www.fgks.org   »   [go: up one dir, main page]

Academia.eduAcademia.edu
Spectral Analysis for Systems of Atoms and Molecules coupled to the Quantized Radiation Field Volker Bach  FB Mathematik MA 7-2; TU Berlin; Str. d. 17 Juni 136; D-10623 Berlin; Germany (bach@math.tu-berlin.de) Jurg Frohlich Inst. f. Theoretische Physik; ETH Honggerberg; CH-8093 Zurich; Switzerland (juerg@itp.phys.ethz.ch) Israel Michael Sigal y Dept. of Math.; Univ. of Toronto; Toronto, M5S 3G3; Canada (sigal@math.toronto.edu) November 20, 1998 Abstract We consider systems of static nuclei and electrons {atoms and molecules{ coupled to the quantized radiation eld. The interactions between electrons and the soft modes of the quantized electromagnetic eld are described by minimal coupling, p~ ! p~ eA~ (~x), where A~ (~x) is the electromagnetic vector potential with an ultraviolet cuto . If the interactions between the electrons and the quantized radiation eld are turned o , the atom or molecule is assumed to have at least one bound state. We prove that, for suciently small values of the feinstructure constant , the interacting system has a ground state corresponding to the bottom of its energy spectrum and that the excited states of the atom or molecule above the ground state turn into metastable states  Heisenberg Fellow of the DFG, supported by SFB 288 of the DFG, the TMR-Network on \PDE and QM". y Supported by NSERC Grant NA 7901 0 01/11/1998 EXT-2000-055 whose life-times we estimate. Furthermore the energy spectrum is absolutely continuous, except, perhaps, in a small interval above the ground state energy and around the threshold energies of the atom or molecule. Keywords: Renormalization Group, Spectrum, Resonances, Fock space, QED. BFS-3, November 20, 1998 I 1 Introduction and Survey of Results I.1 The Quantum Theory of Photons and Nonrelativistic, Bound Electrons In this paper we continue our mathematical analysis of the standard model of nonrelativistic, quantum-mechanical matter interacting with the quantized radiation eld initiated in [4, 5, 6]. The purpose of the present paper is to re ne the techniques developed in [4, 5, 6] in such a way that they can be applied to the study of physically relevant models. The physical system we are analyzing consists of a nite number of nuclei, treated as static sources, and a nite number of nonrelativistic electrons, e.g., atoms, ions, or molecules in a Born-Oppenheimer approximation, interacting with the soft modes of the quantized electromagnetic eld which is cut-o in the ultraviolet. The Hilbert space of pure state vectors of the system is given by H := Hel F (I.1) ; where Hel is the Hilbert space of some nite number, , of electrons, and F is the photon Fock space. Thus, in the Schrodinger (con guration-space) representation, Hel is given by the subspace of totally antisymmetric wave functions in 2 [(R 3  Z2)N ], where R 3 is the con guration space of a single N L electron, and Z2 describes its spin, i.e., Hel := AN 2  L (R  Z2)N 3  ; (I.2) with AN being the orthogonal projection onto the subspace of totally antisymmetric wave functions, as required by the Pauli principle. The one-photon Hilbert space is given by 2 [R 3  Z2], where R 3 is the photon momentum space and Z2 describes the two independent transversal polarizations of a photon. (Here and above, the integration measure on R 3 is Lebesgue measure.) The photon Fock space is then de ned by   1 M 3 n 2 (I.3) F := Sn (R  Z2) L n=0 L ; where Sn is the orthogonal projection onto the subspace of totally symmetric -photon wave functions, in accordance with the fact that photons are bosons. It is convenient to represent the Hilbert space H as the space of antisymmetric, square-integrable wave functions on the -electron con guration n N BFS-3, November 20, 1998 2 space with values in the photon Fock space F , i.e.,   H = AN L2 (R 3  Z2)N ; F : (I.4) The dynamics of the system is generated by the Hamiltonian H 0 :=  N X ~ ~xj ~j  ir j =1 2 21=2 3=2 A~ ( ~xj ) + Vc(x) + Hf ; (I.5) where we use units in which ~ = 1 and the electron mass equals 1=2. In (I.5), ~j = (jx; jy ; jz ) denotes the three Pauli matrices associated with the j th electron, xj is its position (in suitable units of length described below), is the feinstructure constant, and A~ (~x) denotes the quantized vector potential of the transverse modes of the electromagnetic eld in the Coulomb gauge, i.e.,  X Z d3 k (j~kj=K )  ~ i~k~x  ~ q ~"(k)e a(k) + ~"(~k) ei~k~xa (~k) ; A~ (~x) := =1;2  2!(~k) (I.6) where  is an entire function of rapid decrease on the real line, e.g., (r) := exp( r4 ), cutting o the vector potential in the ultraviolet domain, !(~k) := j~kj is the frequency of a photon with wave vector ~k, and ~"(~k),  = 1; 2, are photon polarization vectors satisfying ~"(~k)  ~"(~k) =  ; ~k  ~"(~k) = 0 ; for ;  = 1; 2. (I.7) Moreover, a(~k), a (~k) are standard creation- and annihilation operators (see, e.g., [25]) on F obeying the canonical commutation relations [a# (~k1 ); a# (~k2 )] = 0; [a (~k1); a(~k2)] =  (~k1 ~k2 ) ; (I.8) where a# = a or a . These objects are densely de ned, operator-valued tempered distributions on Fock space F . Fock space F contains a vector , the vacuum vector, uniquely determined, up to a phase, by the properties that k k = 1 and a (~k) = 0, for all  and ~k. A dense set of vectors in F is obtained by applying polynomials in a ( = 1; 2), smeared out with test functions, to the vacuum vector . The term Vc on the right side of Eqn. (I.5) de ning the Hamiltonian is the Coulomb potential describing the electrostatic interactions between electrons and nuclei. In our units, it is given by N X M X Zm + X 1 ; V (x) := (I.9) c j =1 m=1 j~xj R~ m j 1i<j N j~xi ~xj j BFS-3, November 20, 1998 3 where x  (~x ; : : : ; ~xN ) 2 R N , and R~ ; : : : ; R~ M 2 R are the positions of M static nuclei with atomic numbers Z ; : : : ; ZM . Finally, the Hamiltonian of the free quantized electromagnetic eld, Hf , is given by 3 1 3 1 1 Hf := X Z d k a(~k) !(~k) a(~k) : (I.10) 3 =1;2 The r.s. of Eqn. (I.10) de nes a densely de ned, positive, self-adjoint operator on F with absolutely continuous spectrum on the positive half-axis and a simple eigenvalue at 0 corresponding to the eigenvector . Next, we comment on the units chosen in (I.5). Length (and hence the positions, ~xj , of the electrons and, R~ m , of the nuclei) is measured in units of half a Bohr radius, r = m~el2 e2 , expressed in Gaussian units, where mel is the mass of an electron and e its electric charge. Photon wave lengths are measured in units of times half a Bohr radius, r , i.e., the unit for photon wave vectors is 22 r . The energy unit is chosen to e . The ultraviolet cuto of the radiation be 4 Rydberg, with 4 Ry = rBohr eld imposed by the function (jkj=K ) used in the de nition of the vector potential A~ (~x) turns o interactions between electrons and photons with energy large compared to K  4 Ry. The physical value of the feinstructure constant is  1=137. In this paper, plays the r^ole of a small, dimensionless number. Our results hold for suciently small values of . We shall not verify that the radii of convergence in of our analytical constructions cover the physical point =  1=137 (such a veri cation would presumably require numerical work on a computer). Our main concern in this paper is to analyze properties of the energy spectrum of the Hamiltonian H 0 and to study resonances for the dynamics generated by H 0 and estimate their life times. We shall show that, for suciently small values of > 0, H 0 has a normalizable ground state corresponding to the minimum of its spectrum and that \most" of its spectrum is absolutely continuous. Furthermore, we shall show that the excited bound states of the atom or ion (for purely technical reason, we exclude molecules here) with electrons decoupled from the radiation eld, i.e., the bound states of H , turn into metastable states of nite life time when that coupling is turned on. We show that Fermi's golden rule [24] yields an accurate estimate of the life times of metastable states, and we provide a rigorous justi cation of Bethe's formula [10, 9] for the Lamb shift. In order to describe our results and methods more precisely, we start with a discussion of the spectral properties of the unperturbed Hamiltonian 1 2 Bohr 2 1 2 (phys) 0 Bohr 1 2 Bohr BFS-3, November 20, 1998 4 (electrons decoupled from the radiation eld), H0  H 0 =0 = Hel + Hf ; (I.11) where Hel is the usual atomic (or molecular) Hamiltonian de ned by Hel := N X j =1 ~x + Vc(x) : j (I.12) We recall some key properties of Hel , viewed as an operator on Hel . For details and proofs, see [24, 13] and references given there. We rst note that potential Vc is a perturbation of the kinetic energy operator x := PN the j =1 ~x with zero relative bound. Thus Hel is a semibounded, self-adjoint operator on the domain D(Hel) = D(x) = Hel \ H 2[(R 3  Z2)N ], where H 2 is the usual Sobolev space. The essential spectrum of Hel is given by ess (Hel ) = [; 1), where  is the in mum of the spectrum of Hel , with N replaced by N 1, as follows from the HVZ theorem (see, e.g., [13]). Thus,  is the ionization threshold. In what follows, we shall make the following assumption about the atom, ion, or molecule under consideration, represented by the parameters N , ~ 1; : : : ; R ~ M , and Z1 ; : : : ; ZM . For the proof of existence of a ground state of R H in Section II, we assume that Hel has at least one eigenvalue E0 below the ionization threshold , i.e., we require that j 0 E0 := inf (Hel ) <  := inf ess(Hel ) : (I.13) The contents of Section III on the life times of metastable states is nontrivial only if, besides E0 , Hel has at least one further eigenvalue E1, with E0 < E1 < , i.e., in Section III we additionally require that the spectrum  (Hel ) = fE0 ; E1; : : : g [ [; 1) (I.14) of Hel consists of eigenvalues E0 < E1 < E2 <     , of nite multiplicity below , possibly with an accumulation point at  and (absolutely) continuous spectrum in [; 1); see Fig. 1. In Section III we shall also require an assumption saying, roughly speaking, that there are no accidental selection rules in the system described by Hel, which could prevent excited eigenstates of Hel from decaying radiatively (in 2nd order in the relevant coupling constant). For positive ions and (neutral) atoms or molecules our assumption (I.14) is justi ed, as it is known that if N  PMm=1 Zm then Hel has in nitely many eigenvalues of nite multiplicity below  [31]. In the case of negatively BFS-3, November 20, 1998 5 gr. states (abs.) cont. spectrum exc. states E0 E1 E2   Figure 1: The Spectrum of Hel E0 E1 E2   Figure 2: The Spectrum of H0 = Hel 1 + 1 Hf ionized atoms or molecules, i.e., if N > PMm=1 Zm, the question, whether Hel has isolated eigenvalues below the ionization treshold, is more subtle. Indeed, if N  PMm=1(2Zm + 1) then Hel has no eigenvalues at all [23] (see [27, 28] for earlier results). Assuming that the atomic numbers of the nuclei are such that E0 < , i.e., that Hel has isolated eigenvalues below the ionization threshold, it tends to be an open question whether the ground state corresponding to the energy E0 is unique or not, except when N = 1 or = 2, in which case a standard Perron-Frobenius argument proves uniqueness. (Non-uniqueness for N  3 may arise as a consequence of the Pauli principle.) The spectrum of the photon Hamiltonian Hf consists of a simple eigenvalue at 0, corresponding to the vacuum vector 2 F , and absolutely continuous spectrum (of in nite multiplicity) covering the half-axis [0; 1). Consequently, by separation of variables, the unperturbed Hamiltonian H0 = Hel + Hf on F has spectrum  (H0 ) =  (Hel ) +  (Hf ) : (I.15) The point spectrum of H0 is the same as the point spectrum of Hel , i.e., it consists of the eigenvalues fEj gj=0;1;2;::: (corresponding to the eigenvectors 'j;` , where f'j;`g`=1;2;:::;n is an orthonormal basis of eigenvectors of Hel corresponding to the eigenvalue Ej of multiplicity nj ). The continuous spectrum of H0 covers the half-axis [E0; 1) and consists of a union of branches [Ej ; 1) starting at the eigenvalues Ej and the branch [; 1); as indicated in Fig. 2. We note that the ground state energy E0 = inf (Hel ) of the atom or molecule in the absence of the quantized radiation eld coincides with the ground state energy E0  E0 ( = 0) = inf (H0) of the system of an atom j BFS-3, November 20, 1998 6 or molecule in the presence of photons, but decoupled from them. But, while E0 is an isolated eigenvalue of Hel , it lies at the tip of a branch of continuous spectrum of the Hamiltonian H0. Similarly, the energies E1 ; E2; : : : are isolated eigenvalues of Hel; but they are eigenvalues of H0 imbedded in continuous spectrum of H0, and each Ej is the threshold of a branch of continuous spectrum of H0 . These spectral properties of H0 make it a dicult problem to analyze, mathematically rigorously, the fate of the eigenvalues Ej of H0 , and the nature of the energy spectrum of the interacting system described by the Hamiltonian H introduced in (I.5), for > 0. Although the perturbation, W := H H0, is a small perturbation of H0, general analytical methods to deal with this type of problem in perturbation theory do not appear to be available. In [5, 6], we have started to develop such methods, tailor-made to analyze a class of Hamiltonians describing interactions between nonrelativistic quantum-mechanical matter and the radiation eld. In this paper we extend those methods to the Hamiltonian H (= H0 + W ) of Eqn. (I.5), which describes much of the physics of light atoms or molecules interacting with the quantized electromagnetic eld (within the Born-Oppenheimer approximation). For background material concerning the physics described by H , see e.g., [11, 12] and references given there. 0 0 0 0 0 0 I.2 Survey of Main Results In the next section, I.3, we consider the structure and properties of the perturbation W in the Hamiltonian H = H0 + W of (I.5). The strength of the perturbation W relative to H0 is measured by the parameter 0 0 0 0 g := ( K )3=2 ; (I.16) where K is the ultraviolet cuto in the electromagnetic vector potential A~ , as described in Eqn. (I.6). The parameter K is a \dimensionless energy scale" given by the photon energy above which interactions between electrons and the radiation eld are cut o , divided by 4 Ry. In Section I.3 we prove that if g is suciently small then the interaction Wg is bounded by H0 , in the sense of Kato [22, 25], with relative bound strictly smaller than 1. This proves that, for small , H is bounded from below and self-adjoint on the domain of H0 ; see Corollary I.2. (Under somewhat weaker assumptions one can prove that H de nes a semibounded quadratic form on an appropriate core. It is not easy, however, to characterize the domain of the corresponding selfadjoint operator.) Since g := ( K )3=2 is the relevant coupling parameter, we 0 0 BFS-3, November 20, 1998 henceforth write 7 Hg := H 0 =g2=3 =K ; Wg := W 0 =g2=3 =K : (I.17) We should emphasize that our bounds on g = ( K ) = become poor, as the number of electrons, N , becomes large. This does not mean that the Hamiltonian Hg is ill-de ned or unbounded from below, for large values of , as long as the ultraviolet cuto K is kept nite. In fact, using pathspace methods one can rigorously construct the semi-group exp( tHg ), for t  0, and prove that it is self-adjoint, strongly continuous in t > 0 with exp( tHg ) ! 1, as t ! 0, for arbitrary values of and N , as long as K < 1; (see [17, 30] for various ingredients of the proof). However, for the purposes of the analysis presented in Section III, Kato- (or quadratic-form-) boundedness of Wg in terms of H is an important property. In Section II we address the question whether Hg has a ground state corresponding to an eigenvalue E (g) := inf (Hg ). Our main result is Theorem II.1, which answers this question in the armative, as long as the e ective coupling constant g is suciently small. The method used in Section II is "non-perturbative" but non-constructive. For this reason, they do not enable us to estimate the multiplicity of the ground state energy E (g). However, if the number N of electrons is = 1 or = 2, and if the Zeeman terms in the Hamiltonian Hg are set to 0, then we can construct an L -space representation of the photon Fock space F (\electric eld" representation) with the property that exp( tHg ) is positivitypreserving on Hel F , where Hel is taken in the usual Schrodinger con guration-space representation, and, for N = 2, only spin-singlet states are considered. Modulo ergodicity of exp( tHg ) (see [24]), which we have not checked, a Perron-Frobenius argument then establishes uniqueness of the groundstate. It is worthwhile to point out some (fairly standard, but) fundamental consequences of the existence of a ground state of Hg for scattering theory: Using straightforward variants of methods developed in [18, 19, 20, 2, 3], one can construct Moller (wave) operators,  , such that the range of ( ) describes asymptotic states of the system consisting of an atom (or molecule) in a ground state accompanied by an outgoing cloud of freely moving photons. The obvious conjecture is that Ranf g = Ranf g =: Hasy , where Hasy is isomorphic to Hg F , and where Hg is the space of ground states of Hg . This conjecture, called \asymptotic completeness", would imply the unitarity of the scattering matrix for the scattering of photons o an atom or molecule below the ionization threshold. We are miles away from proving this conjecture! But, in a drastically simpler model of massive photons and con ned electrons, it has recently been proven in [16, 14]. 