Abstract
We consider N bosons in a box with volume one, interacting through a two-body potential with scattering length of the order \(N^{-1+\kappa }\), for \(\kappa >0\). Assuming that \(\kappa \in (0;1/43)\), we show that low-energy states exhibit Bose–Einstein condensation and we provide bounds on the expectation and on higher moments of the number of excitations.
1 Introduction
We consider systems of \(N\in {\mathbb {N}}\) bosons trapped in the box \(\Lambda = [0;1]^3\) with periodic boundary conditions (the three-dimensional torus with volume one) and interacting through a repulsive potential with scattering length of the order \(N^{-1+\kappa }\), for \(\kappa \in (0; 1/43)\). We are interested in the limit of large N. The Hamilton operator has the form
and acts on a dense subspace of \(L^2_s (\Lambda ^N)\), the Hilbert space consisting of functions in \(L^2 (\Lambda ^N)\) that are invariant with respect to permutations of the \(N\in {\mathbb {N}}\) particles. Here, we assume the interaction potential \(V \in L^3 ({\mathbb {R}}^3)\) to have compact support and to be nonnegative, ie. \(V(x) \ge 0\) for almost all \(x \in {\mathbb {R}}^3\).
For \(\kappa = 0\), the Hamilton operator (1.1) describes bosons in the so-called Gross–Pitaevskii limit. This regime is frequently used to model trapped Bose gases observed in recent experiments. Another important regime is the thermodynamic limit, where N bosons interacting through a fixed potential V (independent of N) are trapped in the box \(\Lambda _L = [0;L]^3\) and where the limits \(N,L \rightarrow \infty \) are taken, keeping the density \(\rho = N/ L^3\) fixed. After rescaling lengths (introducing new coordinates \(x' = x / L\)), the Hamilton operator of the Bose gas in the thermodynamic limit is given (up to a multiplicative constant) by (1.1), with \(\kappa = 2/3\). Choosing \(0< \kappa < 2/3\), we are interpolating therefore between the Gross–Pitaevskii and the thermodynamic limits.
The goal of this paper is to show that low-energy states of (1.1) exhibit Bose–Einstein condensation in the zero-momentum mode \(\varphi _0 \in L^2 (\Lambda )\) defined by \(\varphi _0 (x) = 1\) for all \(x \in \Lambda \) and to give bounds on the number of excitations of the condensate. To achieve this goal, it is convenient to switch to an equivalent representation of the bosonic system, removing the condensate and focusing instead on its orthogonal excitations. To this end, we notice that every \(\psi _N \in L^2_s (\Lambda ^N)\) can be uniquely decomposed as
where \(\otimes _s\) denotes the symmetric tensor product and \(\alpha _j \in L^2_\perp (\Lambda )^{\otimes _s j}\) for all \(j = 0, \ldots , N\), with \(L^2_\perp (\Lambda )\) the orthogonal complement in \(L^2 (\Lambda )\) of \(\varphi _0\). This observation allows us to define a unitary map \(U_N : L^2_s (\Lambda ^N) \rightarrow {\mathcal {F}}_+^{\le N} = \bigoplus _{j=0}^N L^2_\perp (\Lambda )^{\otimes _s j}\) by setting
The truncated Fock space \({\mathcal {F}}_+^{\le N} = \bigoplus _{j=0}^N L^2_\perp (\Lambda )^{\otimes _s j}\) is used to describe orthogonal excitations of the condensate (some properties of the map \(U_N\) will be discussed in Sect. 2 below). On \({\mathcal {F}}_+^{\le N}\), we introduce the number of particles operator, defining \(({\mathcal {N}}_+ \xi )^{(n)} = n \xi ^{(n)}\) for every \(\xi = \{ \xi ^{(0)} , \ldots \xi ^{(N)} \} \in {\mathcal {F}}_+^{\le N}\).
We are now ready to state our main theorem, which provides estimates of the expectation and on higher moments of the number of orthogonal excitations of the Bose–Einstein condensate for low-energy states of (1.1).
Theorem 1.1
Let \( V\in L^3({\mathbb {R}}^3)\) be pointwise nonnegative and spherically symmetric. Let \(\mathfrak {a}_0 > 0\) denote the scattering length of V. Let \(H_N\) be defined as in (1.1) with \(0< \kappa < 1/43\). Then, for every \(\varepsilon > 0\), there exists a constant \(C>0\) such that
for all \(N \in {\mathbb {N}}\) large enough.
