac{\rho^2}{2T} \biggr) \rho \,d\rho. $$

From Eq. (A.9) we obtain therefore

$$J_{d+2}(z) = 3 n_d 2\pi T^3 I_k^{\frac{d}{2}+3} \biggl(z - \frac{\rho^2}{2T} \biggr) \bigg|^0_{+\infty} = 3 n_{d+2} T^2 I_k^{\frac{d+2}{2}+2}(z), $$

which proves (A.14) by induction. On the other hand, Eq. (A.13) yields

$$\int_{\mathbb {R}^d} \frac{{\vert {p} \vert }^4}{ (\mathrm {e}^{\frac{{\vert {p} \vert }^2}{2T} - z} + \lambda )^k}\,dp = d(d+2)n_d T^2 I_k^{\frac{d}{2}+m}(z) $$

and then, using \({\vert {p} \vert }^{4} = ( \sum_{i} p_{i}^{2} )^{2} = \sum_{i} p_{i}^{4} + \sum_{i\not=j} p_{i}^{2}p_{j}^{2}\) (and symmetry considerations), we obtain

$$ \int_{\mathbb {R}^d} \frac{p_i^2 p_j^2}{ (\mathrm {e}^{\frac{{\vert {p} \vert }^2}{2T} - z} + \lambda )^k}\,dp = n_d T^2 I_k^{\frac{d}{2}+2}(z) $$
(A.15)

for \(i \not= j\). The two cases, (A.14) and (A.15), are summarized by (A.12c). □

Appendix B: Postponed Proofs

2.1 B.1 Proof of Proposition 2.1

According to the short notations adopted from Sect. 3.1 on, let us denote \(\mathcal {G}_{A,B}\) by \(\mathcal {G}\) and \(\frac{1}{T}h_{A,B}\) by h (definition (3.1)). Recalling, moreover, the formalism introduced in Sect. 2.1, we can write

where we used the fact that \(\operatorname {Op}_{\epsilon }(\mathcal {G})\) is, by definition (2.31), a function of \(\operatorname {Op}_{\epsilon }(h )\) and then \(\{h,\mathcal {G}\}_{\#}= 0\), because it is the inverse Weyl quantization of a vanishing commutator. Since, from a direct computation,

$$\frac{i}{\epsilon } \{p_j B_j, \mathcal {G}\}_\# = - p_j \frac{\partial B_j}{\partial x_k} \frac{\partial \mathcal {G}}{\partial p_k} + B_j \frac{\partial \mathcal {G}}{\partial x_j}, $$

then from the previous identity we have

(where Eq. (2.20) was used), which yields Eq. (2.32).  □

2.2 B.2 Proof of Lemma 3.1

Lemma 3.1 follows from elementary manipulations of formal Taylor expansions. In order to shorten the notations, let us introduce the (d+1)-dimensional vectors

The constraint system (3.6), in these notations, reads as follows:

$$ f(\mu) = m. $$
(B.1)

Note that f has a double dependence on ϵ: one is direct (which leads to the expansion (3.2), i.e. to the terms f (k)), and the other is through the Lagrange multipliers, which are expanded as μ=μ (0)+ϵμ (1)+ϵ 2 μ (2)+⋯. Then, we regard f as f(ϵ,μ(ϵ)), whose Taylor expansion at ϵ=0 can be written in this way:

(where we took into account that f (1)=0, from (3.4b), and, for the sake brevity, the third-order term was not shown). Since the moments m do not depend on ϵ, the constraint equation (B.1) is expanded as follows:

The first equation is Eq. (3.8a), the second one implies μ (1)=0 (i.e. A (1)=B (1)=0), the third one is Eq. (3.8b) and, finally, the fourth one (which has been omitted for brevity) implies μ (3)=0 (i.e. A (3)=B (3)=0).  □

2.3 B.3 Proof of Lemma 3.2

By using Eq. (3.4c) with n=1 we obtain

$$ \mathcal {G}_2 = - \frac{\mathcal {G}_0\, \operatorname {\mathcal {E}\!\mathit {xp}}_2(h) + \mathcal {G}_0\mathbin{\#_2}\mathrm {e}^h}{\mathrm {e}^h + \lambda }. $$
(B.2)

The first term contains \(\operatorname {\mathcal {E}\!\mathit {xp}}_{2}(h)\), whose explicit expression can be taken from Eq. (5.14) of Ref. [10] and reads as follows:

$$ \operatorname {\mathcal {E}\!\mathit {xp}}_2(h) = -\frac{\mathrm {e}^h}{8} \biggl( X_{ij} P_{ij} - S_{ij}S_{ji} + \textstyle{\frac{1}{3}} X_{ij}P_iP_j - \textstyle{\frac{2}{3}} S_{ij}P_iX_j + \textstyle{\frac{1}{3}} P_{ij}X_iX_j \biggr), $$
(B.3)

where we introduced the following notations:

(we recall that \(h = \frac{1}{T} h_{A,B}\) is given by (3.1)). The other term, \(\mathcal {G}_{0}\,\#_{2}\,\mathrm {e}^{h}\), is a second-order Moyal product, given by (2.5). Using the notations just introduced, we obtain:

where, for k≥0, we define

$$ F_k = \frac{\mathrm {e}^{kh}}{(\mathrm {e}^h + \lambda )^{k+1}}. $$
(B.4)

Putting together the two terms we obtain the following expression for \(\mathcal {G}_{2}\):

(B.5)

Note that in the Maxwell-Boltzmann case, λ=0, we have F k =eh for all k≥0 and, therefore,

Then, in this case, Eq. (B.5) reduces (as it should) to \(\operatorname {\mathcal {E}\!\mathit {xp}}_{2}(-h)\), which is given by (B.3) with the suitable changes of sign.

Now, defining q=pB, taking the moments \({\langle \mathcal {G}_{2} \rangle }\) and \({\langle q_{i} \mathcal {G}_{2} \rangle }\) from Eq. (B.5), and taking account of vanishing integrals of odd functions, we get

(B.6)

and

(B.7)

The moments of functions F k can be reduced to integrals of type \(I_{k}^{s}\) (see Lemma A.1) by using

$$ F_k = \frac{1}{\mathrm {e}^h + \lambda } \biggl(1 - \frac{\lambda }{\mathrm {e}^h + \lambda } \biggr)^k = \sum_{j=0}^k {k \choose j} \frac{(-\lambda )^j}{(\mathrm {e}^h + \lambda )^{j+1}}. $$
(B.8)

Recalling that

$$h = \frac{{\vert {q} \vert }^2}{2T} - \frac{A}{T}, $$

from (B.8) and (A.12a)–(A.12c) we obtain

Then, by using \({\langle p_{i}\mathcal {G}_{2} \rangle } = B_{i}{\langle \mathcal {G}_{2} \rangle } + {\langle q_{i}\mathcal {G}_{2} \rangle }\) and the identity

(from property (A.4e)), Eqs. (3.12a)–(3.12b) are easily obtained from the expressions (B.6) and (B.7).  □

Rights and permissions

Reprints and permissions

About this article

Cite this article

Barletti, L., Cintolesi, C. Derivation of Isothermal Quantum Fluid Equations with Fermi-Dirac and Bose-Einstein Statistics. J Stat Phys 148, 353–386 (2012). https://doi.org/10.1007/s10955-012-0535-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-012-0535-5

Keywords

Navigation