We slightly extend prior results about the derivation of the Maxwell–Schrödinger equations from the bosonic Pauli–Fierz Hamiltonian. More concretely, we show that the findings from Leopold and Pickl [SIAM J. Math. Anal. 52(5), 4900–4936 (2020)] about the coherence of the quantized electromagnetic field also hold for soft photons with small energies. This is achieved with the help of an estimate from Ammari et al. [arXiv:2202.05015 (2022)], which proves that the domain of the number of photon operator is invariant during the time evolution generated by the Pauli–Fierz Hamiltonian.

In this short paper, we derive the Maxwell–Schrödinger system of equations as an effective model describing a Bose–Einstein condensate of charged particles immersed in a coherent electromagnetic field. More precisely, we prove quantitatively that the Maxwell–Schrödinger system approximates well the many-body quantum evolution generated by the Pauli–Fierz Hamiltonian, provided that the total number of particles N is large, the particles are initially in a Bose–Einstein condensate, and the quantum nature of the field—quantified by the semiclassical parameter ℏ—is negligible. In particular, we focus on the combined regime N1+. Equivalently, the same effective dynamics approximates well N → ∞ bosons weakly interacting with a quantized electromagnetic field; see the discussion below.

This problem has already been studied by one of the authors, together with Pickl, in Ref. 1. The main focus here is to build on the results and techniques introduced there and to strengthen them by studying the convergence for the photons’ reduced density matrix. In the previous work, the quantum fluctuations around the coherent state of photons have been classified only by means of their energy. The extension to the reduced density matrix is physically relevant and mathematically nontrivial because the coherence of photons with small frequencies cannot be shown by the energy of the electromagnetic field due to its massless nature. We often refer to Ref. 1 throughout this paper, hopefully striking a good balance between being concise and being self-contained.

The Maxwell–Schrödinger system of equations describes the wave function φ of a quantum particle (with a nontrivial charge distribution κ) interacting with the classical electromagnetic field, described by the vector potential A and the electric field E=Ȧ. We choose the Coulomb gauge

(1.1)

and this makes, indeed, A and E the only dynamical degrees of freedom of the field. Let us also preliminarily define the current

(1.2)

The Maxwell–Schrödinger system thus takes the form

(1.3)

where V[φ] is an interaction term for the quantum particle and 1(Δ1) is the Helmholtz projection onto the subspace of divergence free vector fields. The choice of gauge, Coulomb’s in this case, can be seen as a constraint, for it is preserved by the Maxwell–Schrödinger flow. A typical example for the particle interaction V[φ] could be

(1.4)

where W is an external potential and (v* |φ|2)φ is a nonlinear term, usually originating from a microscopic pair interaction. With this choice, (1.3) can used as a system of mean-field equations to model the dynamics of identical particles interacting via the potential v and the classical electromagnetic field which they generate. The Cauchy problem associated with (1.3) is obtained by fixing an initial datum (φ0, A0, E0), subjected to the constraint ∇ · A0 = 0. In order to do so, it is convenient to introduce the complex scalar fields α0(,λ)λ=1,2 by defining

(1.5)
(1.6)

where ϵλ(k)λ=1,2 are the polarization vectors satisfying

(1.7)

which implement the Coulomb gauge. In fact, there is such a unique decomposition for any time, i.e.,

(1.8)
(1.9)

which respects both the Coulomb gauge and Ȧ=E. This makes it possible to rewrite Maxwell’s equations (same as in Ref. 2, Chap. 1.C) in terms of αt and to consider the equivalent system

(1.10)

with initial datum (φ0, α0(·, 1), α0(·, 2)). Throughout this article, we use F[f] to denote the Fourier transform of f. As it will be clarified shortly, φt and αt appear naturally as the effective counterparts of the microscopic dynamical variables. Note that the energy functional of the Maxwell–Schrödinger system is given by

(1.11)

with A being defined in analogy to (1.5). Global well-posedness for the Maxwell–Schrödinger system with V[φ]=v*|φ|2φ, κ(x) = e δ(x) (where e is the electric charge of the Schrödinger particle), and v(x)=e2|x| has been proven in Refs. 3 and 4. We will also consider only the case V[φ]=v*|φ|2φ, but we may require the charge distribution κ to be extended, in order to well-define the microscopic system, as discussed below. Typical examples of charge distributions that we will consider are of the form

(1.12)

representing a charged particle with total charge eR, distributed in a Gaussian fashion (“smoothed” spherical distribution of “diameter” σ), or

(1.13)

representing a sharp cutoff in momentum space with total charge eR. Let us remark that globally neutral particles can be considered as long as they have a nontrivial charge distribution: for example,

(1.14)

yields null total charge but nontrivial dipole and quadrupole interactions with the electromagnetic field.