3 2 0 0 0 2 + (0) + (0) BFS-3, November 20, 1998 8 In Section III, we prove that, outside small neighbourhoods of E0 and  and below , the spectrum of Hg is purely absolutely continuous (Outside small neighbourhoods of the thresholds of Hel and above , the spectrum of Hg can be shown to be purely absolutely continuous by using the methods in [5, 7]. In particular, Hg does not have any eigenvalues in the vicinity of the imbedded eigenvalues E1 < E2 <    < En <  , of H0 , where  is a small positive constant depending on g and n. We will, however, make the idea rather precise that imbedded eigenvalues of H0 give rise to metastable states of Hg , and we shall estimate the life time, / g2, of these metastable states up to an error term O(g2+"), for some " > 0. Our notion of resonance is based on dilatation analyticity. In order to state our ideas simply, we assume that there is only one atomic nucleus (M = 1 in Eqn. (I.9)) of atomic number Z = Z1  2N located at R~ 1 = 0; (the general case of an arbitrary, nite number M of nuclei can be studied, too, by borrowing ideas developed in [21]). To describe dilatation analyticity, we start by scaling the positions, ~xj , of the electrons and the momenta, ~k, of the photons by ~ xj 7! ~ k e ~ xj ; 7! e (I.18) ~ k : When  is real, the transformations (I.18) determine a unitary transformation U on the Hilbert space H de ned in Eqns. (I.1){(I.4). It is easy to see that the subspace, D  H, of vectors, , with the property that () := U is analytic in , for jImj < =2, is dense in H. Furthermore, one easily checks that U Hf U = e  (I.19) Hf : These facts, combined with well-known results [1, 8] on dilatation analyticity for Schrodinger operators, show that, for arbitrary ; ' 2 D, the function F 0;' (; z ) :=  D  z H0 ( )  1 E '(  ) ; (I.20) where H0 () = U H0U 1 , is independent of , for jImj < =2, and, for Im =: # xed, F 0;' (; z) is analytic in z in the complement of the shaded region depicted in Fig. 3 In Section III, we construct the function F g ;' (; z ) := D   z Hg ( )  1 E '( ) ; (I.21) where Hg () = U Hg U 1 , with Hg as in (I.5), and we show that, for our choice of an ultraviolet cuto  (see (I.6) and below), and for arbitrary BFS-3, November 20, 1998 E0 9 E1 E2   Im() Figure 3: A projection of the Riemann surface of z 7! F 0;' (; z) onto the energy plane E0 E0 (g ) E1  Ej Ej;` (g ) E1 (g ) Figure 4: A projection of the Riemann surface of z 7! F g;' (; z) onto the energy plane and ' in D, F g;'(; z) is independent of , for jImj < =4 small enough, and, for Im =: # xed, F g;' (; z) is analytic in z in the complement of the shaded region depicted in Fig. 4. It thus provides an analytic continuation of the matrix element F g;' (0; z) of the resolvent of Hg z in z from the upper half plane into the lower half plane outside the spectrum of Hg (). This implies the absolute continuity of the spectrum of Hg for those energies which are contained in the resolvent set of Hg () (see Corollary III.2). In Section III, Eqns. (III.7){(III.10), we introduce a notion of resonance energy Ej (g) corresponding to the energy Ej of the j th excited state of the atom or molecule. We show that the Lamb shift, RefEj (g) Ej g, is given by Bethe's formula and ImfEj (g)g is given by Fermi's Golden Rule, to order g 2 , with error terms that we prove to be O (g 2+" ), for any 0 < " < 1=3. If there are no miraculous selection rules forbidding the decay of the j th excited state of the atom to rst order in g then ImfEj (g)g = jg 2 + O(g2+") ; (I.22) with j strictly positive. It then follows by standard reasoning that, for g > 0 suciently small, the spectrum of Hg is purely absolutely continuous in a neighbourhood of every eigenvalue Ej , j  1, of Hel for which j > 0. What we are really looking after, from a physics point of view, is a precise understanding of the decay of the excited states of theatom under the time  evolution exp[ it Hg ], as t becomes large. Let  := dist Ej ; (Hel) nfEj g > BFS-3, November 20, 1998 10 0, set Ij (=2) := (Ej =2; Ej + =2), and notice that Ij (=2) is an open  interval containing Ej and such that dist Ij ; (Hel ) n fEj g = =2. Let Fj denote a smooth characteristic function of Ij (=2) (see Subsection III.1). We shall identify a "j th excited state" of the atom with a vector of the form j := 1=2 (H Fj g ) Pel;j exp h g 2H i f (I.23)  (see (III.14)), where Pel;j is the projection onto the eigenspace of Hel corh i 2 responding to the eigenvalue Ej , and the operator exp g Hf essentially eliminates high-energy photons of energy larger than g2 in the state , which is assumed to be dilatation analytic. An example for a state  is given by =  'j;` (I.24) ; where 'j;` is an eigenvector of Hel with corresponding to the eigenvalue Ej . We then show that, for any 0 < " < 1=3, there exists a constant C  0 and, for any N 2 N , a constant CN  0 such that  j  exp[ B2   C itHg ] j  ln(1=g) exp   t g2 Cg 2+" ) j  + CN t N g4  ; (I.25) where B := 1 + supjj0 k()k. This estimate implies that, given " > 0, there is a nite constant C" > 0 such that, for t > C" j 1 g 2 ln ln(1=g), D j e itHg E  j ": (I.26) We remark that it is known from other methods (see e.g. [15]) that, given 0, there is a constant D" > 0 such that, for 0  t < D"g 2, "> D itHg j e E  j ": (I.27) Estimate (I:26) shows that the state j decays, with a life-time bounded  2 above by O g ln ln(1=g) . It is a typical example for the kind of estimates we are able to prove with the help of the methods developed in Section III. I.3 Relative Bounds, Self-Adjointness, and Dilatation Analyticity We return to Eqn. (I.5), which we write as Hg = H0 + Wg ; (I.28) BFS-3, November 20, 1998 11 where H0 is de ned in (I.11), and we obtain Wg = N X ( j =1 41=2 + 2 3=2 ~ ~xj ) + 2 3A~ 2( ~xj ) A~ ( ~xj )  ( ir 1=2 5=2 ~j  r~  A~  ( ~xj )   ) =g2=3 =K (I.29) from expanding the square in (I.5). Our rst goal in this subsection is to prove in Corollary I.2 that Wg is de ned on D(H0) and obeys the bound Wg jH0 + iC (N )j 1  C 0(N ) ( K )3=2 ; (I.30) for some constants C (N ); C 0(N )  0. This establishes the semiboundedness and self-adointness of Hg on D(H0) for g := ( K )3=2 < C 0(N ) 1 . Our second goal is to establish the dilatation analyticity of Wg () := U Wg U 1 (see Corollary I.3), where U is the dilatation operator de ned in (I.18), i.e., we prove that  7! Wg () is an analytic function on D(0; 0) := fz 2 C jjzj < 0g with values in B(D(H0 ); H), the bounded operators from D(H0) to H, for some 0 > 0. We establish this property by observing that the coupling functions in Wg () are analytic in , pointwise in the other parameters, and by verifying a bound similar to (I.30), namely, Wg () jH0 + iC (N; )j 1  C 0(N; ) g ; (I.31) for some constants C (N; ); C 0(N; )  0. In fact, Eqn. (I.30) is just the special case  = 0 in Eqn. (I.31). Finally, we establish the dilatation analyticity of Hg () := U Hg U 1 in Corollary I.4, assuming that Hel = x + Vc(x) is the Hamiltonian of an atom, i.e., M = 1. This simplifying assumption could be avoided by using exterior dilatations [21, 29], at the expense of having to deal with more involved estimates; we do not carry out this analysis here. We characterize the coupling functions in Wg () in terms of the following functions, G~ ~x(~k; ; ) := and B~ ~x(~k; ; ) := p 2 eq=2 (e  j~kj=K ) exp[ i ~k  ~x] ~" (~k)   K 3 !(~k) p (I.32) 2 eq3=2 (e  j~kj=K ) exp[ i ~k  ~x] ~k  ~" (~k) ;  i  K 3 !(~k) (I.33) BFS-3, November 20, 1998 12 where we introduce a dilatation parameter  2 D(0; 0)  C , for some 0 > 0 suciently small. Note that, pointwise for every ~x; ~k 2 R 3 and  2 Z2, the maps  7! G~ ~x(~k; ; ) and  7! B~ ~x(~k; ; ) are analytic in D(0; 0). Further~  G~ ~x (~k; ; 0). By means of G ~ ~x (~k; ; ) more, we notice that B~ ~x(~k; ; 0) = r 3 and B~ ~x(~k; ; ) we de ne the following functions on R  Z2 with values in the operators on Hel ,  N X  ~ ~ ~ ~  ~ ~ 2G~xj (k; ; )  ~pj + ~j  B~xj (k; ; ) ; w1;0 (k; ; ) := w0;1 (k; ; ) := j =1 (I.34) w2;0 (~k1 ; 1 ; ~k2; 2 ; ) := w0;2 (~k1 ; 1 ; ~k2; 2 ; ) (I.35)   N X ~ ~ := G~xj (k1 ; 1 ; )  G~ ~xj (~k2 ; 2 ; ) ; j =1 N X ~ ~ ~ ~xj (~k2 ; 2 ; ) G~ ~xj (~k1 ; 1 ; )  G w1;1 (k1 ; 1 ; k2; 2 ; ) := j =1 (I.36)  + G~ ~xj (~k1; 1; )  G~ ~xj (~k2 ; 2; ) ; and these, in turn, serve as coupling functions for the operators de ned by Wm;n () := X 1 ;:::;m+n =1;2 Z d3 k1    d3 km+n wm;n (~k1 ; 1 ; : : : ; ~km+n ; m+n ; ) a1 (~k1 )    am (~km ) am+1 (~km+1 )    am+n (~km+n ) : (I.37) Then we observe that, after normal ordering, the (dilated) interaction Wg () reads Wg () = X m+n2 g m+n Wmn () + g 2 Cno ; (I.38) R where Cno is de ned by Cno := 2N jG~ ~x(~k; 1; 0)j2 d3 k, which is independent of x, and g := ( K )3=2 . Henceforth and consistent with our previous de nitions, we omit  in our notation in the undilated case,  = 0, writing ~ ~x(~k; ) := G~ ~x (~k; ; 0) ; B~ ~x (~k; ) := B~ ~x (~k; ; 0) ; G (I.39) wm;n (~k1 ; 1 ; : : : ; ~km+n ; m+n ) := wm;n (~k1 ; 1 ; : : : ; ~km+n ; m+n ; 0) ; (I.40) Wm;n := Wm;n (0) : (I.41) BFS-3, November 20, 1998 13 Next, we de ne J : R ! R to be the smallest function such that w ; (~k; ; ) ( x + 1) = ; w ; (~k; ; ) ( x + 1) =  J (~k) (I.42) 3 + 1 2 10 1 2 01 holds, for all jj   and (~k; ) 2 R  Z , and such that 3 0 (I.43) w ; (~k ;  ; ~k ;  ; ) ; w ; (~k ;  ; ~k ;  ; )  J (~k ) J (~k ) w ; (~k ;  ; ~k ;  ; ) ; 20 1 1 2 2 11 2 1 1 02 2 1 1 2 2 2 1 2 holds, for all jj   and (~k ;  ); (~k ;  ) 2 R  Z . Note that, due to ir~ ~x  ( ~x + 1) = , we have that 0 1 2 j J (~k)  C ( ) N K 1 = 3 2 0 j~kj 1 = 1 2 2 3 2 2 (e  j~kj=K ) + j~kj (e  j~kj=K ) ; (I.44)   for some constant C ( )  0. The rapid decay of  implies that 0  := Z J (~k) !(~k) d k 2 = !1 2 3 < 1; (I.45) for any > 2. In particular,  is uniform in K  1, for any 1  > 2. This uniformity in K  1 is actually the basic requirement that determines p = 3=2 in the coupling parameter g = = K p. The main relative bound that we use is described in the following lemma. Lemma I.1. For all m; n 2 N with 1  m + n  2 and all  2 D(0;  ), the operators Wm;n() are de ned on D(H ) and obey the bound 3 2 0 0 0 Wm;n() jH + iC (N;  )j 0 0 1   4(1 +  +  ) 2 0 2 m+n)=2 ( 1 ; (I.46) for some constant C (N; ) > 0. Proof: We rst note that the canonical commutation relations (I.8) allow us to convert estimates on W1;0() into those for W0;1(). Indeed, for any 2 D(H0), kW ; () k = kW ; () k + 10 2 01 2 X =1;2 Z d k kw ; (~k; ; ) k 3 2 10  kW ; () k + 2  ( x + 1) = 01 2 0 1 2 2 : (I.47) BFS-3, November 20, 1998 Furthermore, kW0;1() k  14 X Z =1;2   2 d3 k J (~k) ( x + 1)1=2 a (k) 1=2 1 ( x + 1)1=2 Hf ; (I.48) and hence kW1;0() k ; kW0;1() k  2 (0 +  1) ( x + Hf + 1) : (I.49) Similarly, we convert estimates on W2;0 and W1;1 into those for W0;2. For W2;0 , for instance, the canonical commutation relations (I.8) imply that W2;0 () W2;0 () = W0;2Z() W0;2 () (I.50) + 2 d1d2 w2;0(1; 2; )w2;0(1; 2; ) Z +4 d1d2d3 w2;0(1; 3; ) w2;0(2; 3; ) a(1)a(2) ; where we denoted  := (~k; ), d := P=1;2 d3k, and a#( ) := a# (~k). This yields kW2;0() k  3kW0;2() k + 420 k k2 ; (I.51) R R and, as in (I.48), we obtain that kW0;2() k  2 1 kHf k  2 1 ( x + Hf + 1) : (I.52) The bound (I.46) follows from (I.49), (I.51), (I.52), and the additional use of the fact that x is relatively Hel -bounded with relative bound arbitrarily close to 1. Now self-adjointness of Hg on D(H0) and dilatation analyticity of Wg () are just two immediate implications of Lemma I.1. 