Accuracy of the time-dependent Hartree-Fock approximation

Similar documents
Lecture 5. Hartree-Fock Theory. WS2010/11: Introduction to Nuclear and Particle Physics

Mean-Field Limits for Large Particle Systems Lecture 2: From Schrödinger to Hartree

Lecture Models for heavy-ion collisions (Part III): transport models. SS2016: Dynamical models for relativistic heavy-ion collisions

Introduction to determinantal point processes from a quantum probability viewpoint

From the N-body problem to the cubic NLS equation

Systems of Identical Particles

Lecture notes for Mathematics 208 William Arveson. 24 November 1998

Classical and Quantum Propagation of Chaos

Properties of nonfreeness: an entropy measure of electron correlation

Second Quantization: Quantum Fields

08a. Operators on Hilbert spaces. 1. Boundedness, continuity, operator norms

arxiv:quant-ph/ v5 10 Feb 2003

Finite-dimensional spaces. C n is the space of n-tuples x = (x 1,..., x n ) of complex numbers. It is a Hilbert space with the inner product

Lecture 6. Fermion Pairing. WS2010/11: Introduction to Nuclear and Particle Physics

Extensions of representations of the CAR algebra to the Cuntz algebra O 2 the Fock and the infinite wedge

1 Quantum field theory and Green s function

Algebraic Theory of Entanglement

The Phase Space in Quantum Field Theory

Attempts at relativistic QM

Quantum field theory and Green s function

On representations of finite type

Introduction to Electronic Structure Theory

Boolean Inner-Product Spaces and Boolean Matrices

Second quantization (the occupation-number representation)

FREE PROBABILITY THEORY

Derivation of Mean-Field Dynamics for Fermions

TOEPLITZ OPERATORS. Toeplitz studied infinite matrices with NW-SE diagonals constant. f e C :

Elliott s program and descriptive set theory I

Second quantization. Emmanuel Fromager

1 Mathematical preliminaries

2. Introduction to quantum mechanics

Elementary linear algebra

Analysis Preliminary Exam Workshop: Hilbert Spaces

Functional Analysis. Franck Sueur Metric spaces Definitions Completeness Compactness Separability...

W if p = 0; ; W ) if p 1. p times

Mean Field Limits for Interacting Bose Gases and the Cauchy Problem for Gross-Pitaevskii Hierarchies. Thomas Chen University of Texas at Austin

Lecture 10. The Dirac equation. WS2010/11: Introduction to Nuclear and Particle Physics

(1.1) In particular, ψ( q 1, m 1 ; ; q N, m N ) 2 is the probability to find the first particle

SELF-ADJOINTNESS OF DIRAC OPERATORS VIA HARDY-DIRAC INEQUALITIES

Chapter One. The Calderón-Zygmund Theory I: Ellipticity

Ψ ν ), if ν is ω-normal, Ψω Ψ ν

A Concise Course on Stochastic Partial Differential Equations

ON THE UNIQUENESS OF SOLUTIONS TO THE GROSS-PITAEVSKII HIERARCHY. 1. Introduction

The Hartree-Fock Method

Fermionic coherent states in infinite dimensions

Operator norm convergence for sequence of matrices and application to QIT

Multipartite entanglement in fermionic systems via a geometric

The Condensate Equation for non-homogeneous Bosons. André F. Verbeure 1. Institute for Theoretical Fysics, K.U.Leuven (Belgium)

Ir O D = D = ( ) Section 2.6 Example 1. (Bottom of page 119) dim(v ) = dim(l(v, W )) = dim(v ) dim(f ) = dim(v )

The Quantum Heisenberg Ferromagnet

A Brief Introduction to Functional Analysis

Qubits vs. bits: a naive account A bit: admits two values 0 and 1, admits arbitrary transformations. is freely readable,

The X-ray transform for a non-abelian connection in two dimensions

v(r i r j ) = h(r i )+ 1 N

Mean-Field Theory: HF and BCS

Mathematical foundations - linear algebra

Poisson configuration spaces, von Neumann algebras, and harmonic forms

QUANTUM MECHANICS. Franz Schwabl. Translated by Ronald Kates. ff Springer

CONTROLLABILITY OF QUANTUM SYSTEMS. Sonia G. Schirmer

BRST and Dirac Cohomology

msqm 2011/8/14 21:35 page 189 #197

Functional Analysis. James Emery. Edit: 8/7/15

Symmetries, Groups, and Conservation Laws

We can instead solve the problem algebraically by introducing up and down ladder operators b + and b

Ensembles and incomplete information

Quantum Physics II (8.05) Fall 2002 Assignment 3

Spectral Theory, with an Introduction to Operator Means. William L. Green

SECOND QUANTIZATION. notes by Luca G. Molinari. (oct revised oct 2016)

The Spinor Representation

The following definition is fundamental.

New Identities for Weak KAM Theory

Injective semigroup-algebras

Summary lecture II. Graphene exhibits a remarkable linear and gapless band structure

Properties of nonfreeness: an entropy measure of electron correlation

1.2 Fundamental Theorems of Functional Analysis

COMPACT OPERATORS. 1. Definitions

LECTURE OCTOBER, 2016

2.1 Green Functions in Quantum Mechanics

Quantum statistics: properties of the Fermi-Dirac distribution.

