Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 handwiki -- 1228 2022-11-30 01:30:46 |
2 format correct Meta information modification 1228 2022-12-05 05:03:30 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
HandWiki. Lieb–Oxford Inequality. Encyclopedia. Available online: https://encyclopedia.pub/entry/37298 (accessed on 27 July 2024).
HandWiki. Lieb–Oxford Inequality. Encyclopedia. Available at: https://encyclopedia.pub/entry/37298. Accessed July 27, 2024.
HandWiki. "Lieb–Oxford Inequality" Encyclopedia, https://encyclopedia.pub/entry/37298 (accessed July 27, 2024).
HandWiki. (2022, November 30). Lieb–Oxford Inequality. In Encyclopedia. https://encyclopedia.pub/entry/37298
HandWiki. "Lieb–Oxford Inequality." Encyclopedia. Web. 30 November, 2022.
Lieb–Oxford Inequality
Edit

In quantum chemistry and physics, the Lieb–Oxford inequality provides a lower bound for the indirect part of the Coulomb energy of a quantum mechanical system. It is named after Elliott H. Lieb and Stephen Oxford. The inequality is of importance for density functional theory and plays a role in the proof of stability of matter.

quantum chemistry stability coulomb

1. Introduction

In classical physics, one can calculate the Coulomb energy of a configuration of charged particles in the following way. First, calculate the charge density ρ, where ρ is a function of the coordinates x ∈ ℝ3. Second, calculate the Coulomb energy by integrating:

[math]\displaystyle{ \frac{1}{2}\int_{\mathbb{R}^3}\int_{\mathbb{R}^3}\frac{\rho(x)\rho(y)}{|x-y|} \, \mathrm{d}^3 x \, \mathrm{d}^3 y. }[/math]

In other words, for each pair of points x and y, this expression calculates the energy related to the fact that the charge at x is attracted to or repelled from the charge at y. The factor of ​12 corrects for double-counting the pairs of points.

In quantum mechanics, it is also possible to calculate a charge density ρ, which is a function of x ∈ ℝ3. More specifically, ρ is defined as the expectation value of charge density at each point. But in this case, the above formula for Coulomb energy is not correct, due to exchange and correlation effects. The above, classical formula for Coulomb energy is then called the "direct" part of Coulomb energy. To get the actual Coulomb energy, it is necessary to add a correction term, called the "indirect" part of Coulomb energy. The Lieb–Oxford inequality concerns this indirect part. It is relevant in density functional theory, where the expectation value ρ plays a central role.

2. Statement of the Inequality

For a quantum mechanical system of N particles, each with charge e, the N-particle density is denoted by

[math]\displaystyle{ P(x_1,\dots,x_N). }[/math]

The function P is only assumed to be non-negative and normalized. Thus the following applies to particles with any "statistics". For example, if the system is described by a normalised square integrable N-particle wave function

[math]\displaystyle{ \psi\in L^2(\mathbb{R}^{3N}), }[/math]

then

[math]\displaystyle{ P(x_1,\dots,x_N)=|\psi(x_1,\dots,x_N)|^2. }[/math]

More generally, in the case of particles with spin having q spin states per particle and with corresponding wave function

[math]\displaystyle{ \psi(x_1,\sigma_1,\dots,x_N,\sigma_N) }[/math]

the N-particle density is given by

[math]\displaystyle{ P(x_1,\dots,x_N)=\sum_{\sigma_1=1}^q\cdots\sum_{\sigma_N=1}^q|\psi(x_1,\sigma_1,\dots,x_N,\sigma_N)|^2. }[/math]

Alternatively, if the system is described by a density matrix γ, then P is the diagonal

[math]\displaystyle{ \gamma(x_1, ... , x_N; x_1, ..., x_N ). }[/math]

The electrostatic energy of the system is defined as

[math]\displaystyle{ I_P=e^2\sum_{1\le i\lt j\le N}\int_{\mathbb{R}^{3N}}\frac{P(x_1,\dots,x_i,\dots,x_j,\dots,x_N)}{|x_i-x_j|} \, \mathrm{d}^3 x_1\cdots\mathrm{d}^3 x_N. }[/math]

For x ∈ ℝ3, the single particle charge density is given by

[math]\displaystyle{ \rho(x)=|e|\sum_{i=1}^N\int_{\mathbb{R}^{3(N-1)}}P(x_1,\dots,x_{i-1},x,x_{i+1},\dots,x_N) \, \mathrm{d}^3 x_1\cdots\mathrm{d}^3 x_{i-1} \, \mathrm{d}^3 x_{i+1}\cdots\mathrm{d}^3 x_N }[/math]

and the direct part of the Coulomb energy of the system of N particles is defined as the electrostatic energy associated with the charge density ρ, i.e.