If the charge distribution is not concentrated in a single point and the potential v represents an electrostatic mean-field self-interaction, then the form of the latter changes as well: a physically sensible choice would be v=κ*1||*κ. We will allow some liberty in the choices of κ and v; the specific requirements on the two will be made precise in Assumption 2.1. Concerning global well-posedness, let us remark that compared to the literature,3,4 our choices for κ and v will be at most “better” (i.e., more regular) and therefore do not affect the proof in any way since both κ and v act by convolution in the equation.

Proposition 1.1

(Ref.3 ). The Maxwell–Schrödinger system in Coulomb’s gauge(1.3)[with variables (φ, A, E)] is globally well-posed inH1 × H1 × L2. More precisely, we have the following:

  1. (regular solutions4 ) For every
    there exists a unique global solution
    of the Cauchy problem associated with(1.3).
  2. (rough solutions) For every
    there exists a unique global solution
    of the Cauchy problem associated with(1.3), being the unique strong limit of a sequence of regular solutions in (1), whose initial data approximate the rough initial datum inH1 × H1 × L2.
  3. (continuous dependence on initial data) The solutions (φ, A) in (2) continuously depend on the initial datum (φ0, A0, E0) ∈ H1 × H1 × L2.

For mR, let hm denote the weighed L2(R3)C2-space with norm

(1.15)

Throughout this work, we will rely on the following statement, which results almost immediately from Proposition 1.1 (see the  Appendix).

Corollary 1.2.

Let1/2F[κ]L2(R3,C). For every initial datum(φ0,α0)H2(R3,C)×h32such that1/2α0L2(R3)C2, the Maxwell–Schrödinger system(1.10)has a unique global solution inH2(R3,C)×h32.

The microscopic model corresponding to the Maxwell–Schrödinger system with mean-field self-interaction v* |φ|2φ consists of many identical nonrelativistic particles—obeying Bose–Einstein condensation—interacting among themselves by means of a weak pair potential and with a quantized electromagnetic field in Coulomb’s gauge. Contrarily to the “classical” case, the microscopic model is known to be well-defined only for extended charges. Let us start by defining a Hilbert space H(N) depending on two parameters NN,R+ as follows:

(1.16)

where Ls2(R3N) is the natural Hilbert space of N identical bosons (the subscript s indicates symmetry under the interchange of variables) and Γ is the second quantization functor associating with any (pre-)Hilbert space h=L2(R3)C2 the corresponding Fock representation of the canonical commutation relations [a(f),a*(g)]=f,gh, with ℏ being a semiclassical parameter measuring the degree of noncommutativity of the quantum field. The Fock representation is the natural one to describe noninteracting or regularized quantum field theories, the latter being the case here with hL2(R3)C2. With this interpretation, the limits N → ∞ and ℏ → 0 describe, respectively, the regimes in which the bosons are many and the quantum effects of the field are negligible. The time evolution is dictated by the Schrödinger equation,

(1.17)

where the Hamiltonian HN,, called Pauli–Fierz Hamiltonian, is given by

(1.18)

where μN, describes the coupling strength between the particles and the field, gN is the coupling strength between the particles, κ and v are the charge distribution and the pair potential introduced previously,

(1.19)

is the field’s kinetic energy, with a(k,λ) being the polarized creation and annihilation operators satisfying the canonical commutation relations (CCR),

(1.20)

and37 

(1.21)

is the smeared quantized electromagnetic vector potential in Coulomb’s gauge. Let us remark that both Âκ(x) and Hf depend on ℏ through the creation and annihilation operators, which have ℏ-dependent CCRs. The Hamiltonian HN, is self-adjoint on D(HN,)=D(HN,(0)), where HN,(0)=HN,|μN,=gN=0 whenever 1+1/2F[κ]L2(R3) and v is Kato-infinitesimal with respect to −Δ.5–7 