Mathematical Analysis of a Generalized Chiral Quark Soliton Model

Chapter 2. Linear Algebra. rather simple and learning them will eventually allow us to explain the strange results of

Lax Solution Part 4. October 27, 2016

QM and Angular Momentum

Consistent Histories. Chapter Chain Operators and Weights

Quantum Mechanics II

Chapter 8 Integral Operators

where P a is a projector to the eigenspace of A corresponding to a. 4. Time evolution of states is governed by the Schrödinger equation

Inner product on B -algebras of operators on a free Banach space over the Levi-Civita field

Non commutative Khintchine inequalities and Grothendieck s theo

Tutorial 5 Clifford Algebra and so(n)

1 The postulates of quantum mechanics

Bosonic quadratic Hamiltonians. diagonalization

Derivation of Bose-Einstein and Fermi-Dirac statistics from quantum mechanics: Gauge-theoretical structure

An introduction to some aspects of functional analysis

WELL-POSEDNESS FOR HYPERBOLIC PROBLEMS (0.2)

SPECTRAL THEOREM FOR COMPACT SELF-ADJOINT OPERATORS

Hartree-Fock-Roothan Self-Consistent Field Method

Isotropic harmonic oscillator

Wave function methods for the electronic Schrödinger equation

at time t, in dimension d. The index i varies in a countable set I. We call configuration the family, denoted generically by Φ: U (x i (t) x j (t))

Transcription:

Accuracy of the time-dependent Hartree-Fock approximation Claude BARDOS, François GOLSE, Alex D. GOTTLIEB and Norbert J. MAUSER Abstract This article examines the time-dependent Hartree-Fock (TDHF) approximation of single-particle dynamics in systems of interacting fermions. We find the TDHF approximation to be accurate when there are sufficiently many particles and the initial manyparticle state is a Slater determinant, or any Gibbs equilibrium state for noninteracting fermions. Assuming a bounded two-particle interaction, we obtain a bound on the error of the TDHF approximation, valid for short times. We further show that the error of the the TDHF approximation vanishes at all times in the mean field limit. Introduction Dirac [] invented the time-dependent Hartree-Fock equation in 93. The time-dependent Hartree-Fock (TDHF) equation is a nonlinear Schrödinger equation designed to approximate the evolution of the single-electron state in an n-electron system. Dirac noted that the TDHF equation, originally written as a system of n coupled Schrödinger equations for occupied orbitals, may be written as a Liouville-von Neumann equation for the single-particle reduced density operator. We study the TDHF equation in this form, availing ourselves of trace norm techniques inspired by [2] to estimate the error of the TDHF approximation. We will consider Hamiltonian dynamics of fermions interacting through a two body potential V. The energy operator for a solitary particle will be denoted by L, the energy of interaction University of Paris 7 and LAN (Univ. Paris 6), France (bardos@math.jussieu.fr). ENS-Ulm and LAN (Univ. Paris 6), France (Francois.Golse@ens.fr). Wolfgang Pauli Inst. c/o Inst. f. Mathematik, Univ. Wien, Strudlhofg. 4, A 9 Wien, Austria (alex@alexgottlieb.com). Wolfgang Pauli Inst. c/o Inst. f. Mathematik, Univ. Wien, Strudlhofg. 4, A 9 Wien, Austria (mauser@courant.nyu.edu).

of a single pair of particles will be denoted by V, and the total energy operator for a system of particles will be the sum of all single-particle energies and all pair energies. In this article we only consider bounded interaction potentials V ; the case where V represents Coulomb repulsion between electrons will be treated in a forthcoming paper [3]. Although the number of particles does not change under the dynamics just described, we prefer to formulate the dynamics on a fermion Fock space, so that we may consider initial states of indeterminate particle number. We are particularly interested in initial states which are Gibbs equilibrium states (grand-canonical ensembles), for we are going to show that such initial states enhance the accuracy of the TDHF approximation. Let L be a self-adjoint operator on H, and let V be a bounded Hermitian operator on H H that commutes with the transposition operator U defined by U(x y) = y x. We are going to discuss the dynamics on the fermion Fock space F H whose Hamiltonian H may be written in second quantized form as H = j L i a j a i + kl V ij a k a l a ja i () i,j i,j,k,l We will analyze the solutions of the Liouville-von Neumann equation i d D(t) dt = [H, D(t)] D() = D, (2) which is the evolution equation for the density operator in the Schrödinger picture of quantum dynamics. We will see that (2) leads to the following equation for the single-particle density operator N (t): i d dt N (t) = [L, N (t)] + [V, N 2 (t)] : N () = N (D ). (3) (The subscript : denotes a partial trace; see definition () below.) This equation for N (t) is not closed, for its right hand side involves the two-particle density operator N 2 (t). The TDHF approximation to N (t) is the solution of the initial value problem i d dt F (t) = [L, F (t)] + [V, F (t) 2 2A 2 ] : F () = N (D()) (4) where A 2 is the orthogonal projector of H H onto the subspace of antisymmetric vectors. The existence and uniqueness of solutions of (4) were established in [6] for the case where V is a bounded operator, and in [9, 7] for the case where V is a Coulombic interaction. 2