[math]\displaystyle{ D(\rho)=\frac12\int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \frac{\rho(x)\rho(y)}{|x-y|} \, \mathrm{d}^3 x \, \mathrm{d}^3 y. }[/math]

The Lieb–Oxford inequality states that the difference between the true energy IP and its semiclassical approximation D(ρ) is bounded from below as

[math]\displaystyle{ E_P=I_P-D(\rho)\ge -C|e|^\frac23\int_{\mathbb{R}^3}|\rho(x)|^\frac43 \, \mathrm{d}^3 x, }[/math]

 

 

(1)

where C ≤ 1.58 is a constant independent of the particle number N. EP is referred to as the indirect part of the Coulomb energy and in density functional theory more commonly as the exchange plus correlation energy. A similar bound exists if the particles have different charges e1, ... , eN. No upper bound is possible for EP.

3. The Optimal Constant

While the original proof yielded the constant C = 8.52,[1] Lieb and Oxford managed to refine this result to C = 1.68.[2] Later, the same method of proof was used to further improve the constant to C = 1.64.[3] It is only recently that the constant was decreased to C = 1.58.[4] With these constants the inequality holds for any particle number N.

The constant can be further improved if the particle number N is restricted. In the case of a single particle N = 1 the Coulomb energy vanishes, IP = 0, and the smallest possible constant can be computed explicitly as C1 = 1.092.[2] The corresponding variational equation for the optimal ρ is the Lane–Emden equation of order 3. For two particles (N = 2) it is known that the smallest possible constant satisfies C2 ≥ 1.234.[2] In general it can be proved that the optimal constants CN increase with the number of particles, i.e. CNCN + 1,[2] and converge in the limit of large N to the best constant CLO in the inequality (1). Any lower bound on the optimal constant for fixed particle number N is also a lower bound on the optimal constant CLO. The best numerical lower bound was obtained for N = 60 where C60 ≥ 1.41.[5] This bound has been obtained by considering an exponential density. For the same particle number a uniform density gives C60 ≥ 1.34.

The largest proved lower bound on the best constant is CLO ≥ 1.4442. It has been obtained in using a uniform electron gas, melted in the neighborhood of its surface.[6] The same lower bound CLO ≥ 1.4442 had been proved earlier in,[7] and acknowledged as such in.[6] Hence, to summarise, the best known bounds for C are 1.44 ≤ C ≤ 1.64.

4. The Dirac Constant

Historically, the first approximation of the indirect part EP of the Coulomb energy in terms of the single particle charge density was given by Paul Dirac in 1930 for fermions.[8] The wave function under consideration is

[math]\displaystyle{ \psi(x_1,\sigma_1,\dots,x_N,\sigma_N)= \frac{\det(\varphi_i(x_j,\sigma_j))}{\sqrt{N!}}. }[/math]

With the aim of evoking perturbation theory, one considers the eigenfunctions of the Laplacian in a large cubic box of volume |Λ| and sets

[math]\displaystyle{ \varphi_{\alpha,k}(x,\sigma) = \frac{\chi_\alpha(\sigma)\mathrm{e}^{2\pi\mathrm{i} k\cdot x}}{\sqrt{|\Lambda|}}, }[/math]

where χ1, ..., χq forms an orthonormal basis of q. The allowed values of k ∈ ℝ3 are n/|Λ|13 with n ∈ ℤ3+. For large N, |Λ|, and fixed ρ = N |e|/|Λ|, the indirect part of the Coulomb energy can be computed to be

[math]\displaystyle{ E_P(\mathrm{Dirac})=-C |e|^{2/3} q^{-1/3}\rho^{4/3}|\Lambda|, }[/math]

with C = 0.93.

