Lieb–Thirring inequality

In mathematics and physics, Lieb–Thirring inequalities provide an upper bound on the sums of powers of the negative eigenvalues of a Schrödinger operator in terms of integrals of the potential. They are named after E. H. Lieb and W. E. Thirring.

The inequalities are useful in studies of quantum mechanics and differential equations and imply, as a corollary, a lower bound on the kinetic energy of quantum mechanical particles that plays an important role in the proof of stability of matter.[1]

Statement of the inequalities

For the Schrödinger operator on with real-valued potential , the numbers denote the (not necessarily finite) sequence of negative eigenvalues. Then, for and satisfying one of the conditions

there exists a constant , which only depends on and , such that

 

 

 

 

(1)

where is the negative part of the potential . The cases as well as were proven by E. H. Lieb and W. E. Thirring in 1976 [1] and used in their proof of stability of matter. In the case the left-hand side is simply the number of negative eigenvalues, and proofs were given independently by M. Cwikel.,[2] E. H. Lieb [3] and G. V. Rozenbljum.[4] The resulting inequality is thus also called the Cwikel–Lieb–Rosenbljum bound. The remaining critical case was proven to hold by T. Weidl [5] The conditions on and are necessary and cannot be relaxed.

Lieb–Thirring constants

Semiclassical approximation

The Lieb–Thirring inequalities can be compared to the semi-classical limit. The classical phase space consists of pairs . Identifying the momentum operator with and assuming that every quantum state is contained in a volume in the -dimensional phase space, the semi-classical approximation

is derived with the constant

While the semi-classical approximation does not need any assumptions on , the Lieb–Thirring inequalities only hold for suitable .

Weyl asymptotics and sharp constants

Numerous results have been published about the best possible constant in (1) but this problem is still partly open. The semiclassical approximation becomes exact in the limit of large coupling, that is for potentials the Weyl asymptotics

hold. This implies that . Lieb and Thirring[1] were able to show that for . M. Aizenman and E. H. Lieb [6] proved that for fixed dimension the ratio is a monotonic, non-increasing function of . Subsequently was also shown to hold for all when by A. Laptev and T. Weidl.[7] For D. Hundertmark, E. H. Lieb and L. E. Thomas [8] proved that the best constant is given by .

On the other hand, it is known that for [1] and for .[9] In the former case Lieb and Thirring conjectured that the sharp constant is given by

The best known value for the physical relevant constant is [10] and the smallest known constant in the Cwikel–Lieb–Rosenbljum inequality is .[3] A complete survey of the presently best known values for can be found in the literature.[11]

Kinetic energy inequalities

The Lieb–Thirring inequality for is equivalent to a lower bound on the kinetic energy of a given normalised -particle wave function in terms of the one-body density. For an anti-symmetric wave function such that

for all , the one-body density is defined as

The Lieb–Thirring inequality (1) for is equivalent to the statement that

 

 

 

 

(2)

where the sharp constant is defined via

The inequality can be extended to particles with spin states by replacing the one-body density by the spin-summed one-body density. The constant then has to be replaced by where is the number of quantum spin states available to each particle ( for electrons). If the wave function is symmetric, instead of anti-symmetric, such that

for all , the constant has to be replaced by . Inequality (2) describes the minimum kinetic energy necessary to achieve a given density with particles in dimensions. If was proven to hold, the right-hand side of (2) for would be precisely the kinetic energy term in Thomas–Fermi theory.

The inequality can be compared to the Sobolev inequality. M. Rumin[12] derived the kinetic energy inequality (2) (with a smaller constant) directly without the use of the Lieb–Thirring inequality.

The stability of matter

The kinetic energy inequality plays an important role in the proof of stability of matter as presented by Lieb and Thirring.[1] The Hamiltonian under consideration describes a system of particles with spin states and fixed nuclei at locations with charges . The particles and nuclei interact with each other through the electrostatic Coulomb force and an arbitrary magnetic field can be introduced. If the particles under consideration are fermions (i.e. the wave function is antisymmetric), then the kinetic energy inequality (2) holds with the constant (not ). This is a crucial ingredient in the proof of stability of matter for a system of fermions. It ensures that the ground state energy of the system can be bounded from below by a constant depending only on the maximum of the nuclei charges, , times the number of particles,

The system is then stable of the first kind since the ground-state energy is bounded from below and also stable of the second kind, i.e. the energy of decreases linearly with the number of particles and nuclei. In comparison, if the particles are assumed to be bosons (i.e. the wave function is symmetric), then the kinetic energy inequality (2) holds only with the constant and for the ground state energy only a bound of the form holds. Since the power can be shown to be optimal, a system of bosons is stable of the first kind but unstable of the second kind.