The TDHF equation (4) is obtained by closing the single-particle equation (3) with the Ansatz N 2 = (N N )2A 2 (5) at all times. The relation (5) holds for pure states corresponding to Slater determinants, and also for Gibbs densities. However, even supposing that N 2 () satisfies (5), the interaction V is likely to introduce correlations in N 2 (t), that is, departures from (5), and ignoring those correlations in the TDHF equation requires some justification. We are going to prove that the absence of correlations is self-perpetuating in the mean field limit: if N 2 () satisfies (5) then N 2 (t) asymptotically satisfies (5) as the number of particles N tends to infinity and the interaction strength is scaled as /N. This is the content of Theorem 5.2. To prove this theorem we first bound the error of the TDHF approximation in terms of the average particle number and interaction strength. This bound is presented in Theorem 5. without reference to the mean field scaling, and it applies to any system of fermions whose Hamiltonian has the form (). Unfortunately, the bound is valid only at short times, i.e., up to a time inversely proportional to V and the average particle number. Let us advert to some shortcomings of our results. Firstly, we consider only bounded twoparticle potentials V. Fortunately, the challenges presented by the Coulomb potential can be overcome [3, 5] and the mean field limit of Theorem 5.2 also holds for certain electronic systems. Secondly, the explicit error bound of Theorem 5. is valid only at short times, which are too short to be of interest in molecular-electronic problems, even if the Coulomb interaction between electrons is truncated at the Bohr radius. However, the model () does not only apply to molecular-electronic problems, and we hope that Theorem 5. may find other applications, perhaps to certain models of interactions between nucleons [4]. We first published our derivation of the TDHF equation in the mean field limit in [2]. There we assumed that the initial states are Slater determinants. The main theme of this article is that the initial states need not be Slater determinants; the TDHF approximation should work equally well (or badly) for all initial many-fermion states of Gibbs type. Also, in this article we are not only interested in the mean field limit, and we offer the error bound of Theorem 5. for the unscaled problem. The next section gives the background on fermion Fock space, trace class operators, reduced density operators, and Gibbs equilibrium states. Section 3 carefully defines the many-particle dynamics generated by () and derives the single-particle equation (3) for N (t). Section 4 introduces the TDHF approximation of N (t) and its higher-order analogs. Section 5 presents Theorems 5. and 5.2. Finally, the appendix contains the proofs of several Propositions. 3

2 Definitions and notation Consider a quantum particle whose Hilbert space is H, i.e., a particle which, in isolation, would constitute a system whose (pure) quantum states are represented by the rank-one orthogonal projectors on some Hilbert space H. The set of quantum states available to a system of n particles of this kind is determined by their statistics, i.e., whether the particles are fermions, bosons, or distinguishable. If the particles are fermions, the pure states of a system of n of them are represented by the rank-one projectors onto vectors in the antisymmetric subspace H (n) of the tensor power space H n. The Hilbert space H n is the closed span of the simple tensors x x 2 x n x, x 2,..., x n H with the inner product y y 2 y n, x x 2 x n = y, x y 2, x 2 y n, x n. Let Π n denote the group of permutations of {, 2,..., n}. For each π Π n, a unitary operator U π on H n may be defined by extending U π (x x 2 x n ) = x π () x π (2) x π (n) to all of H n. A vector ψ H n is antisymmetric if U π (ψ) = sgn(π)ψ for all π Π n. The antisymmetric vectors in H n form a closed subspace which will be denoted H (n). The orthogonal projector with range H (n) is A n = sgn(π)u π. (6) n! π Π n If x, x 2,..., x n is an orthonormal system in a Hilbert space H, then the vector n! An (x x 2 x n ), (7) is a unit vector in H (n). A vector of the form (7) is called a Slater determinant. If {e α } is a complete orthonormal system in H, then a basis of H (n) may be chosen from the set of all Slater determinants of the form (7) where {x,..., x n } a subset of the single particle basis of cardinality n. If the number of particles in the system is not fixed, the appropriate Hilbert space is the direct sum of the n-particle spaces H (n). This is the (fermion) Fock space F H = H () H () H (2) H (3). (8) 4

