2 Mean-Field Theory: Hartree-Fock and BCS

2 Mean-Field Theory: Hartree-Fock and BCS

2 Mean-Field Theory: Hartree-Fock and BCS Erik Koch Institute for Advanced Simulation Forschungszentrum Julich¨ Contents 1 Many-electron states 2 2 Second quantization 5 2.1 Creation and annihilation operators . 5 2.2 Representation of Slater determinants . 7 2.3 Representation of n-body operators . 9 2.4 Reduced density matrices and Wick’s theorem . 11 3 Variational methods 14 3.1 Non-interacting electrons . 15 3.2 Hartree-Fock theory . 16 3.3 BCS theory . 26 4 Conclusion 29 A Basis orthonormalization 30 B Some useful commutation relations 31 C Pauli matrices and spin rotations 32 E. Pavarini, E. Koch, J. van den Brink, and G. Sawatzky (eds.) Quantum Materials: Experiments and Theory Modeling and Simulation Vol. 6 Forschungszentrum Julich,¨ 2016, ISBN 978-3-95806-159-0 http://www.cond-mat.de/events/correl16 2.2 Erik Koch 1 Many-electron states One of the deepest mysteries of physics is that all the different objects that surround us are built from a small set of indistinguishable particles. The very existence of such elementary particles that have all their properties in common is a direct consequence of quantum physics. Classical objects can always be distinguished by their location in space. This lead Leibniz to formulate his Principle of the Identity of Indiscernibles [1]. For quantum objects, however, the uncertainty principle makes the distinction of particles by their position impossible. Indistinguishability of quantum objects then means that there is no measurement that would let us tell them apart, i.e., all expectation values hΨjMjΨi must remain the same when we change the labeling of the distinct but indistinguishable the particles. The consequences for observables are straightforward: An observable M(x) acting on a single- particle degree of freedom x must operate on all indistinguishable particles in the same way, P 0 i.e., i M(xi). A two-body observable M(x; x ) must operate on all pairs in the same way, P 0 0 i;j M(xi; xj) with M(x; x ) = M(x ; x). We can thus write any observable in the form X 1 X 1 X M(x) = M (0) + M (1)(x ) + M (2)(x ; x ) + M (3)(x ; x ; x ) + ··· (1) i 2! i j 3! i j k i i6=j i6=j6=k (0) X (1) X (2) X (3) = M + M (xi) + M (xi; xj) + M (xi; xj; xk) + ··· ; (2) i i<j i<j<k where the summations can be restricted since the operators must be symmetric in their argu- (2) (2) ments, e.g. M (xi; xj) = M (xj; xi), while for two or more identical coordinates the opera- (2) tor is really one of lower order: M (xi; xi), e.g., only acts on a single coordinate and should be included in M (1). For the many-body wave functions Ψ(x1; x2; ··· ) the situation is slightly more complex. Since 2 the probability density jΨ(x1; x2; ··· )j is an observable, they should transform as one-dimen- sional (irreducible) representations, i.e., either be symmetric or antisymmetric under particle permutations. Which of the two options applies to a given elementary particle is determined by the spin-statistics theorem [2, 3]: The wave functions of particles with integer spin are sym- metric, those of particles with half-integer spin change sign wen arguments are exchanged. From an arbitrary N-particle wave function we thus obtain a many-electron wavefunction by antisymmetrizing 1 X P S− Ψ(x1; : : : ; xN ) := p (−1) Ψ xp(1); : : : ; xp(N) ; (3) N! P where (−1)P is the parity of the permutation P that maps n ! p(n). Since there are N! different permutations, this can easily become an extremely costly operation. Remarkably, a product of 3 N single-electron states 'α can be antisymmetrized much more efficiently (in O(N ) steps) by Hartree-Fock and BCS 2.3 writing it in the form of a determinant ' (x ) ' (x ) ··· ' (x ) α1 1 α2 1 αN 1 1 'α (x2) 'α (x2) ··· 'α (x2) p 1 2 N Φα1,...,αN (x1; : : : ; xN ) := . : (4) N! . .. 'α1 (xN ) 'α2 (xN ) ··· 'αN (xN ) For N = 1 the Slater determinant is simply the one-electron orbital Φα(x) = 'α(x) and for 0 0 N = 2 we get the familiar two-electron Slater determinant Φ 0 (x; x ) = (' (x)' 0 (x ) − p α,α α α 0 'α0 (x)'α(x ))= 2. Slater determinants are important because they can be used to build a basis of the many-electron Hilbert space. To see how, we consider a complete set of orthonormal single-electron states Z X 0 0 dx 'n(x) 'm(x) = δn;m (orthonormal) 'n(x) 'n(x ) = δ(x − x ) (complete) : (5) n To expand an arbitrary N-particle function a(x1; : : : ; xN ), we start by considering it as a func- tion of x1 with x2; : : : ; xN kept fixed. We can then expand it in the complete set f'ng as X a(x1; : : : ; xN ) = an1 (x2; : : : ; xN ) 'n1 (x1) n1 with expansion coefficients Z an1 (x2; : : : ; xN ) = dx 'n1 (x1) a(x1; x2; : : : ; xN ) : These, in turn, can be expanded as a functions of x2 X an1 (x2; : : : ; xN ) = an1;n2 (x3; : : : ; xN ) 'n2 (x2) : n2 Repeating this, we obtain the expansion of a in product states X a(x1; : : : ; xN ) = an1;:::;nN 'n1 (x1) ··· 'nN (xN ) : n1;:::;nN When the N-particle function Ψ is antisymmetric, the expansion coefficients will be antisym- P metric under permutation of the indices anp(1);:::;np(N) = (−1) an1;:::;nN . Fixing an order of the indices, e.g., n1 < n2 < : : : < nN , we thus get an expansion in Slater determinants X p Ψ(x1; : : : ; xN ) = N! an1;:::;nN Φn1;:::;nN (x1; : : : ; xN ) : n1<:::<nN Since we can write any antisymmetric function as such a configuration-interaction expansion, the set of Slater determinants n o Φ (x ; : : : ; x ) n < n < ··· < n (6) n1;:::;nN 1 N 1 2 N 2.4 Erik Koch forms a basis of the N-electron Hilbert space. Since the overlap of two Slater determinants Z Z 1 X 0 Y dx Φ (x) Φ (x) = (−1)P +P dx ' (x ) ' (x ) α1,...,αN β1,...,βN N! n αp(n) n αp0(n) n P;P 0 n h'α1 j'β1 i · · · h'α1 j'βN i = hΦ jΦ i = . .. (7) α1,...,αN β1,...,βN . h'αN j'β1 i · · · h'αN j'βN i is the determinant of the overlap of the constituent orbitals, the Slater determinants (6) form a complete orthonormal basis of the N-electron Hilbert space when the orbitals 'n(x) are a complete orthonormal basis of the one-electron Hilbert space. While we use a set of N one-electron orbitals 'n(x) to define an N-electron Slater determinant Φα1,...,αN (x) (4), this representation is not unique: Any unitary transformation among the N occupied orbitals will not change the determinant. Thus, strictly, a Slater determinant is not determined by the set of indices we usually give, but, up to a phase, by the N-dimensional subspace spanned by the orbitals '1;:::;'N in the single-electron Hilbert space. The projector to this space is the one-body density matrix Z (1) 0 0 Γ (x; x ) = N dx2 ··· dxN Φ(x; x2; : : : ; xN ) Φ(x ; x2; : : : ; xN ) : (8) To see this, we expand the Slater determinant along its first row N 1 X 1+n Φα1···αN (x1; : : : ; xN ) = p (−1) 'αn (x1) Φαi6=n (x2; : : : ; xN ) ; (9) N n=1 where Φαi6=n (x2; : : : ; xN ) is the determinant with the first row and the n-th column removed, which can be written as N −1-electron Slater determinants with orbital αn removed. Inserting this into (8) we find N (1) 0 X 0 ΓΦ (x; x ) = 'αn (x) 'αn (x ) ; (10) n=1 which is the expansion of the one-body density matrix in eigenfunctions (natural orbitals) show- ing that its eigenvalues (natural occupation numbers) are one. Any many-electron wave function (1) Ψ(x) with the same one-body density matrix ΓΦ equals Φ(x) up to a phase, i.e., jhΨjΦij = 1. We can generalize this procedure and calculate higher order density matrices by introducing the generalized Laplace expansion 1 P X 1+ i ni Φα1···α (x) = (−1) Φαn ···αn (x1; : : : ; xp)Φα (xp+1; : : : ; xN ); N qN 1 p i62fn1;:::;npg p n1<···<np which is obtained by writing the permutation of all N indices as a permutation of N − p indices and the remaining p indices separately summing over all distinct sets of p indices. This allows us to evaluate arbitrary matrix elements and higher order density matrices [4]. But as can be seen from the above expansion, the expressions very quickly get quite cumbersome. Fortunately there is a representation that is much better suited to handling antisymmetric wave functions. It is called second quantization. Hartree-Fock and BCS 2.5 2 Second quantization While originally introduced for quantizing the electromagnetic field, we can use the formalism of second quantization just as a convenient way of handling antisymmetric wave functions [5,6]. The idea behind this approach is remarkably simple: When writing Slater determinants in the form (4) we are working in a real-space basis. It is, however, often better not to work in a specific basis but to consider abstract states: Instead of a wave function 'α(x), we write a Dirac state jαi. Second quantization allows us to do the same for Slater determinants. Let us consider a Slater determinant for two electrons, one in state 'α(x), the other in state 'β(x).

View Full Text

Details

  • File Type
    pdf
  • Upload Time
    -
  • Content Languages
    English
  • Upload User
    Anonymous/Not logged-in
  • File Pages
    34 Page
  • File Size
    -

Download

Channel Download Status
Express Download Enable

Copyright

We respect the copyrights and intellectual property rights of all users. All uploaded documents are either original works of the uploader or authorized works of the rightful owners.

  • Not to be reproduced or distributed without explicit permission.
  • Not used for commercial purposes outside of approved use cases.
  • Not used to infringe on the rights of the original creators.
  • If you believe any content infringes your copyright, please contact us immediately.

Support

For help with questions, suggestions, or problems, please contact us