3=2 < (1 + 2 + 2 ) 1=2 =10 then Hg is Corollary I.2. If 0 < g = ( K ) 0 1 self-adjoint and semibounded on D(H0 ). Corollary I.3. The map Wg : D (0; 0 ) ! B (D (H0 ); H),  7! Wg ( ) is analytic. Finally, we establish the dilatation analyticity of Hg () := U Hg U 1 , assuming that the potential Vc(x) is dilatation analytic, i.e., D(0; 0) 3  ! Vc (e x) 2 B(D(H0 ); H) is an analytic function. This property holds in case BFS-3, November 20, 1998 15 that Hel = x + Vc(x) is the Hamiltonian of an atom, i.e., M = 1, for arbitrary . Indeed, in the atomic case we may choose without loss of generality the position of the nucleus to be the origin of the one-electron con guration space, and then we obtain  Hel () := U Hel U 1 = e 2 x ) + e  Vc(x) : (I.53) H0 () := U H0 U 1 = Hel () + e  Hf (I.54) Therefore, is an analytic family of type A and, by Corollary I.3, so is Hg () = H0() + Wg (). We summarize this discussion and a simple consequence of (I.54) in the following corollary. Corollary I.4. The family fHg ()j 2 D(0; 0)g is dilatation analytic, i.e., the map Hg : D(0; 0 ) ! B(D(H0 ); H),  7! Hg () is analytic. Moreover, there exists a constant b  0 such that Hel () (Hel  i) 1  b jj ; (I.55) where Hel () := Hel() Hel , for all  2 D(0; 0 ). II Soft Photon Bound and Existence of a Ground State In this section we derive a new soft photon bound; see Inequalities (II.5){(II.6) and Theorem II.3 below. It is taylored for the minimal coupling model, and we use this bound to prove the existence of a ground state. Theorem II.1. There exists a constant C (N;  E0 )  0 such that, for all 0 < g = ( K )3=2  C (N;  E0), the Hamiltonian Hg has a ground state, i.e., E0 ( ) := inf  (Hg ) is an eigenvalue.  T Proof: We introduce the notation a# (F~ ) := a# (F1 ); a# (F2 ); a# (F3) , R R P P a (F ) := =1;2 d3 k F (~k; )a (~k), and a(F ) := =1;2 d3 k F (~k; ) a (~k). Here, F is a function on R 3  Z2 with values in the operators on Hel such 2 R that P=1;2 d3k F (~k; )( x + 1) 1=2 < 1. We further denote (F ) := BFS-3, November 20, 1998 16 ~ ~x . In this notation, the a (F ) + a(F ), (F~ ) := a  (F~ ) + a(F~ ), and ~pj := ir interaction Wg reads j Wg = N X  2g(G~ ~x )  p~j + g (G~ ~x ) + ~j  g(B~ ~x ) ; 2 j j =1 2 j (II.1) j where G~ and B~ are de ned in Eqns. (I.32), (I.33), and (I.39). Next, we introduce an infrared regularization by switching o the interaction for photons of small momenta. Speci cally, we pick a \photon mass", m > 0, that is, we replace G~ ~x(~k; ), B~ ~x(~k; ), and Wg in (I.32), (I.33), and (II.1) by Wg m := ( ) PN j =1 G~ ~xm (~k; ) := [!(k)  m] G~ ~x(~k; ) ; (II.2) m B~ (~k; ) := [!(k)  m] B~ ~x(~k; ) ; (II.3)  ~x  2g (G~ ~xm )  p~j + g (G~ ~xm ) + g ~j  (B~ ~xm ) ;(II.4) ( ) ( ) ( ) ( 2 j ) 2 ( j ) j and we denote Hgm := H + Wg m and E m (g) := inf (Hgm ). We remark that Hgm ! Hg = Hg in norm resolvent sense, as m ! 0. This easily follows from an estimate similar to Lemma I.1 (see [5]). In Theorem II.2 below we show that, for g = ( K ) = suciently small, Hgm has a ground state, m , i.e., there exists a normalized solution of Hgm m = E m (g)m, for all m > 0. Since kmk = 1, the family fmgm> contains a weakly convergent subsequence, fm n gn2N , where limn!1 m(n) = 0. We put  := w limn!1 m n . Then one easily shows [5], that  2 D(Hg ) and that Hg  = E ( ) . To conclude, it remains to show that  6= 0. To show that  6= 0, we employ a soft photon bound, as in [5]. There is an important di erence, though. The soft photon boundPin [5]Restimated the photon number expectation hmjNf m i, where Nf :=  ; a (~k)a(~k)d k, in terms of sup~x k! G~ ~xk . It was derived from a virial type argument, using the commutator of a (~k) and Hg . This bound does not directly apply to the present problem because k! G~ ~xk = 1, for all ~x. Modifying the argument slightly by using the commutator of a (~k) iFx(~k; ) and Hg , for a suitably chosen Fx, we avoid the appearance of sup~x k! G~ ~xk on the right side of the estimate, which we trade for a factor of k jxjm k. More precisely, in Theorem II.3 below we show that ( ( ) ) ( 0 ( 0 ) ) ( ) ( ) (0) 3 2 ( ) ( 0 ) 0 0 ( ) ( ) 0 0 0 0 0 0 3 =1 2 1 2 1 2 1 2 2 hmj Nf mi  C (N ) g (1 + jxj) m ; 1 2 (II.5) BFS-3, November 20, 1998 17 for some constant C1(N )  0. In [5] we showed that m is exponentially localized in the electron variables. More generally, there exists an " > 0 such that e"jxj (Hg(m) )  C2 < 1 ; (II.6) where (Hg(m) ) is the spectral projection onto  := ( 1 ; ( + E0 )=2), provided that g2(1+ E0 ) is suciently small (with C2 and " independent of g). Since m 2 Ranf(Hg(m) )g, this implies the boundedness of k jxj mk. Thus there exists a constant, C3  C3(N;  E0 ), such that hmj Nf mi  C3 g2 ; (II.7) for all m > 0. Next, we introduce the projection Pel onto all bound states of Hel below 21 (+E0) < . Note that Pel has nite rank and that ( E0 )Pel?  2(H0 E0 ). The latter implies that hmj Pel? mi  2( E0 ) 1 hmj (E0(m) (g) E0 Wg(m) ) mi  C3 g 2 : (II.8) From (II.7){(II.8) and Pel P  1 Pel? Nf we draw the important consequence that hmj Pel P m i  1 2C3 g2 ; (II.9) where P := j ih j denotes the rank-one projection onto the photon vacuum vector . Thus, if g = ( K )3=2 is suciently small then hmj Pel P mi  1=2, for all m > 0. Since Pel P has nite rank, it follows that 0 6= 0. In the following theorem, we review the proof in [5] of the existence of a ground state of Hg(m) , for m > 0. Theorem II.2. There exists a constant C  C (N;  E0 ) > 0 such that, for all 0 < g = ( K )3=2  C , the Hamiltonian Hg(m) has a ground state, i.e., E0(m) (g) := inf (Hg(m) ) is an eigenvalue, for any m > 0. Proof: We only sketch the argument, see [5, Section II.2] for details. Alternatively, one may proceed as in [16]. The assertion is proven if we can nd some m~ > 0 such that the sum of the negative eigenvalues of Hg(m) n o E0(m) (g) m~ is nite, i.e., Tr [Hg(m) E0(m) (g) m~ ] > 1, where the negative part of a real number  is de ned as [] := minf; 0g. To this BFS-3, November 20, 1998 18 end, we rst employ a discretization. Given " > 0 and a locally integrable function F , we de ne its "-average by hF (~k)i" := " 3 Z ~n(~k)+Q" F (~k0; )d3k0 ; (II.10) where Q" = [ "=2; "=2)3 and ~n(~k) 2 ("Z)3 is the integer part of ~k, i.e., ~k ~n(~k) 2 Q". We de ne the corresponding interaction, Wg(m;"), by replacing     ~G~x(m) (~k; ) and B~ ~x(m) (~k; ) in (II.2){(II.4) by G~ ~x(m) (~k; ) and B~ ~x(m) (~k; ) , " " respectively. Then, by the bound (I.46) it follows that (Wg(m;") Wg(m) ) (II.11)    o("0) (Hg(m) E0(m) (g) + 1)1=2 1 + jxj (Hg(m) E0(m) (g) + 1)1=2 ; where o("0) denotes a function which possibly depends on g, N , , m and tends to zero as " ! 0. Here, our original manuscript contained a small mistake in that the factor 1 + jxj was missing, as was kindly pointed out to us by F. Hiroshima. Next, we de ne Hf(m;") by replacing !(~k) in Hf(m) by h!(~k)i". Since j!(~k) h!(~k)i"j = O(") and !(k)  m, we obtain that  1  o("0) H (m) f  H (m;") f   1 + o("0)  Hf(m) : (II.12) Denoting Hg(m;") := Hel + Hf(m;") + Wg(m;") , we hence obtain that Hg(m)    1 o("0) Hg(m;") (II.13) o("0) (Hg(m) E0(m) (g) + 1)1=2 (1 + jxj) (Hg(m) E0(m) (g) + 1)1=2 : Next, we introduce the interval e := ( 1 ; E0(m) (g) + m~ ) and observe that   [Hg(m) E0(m) (g) m~ ] = e (Hg(m) ) Hg(m) E0(m) (g) m~ e (Hg(m) ) : (II.14) Furthermore, we note that, thanks to (II.6) and e = e , for m~ > 0 suciently small, we have e (Hg(m) ) (Hg(m) E0(m) (g) + 1)1=2 (II.15)   1 + jxj (Hg(m) E0(m) (g) + 1)1=2 e (Hg(m) ) < 1 : BFS-3, November 20, 1998 Thus, n 19 o Tr [Hg(m) E0(m) (g) m~ ] n   o  Tr e (Hg(m) ) Hg(m;") E0(m;")(g) m~ o("0) e (Hg(m)) n h i o  Tr e (Hg(m) ) Hg(m;") E0(m;") (g) m~ o("0) e (Hg(m) ) nh i o  Tr Hg(m;") E0(m;") (g) m= ~ 2 ; (II.16) for " > 0 suciently small, and the niteness of the right side in (II.16) (and hence the claim) follows if we can show that, for any " > 0, the discretized Hamiltonian Hg(m;") has only nitely many eigenvalues below E0(m;") (g)+ m= ~ 2. The key point of the discretization by means of the "-average is the tensor ? ], where Hdisc is spanned product representation F  = F [Hdisc] F [Hdisc by Q"+~n, ~n 2 ("Z)3 (see [5] for details). Note that, with respect to this representation, we have Hf(m;")  = Wg(m;") 1, = Hf(m;") 1 + 1 Hf(m;") , Wg(m;")  and hence (II.17) Hg(m;")  = Hg(m;") 1 + 1 Hf(m;")    E0(m;") (g) + m P ? + Hg(m;") P ; ? ] by P . Hence, for denoting the projection onto the vacuum in F [Hdisc m~  m,   Hg(m;") E0(m;") (g) m= ~ 2  Hg(m;") E0(m;") (g) m= ~ 2 P : (II.18) Next, for suciently small " > 0, an estimate similar to Lemma I.1 together with an interpolation argument implies that, on Hel F [Hdisc], we have   Wg(m;")  C10 g Hel + Hf(m;") + C (N ) ; (II.19) for some constants C10 ; C (N )  0. Thus, we obtain that Hg(m;") E0(m;") (g) (II.20)  (1 C10 g)Hel (1 + C10 g)E0 + (1 C10 g)Hf(m;") 2C10 C (N ) n o  21 ( E0 ) 2C10 g(1 + 2jE0j) o("0) Pel? n o + (1 C10 g)Hf(m;") 2C10 g(1 + 2jE0j) o("0) Pel ; where Pel is the nite dimensional projection onto the bound states of Hel of energy < 21 ( + E0 ). Now, if g is suciently small such that m^ := 21 ( E0 ) 2C10 g(1 + 2jE0j) > 0 (II.21) BFS-3, November 20, 1998 20 then, for any 0 < m~ < minfm; m^ g, we obtain from inserting (II.20) into (II.18) that Hg(m) E0(m) (g) m= ~ 2 n  (1 o("0) 2C10 g)Hf(m;") (II.22) P ; o 2C10 g(1 + 2jE0j) o("0) Pel for " > 0 suciently small. The right side, however, has clearly only nitely many negative eigenvalues, for any " > 0, which proves that, for any " > 0, the discretized Hamiltonian Hg(m;") has only nitely many eigenvalues below E0(m;") (g) + m= ~ 2. TheoremPII.3. RLet m be a ground state of Hg(m) , for m > 0, and denote by Nf := =1;2 a(~k)a (~k)d3 k the photon number operator. Assume that (K + 1)  1. Then there exist constants C (N ); C1 (N )  0 such that (Z   1 2 ~ ~ ~ h m j N f  m i  sup Tr33 jr~xG~x(k; )j 1 + ~ 2 d3k ~x; !