The possibility of a zero-particle state is accommodated by H () C, a one-dimensional space spanned by the vacuum vector Ω. The number operator N on F H is the self-adjoint operator whose restriction to H (n) equals multiplication by n. Annihilation and creation operators a x and a x are explicitly represented on F H as discussed in [8]. The statistical state of a system of fermions with single-particle space H determines a bounded positive continuous functional ω on the bounded operators on F H with ω(i) =. We are only going to consider normal states, i.e., those ω that satisfy ω(a) = Tr(D ω A) for some nonnegative trace class operator D ω of unit trace, which may be called the (statistical) density operator. Density operators on any Hilbert space K are trace class operators in particular. Let T (K) denote the real Banach space of Hermitian trace class operators T with the norm T = Tr( T ). We often use the inequality that, for bounded operators B on K, Tr(T B) T B. (9) A linear functional defined on the space of compact Hermitian operators by Φ T (K) = Tr(T K) () is continuous by (9), and indeed the space T (K) is isomorphic to the Banach dual of the space of compact Hermitian operators on K, via the isomorphism T Φ T. 2. Reduced density operators We now restrict our attention to density operators D on F H that commute with the number operator and such that N m D is trace class for all m N. For such D we will define the m th order reduced number density operators N m (D) and explain their physical significance. We begin by defining the reduction of an n-particle statistical density operator, i.e., a positive operator of unit trace D on H (n), to an m-particle density operator denoted D :m. If A is an operator on H (n), let A denote the extension of A to H n defined by first projecting onto H (n) and then applying A, i.e., A = A H n H (n). For each m n there exists a positive contraction from T (H (n) ) onto T (H (m) ) known as the partial trace. This partial trace map, which we denote by T T :m, is defined implicitly through the duality () by the requirement that Tr((K I I)T ) = Tr(KT :m ) 5

for all compact Hermitian operators K on H (m). It follows that, for any orthonormal basis O of H n m, ξ, T :m ψ = φ O (ξ ) φ, T ((ψ ) φ). () In case D is a density operator on H (n), and if m n, the operator D :m is known as the m-particle reduced density operator []. It is used to determine the expected values of m- particle observables when the n-particle system is in the statistical state D. The m-particle reduced density operators obtained from P ψ will be denoted (P ψ ) :m. If ψ is an n-particle Slater determinant, then the reduced density operators (P ψ ) :m satisfy (P ψ ) : = n n P xj (2) j= (P ψ ) :m = ( nm n ) (P ψ ) m : A m. (3) m Now we can define the reduced number densities N m, which serve to describe the m-particle correlations in a system of many particles. Let D be a density operator that commutes with N. Then D = D n (4) where each D n is a nonnegative trace class operator on H (n). Assume that n= n m Tr(D n ) <. (5) n= For such D, define the m th order reduced number density N m (D) = n=m n! (n m)! D n:m. (6) The operators N m (D) are called reduced density operators in Section 6.3.3 of [8], but we prefer to call them reduced number density operators because the trace of N is the average particle number. They are trace class by (5). An n-particle density operator D n extends to a density operator D n = D n (7) on all of Fock space. It is clear that N m ( D n ) equals N m ( D n ) = n(n ) (n m + )D n:m if m n but it equals the zero operator if m > n. Formula (9) below may clarify the sense in which N determines expected values of singleparticle observables. Following [8], we define the second quantization dγ(h) of a self-adjoint 6

operator H on H to be the closure of the essentially self-adjoint operator dγ n (H) on F H, where dγ n (H) denotes the restriction to H (n) of the operator dγ n (H) = n j times I I H I n j times I (8) j= on H n. The operator dγ(h) is called the second quantization of H; we might also call it a single-particle observable. Suppose that ω is a normal state on F H whose density D ω satisfies (4) and (5). From (6) it then follows that ω(dγ(b)) = Tr(N (D ω )B) (9) for any bounded Hermitian operator B on H. For example, the number operator N on F H is the second quantization of the identity operator, i.e., N = dγ(i). From (9) we see that the trace of N (D ω ) is the average number of particles in the system when it is in the state ω. Another example: if x is a unit vector in H, the operator a xa x on F H is the second quantization dγ(p x ) of the rank-one projector P x, and (9) tells us that ω(a xa x ) = Tr(N (D ω )P x ). Theorems 5. and 5.2 rely on the following fact, which is proved in the appendix: Proposition 2. If D is a density operator on F H that commutes with N and such that Tr(ND) <, then the operator norm of N (D) is not greater than. 2.2 Gibbs equilibrium states A Gibbs equilibrium state is that of a system of noninteracting fermions in thermal and chemical equilibrium with its environment, but only in case e βh is trace class, where H is the singleparticle Hamiltonian and /β is Boltzmann s constant times the temperature. In this case the density operator is proportional to exp( β dγ(h µi)), where µ R is the chemical potential. This density operator is diagonalizable with respect to an occupation number basis of Fock space, and occupation numbers are stochastically independent. If the single-particle Hamiltonian has discrete eigenvalues ɛ ɛ 2, the probability of occupation of the j th level equals /( + e β(ɛj µ) ). When D is a Gibbs equilibrium density then N m (D) = N (D) m m!a m. The same is true when D is a pure state corresponding to a Slater determinant. Indeed, this relation characterizes the gauge-invariant quasifree states of the CAR algebra that have a trace class single-particle operator, and both Slater densities and Gibbs equilibrium densities are of this type. We describe these states in probabilistic language. Let {φ j } j J be a basis of H. One basis of F H consists of the vacuum vector Ω and all Slater determinants made of vectors from {φ j }. This basis is indexed by F, the set of finite subsets of J including the empty set. A nonempty subset s = {j,..., j n } of J corresponds to a Slater determinant Ψ(s) formed from the vectors 7