Let \(\psi _N \in L^2_s (\Lambda ^N)\) with \(\Vert \psi _N \Vert = 1\) and
for a \(\zeta > 0\). Then, for every \(\varepsilon > 0\) there exists a constant \(C > 0\) such that
for all \(N \in {\mathbb {N}}\) large enough. If moreover \(\psi _N = \chi (H_N \le E_N + \zeta ) \psi _N\), then for all \(k \in {\mathbb {N}}\) and all \(\varepsilon > 0\) there exists \(C > 0\) such that
for all \(N \in {\mathbb {N}}\) large enough.
The convergence \(E_N/4\pi \mathfrak {a}_0 N^{1+\kappa } \rightarrow 1\), as \(N \rightarrow \infty \), has been first established, for Bose gases trapped by an external potential, in [19] (the choice \(\kappa > 0\) corresponds, in the terminology of [19], to the Thomas–Fermi limit).
It follows from (1.5) that the one-particle density matrix \(\gamma _N = \mathrm{tr}_{2,\ldots , N} |\psi _N \rangle \langle \psi _N |\) associated with a normalized \(\psi _N \in L^2_s (\Lambda ^N)\) satisfying (1.4) is such that
as \(N \rightarrow \infty \). Here, we used the formula \(U_N a^* (\varphi _0) a(\varphi _0) U_N = N - {\mathcal {N}}_+\); see (2.5). Equation (1.7) implies that low-energy states of (1.1) exhibit complete Bose–Einstein condensation, if \(\kappa < 1/43\).
We remark that the estimate (1.6) follows, in our analysis, from a stronger bound controlling not only the number but also the energy of the excitations of the condensate. As we will explain in Sect. 3, in order to estimate the energy of excitations in low-energy states, we first need to remove (at least part of) their correlations. If we choose, as we do in (1.6), \(\psi _N \in L^2_s (\Lambda ^N)\) with \(\Vert \psi _N \Vert = 1\) and \(\psi _N = \chi (H_N \le E_N + \zeta ) \psi _N\), we can introduce the corresponding renormalized excitation vector \(\xi _N = e^B U_N \psi _N \in {\mathcal {F}}_+^{\le N}\), with the antisymmetric operator B defined as in (3.21) (the unitary operator \(e^B\) will be referred to as a generalized Bogoliubov transformation). We will show in Sect. 6 that for every \(k \in {\mathbb {N}}\), there exists \(C > 0\) such that
for all N large enough. Here \({\mathcal {H}}_N = {\mathcal {K}}+ {\mathcal {V}}_N\), where
are the kinetic and potential energy operators, restricted to \({\mathcal {F}}_+^{\le N}\). (Here, \({\widehat{V}}\) is the Fourier transform of the potential V, defined as in (2.4).) Equation (1.6) follows then from (1.8), because \({\mathcal {N}}_+\) commutes with \({\mathcal {H}}_N\), \({\mathcal {N}}_+ \le {\mathcal {K}}\le {\mathcal {H}}_N\) and because conjugation with the generalized Bogoliubov transformation \(e^B\) does not change the number of particles substantially; see Lemma 3.2 (for \(k \in {\mathbb {N}}\) even, we also use simple interpolation).
In the Gross–Pitaevskii regime corresponding to \(\kappa = 0\) the convergence \(\gamma _N \rightarrow |\varphi _0 \rangle \langle \varphi _0|\) has been first established in [16,17,18] and later, using a different approach, in [21].Footnote 1 In this case (ie. \(\kappa = 0\)), the bounds (1.3), (1.5) and (1.6) with \(\varepsilon = 0\) (which are optimal in their N-dependence) have been shown in [4]. Previously, they have been established in [2], under the additional assumption of small potential. A simpler proof of the results of [2], extended also to systems of bosons trapped by an external potential, has been recently given in [20]. The result of [4] was used in [3] to determine the second order corrections to the ground state energy and the low-energy excitation spectrum of the Bose gas in the Gross–Pitaevskii regime. Note that our approach in the present paper could be easily extended to the case \(\kappa = 0\), leading to the same bounds obtained in [4]. We exclude the case \(\kappa = 0\) because we would have to modify certain definitions, making the notation more complicated (for example, the sets \(P_H\) in (3.14) and \(P_L\) in (4.2) would have to be defined in terms of cutoffs independent of N).