Our aim is to prove that the Maxwell–Schrödinger system emerges in some limit N → ∞ and/or ℏ → 0 as an effective model of the microscopic Pauli–Fierz dynamics. This is true only if we couple the parameters N, ℏ suitably and choose the coupling constants μN,, gN accordingly. A possible choice is given by N → ∞, =1N, μN,=1, and gN=1N. In this regime, the electromagnetic field becomes classical inverse proportionally to the increasing number of bosons. At the same time, the coupling between the particles and field is of order one, while the coupling between pairs of particles becomes weak (of order 1N). Defining N1/2aN1(k,λ) and N1/2aN1*(k,λ) as new creation and annihilation operators leads to the mathematically equivalent but physically different choice N → ∞, ℏ = 1, μN,=1N, and gN=1N. Here, the physical interpretation is of many bosons that interact weakly both with the quantized electromagnetic field (coupling of order 1N) and among themselves (pair coupling of order 1N). Our result reads as follows.

Provided that we choose an initial microscopic state that is “close enough” to a non-interacting state representing a complete condensate and a coherent field of minimal uncertainty, then at any timet ≥ 0, the evolution keeps the state “close” (in the same sense as above) to an analogous configuration in which the one-particle wave function and the argument of the coherent field have been evolved by the coupled Maxwell–Schrödinger equations.

The described scaling regime has been considered in earlier works for the Nelson model with an ultraviolet cutoff,8–11 the renormalized Nelson model,12 and the Fröhlich model.13,14 In Ref. 15, the Nelson model with an ultraviolet cutoff has been studied in a limit of many weakly interacting fermions. The classical behavior of quantum fields has also been proven in different scaling regimes.16–31 We also would like to mention Ref. 32, which derives the Maxwell–Schrödinger equations in a nonrigorous manner by neglecting certain terms in the Pauli–Fierz Hamiltonian.

From now on, we will keep N as the single parameter and choose ℏ = 1, μN,=1N, and gN=1N. We will use the notations H(N)=H(N), Fp=Γ1L2(R3)C2 with vacuum Ω, HN = HN,1, and ΨN = ΨN,1. Concerning the interaction potential and charge distribution, we will make the assumptions.

Assumption 2.1.
The (repulsive) interaction potentialvis a positive, real, and even function satisfying
(2.1)
The charge distributionκwith Fourier transformF[κ]satisfies
(2.2)

In order to state our result, we define for ΨNH(N) the one-particle reduced density matrix of the charged particles γΨN(1,0):L2(R3)L2(R3) by

(2.3)

where Tr2,…,N denotes the partial trace over the coordinates x2, …, xN and TrFp is the trace over Fock space. In addition, we introduce the number of photon operators,

(2.4)

and the unitary Weyl operator (fh),

(2.5)

satisfying

(2.6)

Our result is as follows.

Theorem 2.2.
Letvandκsatisfy Assumption 2.1,(φ0,α0)H2(R3,C)×h32such that1/2α0handφ0L2(R3)=1andΨN,0DHNDN1/2such thatΨN,0H(N)=1. Define
(2.7)
(2.8)
(2.9)
Let (φt, αt) and ΨN,tbe the unique solutions of(1.10)and(1.17), respectively. Then, there exists a functionC(s) depending monotonically on the normsφsH2(R3),1/2αsh,vL2+L(R3), and1+1/2F[κ]L2(R3)such that
(2.10)
(2.11)
for anyt ≥ 0. In particular, forΨN,0=φ0NW(Nα0)Ω, one obtains
(2.12)
(2.13)

Remark 2.3.
LetγΨN,t(0,1)be the one-particle reduced density matrix of the photons with the integral kernel
(2.14)
By similar means as in Ref.1 , Lemma 5.3, one obtains
(2.15)
from(2.11)and
(2.16)
for initial product statesΨN,0=φ0NW(Nα0)Ωfrom(2.13).

Remark 2.4.
In Ref.1 , Theorem 2.2 was proven for the charge distribution(1.13)and with the number operatorNin(2.8),(2.11), and(2.13)being replaced by the field energyHf. Because of Markov’s inequality,
one can use the field energy to conclude that the quantum fluctuations around the coherent state are subleading for all photons withkI. For sufficiently smallaN,bN, andcN, one can chooseINawitha < 1. However, this choice does not provide information about the coherence of soft photons with frequencies below this threshold.