φ j,..., φ jn. (There are in fact two such Slater determinants, opposite in sign, and we choose one of them.) The empty subset of J corresponds to the vacuum vector Ω, i.e., P ( ) = Ω. Let P be a probability measure on F with the σ-field of all its subsets. That is, let P : F [, ] be equal to except on a countable subset of F, on which subset P(s) =. Define the random variables N j on F by N j (s) = { if j s if j / s and define N = N j. Suppose that the N j are independent with respect to P, and define P(N j = ) = p(j). (2) Then P(s) = P((N j = ) (N jn = )) = s j,...,j n n p(j i ) (2) when j,..., j n are distinct. Note that E(N) = p(j) < since P(N < ) =. j J Given a probability measure P on F, we define a density operator on F H by G[P] = P(s)P s, (22) s F i= where P s denotes the orthogonal projector onto the span of Ψ(s). Proposition 2.2 Let P be as in (2) and (2). Let G denote the density operator G[P] of (22). Then for all n N N n (G) = N (G) n n!a n. (23) The preceding proposition is proved in the appendix. Proposition 2.2 and the following proposition (also proved in the appendix) together imply Proposition 2.4 below: Proposition 2.3 If T is a Hermitian trace class operator then T n n!a n T n. (24) Proposition 2.4 Let G be as in Proposition 2.2. Then Nn (G) N (G) n. 8

3 Definition of the dynamics on Fock space We now define the dynamics (2) more carefully, and derive the reduced dynamics (3) from (2). Let H be a self-adjoint operator on H, and let V be a bounded Hermitian operator on H H that commutes with the transposition operator U(x y) = y x. For j n, let L j denote the operator j times I I L I n j times I on H n (the value of n j is not explicit in the notation L j but it will always be clear from context). For i < j n, let U (ij) denote the permutation operator on H n that transposes the i th and j th factors of any simple tensor x x n, and let V ij = U (i) U (2j) ( V I n 2 ) U (2j) U (i) (again, the domain H n of V ij will always be clear from context). For each n, define the operators L (n) = n j= L j H (n) = L (n) + i<j n on H (n) (these operators are defined on all of H n but we are only considering their restrictions to the invariant subspace H (n) ). The Hamiltonian operator H, which we had formally represented above by (), is the direct sum H = H (n) defined on the domain { D(H) = x = x n F H : H (n) x n 2 < }. This operator is closed and self-adjoint (see Section 6.3. of [8]), and i H is the generator of the strongly continuous group W t = n= n W (n) t of unitary operators on F H, where W (n) t = exp ( ith(n)). It is convenient to have some notation for the free part of the dynamics, and we will subsequently use V ij U (n) t = exp ( it L(n)) and U t = n= U (n) t. 9

The Liouville-von Neumann dynamics corresponding to (2) are given by the group W t (D) = W t DW t (25) of isometries of T (F H ). (See Proposition 3.4 of [6] for a proof that groups of isometries defined in this way are strongly continuous.) Define the subspaces T n T (H) consisting of all n-particle trace class operators: { T n = T as in (7) : Tn T (H )} (n). These subspaces are invariant under W t, and the restriction of W t to T n is W (n) t The generator of this group of isometries is (T ) = W (n) t T W (n) t. L n + where L (n) is the the generator of the group U (n) t i<j n [V ij, ], (26) (T ) = U (n) t T U (n) t, (27) n [Lj, ]. (See [6] and references therein for a proof.) which may be written formally as i Since (26) is a bounded perturbation of L (n), it follows that W t ( T n ) equals T n (t), where T n (t) = U (n) t T n () i U (n) i<j n when T n T n. Taking the m th partial trace of both sides of (28) yields T n:m (t) = t T n:m () i (n m) i i<j m [V ij, T n (s)] ds (28) [V ij, T n:m (s)] m [V j,m+, T n:m+ (s)] :m ds because of the symmetry properties of T and V. Multiplying both sides of the last equation by n!/(n m)! we obtain j= n! (n m)! T n:m(t) = n! t (n m)! T n:m() i i m [ V j,m+, j= i<j m n! (n m )! T n:m+(s) [ n! ] V ij, (n m)! T n:m(s) ] :m ds. (29)

Now let D = D n be a density operator on F H. Then W t D = D n (t), where D n (t) is the solution of (28) with initial condition D n () = D n. If D satisfies moment condition (5) then N m (D(t)) = n=m n! (n m)! D(t) n:m is trace class. We abbreviate N m (D(t)) by N m (t). Summing the right-hand sides of (29) with D in place of T yields N m (t) = t N m () i i i<j m [V ij, N m (s)] m [V j,m+, N m+ (s)] :m ds, (3) j= for (5) permits the interchange of the sum and the integral. This is equation (3) in integral form when m =. To summarize: Proposition 3. Let U t and W t be as defined in (27) and (25). Suppose that D is a density operator on F H of the form D = D n such that moment condition (5) holds for some m N. Let N m (t) denote N m (W t (D)). Then N m (t) satisfies equation (3). 4 The TDHF hierarchy The existence and uniqueness of mild solutions of the TDHF equation (4) is established in [6]. There it is shown that the integral equation F (t) = U () t F () i U () [ ] V, F (s) 2 2A 2 ds (3) : has a unique solution F (t) for any Hermitian trace class operator F (). Define F (t) = F (t) and, for m >, define F m (t) = F (t) m m!a m. (32) We proceed to derive equations for the F m (t) from (3). First, set G(t) = U () t F (t), so that G(t) = F () i U () [ ] s V, F (s) 2 2A 2 ds. :