The methods of [16,17,18] can also be extended to show Bose–Einstein condensation for low-energy states of (1.1), for some \(\kappa > 0\). In fact, following the proof of [18, Theorem 5.1], it is possible to show that for a normalized \(\psi _N \in L^2_s (\Lambda ^N)\) with \(\Vert \psi _N \Vert =1\) and such that \(\langle \psi _N, H_N \psi _N \rangle \le E_N + \zeta \), the expectation of the number of excitations is bounded by
which implies complete Bose–Einstein condensation for low-energy states, for all \(\kappa < 1/10\). For sufficiently small \(\kappa > 0\), Theorem 1.1 improves (1.10) because it gives a better rateFootnote 2 (if \(\kappa < 15/711\)) and because, through (1.6), it also provides (under stronger conditions on \(\psi _N\)) bounds for higher moments of the number of excitations \({\mathcal {N}}_+\).
In [10], in a slightly different setting, the authors obtain a bound of the form (1.6) for \(k=1\), for the choice \(\kappa = 1/(55+1/3) \) (for normalized \(\psi _N \in L^2_s (\Lambda ^N)\) that satisfy \( \langle \psi _N, H_N\psi _N\rangle \le E_N + \zeta \)). They use this result to show a lower bound on the ground state energy of the dilute Bose gas in the thermodynamic limit matching the prediction of Lee–Yang and Lee–Huang–Yang [13, 14].
After completion of our work, two more papers have appeared whose results are related with Theorem 1.1. Based on localization arguments from [8, 10], a bound for the expectation of \({\mathcal {N}}_+\) in low-energy states has been shown in [9], establishing Bose–Einstein condensation for all \(\kappa < 2/5\) (as pointed out there, using a refined analysis similar to that of [10], the range of \(\kappa \) can be slightly improved). On the other hand, following an approach similar to [2], but with substantial simplifications (partly due to the fact that the author works in the grand canonical, rather than the canonical, ensemble), a new proof of Bose–Einstein condensation was obtained in [11], in the Gross–Pitaevskii regime, under the assumption of small potential. There is hope that the approach of [11] can be extended beyond the Gross–Pitaevskii regime, providing a simplified proof of Theorem 1.1, potentially allowing for larger values of \(\kappa \).
The derivation of the bounds (1.5), (1.6), (1.8) is crucial to resolve the low-energy spectrum of the Hamiltonian (1.1). The extension of estimates on the ground state energy and on the excitation spectrum obtained in [3] for the Gross–Pitaevskii limit, to regimes with \(\kappa > 0\) small enough will be addressed in a separate paper [6], using the results of Theorem 1.1. With our techniques, it does not seem possible to obtain such precise information on the spectrum of (1.1) using only previously available bounds like (1.10).
Let us now briefly explain the strategy we use to prove Theorem 1.1. The first part of our analysis follows closely [4]. We start in Sect. 2 by introducing the excitation Hamiltonian \({\mathcal {L}}_N = U_N H_N U_N^*\), acting on the truncated Fock space \({\mathcal {F}}_+^{\le N}\); the result is given in (2.6), (2.7). The vacuum expectation \(\langle \Omega , {\mathcal {L}}_N \Omega \rangle = N^{1+\kappa } {\widehat{V}} (0)/2\) is still very far from the correct ground state energy of \({\mathcal {L}}_N\) (and thus of \(H_N\)); the difference is of order \(N^{1+\kappa }\). This is a consequence of the definition (1.2) of the unitary map \(U_N\), whose action removes products of the condensate wave function \(\varphi _0\), leaving however all correlations among particles in the wave functions \(\alpha _j \in L^2_\perp (\Lambda )^{\otimes _s j}\), \(j=1, \ldots , N\).
To factor out correlations, we introduce in Sect. 3 a renormalized excitation Hamiltonian \({\mathcal {G}}_N = e^{-B} {\mathcal {L}}_N e^B\), defined through unitary conjugation of \({\mathcal {L}}_N\) with a generalized Bogoliubov transformation \(e^B\). The antisymmetric operator \(B: {\mathcal {F}}_+^{\le N} \rightarrow {\mathcal {F}}_+^{\le N}\) is quadratic in the modified creation and annihilation operators \(b_p, b_p^*\) defined, for every momentum \(p \in \Lambda ^*_+ = 2\pi {\mathbb {Z}}^3 \backslash \{0 \}\), in (2.8) (\(b^*_p\) creates a particle with momentum p annihilating, at the same time, a particle with momentum zero; in other words, \(b_p^*\) creates an excitation, moving a particle out of the condensate). The properties of \({\mathcal {G}}_N\) are listed in Prop. 3.3. In particular, Proposition 3.3 implies that to leading order, \(\langle \Omega , {\mathcal {G}}_N \Omega \rangle \simeq 4\pi \mathfrak {a}_0 N^{1+\kappa }\), if \(\kappa \) is small enough.