Generalisations

If the Laplacian is replaced by , where is a magnetic field vector potential in , the Lieb–Thirring inequality (1) remains true. The proof of this statement uses the diamagnetic inequality. Although all presently known constants remain unchanged, it is not known whether this is true in general for the best possible constant.

The Laplacian can also be replaced by other powers of . In particular for the operator , a Lieb–Thirring inequality similar to (1) holds with a different constant and with the power on the right-hand side replaced by . Analogously a kinetic inequality similar to (2) holds, with replaced by , which can be used to prove stability of matter for the relativistic Schrödinger operator under additional assumptions on the charges .[13]

In essence, the Lieb–Thirring inequality (1) gives an upper bound on the distances of the eigenvalues to the essential spectrum in terms of the perturbation . Similar inequalities can be proved for Jacobi operators.[14]

References

  1. 1 2 3 4 5 E. H. Lieb, W. E. Thirring, Inequalities for the moments of the eigenvalues of the Schrödinger hamiltonian and their relation to Sobolev inequalities, Studies in Mathematical Physics, Princeton University Press (1976), 269–303
  2. M. Cwikel, Weak type estimates for singular values and the number of bound states of Schrödinger operators, Ann. of Math. (2) 106 (1977), no. 1, 93–100
  3. 1 2 E. H. Lieb, Bounds on the eigenvalues of the Laplace and Schroedinger operators, Bull. Amer. Math. Soc. 82 (1976), no. 5, 751–753
  4. G. V. Rozenbljum, Distribution of the discrete spectrum of singular differential operators, Izv. Vysš. Učebn. Zaved. Matematika (1976), no. 1(164), 75–86
  5. T. Weidl, On the Lieb–Thirring constants for , Comm. Math. Phys. 178 (1996), no. 1, 135–146
  6. M. Aizenman and E. H. Lieb, On semiclassical bounds for eigenvalues of Schrödinger operators, Phys. Lett. A 66 (1978), no. 6, 427–429
  7. A. Laptev and T. Weidl, Sharp Lieb–Thirring inequalities in high dimensions, Acta Math. 184 (2000), no. 1, 87–111
  8. D. Hundertmark, E. H. Lieb and L. E. Thomas, A sharp bound for an eigenvalue moment of the one-dimensional Schrödinger operator, Adv. Theor. Math. Phys. 2 (1998), no. 4, 719–731
  9. B. Helffer and D. Robert, Riesz means of bounded states and semi-classical limit connected with a Lieb–Thirring conjecture. II, Ann. Inst. H. Poincaré Phys. Théor. 53 (1990), no. 2, 139–147
  10. J. Dolbeault, A. Laptev, and M. Loss, Lieb–Thirring inequalities with improved constants, J. Eur. Math. Soc. (JEMS) 10 (2008), no. 4, 1121–1126
  11. A. Laptev, Spectral inequalities for partial differential equations and their applications, Fifth International Congress of Chinese Mathematicians. Part 1, 2, AMS/IP Stud. Adv. Math., 51, pt. 1, vol. 2, Amer. Math. Soc., Providence, RI, 2012, pp. 629–643
  12. M. Rumin, Balanced distribution-energy inequalities and related entropy bounds, Duke Math. J. 160 (2011), no. 3, 567–597
  13. R. L. Frank, E. H. Lieb, and R. Seiringer, Hardy–Lieb–Thirring inequalities for fractional Schrödinger operators, J. Amer. Math. Soc. 21 (2008), no. 4, 925–950
  14. D. Hundertmark and B. Simon, Lieb–Thirring inequalities for Jacobi matrices, J. Approx. Theory 118 (2002), no. 1, 106–130

Literature

  • Lieb, E.H.; Seiringer, R. (2010). The stability of matter in quantum mechanics (1st ed.). Cambridge University Press, Cambridge. ISBN 9780521191180.
  • Hundertmark, D. (2007). "Some bound state problems in quantum mechanics". Proc. Sympos. Pure Math. Providence, RI: Amer. Math. Soc. 76 (Spectral theory and mathematical physics: a Festschrift in honor of Barry Simon’s 60th birthday): 463–496.
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.