(k)  3 ) Z Z  2 + jG~ ~x(~k; )j2d3k + jB~ ~x(~k; )j2 + j~xG~ ~x(~k; )j2 d~k 2 (1 + jxj)m ! (k ) 2 (II.23)  C1(N ) g2 (1 + jxj) m : C (N ) g2  Before giving the proof of Theorem II.3, we remark that using the de nitions (I.32), (I.33), and (I.39) of G~ and B~ , one easily checks the integrals on the right side of Eqn. (II.23) to be bounded by a constant, uniformly in K  1. Proof of Theorem II.3: Throughout the proof we omit the superscript \(m)". To prove the asserted bound, we rst observe the following commutation relations, a (~k) Hg =   Hg + !(~k) a(~k) + N X ( j =1 (II.24) )  ~j  gB~ ~x (~k; )  2gG~ ~x (~k; )  ~pj g(G~ ~x ) ; j j j F~x (~k; ) Hg = Hg F~x (~k; ) (II.25)   ~ ~x F~x (~k; )  p~j g(G~ ~x ) ; + ~x F~x (~k; ) + 2i r j j j j j j j 21 BFS-3, November 20, 1998 on C0(R 3N ; F ) \D(H0), for any F 2 C 2(R 3 ; L2(R 3  Z2)). We therefore have N a (~k) i F~x (~k; ) Hg E0 ( ) (II.26) ! X j j =1 =  E0 ( Hg +  ) + !(~k)   a (~k) E0 ( i Hg  ) N( X j =1 N X j =1 F~xj (~k; ) ! ~j  g B~ ~xj (~k; ) + ~xj F~xj (~k; )     N + R ! (k )    X 2  j =1 r~ ~xj F~xj (~k; ) ! X ( j ~j  g B~ ~xj (~k; ) + ~xj F~xj (~k; ) j =1  ~ ~xj ) g (G ~pj ~ ~xj (~k; ) gG ) 2 p~j  ; We apply (II.26) to (jxj=R)m, where  2 C01(R ; [0; 1]) and  1 on [ 1; 1]. Letting R ! 1 and using (Hg E0 ( )) m = 0, we derive that N F~x (~k; ) m (II.27) a (~k) m = iR ! (k) Hg E0 ( )  ~ ~xj ) g (G )   gG~ ~x (~k; ) r~ ~x F~x (~k; ) m ; j j j where R(!) := Hg E0 ( ) + ! 1 . We choose F~x(~k; ) := g ~x  G~ ~x(~k; ). Then (II.27) reads N ~ ~xj  G~ ~x (~k; ) m (II.28) a (k) m = igR ! (k) Hg E0 ( )     + R !(k)   N j =1 (  X j =1 ! X j ~j  g B~ ~xj (~k; ) + ig~xj  ~xj G~ ~xj (~k; )  + 2g (p~j  ) ~ ~xj G ~ ~xj (~k; )  ~xj m : g (G~ ~xj ))  r  We observe that R !(k) Hg E0( )  1 and that Nj=1 ~xj  G~ ~x  1=2 jxj Nj=1 jG~ ~x j2 , denoting jxj2 := Nj=1(~xj )2 . Hence, for  2 f1; 2g,    P j   P P j Z d3 k R ! (k) Hg   E0 ( ) N X j =1  N sup ~x ~ ~xj (~k; ) m ~xj  G ( Z jG~ ~x(~k; )j2d3 k (II.29) =2 )1 jxj m : BFS-3, November 20, 1998 22   Additionally using k R !(k) k  !(~k) 1, we similarly obtain Z  d k R ! (k) 3  N sup N X  ~ ~xj (~k; ) m ~j  B~ ~xj (~k; ) + i~xj  ~xj G j =1 (Z  jB~ ~x(~k; )j + j~xG~ ~x(~k; )j ~x 2 2  (II.30) ) d3 k 1=2 (1 + jxj) m : ! (~k)2 Furthermore, we note that   R ! (k) (Hg  R ! (k) N X j =1 2 g (G~ ~xj ) p~j   E0 ( ) + 1)  1 + !(~k) 1 ; (II.31)  C (N ) ; (II.32) for some constant C (N )  0 which only depends on N . These bounds yield the following estimate, Z  d k R ! (k ) 3 N  X j =1  N sup ~x (Z (p~j  )  ~ ~xj G ~ ~xj (~k; )  ~xj m g (G~ ~xj ))  r  ~ ~xG~ ~x (~k; )j2 Tr33 jr  1 + !(~k) 2  d3 k (II.33) )1=2 jxj m : Thus we arrive at the assertion. III Resonances and Time-Decay Estimates In the present section, we study spectral properties of the Hamiltonian Hg . We also study the propagator exp[ itHg ] applied to states whose spectral support is localized about the excited atomic energy level Ej , j  1. As we describe in the introduction, our main tool for this analysis is the complex dilatation, Hg () = U Hg U 1 , of the Hamiltonian Hg , where U is the dilatation de ned in (I.18). We prove below that, for  = i#, 0 < #, and g  #, a complex neighbourhood of an interval Ij about Ej does not contain any spectrum of Hg (). By the dilatation analyticity of Hg () in , this implies that the spectrum of Hg in Ij is purely absolutely continuous,nand it allows o for an estimate of the time decay rate of certain states in Ran I (Hg ) . j BFS-3, November 20, 1998 23 To state this estimate more precisely, we recall from (I.32){(I.38) that Hg () = H0 () + Wg () ; (III.1)  H0 () = Hel () + e Hf ; (III.2) 2 X Wg () = m+n=1 g m+n Wm;n () + g 2 Cno ; (III.3) R where g := ( K )3=2 , and Cno := 2N jG~ ~x(~k; 1)j2 d3k is an energy shift resulting from normal-ordering Wg . We absorb this constant by rede ning Wg () ! Wg () + g 2 Cno ; Hel () ! Hel () g 2 Cno ; (III.4) and since it only shifts all energies by Cno, we henceforth ignore the constant Cno by setting it equal to zero. Thus we obtain Hg () := U Hg U 1 = H0 () + Wg () = H0 () + X m+n2 g m+n Wm;n () : (III.5) Recall that we assumed in Eqn. (I.14) that the j th atomic energy level is of nite degeneracy nj and isolated from the rest of the spectrum of Hel by a positive distance n o  := dist Ej ;  (Hel ) n fEj g > 0 : (III.6) As in [5, Eqns. (IV.84), (IV.85)], we de ne two nj  nj matrices by := j Z d ( ) j X Z ? U Pel;j w0;1 (~k; )Pel;j (III.7) =1;2   1 ? w (~k; )P U 1 d3 k ; ~ Hel Ej + ! (k) i0 Pel;j 1;0 el;j  Z od () := X Z d3 k U Pel;j w0;1 (~k; ) Pel;j w1;0 (~k; )Pel;j U 1 ~ ; (III.8) ! (k) =1;2 Pnj `=1 j'j;`ih'j;`j where Pel;j := is the projection onto the eigenspace of Hel corresponding to the eigenvalue Ej . Note that the matrices Zjod () and Zjd() are similar to Zjod(0) and Zjd (0), respectively, for all  2 D(0; 0). We remark that Zjd(0) and ImfZjod(0)g  (Zjod (0) Zjod(0))=2i are self-adjoint. We require ImfZjod()g to be invertible and to obey 0 < c  j n o := min (ImfZjod (0)g)  c0 ; (III.9) BFS-3, November 20, 1998 24 for some constants 0 < c  c0. Given a small constant " > 0, and a large constant C > 0, we de ne the following union of \comet-shaped" sets Rj  Rj ("; C )  C in the lower halfplane, Rj := Ej + g2(Zjd()) ig 2 j +e  + D(0; Cg2+") ; R+ (III.10) Furthermore, for any 0 <  < , we de ne the interval Ij () := ( Ej  ; Ej +  ) ; and we remark that distfIj (=2) ; (Hel ) n fEj g g  =2. (III.11) Now we are ready to formulate the rst main spectral result of this section Theorem III.1. Let 0 < " < 1=3. For  = i# and # > 0 and g > 0 suciently small, there exist constants c; C > 0 such that  Aj := Ij (=2) + i[  where  that Hg ( )  cg 2 " ;   1) n Rj ("; C )    Hg ( ) ; (III.12) is the resolvent set of Hg (), and for any z 2 Aj , we have  Hg ( ) z  1  n C 0 dist z ; Rj o 1 (III.13) ; for some C 0 > 0. Theorem III.1 has two important consequences. The rst is immediate from the analytic continuation in , and we state it as a corollary. Corollary III.2. For g > 0 suciently small, the spectrum of Hg in Ij (=2) is absolutely continuous, Ij (=2)  ac (Hg ). III.1 Time-Decay Estimates To formulate the second main result of this section, of  another consequence  1 Theorem III.1, we pick a smooth function F 2 C0 [0; 1=2); [0; 1] such that F  1 on [0; 1=4], and we de ne Fj 2 C01 Ij (=2) ; [0; 1] , with Fj  1 on Ij (=4), by Fj () := F ( 1j Ej j). Furthermore, for  2 H, we set j  j ( ) := Pel;j exp h g 2H i  : (III.14) f Theorem III.3. Let  2 H be normalized, dilatation analytic in D(0; 0). Denote  () := U  , and assume that B := 1 + supjj k ()k < 1. 0 Moreover, assume that g > 0 is suciently small, t > 1, and 0 < " < 1=3. BFS-3, November 20, 1998 25 Then there exist a constant C  0 and, for any N 2 N , a constant CN  0 such that   exp[ itHg ] Fj (Hg ) j  j    B C ln(1=g) exp t g 2 2 j Cg  2+" ) + CN t N g  4 : (III.15) Proof of Theorem III.3, given Theorem III.1: We rst use the fact that Fj () =  1 Z =2 F 0 (!= ) Ij (!) () d! =4 (III.16) to rewrite the matrix element on the left side of (III.15) as  j exp[ itHg ] Fj (Hg ) = =   1 Z =2 =4 Z =2 =4 j F 0 (!= )  F 0 (!= ) Z e   Ij ( ! ) exp[ itHg ] I j ! (Hg ) j( ) it D  Hg ()  j ( )  Im  j (III.17) d!  1 j ( ) E d d! ; where we use Stone's formula (see, e.g., [26]) together with Theorem III.1, which implies that the limit lim"& (Hg ()  i") = (Hg () ) exists and is bounded. Indeed, Theorem III.1 even implies that  is 7! (Hg ()  + is) is bounded analytic, provided s 2 R is not too large. We exploit this fact by deforming the integration contour Ij (!)  C into the lower half-plane. To this end, we de ne a number 1 0 1 1 S := g 2 (III.18) CS g 2+" ; j where CS > 0 is later chosen suciently large, and we assume that  2 Ij (=2) and s 2 [ S; g ]. Then, Theorem III.1 yields that 2  Hg ()  is  1  C0  n dist  + is ; Rj o 1 ; (III.19) for g = ( K ) = suciently small and some C 0  0, where Rj is de ned in (III.10) above. Since Zj () is similar to a self-adjoint nj  nj -matrix, we may write 3 2 d Rj := [ 2 e  (Zjd ( )) R j ( e) ; (III.20) BFS-3, November 20, 1998 26 with Rj (e) := Ej + g e ig 2 2 j + e  R + D(0; Cg ") : (III.21) 2+ + Hence  Hg ()  is   1  2 o  1 n 2 min dist  + is ; Rj (e) e  (Zjd ( )) n  2 e2max dist  Ej eg + iCS g " ; e Z  2 2+ i# ( jd ( )) R+ o Cg (III.22) 2+"  1 ; where Cg " takes the disc D(0; Cg ") in the de nition of Rj (e) into account. Thus, choosing CS in (III.18) larger than 4C , we obtain that 2+ 2+   1 Hg ()  is ( Ej eg ) + (CS g "=4)  2# e2max Z   2 2 2+ 2 ( jd ( ))  1=2 : (III.23) Therefore, z 7! (Hg () z) is analytic in the rectangular domain Ij (=2)+ i[ S; g ], and by Cauchy's integral formula, Eqn. (III.17) can be written as 1 2  j exp[ itHg ] Fj (Hg ) where A := Z =2 =4 D Z =2 =4 D j = A A + Ak ; (III.24) F 0 (!