Now apply the product rule (in integral form) to G(t) m : Apply t G(t) m = F () m i F (t) m = F () m i m j= s G(s) j U () [ ] s V, F (s) 2 2A 2 : G(s) n j ds m j= to both sides of the preceding equation to obtain = t F () m i [ V j,m+, F (s) m+( I U (j,m+) ) ] :mds. m j= Multiply both sides of the last equation by m!a m U s (m) (XA m ) and that A m commutes with V j,m+ : F m (t) = t F m () i [ V j,m+, F (s) m+( I U (j,m+) ) ] :mds. on the left, noting that s (X)A m = m [V j,m+, F (s) m+( ) ] I U (j,m+) m!am j= :m ds. Since (m + )!A m+ = ( I U (,m+) U (m,m+) ) m!am, the last equation may be rewritten F m (t) = t F m () i + j k m i m [V j,m+, F m+ (s)] :m ds j= [V j,m+, F (s) m+ U (k,m+) m!a m ] :m ds. (33) We call these equations for the F m (t) the TDHF hierarchy. The trace norm of the last term in (33) is bounded by m(m ) 2 { ( V m,m+ U (m,m+) Fm (s) F (s) )} ds. :m It can be verified that ( Vm,m+ U (m,m+) (F m (s) F (s)) ) :m = ( I m F (s) ) V m,m F m (s), whence the trace norm of the last term in (33) does not exceed m(m ) 2 V F (s) F m (s) ds. (34) 2

Now F m (s) F (s) m by Proposition 2.3 since F m = F m m!a m. Furthermore, F (s) = F () and F (s) = F () for all s > by Proposition 4.3 of [6]. Substituting into (34) produces the bound m(m ) 2 V F () F () m t (35) on the trace norm of the last term in (33). 5 Accuracy of the TDHF approximation We have shown that if D is a density operator on F H of the form D = D n such that n 2 Tr(D n ) <, then N (t) = U () t N () i U () [V j,2, N 2 (s)] : ds, where N and N 2 are the one-particle and two-particle reduced number density operators for a system which evolved under the dynamics () from the initial state D. In this section we will compare N (t) to the solution of the TDHF equation F (t) = U () t F () i U () [ ] V, F (s) 2 2A 2 ds : F () = N () (36) whose initial condition is N (). When the initial state D is a Gibbs density for noninteracting fermions, then the distance between N (t) and F (t) in trace norm can be controlled at short times (Theorem 5.). In the mean field limit, F (t) is an asymptotically accurate approximation to N (t) at all times t, provided that the initial states are Gibbs states for noninteracting fermions (Theorem 5.2). We state and discuss Theorems 5. and 5.2 before going on to prove them: Theorem 5. Let G be the density operator on F H of a Gibbs equilibrium state for noninteracting fermions, as in Proposition 2.2. Let N (t) and N 2 (t) denote N (W t (G)) and N 2 (W t (G)), respectively, where W t is the dynamics with two-particle interactions defined in (25). Let F (t) be the solution of the TDHF equation (36). Let τ denote (2 V N ). Then for t < τ. N (t) F (t) 3 ( ) 2 t (37) 2 τ t 3

This theorem implies, for instance, that N (τ/2) F (τ/2) 3 2. This is remarkable because there are about 2 N 2 interactions driving the dynamics () and the error could be much larger prima facie: it could be proportional to N. Unfortunately, the bound (37) on the error of the TDHF approximation is valid only when t < τ, which is inversely proportional to N, and we have no explicit bounds for larger t. If N is too large, the bound (37) is useless, for then it is valid for too short a time. The time-of-validity of (37) ends up being inversely proportional to N because the number of two-particle interactions is proportional to N 2, and none of these interactions is weaker than any other a priori. In the thermodynamic limit, where N with constant spatial density, the total interaction energy grows like N (rather than the square of N ) if the interaction potential is short-ranged. We do not know how to derive the TDHF equation in the thermodynamic limit, but we can derive it in the mean field limit, where the strength of the interaction is scaled in inverse proportion to N. For each value of the parameter λ >, consider the Hamiltonian H λ = j L i a j a i + λ kl V ij a k a l a ja i. (38) i,j If the initial state is given by a Gibbs equilibrium density D λ for noninteracting fermions, then Proposition 3. implies that all reduced number density operators exist and satisfy Nm(t) λ = t Nm() λ λ i λ i i,j,k,l i<j m m [ Vj,m+, Nm+(s) ] λ ds :m j= [ Vij, N λ m(s) ] N λ m() = N m (D λ ()). (39) We will show that if λ is inversely proportional to N λ, then Nm(t) λ is close to F λ (t) m m!a m in trace norm, where F λ (t) = U () t F λ () i U () [ ] V, F λ (s) 2 2A 2 ds : F λ () = N λ (). (4) Theorem 5.2 Let { D λ} be a family of Gibbs equilibrium densities for noninteracting fermions, λ> as in Proposition 2.2, with lim sup λ N (D λ ) <. λ 4