The rest of the article outlines the Proof of Theorem 2.2. We will proceed as follows:

  1. We define a functional βN, φ, α], which measures if the charges of the many-body state ΨN form a Bose–Einstein condensate with condensate wave function φ and if the photons are in a coherent state with mean photon number Nαh2.

  2. Next, we show that the domain of β contains the solutions (φt, αt) of (1.10) and ΨN,t of (1.17) from Theorem 2.2 for all t ≥ 0.

  3. Afterward, we compute the change of βN,t, φt, αt] in time.

  4. Finally, we control the growth of βN,t, φt, αt] with the help of Grönwall’s inequality. This concludes the proof.

In doing so, we will rely on the findings from Ref. 1 and rather explain how the original proof of Ref. 1 has to be adapted. Most of the modifications are necessary to show the invariance of the domain in step 2 and to compute the time derivative of β in step 3.

We define a functional that consists of three parts.

Definition 3.1.
ForφL2(R3), we definep1φ:L2(R3N)L2(R3N)by
(3.1)
andq1φ:=1L2(R3N)p1φ. Now, letΨND(HN)DN1/2,φH1(R3), andαh12. Then,
(3.2)
and the functionalβ:D(HN)DN1/2×H1(R3)×h12R0+is defined asβ: = βa + βb + βc.

The functional βa measures if the charges of the many-body state are in a Bose–Einstein condensate (we refer to Refs. 33 and 34 for a comprehensive introduction). Its relation to the trace norm distance of the one-particle reduced density matrix is given by (see, e.g., Ref. 1, Lemma 5.3)

(3.3)

The functional βb quantifies the fluctuations of ΨN around the coherent state W(Nα)Ω. Using property (2.6) of the Weyl operators, it can be written as

(3.4)

showing that it is the same quantity as usually considered in the coherent state approach.35 While βa and βb measure the deviation of ΨN from the product state φNW(Nα)Ω, the functional βc is introduced for technical reasons. It quantifies the fluctuations of the many-body energy per particle around the energy of the Maxwell–Schrödinger system.

In the original proof of Ref. 1, the functional β was considered with βbΨN,α being replaced by

(3.5)

This definition has the advantage that it can be defined for many-body states ΨN in the domain DHN=H2(R3N,C)FDHf, which is invariant under the time evolution eiHNt. The additional difficulties with respect to Ref. 1 actually originate from the fact that DN1/2 [in contrast to DHf] is not contained in the domain of the Pauli–Fierz Hamiltonian. On the contrary, β̃b does not allow us to investigate the coherence of photons with small frequencies because the factor k in the integral on the right-hand side of (3.5) suppresses contributions from photons with small energies.

Throughout the rest of this article, (φt, αt) and ΨN,t denote the solutions of (1.10) and (1.17) from Theorem 2.2. In this section, we show that (ΨN,t,φt,αt)D(HN)DN1/2×H2(R3)×h32 for all t ≥ 0. The condition on the Maxwell–Schrödinger solutions is satisfied because of Corollary 1.2. While DHN is invariant under the evolution of the Pauli–Fierz Hamiltonian, due to Stone’s theorem, the invariance of DN1/2 is less clear because the photon number is not conserved during the time evolution. The next statement, however, displays that the number of photons can be controlled by the energy of the system.38 

Lemma 3.2.
LetΨN,0DHN1/2DN1/2. Then, there exists a constantC(depending onNand the choice ofκ) such that
(3.6)
This implieseiHNtDHN1/2DN1/2=DHN1/2DN1/2.

Proof of Lemma 3.2.
In the following, we use the notation
(3.7)
and
(3.8)
The vector potential can then be written as Âκ(x)=aGx+a*Gx. Recall the standard estimates for the annihilation and creation operators,
(3.9)
Let ΨN,0DHN1/2DN1/2, ΨN,t=eiHNtΨN,0, and δ ≥ 0, and consider the bounded operator Nδ=NeδN. Using j,Âκ(xj)=0, we obtain
(3.10)
Together with the Cauchy–Schwarz inequality, let us estimate
(3.11)
By means of the canonical commutation relations and the shifting property of the number operator, we get
(3.12)
Using eδN1, N(1eδ)eδN1, N(1eδ)eδN1N12, and (3.9), we obtain
(3.13)
Hence,
(3.14)
with HN(0)=j=1NΔj+Hf. Note that there exists a constant C(N, κ) dependent on the number of particles and the choice of κ such that HN(0)+11/2ΨC(N,κ)HN+11/2Ψ holds for all ΨDHN1/2. This fact follows from DHN(0)=DHN=H2(R3N,C)FpDHf and the closed graph theorem (Ref. 5, Theorem 1.3 and Corollary 1.4). In that regard, note that HN(0)1/2HN1/2+i1 is a closed operator. We consequently obtain
(3.15)
By the spectral theorem and monotone convergence,
(3.16)
Together with Stone’s theorem, this shows the claim.□