=) S exp[ it(E  ! is)] j 4i    Hg () Ej  ! + is j ( ) j ( ) Z (III.25) 0   1 E Hg () Ej  ! + is j ( ) F 0 (!=) Z exp[ it( iS )] 4i   Ij !   E  Hg ()  + iS j ( ) j ( ) D Ak :=    1  j ( ) E ds d! ; (III.26) ( ) 1 D Hg ()  + iS j ( )   1  j ( ) E d d! : We rst estimate Ak. Since Pel;j and e Hf =g2 are dilatation analytic and k()k  B , we have for all jj   that k j ()kk j ()k  C B , for some constant C depending on  . Inserting this and (III.23) into the de nition of 0 0 2 BFS-3, November 20, 1998 27 Ak , we obtain that 0 = ) jAkj = = F 2(!=  I ! exp[ it( iS )]   + iS j ( ) Hg ( ) j ( ) Z Z 2 4 j( ) D  Hg ()  + iS j ( ) D  C B e 2 tS 2  1  max Ij (=2) h E  j ( )  1 d d!  Z e  (Zjd ( )) E (Ej + g e ) + (CS g =4) 2 2+" 2 2  1=2 d  i  C 0 B ln(1=g) exp t(g j CS g ") ; (III.27) for some constants C; C 0  0 which depend on  , , and nj . Note that 2 2 2+ 0 Hg ()  + iS is bounded invertible because the spectrum of Hg () = Hg () lies in the upper half plane. Second, we estimate A . We omit the similar estimate on A . We use the fact that, for any N 2 N and t > 0, + h i exp it(Ej + ! is) = ( it) Thus, an integration by parts yields N dN exp it(E + ! is) : (III.28) j d!N  i h i e it E ! is (III.29) A = 2i tN = dN F 0(!=)  Hg () Ej ! + is j ( ) j ( ) d!N  Hg () Ej ! + is ds d! j ( ) j ( ) i N S = e it E ! is N N !  N k F N k (!=) = 2i tN (N k)! = k k  Ej ! + is j ( ) Hg ( ) j ( ) Z S Z =2 N + ( j+ 4 D 0  )   1  D Z Z 0 D ( j+ ) 4 X =0  +  Hg () Ej ! + is j ( ) D E  1 2  ( +1) E 1  k E 1 E  j ( ) ds d! : Since all derivatives of F are bounded and S  g j , there exists a constant CN  0 such that  D   E jA j  CNNg sup j () Hg () Ej ! + is k j () 2 2 + t 1 D Hg () Ej ! + is j ( )   k 1  j ( ) E 0  k  N ;   !   ; 0  s  S : (III.30) 4 2  BFS-3, November 20, 1998 28 Thus, Theorem III.3 follows if we can nd a constant C > 0 such that D j () Rg ( )k+1 j ( ) E D j () Rg ()k+1 j () E  C B2 g 2 ; (III.31) for all k 2 f0; 1; : : : ; N g, ! 2 [=4; =2], and s 2 [0; S ], where we denote Rg (  )  := H g ( ) Ej ! + is  1  1 : (III.32) ; (III.33) To this end we introduce an unperturbed resolvent, Q0 ( ) :=  H0 ( ) Ej ig 2 ! and we observe that, for g=jj suciently small, j j 1; Wg ( ) Q0 ( )  Cgjj 1 ; Q 0 ( ) Rg ( ) ; Wg ( ) Rg ( ) ;  (III.34) (III.35) C  and some constant C > 0, which is uniform in ! 2 [=4; =2] and s 2 [0; S ]. Using the second resolvent equation, we obtain that Rg ( ) = i h Q0 ( ) Wg ( ) + is + ig 2 Q0 ( ) (III.36) + Q0 () Wg () + is + ig2 Rg () Wg () + is + ig2 Q0 () : Q 0 ( ) h i h i We expand Rg ()k+1 by means of (III.36), Rg ( )k+1 = Q0 ( )k+1 kX +1  =1 Q0 ( ) Wg ( ) Q0 ( )k+2  + Rem ; (III.37) and (III.34){(III.35) show that there is a constant C  0, depending on k and , such that k Rem k  C g2 : (III.38) Similarly, we nd that Rg ()k+1 = Q0 ( )k+1 kX +1  =1 Q0 ( ) Wg () Q0 ( )k+2  k Rem0 k  C g2 + Rem0 ; (III.39) (III.40) BFS-3, November 20, 1998 29 (note that Q0() 6= Q0 ()). Inserting the two identities (III.37) and (III.39) into (III.31) and using (III.38) and (III.40), we observe that it suces to prove that there is a constant C  0 such that D Im j () Q0 ( )k+1 j ( ) E  k+2  () j () Q0 () Wg () Q0 () j D E  C B2 g 2 ; (III.41)  C B2 g 2 ; (III.42) for all k 2 f0; 1; : : : ; N g,  2 f1; 2; : : : ; k + 1g, and ! 2 [=4; =2]. We remark that (III.41) and (III.42) hold trivially for j = 'j , where 'j is an normalized eigenvector of Hel corresponding to the eigenvalue Ej . To prove (III.42), we observe that, for any two vectors 2 D(Hfm=2), n=2 ' 2 D (Hf ), we have D E Pel;j ( ) Wg ( ) Pel;j ( ) ' 2 X  m+n=1 g m+n C  C0 (III.43) Pel;j ( ) wm;n ( (m) ; ~(n) ;  ) Pel;j ( ) ka( (m) ) k ka(~(n))'kd (m) d~(n) 2 X  Z m+n=1 Z g m+n 2 X ( m+n)=2 D !( ! (k ) 1g m+n=1 J (k )2 d3 k )m+n Hfm E D ' Hfn ' E n=2 ' ; m=2 Hf Hf P2 P2 by Schwarz' inequality. Here we abbreviate the summation     =1 m =1 1 R d3 k1    d3 km by d (m) , and  (m) := (k1 ; 1 ; : : : ; km ; m ). Thus we have  k+2  () (III.44) j j () Q0 () Wg () Q0 ()  D E k+2  () = Q0 ( ) j () Pel;j () Wg () Pel;j () Q0 () j     n= 2 m= 2 k +2  m +n   g Hf Q0 ( )  C mmax j () ; j () Hf Q0 () +n=1;2 E D Next, we observe that  Q0 ( )   j () =  e H f ! + ig 2   exp h i  g 2 e  Hf Pel;j ()  () ; (III.45) BFS-3, November 20, 1998 30 and hence, for suitable constants C; C 0; C 00  0, m=2 Q ()  () Hf 0 j  C B sup  C0 B 1 r0 #  r m=2 ei# r sup r0  r m=2 ! + ig 2 exp h   exp g e# r 2 i h (III.46) g 2 ei# r  i  C 00 B # 1 g 2 e# m=2  : Inserting this and a similar estimate for k Hfn=2 Q0()k+2  j ()k into (III.44), we obtain that o n D E 2 m+n  W () Q ()k+2  () 2 2  ( Cg ) (  ) Q (  )  j  j B max j 0 g 0 j  m+n=1;2 = C jj 2 B2 g2 ; (III.47) for some constant C  0 and g > 0 suciently small. This proves (III.42). Finally, we establish (III.41) by using the fact that we may analytically continue in , since the spectral parameter Ej + ! + ig2 in Q0 () is in the upper half-plane. Thus E E D D  2 k 1 k+1 () = j () j j () (e Hf ! ig ) j () Q0 () D E = j (0) (Hf ! ig2) k 1 j (0) (III.48) E D = Pel;j  e 2Hf =g2 (Hf ! ig2) k 1 Pel;j  : Therefore,  o  n E D 2r=g2 2 k 1 k +1  Im (r ! ig ) : Im j () Q0 ()  sup e j ( ) r0 (III.49) Now, we use that !  =4. If r  !=2 then ! r  =8 and thus 8g2 : r ! ig 2 )j  (III.50) j arg(r ! ig2)j  jjIm( Re(r ! ig2)j  Hence, for r  !=2, n o  k+2  exp[ 2g 2r] Im (r ! ig2) k 1  (k + 1) 8= g2 : (III.51) We point out that only for the derivation of (III.51) we need to estimate the imaginary part of a matrix element rather than its magnitude. It remains to consider the case r  !=2  =8. We estimate as follows, o n exp[ 2g 2r] Im (r ! ig2) k 1  g 2k 2 exp[ g 2=8]  C g2 ; (III.52) BFS-3, November 20, 1998 31 for some constant C  0. Inserting (III.51), (III.52) into (III.49), we obtain (III.41) which, together with (III.42), nishes the proof. III.2 Resolvent Norm Estimates and the Proof of Theorem III.1 The purpose of this subsection is to prove Theorem III.1, i.e., that Hg () z is invertible for z 2 Aj , where Aj is de ned in (III.12), and its inverse obeys the norm estimate  Hg ( ) z  n  C dist 1 z; Rj o ; (III.53) for some C > 0 and with Rj as de ned in (III.10). Before we turn to the actual proof, we outline our strategy. We distinguish the spectral parameters z close to the eigenvalue Ej of Hel ( ) from those which are away from Ej by a certain minimal distance,  =2, where 0 0 := 2" g2 (III.54) ; and 0 < " < 1=3 is arbitrary but xed. Note that, given any ; #; c > 0, we have 0   sin(#=2) and 0  cg 2 " ; (III.55) provided g  0 is suciently small. For z 2 Aj n D(Ej ;  =2), a Neumann series expansion shows the invertibility of Hg () z and proves (III.53) for that region. This is proven in Lemma III.12 below. For z 2 D(Ej ; =2), the situation is more delicate, and we contruct the inverse of Hg () z by means of the Feshbach map discussed in detail in [5, 6]. For this construction, we specify a partition of unity given by the (non-orthogonal) projections 0 P ( ) := Pel;j ( ) Hf <0 and P ( ) := 1 P ( ) ; (III.56) where Pel;j () := U Pel;j U and Pel;j is the (orthogonal) projection onto the eigenspace of Hel corresponding to the eigenvalue Ej . In Lemma III.9 below we prove that 1 Hg ( )P () z is invertible on RanfP ()g ; (III.57) BFS-3, November 20, 1998 32 where we denote AP := PAP . This property and some further relative bounds of more technical nature stated in Lemma III.10, choosing := 2"(1 ") 2 (0; 1), insure the existence of the Feshbach operator de ned by 1  FP  := FP  Hg () z ( )  ( ) := Hg ()P   P ()Wg P () Hg ()P  z P ( ) ( ) ( )  1  1 z (III.58) P ()Wg P () : Moreover, the following identity holds,  Hg () z  1 F   P () Hg ()P  = P () 1 P ( )  ( )  z P ()Wg P () Hg ()P  P ( )  z + P () Hg ()P  ( )  ( ) 1  Wg P () z  P ( ) : 1  (III.59) Thus using bounds collected in Lemma III.10 below, we obtain that, for z 2 D(Ej ;  =2), 0  Hg () z    = 1 + O(g # ) FP  + O(1) : 1 0 1 1 1 ( ) (III.60) Next, a careful analysis of the Feshbach operator in Lemma III.11 yields that FP  ( ) (Ej + g Zj () + g Zj () + e  Hf )P ()  O(g ") ; (III.61) 2 d 2 od 2+ for any 0 < " < 1=3, which immediately implies (III.53). III.2.1 Estimates on the Dilated Atomic Hamiltonian In this subsection, we start to implement the strategy of the proof of Theorem III.1 outlined above. To begin with, we recall some de nitions and notation. The pure point spectrum of Hel is given by the set fE ; E ; : : : ; Ej ; : : : g contained in ( 1; ), and its essential spectrum is contained in [; 1). We  assume that   0 and, we denote Ri := dist Ei ; (Hel) nfEig > 0. Fixing j  1 { and we shall keep j  1 xed henceforth { we set  := Rj . So denoting Hel() := U HelU , we can construct the projection Pel;i() onto the eigenspace of Hel() corresponding to the eigenvalue Ei by using the Dunford integral, 0 1 1 Z i Pel;i() = 2 jz j Ei =Ri =2 dz ; Hel() z (III.62) BFS-3, November 20, 1998 33 Next, we de ne a nite-rank projection Pdisc() by X Pdisc ( ) := Pel;i ( ) ; (III.63) i:Ei   where  is some xed, strictly positive number. Note that k Pdisc() Pdisc(0) k  C b jj ; thanks to the relative bound (I.55),  Hel () Hel + i 1  (III.64)  b j j ; (III.65) where Hel () := Hel () Hel. Our rst result is the following bound.  