Unfortunately, \({\mathcal {G}}_N\) is not coercive enough to prove directly that low-energy states exhibit condensation (in the sense that it is not clear how to estimate the difference between \({\mathcal {G}}_N\) and its vacuum expectation from below by the number of particle operator \({\mathcal {N}}_+\)). For this reason, in Sect. 4, we define yet another renormalized excitation Hamiltonian \({\mathcal {J}}_N = e^{-A} {\mathcal {G}}_N e^A\), where now A is the antisymmetric operator (4.1), cubic in (modified) creation and annihilation operators (to be more precise, we only conjugate the main part of \({\mathcal {G}}_N\) with \(e^A\); see (4.3)). Important properties of \({\mathcal {J}}_N\) are stated in Proposition 4.1. Up to negligible errors, the conjugation with \(e^A\) completes the renormalization of quadratic and cubic terms; in (4.5), these terms have the same form they would have for particles interacting through a mean-field potential with Fourier transform \(8\pi \mathfrak {a}_0 N^\kappa {\mathbf{1}} (|p| < N^\alpha )\), with a parameter \(\alpha > 0\) that will be chosen small enough, depending on \(\kappa \) (in other words, the renormalization procedure allows us to replace, in all quadratic and cubic terms, the original interaction with Fourier transform \(N^{-1+\kappa } {\widehat{V}} (p/ N^{1-\kappa })\) decaying only for momenta \(|p| > N^{1-\kappa }\), with a potential whose Fourier transform already decays on scales \(N^{\alpha } \ll N^{1-\kappa }\)).
The main problem with \({\mathcal {J}}_N\) is that its quartic terms (the restriction of the initial potential energy on the orthogonal complement of the condensate wave function) are still proportional to the local interaction with Fourier transform \(N^{-1+\kappa } {\widehat{V}} (p / N^{1-\kappa })\).
One possibility to solve this problem is to neglect the original quartic terms (they are positive) and insert instead quartic terms proportional to the renormalized mean-field potential \(8\pi \mathfrak {a}_0 N^\kappa \mathbf{1} (|p| < N^\alpha )\), so that Bose–Einstein condensation follows as it does for mean-field systems (see [22]). Since (with the notation \({\check{\chi }}\) for the inverse Fourier transform of the characteristic function on the ball of radius one)
and since we know from (1.10) that \({\mathcal {N}}_+ \lesssim N^{\frac{15 + 20 \kappa }{17}}\) in low-energy states, the insertion of the renormalized quartic terms produces an error that can be controlled by localization in the number of particles, if
This strategy was used in [4] to prove Bose–Einstein condensation with optimal rate in the Gross–Pitaevskii regime \(\kappa = 0\) (in this case, one can choose \(\alpha = 0\)).
Here, we follow a different approach. We perform a last renormalization step, conjugating \({\mathcal {J}}_N\) through a unitary operator \(e^D\), with D quartic in creation and annihilation operators. This leads to a new Hamiltonian \({\mathcal {M}}_N= e^{-D} {\mathcal {J}}_N e^D\) (in fact, it is more convenient to conjugate only the main part of \({\mathcal {J}}_N\), ignoring small contributions that can be controlled by other means; see (5.5)), where the original interaction \(N^{-1+\kappa } {\widehat{V}} (p/ N^{1-\kappa })\) is replaced by the mean-field potential \(8\pi \mathfrak {a}_0 N^\kappa \mathbf{1 } (|p| < N^\alpha )\) in all relevant terms.Footnote 3 Condensation can then be shown as it is done for mean-field systems, with no need for localization. This is the main novelty of our analysis, compared with [4]. In Sect. 5, we define the final Hamiltonian \({\mathcal {M}}_N\) and in Proposition 5.1 we bound it from below. The proof of Proposition 5.1, which is technically the main part of our paper, is deferred to Sect. 7. In Sect. 6, we combine the results of the previous sections to conclude the proof of Theorem 1.1.