In order to estimate the emerging correlations between the particles and the photons during the evolution of the system, we compute the change of βΨN,t,φt,αt in time.

Lemma 3.3.
Let (φt, αt) and ΨN,tbe the solutions of(1.10)and(1.17)from Theorem 2.2 andβbbe defined as in Definition 3.1. Then,
(3.17)

Proof of Lemma 3.3.
Let
(3.18)
with Nδ being defined as in the Proof of Lemma 3.2. Using that W1(Nαt) is strongly differentiable in t from DN1/2 to H(N) with (see Ref. 36, Lemma 3.1)
(3.19)
and (2.6), we obtain that the time derivative of the fluctuation vector ξN,t=W1(Nαt)ΨN,t is given by ddtξN,t=iG(t)ξN,t with
(3.20)
In analogy to (3.12) and the subsequent discussion, one shows
(3.21)
and
(3.22)
for fh and ξN,tDHNDN1/2. Together with
(3.23)
this leads to
(3.24)
The claim then follows by Duhamel’s formula, monotone convergence, and straightforward manipulations using (2.6), j,Âκ(xj)=0, and j,Gxj(k,λ)=0.□

From Ref. 1, Secs. 6.2 and 6.4 and Duhamel’s formula, we immediately obtain the following.

Lemma 3.4.
Let (φt, αt) and ΨN,tbe the solutions of(1.10)and(1.17)from Theorem 2.2 andβa,βcbe defined as in Definition 3.1. Then,
(3.25)
Moreover,
(3.26)
due to energy conservation.

In this section, we classify the growth of βΨN,t,φt,αt in time. The first two inequalities of Theorem 2.2 then follow from (3.3) and (3.4) and the statement below. By similar estimates as in Ref. 1, Chap. 7, one obtains (2.12) and (2.13).

Lemma 3.5.
Let (φt, αt) and ΨN,tbe the solutions of(1.10)and(1.17)from Theorem 2.2. Then, there exists a monotone increasing functionC(s) of the normsφsH2(R3),1/2αsh,vL2+L(R3), and1/2+1F[κ]L2(R3)such that
(3.27)
(3.28)
holds for anyt ≥ 0.

Sketch of the Proof of Lemma 3.5.
Inequality (3.27) is proven analogously to Ref. 1, Lemma 6.10. The similarity becomes obvious if one defines the auxiliary fields
(3.29)
By means of the cutoff function η with Fourier transform,
(3.30)
we can write the quantum and classical vector potentials as
(3.31)
These relations are the analog of Ref. 1, Lemma 6.1. Thus, if we replace Êκ, Eκ in the original estimates of Ref. 1 by F̂, F with ∈ {−, +} and use
(3.32)
we obtain (3.27) by similar means. This implies (3.28) because of Grönwall’s inequality.□

N.L. would like to thank Peter Pickl for many fruitful discussions within the project.1 M.F. acknowledges the support from “Istituto Nazionale di Alta Matematica (INdAM)” through the “Progetto Giovani GNFM 2020: Emergent Features in Quantum Bosonic Theories and Semiclassical Analysis.” N.L. acknowledges the support from the Swiss National Science Foundation through the NCCR SwissMap and funding from the European Union’s Horizon 2020 Research and Innovation Programme under Marie Skłodowska-Curie Grant Agreement No. 101024712.

The authors have no conflicts to disclose.