Lemma III.4. Let z 2 C with Refz g <  . Then, for j j 1 + (   Refzg) 1 suciently small, Hel () z is invertible on RanfP disc()g and (III.66) (Hel () z) 1 P disc()    2 Refzg : Proof: We rst observe that Q := P disc(0)Hel (0) z is globally invertible on Hel , and since   0 we have o   n kQ 1k  max j  zj 1 ; j zj 1    Refzg 1 : (III.67) Similarly, we obtain   r+i r+ +i 1 k (Hel(0) + i) Q k = max sup r +   z ; sup z r0 E0 r     1  C1 1 +   Refzg ; (III.68) for some constant C1  0. Inserting this and (III.64){(III.65), we obtain   P disc (0)Hel (0) P disc ( )Hel ( ) Q 1 (III.69)    jj 1+  C2    Refzg 1 (III.70) ; for some constant C2  0. Thus a Neumann series expansion yields (III.71) (Hel() z) 1 P disc() = (P disc()Hel () z) 1 P disc()  Q 1   1 X n=0 1  (P disc(0)Hel(0) Refzg 1 X n=0 n P disc ( )Hel ( )) Q 1 P disc ( ) jj+ C2  jj C2   Refzg n   2  Refzg ; BFS-3, November 20, 1998  for jj 1 + ( Refzg)  34 1   (2C2) 1. Next, we extend Lemma III.4 to a global bound for the resolvent of Hel. Lemma III.5. Let  > 0. For  = i# and # > 0 suciently small, there exists a constant C  0 such that Hel () Ej + e   is invertible on Hel and (Hel() Ej +e   ) 1 C (#) 1 : (III.72) Proof: We observe that Hel () commutes with Pdisc() = Pi:E  and that i (Hel () Ej +e  ) 1 Pdisc ( ) = X i:Ei   (Ei Ej +e   Pel;i ( ) ) 1 Pel;i ( ) : (III.73) Thus, for some constant C 0  0, (Hel ()  Ej +e max i:E     ) 1 Pdisc ( ) C0 # i Pel;i ( )  (III.74) n # Ei 2 (Hel ) Ei   o  :   1 Using the integral representation Pel;i() = (2i) 1 jz E j=R =2 z Hel () dz together with the relative bound (III.65), we obtain that Pel;i()  1 + O(jj). Conversely, on RanfP disc()g we apply Lemma III.4 and obtain R i (Hel() Ej +e  ) 1 P disc ( ) i    Ej +   Refe  g 1  2 cos(#) 1 : (III.75)     Lemma III.6. Let  = i#, 0 < # < 0 and 0 < 0  (=3) sin #. There exists a constant C  0 such that, for all z 2 D(Ej ; 0 =2) and all r  0,  Hel ( ) Ej + e (0 + r)   P el;j ( ) Hel ( ) z +e  r  C: (III.76) BFS-3, November 20, 1998 35 Using 1 = Pdisc() + P disc() as in the proof of Lemma III.5, we obtain that Proof: X P el;j ( )  X P disc ( )  X P disc ( ) Ei X + i:Ei  Ej Ei 6 ";i=j + t + e  (0 + r ) C sup t +  + e r ( t (III.77) + e  (0 + r) z + e r 2R n( ;  ) ;  Pel;j ( ) 2 D(0; 0=2) ) ; where we denote X :=  Hel ( ) Ej + e  (0 + r)  Hel ( ) z +e  r 1  (III.78) : We observe that minimization with respect to r  0 yields t+ + e r = 1 t + e  (0 + r )  1  1  e t+e 0   (III.79) (0 + r) 30 t + ( + r) cos #2 + ( + r)2 sin2 # 0 0 2  30 (1 cos #)t2 + ( + r)3 1=2  1 ; 0 2 2 1=2 additionally taking into account that 0  (=3) sin #. Hence X P el;j ( )  X P disc ( ) + 2C : (III.80) Next, we write X P disc ( ) = P disc ( ) + (e  0 +z  Ej ) Hel ( ) and we obtain from Lemma III.4 that  30  X P disc ( )  P disc ( ) 1 + 2 and hence we arrive at the claim.  +e   1  Refzg z r  1 P disc ( ) ; (III.81) C; (III.82) BFS-3, November 20, 1998 36 III.2.2 Relative Bounds on the Interaction In this subsection we use the estimates on the dilated electron Hamiltonian derived in the previous subsection to obtain suitable relative bounds on the interaction. To this end, we recall Eqns. (I.37){(I.45), and we introduce the operator B () := H () Ej + e   = Hel() Ej + e  (Hf + ) : (III.83) We shall generally assume that 0 < #   and that 0 <   (=3) sin #. As before, we denote Hg ()P  := P ()Hg ()P () and pH ()P  := H ()P (). Finally, for a closed operator A we denote jAj := AA. We start with a 0 0 0 0 ( ) 0 ( ) preparatory lemma. Lemma III.7. For  = i# and # 2 (0;  ) suciently small there exists a constant C  0 such that 0   C !  # 1+  ;    C# 1 + 1 ; jB ()j (Hf + !) 1 jB ()j (Hel(0) + i) 1 (III.84) (III.85) for all  > 0 and !  0. Proof: By the functional calculus and Lemma III.5 we have jB ()j (Hf + !) = sup     Hel() Ej + e  (r + ) (r + !) r  r + !   C max1 ; ! ; (III.86)  C sup #(r + ) #  r> 1 1 0 0 which implies (III.84). To establish (III.85) we start with a similar observation, namely, that jB ()j (Hel(0) + i) = sup Y 1  r 0 where   Y := Hel () Ej + e  (r + ) 1 Hel ; (III.87) (Hel (0) + i) : (III.88) We observe the following identity, Y = 1 Y (Hel(0) + i) Hel () (III.89)    + i + Ej e  (r + ) Hel () Ej + e  (r + ) : 1 1 BFS-3, November 20, 1998 37 Solving for Y and applying Lemma III.5, we obtain      kX ()k  1 1bjj 1 + C ji + Ej e  j 1 + #1  C 1 + #1 : (III.90) Now, we come to the main relative bound used in this section. Lemma III.8. For ; 1 ; 2 2 fi#g, 0 < < 1, and # 2 (0; 0) suciently small there exists a constant C  0 such that jB ()j 1=2 1 jB ()j Wm;n () jB2 ()j 1=2  P () Wm;n() jB ()j 1=2  P () Wm;n() P ()  1=2 1 Wm;n () P () 2 C C C  C 1+ 1=2 1+ 1=2 1+ 1=2     1+    1=2  (III.91) ; ( + 0 )1=2 n0 =2 ; (III.92) ( + 0 )1=2 m0 =2 ; (III.93) ( + 0 ) (0m+n) =2 ; (III.94) for all  > 0. Proof: We rst observe that a (~k)B () = B ( + ! (k))a(~k) and thus a (~k) jB ()j2 = a (~k) B() B () = B( + !(k)) B ( + !(k)) a(~k) (III.95) = jB ( + !(k))j2 a (~k) : Thus, functional calculus implies the Pull-Through Formulae = jB ( + !(k))j 1=2 a(~k) ; = a (~k) jB ( + !(k))j 1=2 : Using (III.96){(III.97), we observe that, for any 2 H, jB ()j 1 = 1=2 2 Z X =1 = sup =1;2 a (~k) jB ()j 1=2 jB ()j 1=2 a(~k) (III.96) (III.97) W0;1 () jB2 ()j 1=2 (III.98) d3 k jB1 ()j 1=2 w0;1 (~k; ; ) jB2 ( + ! )j 1=2 a (~k) (Z jB ()j 1 1=2 (  2 Z X =1  Hf w0;1 (~k; ; ) jB2 ( + ! )j +!  1=2 a (~k) 2 3 !d k 1=2 )1=2  Hf ; +! 1=2 2 d3 k ! )1=2 BFS-3, November 20, 1998 38 where here and henceforth we denote ! := !(k) and !0 := !(k0). Note that there is an additional constraint !(k)  0 in the integrals on the right side of (III.98) if we require that 2 RanHf <0 . The last factor in (III.98) equals ( 2 X Z   1=2  ~ 1=2 ? P ? Hf a (k )a (~k ) Hf P =1 ! d3 k 1=2 )  k k (III.99) : Since furthermore Hf + ! commutes with and B1 () and w0;1(~k; ; ), we may use (I.42) and Lemma III.7 to estimate  B1 ()j 1=2 w0;1 (~k; ;  ) jB2 ( + ! )j 1=2 Hf j  j  B1 ()j 1=2 Hf +! 1=2   Hel (0) + ij1=2 jB2 ( + ! )j 1=2 Hf 1=2 1 1+  1+ +! J (k ) #  1=2 C 1 1 +  (1 + !)1=2 J (k) ; #  C    ! 1=2  for some constant obtain that C  1=2  +! 1=2  0. Inserting (III.100) and (III.99) into (III.98), we 2 B1 ()j 1=2 W0;1 ( ) jB2 ()j 1=2 (III.101) j  C #2  1+ 1 (III.100) w0;1 (~k; ;  ) jHel (0) + ij 1=2   j +! Z  1 + ! (k )  C (0 +  1 )  1 ; 1 +  #2  1 J (k)2 d3 k  for some constant C  0. Similarly, by additionally requiring that RanHf <0 , we obtain that j 2 2 B1 ()j 1=2 W0;1 ( ) jB2 ()j 1=2 Hf <0    C  #2 1+ C 0 #2 C ( 1  Z !(k)0  1+ 1 + #2  1  Z  1 + ! (k ) ! (k )  1 J (k)2 d3 k + ! (k ) ) 0 1 +   1 : 1  J (k )2 d3 k (III.102) The estimate for W1;0() is similar. Next, we derive (III.91) in the case of BFS-3, November 20, 1998 W0;2 ( ). j Picking B1 ()j  2 H 39 , we observe that W0;2 ( ) jB2 ()j 1=2 (Z 1=2 jB1 ()j w0;2 (~k; ~k 0 ; ; 0 ;  ) (III.103) 1=2 sup ;0 =1;2 0 jB2 ( + ! + ! )j ( Z 2 X  ;0 =1  Hf 1=2 + ! + !0  1  Hf + ! + ! 0  a (~k 0 )a (~k ) 0 2 2 )1=2 d3 k d3 k 0 ! !0 )1=2 ! ! 0 d3 k : Again, we have the additional constraint !(k)  0 in the integrals on the right side of (III.103) if we require that 2 RanHf <0 . The last factor is bounded by k k. Thus Eqn. (I.43) and Lemma III.7 imply that, for some constants C  0, B1 ()j Z j  2 W0;2 ( ) jB2 ()j 1=2 1=2  1=2 Hf + ! + ! 0 jB1 ()j 1=2 (III.104) 2 1=2 2 J (k )2 J (k 0 )2 d3 k d3 k 0  B2 ( + ! + ! 0 )j 1=2 Hf + ! + ! 0 ! !0    Z  ! + !0 ! + !0 J (k )2 J (k 0 )2 d3 k d3 k 0 C  j   #2 C (0 1+ 1+   + ! + !0 +  1)2 1 +  1  ; #2 ! !0 and j B1 ()j 1=2 W0;2 ( ) jB2 ()j  C ( 1=2 Hf <0 + 1 #2 2 ) 20 1 +  (III.105) 1  : Estimates similar to (III.103){(III.104) establish (III.91) in the remaining cases, i.e., for W1;1() and W2;0 (). Finally, we observe that P () = Hf <0 P () and B () P ( ) = (Hf + ) P () = 0 +  ; which together with (III.102) and (III.105) yields (III.92){(III.94). (III.106) BFS-3, November 20, 1998 40 III.2.3 Domain of the Feshbach Map In the following subsection we apply the relative bounds from Lemma III.8 to prove that, for z suciently close to Ej , the Feshbach map with projection P ( ) is applicable to Hg ( ) z . Lemma III.9. Let 0 < (=3) sin #, and assume that # 2 (0; 0 ) is sufciently small. Then, for g0 1=2 > 0 suciently small and for all z 2 D (Ej ; 0 =2), the operator Hg ( )P () z is invertible on RanfP ( )g, and  Hg ( )P () z  1 P ( ) C  # 0 (III.107) ; for some constant C  0. Proof: We construct mann series,  Hg ( )P () z  1  Hg ( )P () P ( ) = z 1 1 P ( ) !" P ( ) X n=0  H0 ( ) z by a norm-convergent NeuWg ( ) ! #n P ( ) H0 ( ) z : (III.108) We estimate the norm of the term in nth order by means of Lemma III.8, !" P ( ) H 0 ( )  B (0 ) 1=2 jB(0)j  C # 0 Wg ( ) z   C g 0 H0 ( ) P ( ) H0 ( ) 1=2  z  jB (0)j  1=2 n ! #n P ( ) (III.109) z B(0 ) 1=2 1=2 n+1  jB(0)j 1=2 j j Wg ( ) B (0 ) 1=2 n : This proves the convergence of the Neumann series (III.108) in norm. Lemma III.9 is the main ingredient used to prove the existence of the Feshbach operator de ned in (III.58){(III.60). Lemma III.10. Let 0 < (=3) sin #, and assume that # 2 (0; 0) and g > 0 are suciently small. Then, for all z 2 D(Ej ; 0 =2), the Feshbach operator de ned in (III.58) exists and obeys Eqn. (III.59). Moreover, for some BFS-3, November 20, 1998 41 constant C  0, we have  P () Wg () P () Hg ()P ()  P () Hg ()P () z  1 z  P ( ) 1 ; P () Wg () P ()  > ; Cg : # 10=2 Cg 10=2 : #  P () Wg () P () 9 > = (III.110) (III.111) Proof: The proof of (III.110) is similar to the one for Lemma III.9. Then Lemma III.9 and (III.108) imply the existence of the Feshbach operator dened in (III.58) and that it obeys Eqn. (III.59) (see, e.g., [5, 6]). We nally establish two lemmata which, inserted into (III.61) and (III.53), prove Theorem III.1. Lemma III.11. Let 0 < " < 1=3,  := g ", 0 < < 1, and assume that # 2 (0;  ) and g > 0 are suciently small and such that  < (=3) sin #. Then, for all z 2 D(Ej ;  =2), 2 0 2 0 0 0 FP  (Ej + g Zj () + g Zj () + e  Hf )P () (III.112)   " "  C g "+g +g +g " ; for some constant C  0. 