The results we prove with our new technique are stronger than what we would obtain using the approach of [4] in the sense that they allow for larger values of \(\kappa \) and better rates. More importantly, we believe that the approach we propose here is more natural and that it leaves more space for extensions. In particular, with the final quartic renormalization step, we map the original Hamilton operator (1.1), with an interaction varying on momenta of order \(N^{1-\kappa }\), into a new Hamiltonian having the same form, but now with an interaction restricted to momenta smaller than \(N^{\alpha }\). If \(\alpha < 1-\kappa \), this leads to an effective regularization of the potential and it suggests that further improvements may be achieved by iteration; we plan to follow this strategy, which bears some similarities to the renormalization group analysis developed in [1], in future work.
In order to control errors arising from the quartic conjugation, it is important to use observables that were not employed in [4]. In particular, the expectation of the number of excitations with large momenta
and of its powers \({\mathcal {N}}^2_{\ge N^\gamma }, {\mathcal {N}}^3_{\ge N^\gamma }\), as well as the expectation of products of the form \({\mathcal {K}}_L {\mathcal {N}}_{\ge N^\gamma }\) and \({\mathcal {K}}_L {\mathcal {N}}_{\ge N^\gamma }^2\), involving the kinetic energy operator restricted to low momenta \({\mathcal {K}}_L\), will play a crucial role in our analysis. It will therefore be important to establish bounds for the growth of these observables through all steps of the renormalization procedure (Lemmas 4.2, 4.3, 7.1, 7.2). In Sect. 6, an important step in the proof of Theorem 1.1 will consist in controlling the expectation of these observables on low-energy states of the renormalized Hamiltonian \({\mathcal {G}}_N\).
2 The Excitation Hamiltonian
We denote by \({\mathcal {F}}= \bigoplus _{n \ge 0} L^2 (\Lambda )^{\otimes _s n}\) the bosonic Fock space over the one-particle space \(L^2 (\Lambda )\) and by \(\Omega = \{ 1, 0, \ldots \}\) the vacuum vector. We can define the number of particles operator \({\mathcal {N}}\) by setting \(({\mathcal {N}}\psi )^{(n)} = n \psi ^{(n)}\) for all \(\psi = \{ \psi ^{(0)}, \psi ^{(1)}, \ldots \}\) in a dense subspace of \({\mathcal {F}}\). For every one-particle wave function \(g \in L^2 (\Lambda )\), we define the creation operator \(a^* (g)\) and its hermitian conjugate, the annihilation operator a(g), through
Creation and annihilation operators are defined on the domain of \({\mathcal {N}}^{1/2}\), where they satisfy the bounds
and the canonical commutation relations
for all \(g,h \in L^2 (\Lambda )\) (\(\langle . , . \rangle \) denotes here the inner product on \(L^2 (\Lambda )\)). For \(p \in \Lambda ^* = 2\pi {\mathbb {Z}}^3\), we define the plane wave \(\varphi _p \in L^2 (\Lambda )\) through \(\varphi _p (x) = e^{-i p \cdot x}\) for all \(x \in \Lambda \), and the operators \(a^*_p = a (\varphi _p)\) and \(a_p = a (\varphi _p)\) creating and, respectively, annihilating a particle with momentum p. It is sometimes convenient to switch to position space, introducing operator valued distributions \(\check{a}_x, \check{a}_x^*\) such that
In terms of creation and annihilation operators, the number of particles operator can be written as
We will describe excitations of the Bose–Einstein condensate on the truncated Fock space
constructed over the orthogonal complement \(L^2_\perp (\Lambda )\) of the condensate wave function \(\varphi _0\). On \({\mathcal {F}}_+^{\le N}\), we denote the number of particles operator by \({\mathcal {N}}_+\). It is given by \({\mathcal {N}}_+ = \sum _{p \in \Lambda ^*_+} a_p^* a_p \), where \(\Lambda ^*_+ = \Lambda ^* \backslash \{ 0 \} = 2\pi {\mathbb {Z}}^3 \backslash \{ 0 \}\) is the momentum space for excitations. Given \(\Theta \ge 0\), we also introduce the restricted number of particles operators
measuring the number of excitations with momentum larger or equal to \(\Theta \), and \( {\mathcal {N}}_{< \Theta } = {\mathcal {N}}_+ - {\mathcal {N}}_{\ge \Theta }\).
Consider the operator \(U_N : L^2_s (\Lambda ^N) \rightarrow {\mathcal {F}}_+^{\le N}\) defined in (1.2). Identifying \(\psi _N \in L^2_s (\Lambda ^N)\) with the Fock space vector \(\{ 0, \ldots , 0, \psi _N, 0, \ldots \}\), we can also express \(U_N\) in terms of creation and annihilation operators; we obtain
It is then easy to check that \(U_N^* : {\mathcal {F}}_{+}^{\le N} \rightarrow L^2_s (\Lambda ^N)\) is given by
and that \(U_N^* U_N = 1\), ie. \(U_N\) is unitary.