Marco Falconi: Writing – original draft (equal); Writing – review & editing (equal). Nikolai Leopold: Writing – original draft (equal); Writing – review & editing (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

Proof of Corollary 1.2.
For the initial data of the corollary, we have that (φ0, A0, E0) ∈ H2 × H2 × H1. The existence of a unique global solution φ,α with φtH2(R3) and 1/2+3/2αth then follows from Proposition 1.1. In order to see that αth, we bound the integral version of (1.10) by
(A1)
Using Hölder’s inequality and Young’s inequality, we get
(A2)
With the help of Proposition 1.1, we conclude that the right-hand side of (A1) is finite. This shows the claim.□

1.
N.
Leopold
and
P.
Pickl
, “
Derivation of the Maxwell–Schrödinger equations from the Pauli–Fierz Hamiltonian
,”
SIAM J. Math. Anal.
52
(
5
),
4900
4936
(
2020
).
2.
C.
Cohen-Tannoudji
,
J.
Dupont-Roc
, and
G.
Grynberg
,
Photons and Atoms: Introduction to Quantum Electrodynamics
(
Wiley
,
1997
), ISBN: 978-0-471-18433-1.
3.
I.
Bejenaru
and
D.
Tataru
, “
Global wellposedness in the energy space for the Maxwell-Schrödinger system
,”
Commun. Math. Phys.
288
(
1
),
145
198
(
2009
).
4.
M.
Nakamura
and
T.
Wada
, “
Global existence and uniqueness of solutions to the Maxwell-Schrödinger equations
,”
Commun. Math. Phys.
276
,
315
339
(
2007
).
5.
F.
Hiroshima
, “
Self-adjointness of the Pauli-Fierz Hamiltonian for arbitrary values of coupling constants
,”
Ann. Henri Poincare
3
,
171
201
(
2002
).
6.
O.
Matte
, “
Pauli-Fierz type operators with singular electromagnetic potentials on general domains
,”
Math. Phys., Anal. Geom.
20
,
18
(
2017
).
7.
H.
Spohn
,
Dynamics of Charged Particles and Their Radiation Field
(
Cambridge University Press
,
Cambridge
,
2004
), ISBN: 0-521-83697-2.
8.
Z.
Ammari
and
M.
Falconi
, “
Wigner measures approach to the classical limit of the Nelson model: Convergence of dynamics and ground state energy
,”
J. Stat. Phys.
157
(
2
),
330
362
(
2014
).
9.
M.
Falconi
, “
Classical limit of the Nelson model with cutoff
,”
J. Math. Phys.
54
(
1
),
012303
(
2013
).
10.
M.
Falconi
,
N.
Leopold
,
D.
Mitrouskas
, and
S.
Petrat
, “
Bogoliubov dynamics and higher-order corrections for the regularized Nelson model
,”
Rev. Math. Phys.
(published online
2022
).
11.
N.
Leopold
and
P.
Pickl
, “
Mean-field limits of particles in interaction with quantized radiation fields
,” in
Macroscopic Limits of Quantum Systems
, Springer Proceedings in Mathematics and Statistics Vol. 270, edited by
D.
Cadamuro
,
M.
Duell
,
W.
Dybalski
, and
S.
Simonella
(
Springer
,
2018
), pp.
185
214
.
12.
Z.
Ammari
and
M.
Falconi
, “
Bohr’s correspondence principle for the renormalized Nelson model
,”
SIAM J. Math. Anal.
49
(
6
),
5031
5095
(
2017
).
13.
N.
Leopold
, “
Norm approximation for the Fröhlich dynamics in the mean-field regime
,” arXiv:2207.01598 (
2022
).
14.
N.
Leopold
,
D.
Mitrouskas
, and
R.
Seiringer
, “
Derivation of the Landau–Pekar equations in a many-body mean-field limit
,”
Arch. Ration. Mech. Anal.
240
,
383
417
(
2021
).
15.
N.
Leopold
and
S.
Petrat
, “
Mean-field dynamics for the Nelson model with fermions
,”
Ann. Henri Poincare
20
(
10
),
3471
3508
(
2019
).
16.
Z.
Ammari
,
M.
Falconi
, and
F.
Hiroshima
, “
Towards a derivation of Classical ElectroDynamics of charges and fields from QED
,” arXiv:2202.05015 (
2022
).
17.
R.
Carlone
,
M.
Correggi
,
M.
Falconi
, and
M.
Olivieri
, “
Microscopic derivation of time-dependent point interactions