2 ( ) 2+ d 2+2 (1 2 od ) 1+(1+ )(1 ) 4 2 Proof: Recall from (III.58) and (III.7){(III.8) that FP  := FP  ( )  Hg () ( ) := Hg ()P  z P ( ) ( ) and Z j ( ) od X Z :=  =1;2 Hel Zj () d := z X Z =1;2   P ()Wg P () Hg ()P () z  1 (III.113) P ()Wg P () ? U Pel;j w0;1 (~k; )Pel;j Ej + ! (~k) i0  1 (III.114) ? w (~k; )P U 1 d3 k ; Pel;j 1;0 el;j  d3 k ; U Pel;j w0;1 (~k; ) Pel;j w1;0 (~k; )Pel;j U 1 ! (~k) (III.115) BFS-3, November 20, 1998 42 where Pel;j = Pel;j ( = 0) = Pn` j j'j;`ih'j;`j is the orthogonal projection onto the eigenspace of Hel corresponding to the eigenvalue Ej . As in [5], we write the di erence to be estimated as a sum of six error terms, =1 FP  (Ej + g ( ) 2 Zjd ( ) + g 2 Zjod ( ) + e  Hf )P ( ) 5 X = =0 Rem ; (III.116) where (compare to [5, (IV.58), (IV.60), (IV.68), (IV.77), (IV.101), (IV.86), and (IV.87)]) Rem := P ( )Wg ( )P ( ) 0 " P ( )Hg ( )P ( ) z   P ( )H0 ( )P ( )  P ( )Wg ( )P ( ) P ( )Hg P ( ) Rem := 1   g 2 P ( ) W0;1 ( ) + W1;0 ( ) Rem := g P ( ) W0;1 ( ) + W1;0 ( ) 2 g 2  2 2 Z X P ( )w0;1 (~k; ) =1 " z  P ( ) H0 z z H0 +e P ( )Wg ( )P ( ) ; ! (III.118)  W0;1 ( ) + W1;0 ( ) P ( ) ; ! (k )  W0;1 ( ) + W1;0 ( ) P ( ) # P ( ; ! (k )) i# !  1# P ( )Wg ( )P ( ) 1 P ( ) H0 z  (III.117) 1 z w1;0 (~k; )P ( ) d3k ; (III.119) where P (; !) := Pel;j () Hf !<0 , Rem := P () Wg () P () ; + (III.120) 3 Rem := g 4 2 Z X =1  2 4 2 e i# Hf dk Pel;j w0;1 (~k; ) + B (! (k )) + Ej Rem := 5 g2 h e Z  X =1;2 i# Ej z (III.121)   z P el;j Hf <0 Hel Ej +e i# ! 3  5 w1;0 (~k; ) Pel;j Hf <0 ; (k ) Pel;j w0;1 (~k; ) Pel;j w1;0 (~k; )Pel;j (Hf + !(k)) + Ej z i 1  Hf +!(k)0 dk (III.122)  Zj ( ) Hf <0 : d BFS-3, November 20, 1998 43 We rst rewrite Rem using the second resolvent equation which yields 0  Rem = P ( )Wg ( )P ( ) P ( )Hg ( )P ( ) 0  P ( )H0 ( )P ( ) z  1 z  1 P ( )Wg ( )P ( ) P ( )Wg ( )P ( ) : (III.123) Then an application of Lemma III.8 gives Cg 3 k Rem k  1=2 0 #2 0 = O   = O   (III.124) g 2+" : Second, a similar estimate yields Cg 3 k Rem k  1=2 1 #2 0 (III.125) g 2+" : The derivation of these two estimates, (III.124) and (III.125), is similar to [5, (IV.58){(IV.62)]. Third, we observe that Rem = 2 Z 2 X ;0 =1 dk dk 0 ( (III.126) P ( ) w1;0 (k; ;  ) a (k )a (k 0 ) + P () w ; (k; ; ) " 01 " # P (; ! (k )) H0 + e i# ! (k ) + P (") w ; (k; ; ) a(k0) P (; ! (k ) + ! (k 0 )) H + e i# (! (k ) + ! (k 0 )) "  + P () w (k; ; ) a (k) P () 01 0 1;0 # P (! (k 0 )) H0 + e i# ! (k 0 ) H0 # z # z z w1;0 (k 0 ; 0 ;  ) P ( ) w0;1 (k 0 ; 0 ;  ) a(k )a(k 0 ) P ( ) w1;0 (k 0 ; 0 ;  ) a(k ) P ( ) z w0;1 (k 0 ; 0 ;  ) a(k 0 ) P ( ) )  f ; ( ) + W f ; ( ) + (compare to [5, (IV.66)]), which is of the form P () W  f ; () P (). A somewhat lengthy estimate analogous to [5, Lemma IV.9] W yields, after using (III.92){(III.93), 20 02 11 k Rem k = O 2  g 2+2 (1 ")  (III.127) : Fourth, we apply (III.94) and directly obtain k Rem k = O 3  (1+ )=2 g 0  = O  g 1+(1+ )(1 ")  : (III.128) BFS-3, November 20, 1998 44 In order to estimate Rem4, we observe that when restricted to RanP el;j , the resolvents of B (!(k)) + Ej z and Hel Ej + e i#!(k) are bounded by a constant C  0. Since, furthermore, z Ej 2 D(Ej ; 0 =2) and k Hf Hf <0 k = 0 , the fraction in the integrand on the right side of (III.121) is bounded in norm by 2C0, and we thus obtain     (III.129) k Rem4 k  O g2 0  O g4 2" : Finally, a similar argument, which is along the lines of [5, Lemma IV.12], yields     (III.130) k Rem5 k  O g2 0  O g2+2 (1 ") : Adding up all error terms, taking into account that 0 < < 1, we arrive at (III.112). and assume that # 2 (0; 0 ) and g > 0 are suciently small and such that 0 < (=3) sin # and 0  2# 1 minfImAj g. Then, for all z 2 Aj n D(Ej ; 0 =2), Lemma III.12. Let 0 < " < 1, 0 := g2  Hg () z  1 2" h  C dist z ; Rj i 1 (III.131) ; for some constant C  0. Proof: We rst consider all eigenvalues Ei of Hel() below  . Since the eigenvalue closest to z is Ej , there exists a constant c > 0 such that  h min dist z ; Ei + e #R + E   i i  h i c # dist z ; Ej + e # R + ; (III.132) for any z 2 Aj n D(Ej ; 0=2). Thus, on RanfPdisc()g (see Eqns. (III.62){ (III.62)) we have  h i 1 Pdisc () # dist z ; Ei + e R +  C Emax H0 () z i   h i  C dist z ; Ej + e # R + 1 : (III.133) c# Moreover, since 0  2# 1 minfImAj g, for g > 0 suciently small, we have #=2  arg(z Ej )   + #=2, for any z 2 Aj n D(Ej ; 0 =2). Therefore, we also have the estimate ( ) r + 0 B (0 ) Pdisc()  C sup max  C# 2; i# E    H 0 ( ) z jEi z + e rj r>0 i (III.134) BFS-3, November 20, 1998 45 for some C  0. Second, on RanfP disc()g we employ Lemma III.4 to prove a similar bound, h i 1 P disc ( ) C0 C B (0 )P disc ( )  dist z ; Ej + e # R + ; ;  H0 ( ) z H0 ( ) z # # (III.135) for some constants C; Ch0  0. Summarizing (III.133), (III.134), (III.135), i h i # and observing that dist z ; Ej + e R +  c dist z ; Rj , for some c > 0, we obtain B (0 ) C# 1 1 h i and   C # 1 ; (III.136) H0 ( ) z H (  ) z 0 dist z ; Rj for some constant C  0 and all z 2 Aj n D(Ej ; 0 =2). So, nally, we obtain the invertibility of Hg () z and (III.131) from an expansion in a Neumann series as in (III.108){(III.109). Acknowledgements: We thank T. Chen and A. So er for numerous very helpful discussions on the material presented in this paper and H. Zenk for careful proofreading. We are also grateful to D. Buchholz, F. Hiroshima, F. Klopp, Y. Last, and H. Spohn for valuable comments. References [1] J. Aguilar and J. M. Combes. A class of analytic pertubations for onebody Schrodinger Hamiltonians. Commun. Math. Phys. , 22:269{279, 1971. [2] S. Albeverio. Scattering theory in a model of quantum elds. I. J. Math. Phys., 14(2):1800{1816, 1972. [3] S. Albeverio. Scattering theory in a model of quantum elds. II. Helv. Phys. Acta, 45:303{321, 1972. [4] V. Bach, J. Frohlich, and I. M. Sigal. Mathematical theory of nonrelativistic matter and radiation. Lett. Math. Phys. , 34:183{201, 1995. [5] V. Bach, J. Frohlich, and I. M. Sigal. Quantum electrodynamics of con ned non-relativistic particles. Adv. in Math. , 137:299{395, 1998. BFS-3, November 20, 1998 46 [6] V. Bach, J. Frohlich, and I. M. Sigal. Renormalization group analysis of spectral problems in quantum eld theory. Adv. in Math. , 137:205{298, 1998. [7] V. Bach, J. Frohlich, I. M. Sigal, and A. So er. Positive commutators and spectrum of nonrelativistic QED. Preprint, 1997. [8] E. Balslev and J. M. Combes. Spectral properties of Schrodinger operators with dilatation analytic potentials. Commun. Math. Phys. , 22:280{ 294, 1971. [9] H. Bethe and E. Salpeter. Quantum mechanics of one- and two-electron atoms. In S. Flugge, editor, Handbuch der Physik, XXXV, pages 88{436. Springer, Berlin, 1957. [10] H. A. Bethe. The electromagnetic shift of energy levels. Phys. Rev. , 72:339, 1947. [11] C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg. Photons and Atoms { Introduction to Quantum Electrodynamics. John Wiley, New York, 1991. [12] C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg. Atom-Photon Interaction. John Wiley, New York, 1992. [13] H. Cycon, R. Froese, W. Kirsch, and B. Simon. Schrodinger Operators. Springer, Berlin, Heidelberg, New York, 1 edition, 1987. [14] J. Derezinski and C. Gerard. Asymptotic completeness in quantum eld theory. massive Pauli-Fierz Hamiltonians. Preprint, June 1997. [15] J. Frohlich and P. Pfeifer. Generalized time-energy uncertainty relations and bounds on lifetimes of resonances. Rev. Mod. Phys., 67:795, 1995. [16] Ch. Gerard. Asymptotic completeness for the spin-boson model with a particle number cuto . Rev. Math. Phys., 8:549{589, 1996. [17] F. Hiroshima. Functional integral representation of a model in QED. Rev. Math. Phys., 9(4):489{530, 1997. [18] R. Hoegh-Krohn. Asymptotic elds in some models of quantum eld theory. III. J. Math. Phys., 11(1):185{189, 1969. [19] R. Hoegh-Krohn. Boson elds under a general class of cut-o interactions with bounded interaction densities. Commun. Math. Phys., 12:216{ 225, 1969. BFS-3, November 20, 1998 47 [20] R. Hoegh-Krohn. Boson elds with bounded interaction densities. Commun. Math. Phys., 17:179{193, 1970. [21] W. Hunziker. Distortion analyticity and molecular resonance curves. Ann. Inst. H. Poincare, 45:339{358, 1986. [22] T. Kato. Perturbation Theory of Linear Operators, volume 132 of Grundlehren der mathematischen Wissenschaften. Springer-Verlag, 2 edition, 1976. [23] E. H. Lieb. Bound on the maximum negative ionization of atoms and molecules. Phys. Rev. A, 29:3018{3028, 1984. [24] M. Reed and B. Simon. Methods of Modern Mathematical Physics: Analysis of Operators, volume 4. Academic Press, San Diego, 1 edition, 1978. [25] M. Reed and B. Simon. Methods of Modern Mathematical Physics: Fourier Analysis and Self-Adjointness, volume 2. Academic Press, San Diego, 2 edition, 1980. [26] M. Reed and B. Simon. Methods of Modern Mathematical Physics: Functional Analysis, volume 1. Academic Press, San Diego, 2 edition, 1980. [27] Mary Beth Ruskai. Absence of discrete spectrum in highly negative ions II. Extension to Fermions. Commun. Math. Phys. , 85:325{327, 1982. [28] I. M. Sigal. Geometric methods in the quantum many-body problem. nonexistence of very negative ions. Commun. Math. Phys. , 85:309{324, 1982. [29] B. Simon. The de nition of molecular resonance curves by the method of exterior complex scaling. Phys. Lett. , A 71:211{214, 1979. [30] B. Simon. Functional Integration and Quantum Physics. Pure and applied mathematics. Academic Press, New York, 1979. [31] G. M. Zishlin. Discussion of the spectrum of the Schrodinger operator for systems of many particles. Tr. Mosk. Mat. O. -va, 9:81{120, 1960.