Let Nm(t) λ be the solution of (39), let F λ (t) be the solution of the TDHF equation (4) and let Fm(t) λ = F λ (t) m m!a m. Then lim Nm(t) λ Fm(t) / λ N λ m = for all t > and all m N. λ We will derive Theorems 5. and 5.2 from Lemma 5. below. Let N m (t) be as in Proposition 3., and let F m (t) satisfy the TDHF hierarchy. In the hypotheses Theorems 5. and 5.2 we suppose that F () = N (D()), but for now let us only assume that F () and F () = N (D()). (4) The trace norm of N (t) is independent of t, and we shall denote it simply by N. Assuming (4), the bound (35) is itself bounded by m(m ) 2 V N m t. (42) Subtracting equations (33) from equations (3) and using (42) leads to the estimates Nm (t) F m (t) Nm () F m () m(m ) ) + V (2 N m + N m t + m 2 V N m+ (s) F m+ (s) ds. Iterating this estimate n times, one obtains Nm (t) F m (t) with C = 2 V / and n ( ) m + j a m+j C j t j + j j= n (m + n)! + C (m )! n j= b m+j j + ( m + j j ) C j t j+ n N m+n+ (s) F m+n+ (s) dsdt n dt (43) a m = Nm () F m () m(m ) ) b m = V (2 N m + N m. (44) 5

To make use of these estimates we need some control over the size of the integrand in (43). We will assume that there exists a constant B such that N m B N m (45) for all m. Then the last term on the right hand side of (43) is bounded by ( ) m + n (B + ) N m+n C n t n, m which tends to as n tends to infinity if m fixed and C N t <. Furthermore, assuming (45), we can bound b m of (44) by C(B/2 + ) N m m(m ) and we find the following: Lemma 5. Suppose that D is a density operator on F H of the form D = D n, such that (45) holds for all m N. Let W t be as defined in (25) and let N m (t) denote N m (W t (D)). Let F (t) be the solution of a TDHF equation (3) whose initial condition F () satisfies (4), and let F m (t) be as in (32). Then, with C = 2 V /, Nm (t) F m (t) N m when C N t <. j= ( ) m + j (C N t ) j m ( ) m + j (C N (m + j ) t ) j+ m Nm+j () F m+j () N m+j + N B + 2 2 Proof of Theorem 5. When G is the density operator of a Gibbs equilibrium state and F () = N (G), then F () = N (G) by Proposition 2., Nm () F m () = for all m by Proposition 2.2, and B = in (45) by Proposition 2.4. Upon simplifying the inequality in Lemma 5., one obtains Theorem 5.. Proof of Theorem 5.2 For N λ m(t) and F λ m(t) as in the hypothesis of Theorem 5.2, observe that F (t) and F (t) = N (D(t)) at all times t, and N λ m satisfies (45) with B =. Thus, we may apply Lemma 5., with a few changes: t and t + t may be substituted for and t, and C should be replaced by Cλ. These substitutions yield N λ m (t + t) F λ m(t + t) N λ m + j= N λ m+j (t) Fm+j(t) λ ( ) m + j (Cλ N λ N λ j= m+j t ) j m 3 ( ) m + j (Cλ N λ (m + j ) N λ t ) j+ 2 m j= 6 (46)

for t < (Cλ N λ ). Since u = lim sup Cλ N λ is finite by hypothesis, taking the lim sup λ of both sides of (46) implies that lim N λ m (s) Fm(s) / λ N λ m = m N (47) λ holds at time s = t + t if it holds at time s = t and t < /u. Since (47) holds at s =, an inductive argument proves that it holds at all times s >. 6 Appendix: the proofs of Propositions 2., 2.2, and 2.3 6. Proof of Proposition 2. We begin by proving Proposition 6. If D n is an n-particle fermionic density operator then Dn: /n. (48) Proof: Thanks to the convexity of the norm and the linearity of the partial trace, it suffices to prove (48) for fermionic pure states. Let H denote the single-particle Hilbert space, and let Ψ be a unit vector in H (n). Since P Ψ: is a compact Hermitian operator, there exists a unit vector u H such that PΨ: = u, PΨ: (u). (49) Let {φ j } j J be a basis of H containing u. For each subset s = {j,..., j n } of J, let Ψ(s) denote one of the two Slater determinants that may be formed from the vectors φ j,..., φ jn (the two choices differ only in sign). The set of vectors Ψ(s) is an orthonormal basis of H (n) and so Ψ = s Ψ(s), Ψ Ψ(s). By definition of the partial trace, u, P Ψ:(u) equals u φ j φ jn, P Ψ (u φ j φ jn ). (5) j,...,j n J From (49) and (5) P Ψ: = = j,...,j n J j,...,j n J PΨ (u φ j φ jn ) 2 Ψ(s), Ψ 2 Ψ(s), u φ j φ jn 2. s (5) 7