This result can be compared to the lower bound (1). In contrast to Dirac's approximation the Lieb–Oxford inequality does not include the number q of spin states on the right-hand side. The dependence on q in Dirac's formula is a consequence of his specific choice of wave functions and not a general feature.

5. Generalisations

The constant C in (1) can be made smaller at the price of adding another term to the right-hand side. By including a term that involves the gradient of a power of the single particle charge density ρ, the constant C can be improved to 1.45.[9][10] Thus, for a uniform density system C ≤ 1.45.

References

  1. Lieb, E. H. (1979). "A lower bound for Coulomb energies". Physics Letters A 70 (5–6): 444–446. doi:10.1016/0375-9601(79)90358-X. Bibcode: 1979PhLA...70..444L.  https://dx.doi.org/10.1016%2F0375-9601%2879%2990358-X
  2. Lieb, E. H.; Oxford, S. (1981). "Improved lower bound on the indirect Coulomb energy". International Journal of Quantum Chemistry 19 (3): 427. doi:10.1002/qua.560190306.  https://dx.doi.org/10.1002%2Fqua.560190306
  3. Kin-Lic Chan, G.; Handy, N. C. (1999). "Optimized Lieb-Oxford bound for the exchange-correlation energy". Physical Review A 59 (4): 3075. doi:10.1103/PhysRevA.59.3075. Bibcode: 1999PhRvA..59.3075K. https://authors.library.caltech.edu/69066/1/PhysRevA.59.3075.pdf. 
  4. Lewin, Mathieu; Lieb, Elliott H.; Seiringer, Robert (October 2022). "Improved Lieb–Oxford bound on the indirect and exchange energies". Letters in Mathematical Physics 112 (5): 92. doi:10.1007/s11005-022-01584-5.  https://dx.doi.org/10.1007%2Fs11005-022-01584-5
  5. Seidl, M.; Vuckovic, S.; Gori-Giorgi, P. (2016). "Challenging the Lieb–Oxford bound in a systematic way. Molecular Physics". Molecular Physics 114 (7–8): 1076–1085. doi:10.1080/00268976.2015.1136440. Bibcode: 2016MolPh.114.1076S.  https://dx.doi.org/10.1080%2F00268976.2015.1136440
  6. Lewin, M.; Lieb, E.H.; Seiringer, R. (2019). "Floating Wigner crystal with no boundary charge fluctuations". Phys. Rev. B 100 (3): 035127. doi:10.1103/PhysRevB.100.035127. Bibcode: 2019PhRvB.100c5127L.  https://dx.doi.org/10.1103%2FPhysRevB.100.035127
  7. Cotar, C.; Petrache, M. (2019). "Equality of the Jellium and Uniform Electron Gas next-order asymptotic terms for Coulomb and Riesz potentials". arXiv:1707.07664 [math-ph]. //arxiv.org/archive/math-ph
  8. Dirac, P. A. M. (2008). "Note on Exchange Phenomena in the Thomas Atom". Mathematical Proceedings of the Cambridge Philosophical Society 26 (3): 376–385. doi:10.1017/S0305004100016108. Bibcode: 1930PCPS...26..376D.  https://dx.doi.org/10.1017%2FS0305004100016108
  9. Benguria, R. D.; Gallegos, P.; Tušek, M. (2012). "A New Estimate on the Two-Dimensional Indirect Coulomb Energy". Annales Henri Poincaré 13 (8): 1733. doi:10.1007/s00023-012-0176-x. Bibcode: 2012AnHP...13.1733B.  https://dx.doi.org/10.1007%2Fs00023-012-0176-x
  10. Lewin, Mathieu; Lieb, Elliott H. (2015). "Improved Lieb-Oxford exchange-correlation inequality with a gradient correction". Physical Review A 91 (2): 022507. doi:10.1103/PhysRevA.91.022507. Bibcode: 2015PhRvA..91b2507L.  https://dx.doi.org/10.1103%2FPhysRevA.91.022507
More
Information
Subjects: Others
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 517
Entry Collection: HandWiki
Revisions: 2 times (View History)
Update Date: 05 Dec 2022
1000/1000
Video Production Service