,”
SIAM J. Math. Anal.
53
(
4
),
4657
4691
(
2021
).
18.
M.
Correggi
and
M.
Falconi
, “
Effective potentials generated by field interaction in the quasi-classical limit
,”
Ann. Henri Poincare
19
(
1
),
189
235
(
2018
).
19.
M.
Correggi
,
M.
Falconi
, and
M.
Olivieri
, “
Quasi-classical dynamics
,”
J. Eur. Math. Soc.
(published online
2022
).
20.
M.
Correggi
,
M.
Falconi
, and
M.
Olivieri
, “
Ground state properties in the quasi-classical regime
,” Anal. PDE (to be published); arXiv:2007.09442 (
2020
).
21.
E. B.
Davies
, “
Particle–boson interactions and the weak coupling limit
,”
J. Math. Phys.
20
,
345
351
(
1979
).
22.
D.
Feliciangeli
,
S.
Rademacher
, and
R.
Seiringer
, “
Persistence of the spectral gap for the Landau–Pekar equations
,”
Lett. Math. Phys.
111
,
19
(
2021
).
23.
R. L.
Frank
and
G.
Zhou
, “
Derivation of an effective evolution equation for a strongly coupled polaron
,”
Anal. PDE
10
(
2
),
379
422
(
2017
).
24.
R. L.
Frank
and
B.
Schlein
, “
Dynamics of a strongly coupled polaron
,”
Lett. Math. Phys.
104
,
911
929
(
2014
).
25.
J.
Ginibre
,
F.
Nironi
, and
G.
Velo
, “
Partially classical limit of the Nelson model
,”
Ann. Henri Poincare
7
,
21
43
(
2006
).
26.
M.
Griesemer
, “
On the dynamics of polarons in the strong-coupling limit
,”
Rev. Math. Phys.
29
(
10
),
1750030
(
2017
).
27.
A.
Knowles
, “
Limiting dynamics in large quantum systems
,” Ph.D. thesis,
ETH Zürich
,
2009
; available at http://www.unige.ch/knowles/thesis.pdf.
28.
N.
Leopold
,
D.
Mitrouskas
,
S.
Rademacher
,
B.
Schlein
, and
R.
Seiringer
, “
Landau–Pekar equations and quantum fluctuations for the dynamics of a strongly coupled polaron
,”
Pure Appl. Anal.
3
(
4
),
653
676
(
2021
).
29.
N.
Leopold
,
S.
Rademacher
,
B.
Schlein
, and
R.
Seiringer
, “
The Landau–Pekar equations: Adiabatic theorem and accuracy
,”
Anal. PDE
14
(
7
),
2079
2100
(
2021
).
30.
D.
Mitrouskas
, “
A note on the Fröhlich dynamics in the strong coupling limit
,”
Lett. Math. Phys.
111
,
45
(
2021
).
31.
S.
Teufel
, “
Effective N-body dynamics for the massless Nelson model and adiabatic decoupling without spectral gap
,”
Ann. Henri Poincare
3
,
939
965
(
2002
).
32.
M.
Šindelka
, “
Derivation of coupled Maxwell-Schrödinger equations describing matter-laser interaction from first principles of quantum electrodynamics
,”
Phys. Rev. A
81
(
3
),
033833
(
2010
).
33.
A.
Knowles
and
P.
Pickl
, “
Mean-field dynamics: Singular potentials and rate of convergence
,”
Commun. Math. Phys.
298
,
101
138
(
2009
).
34.
P.
Pickl
, “
A simple derivation of mean field limits for quantum systems
,”
Lett. Math. Phys.
97
,
151
164
(
2011
).
35.
I.
Rodnianski
and
B.
Schlein
, “
Quantum fluctuations and rate of convergence towards mean field dynamics
,”
Commun. Math. Phys.
291
(
1
),
31
61
(
2009
).
36.
J.
Ginibre
and
G.
Velo
, “
The classical field limit of scattering theory for non-relativistic many-boson systems. I
,”
Commun. Math. Phys.
66
,
37
76
(
1979
).
37.

To simplify the notation we assume F[κ](k)R for all kR3. Theorem 2.2 equally applies if F[κ] is complex valued. In this case, Âκ(x)=λ=1,2d3k12kϵλ(k)F[κ](k)̄eikxa(k,λ)+F[κ](k)eikxa*(k,λ).

38.

Inequality (3.6) was originally proven by Hiroshima and appeared in a slightly different form (for the second instead of the first moment of the number operator) in Ref. 16, Proposition 3.11. We would like to thank Hiroshima for sharing his notes with us. The proof is presented again for the convenience of the reader.