But whence by (5). Ψ(s), u φ j φ jn 2 = { /n! if s = {u, φj,..., φ jn } otherwise P Ψ: s Ψ(s), Ψ 2 (n )! n! Now we return to the proof of Proposition 2.. = Ψ(s), Ψ 2 = n n Let D be a density operator on F H that commutes with the number operator N and satisfies Tr(ND) <. Then D has the form (4) and ntr(d n ) <, so that N (D) = nd n: is defined and N (D) n Dn:. By Proposition 6. the operator norm of D n is less than or equal to Tr(D n )/n, whence n= s N (D) ntr(d n )/n = Tr(D) =. n= 6.2 Proof of Proposition 2.2 From (22), with P(s) as in (2), N n (G) = s:n(s) n N(s)! (N(s) n)! P(s) ( ) P Ψ(s). :n Substituting the expressions (3) and (2) for ( P Ψ(s) and collecting terms, we find that ):n N n (G) = = s:n(s) n distinct j,...,j n J P(s) ( N(s)P Ψ(s) ) n : n!a n [ s j,...,j n P(s) ] (Pφj P φjn ) n!an, (52) where N(s) is the size of s. The sum in (52) is made over distinct j,..., j n since (P φj P φjn )A n equals the zero operator if j r = j s for any r s. In particular, N (G) = j J [ s j ] P(s) P φj = j J p(j)p φj. (53) 8

Thus, N (G) n n!a n = distinct j,...,j n J n p(j i ) ( ) P φj P φjn n!an. (54) i= The sum in (54) is again restricted to distinct j,..., j n because of the presence of the antisymmetrizer A n. Substituting (2) into (54) yields (52), proving the proposition. 6.3 Proof of Proposition 2.3 Suppose T is a Hermitian trace class operator. There exists an orthonormal basis {e j } j J of H such that T = λ j P ej with λ j = T. j j The operator T n n!a n is diagonalizable with respect to the basis of Slater determinants formed from n distinct members of {e j }. Indeed, ( T n n!a n n! An (e j e jn ) ) = n!a n n!t n (e j e jn ) ( n ) = n! λ js n! An (e j e jn ). The trace norm of T n n!a n is the sum of the absolute values of its eigenvalues, whence T n n!a n = s= distinct j,...,j n J n λ js. (55) Note that the sum in (55) is over ordered sequences j, j 2,..., j n rather than subsets {j,..., j n }. But n λ js n n λ js = λ j = T n proving (24). distinct j,...,j n J s= j,...,j n J s= s= s= j J 9

References [] C. Bardos, F. Golse and N.J. Mauser, Weak coupling limit of the N-particle Schrödinger equation, Mathematical Analysis and Applications 7 no. 2 (2) 275-293 [2] C. Bardos, F. Golse, A.D. Gottlieb and N.J. Mauser, Mean field dynamics of fermions and the time-dependent Hartree-Fock equation, to appear in Journal de Mathématiques Pures et Appliqueés (22) [3] C. Bardos, F. Golse, A.D. Gottlieb and N.J. Mauser, Mean field dynamics of electrons and the time-dependent Hartree-Fock equation (manuscript) [4] C. Bardos, B. Ducomet, F. Golse, A.D. Gottlieb and N.J. Mauser, Weak interaction limit for nuclear matter and the time-dependent Hartree-Fock equation (manuscript) [5] C. Bardos, L. Erdös, F. Golse, N.J. Mauser and H.-T. Yau, Derivation of the Schrödinger- Poisson equation from the quantum N-particle Coulomb problem, C. R. Acad. Sci., t 334 (6) Série I Math. (22) 55-52 [6] A. Bove, G. Da Prato, and G. Fano, An existence proof for the Hartree-Fock time-dependent problem with bounded two-body interaction, Comm. Math. Phys. 37 (974) 83-9 [7] A. Bove, G. Da Prato, and G. Fano, On the Hartree-Fock time-dependent problem, Comm. Math. Phys. 49 (976) 25-33 [8] O. Bratteli and D. W. Robinson, Operator Algebras and Quantum Statistical Mechanics 2, Springer-Verlag, 997 [9] J. M. Chadam and R. T. Glassey, Global existence of solutions to the Cauchy problem for time-dependent Hartree equations, Journal of Mathematical Physics, 6 no. 5 (976) 22-3 [] P. Dirac, Note on exchange phenomena in the Thomas atom, Proc. Cambridge Philosophical Society 26 (93) pp. 376-385 [] D. ter Haar, On the density matrices used in Hartree-Fock calculations, Physica 26 (96) 4-44 [2] H. Spohn, Kinetic Equations from Hamiltonian Dynamics, Reviews of Modern Physics 53 (98) 6-64 2