Using \(U_N\), we can define the excitation Hamiltonian \({\mathcal {L}}_N := U_N H_N U_N^*\), acting on a dense subspace of \({\mathcal {F}}_+^{\le N}\). To compute \({\mathcal {L}}_N\), we first write the Hamiltonian (1.1) in momentum space, in terms of creation and annihilation operators. We find
where
is the Fourier transform of V, defined for all \(k \in {\mathbb {R}}^3\) (in fact, (1.1) is the restriction of (2.3) to the \(N\in {\mathbb {N}}\)-particle sector of the Fock space \({\mathcal {F}}\)). We can now determine the excitation Hamiltonian \({\mathcal {L}}_N\) using the following rules, describing the action of the unitary operator \(U_N\) on products of a creation and an annihilation operator (products of the form \(a_p^* a_q\) can be thought of as operators mapping \(L^2_s (\Lambda ^N)\) to itself). For any \(p,q \in \Lambda ^*_+ = 2\pi {\mathbb {Z}}^3 \backslash \{ 0 \}\), we find (see [15]):
We conclude that
with
where we introduced generalized creation and annihilation operators
for all \(p \in \Lambda ^*_+\). Observe that by (2.5),
In other words, \(b_p^*\) creates a particle with momentum \(p \in \Lambda ^*_+\) but, at the same time, it annihilates a particle from the condensate; it creates an excitation, preserving the total number of particles in the system. On states exhibiting complete Bose–Einstein condensation in the zero-momentum mode \(\varphi _0\), we have \(a_0 , a_0^* \simeq \sqrt{N}\) and we can therefore expect that \(b_p^* \simeq a_p^*\) and that \(b_p \simeq a_p\). Modified creation and annihilation operators satisfy the commutation relations
Furthermore, we find
for all \(p,q,r \in \Lambda _+^*\); this implies in particular that \([b_p , {\mathcal {N}}_+] = b_p\), \([b_p^*, {\mathcal {N}}_+] = - b_p^*\). It is also useful to notice that the operators \(b^*_p, b_p\), like the standard creation and annihilation operators \(a_p^*, a_p\), can be bounded by the square root of the number of particles operators; we find
for all \(\xi \in {\mathcal {F}}^{\le N}_+\). Since \({\mathcal {N}}_+ \le N\) on \({\mathcal {F}}_+^{\le N}\), the operators \(b_p^* , b_p\) are bounded, with \(\Vert b_p \Vert , \Vert b^*_p \Vert \le (N+1)^{1/2}\).
We can also define modified operator valued distributions
in position space, for \(x \in \Lambda \). The commutation relations (2.9) take the form
Moreover, (2.10) translates to
which also implies that \([ \check{b}_x, {\mathcal {N}}_+ ] = \check{b}_x\), \([ \check{b}_x^* , {\mathcal {N}}_+ ] = - \check{b}_x^*\).
3 Renormalized Excitation Hamiltonian
Conjugation with \(U_N\) extracts, from the original quartic interaction in (2.3), some constant and some quadratic contributions, collected in \({\mathcal {L}}^{(0)}_N\) and \({\mathcal {L}}^{(2)}_N\) in (2.7). For bosons described by the Hamiltonian (1.1), this is not enough; there are still large contributions to the energy that are hidden in \({\mathcal {L}}^{(3)}_N\) and \({\mathcal {L}}^{(4)}_N\).
To extract the missing energy, we have to take into account correlations. To this end, we consider the ground state solution \(f_\ell \) of the Neumann problem
on the ball \(|x| \le N^{1-\kappa }\ell \) (we omit the \(N\in {\mathbb {N}}\)-dependence in the notation for \(f_\ell \) and for \(\lambda _\ell \); notice that \(\lambda _\ell \) scales as \(N^{3\kappa -3}\)), with the normalization \(f_\ell (x) = 1\) if \(|x| = N^{1-\kappa } \ell \). By scaling, we observe that \(f_\ell (N^{1-\kappa }.)\) satisfies the equation
on the ball \(|x| \le \ell \). From now on, we fix some \(0< \ell < 1/2\), so that the ball of radius \(\ell \) is contained in the box \(\Lambda = [-1/2 ; 1/2]^3\). We then extend \(f_\ell (N^{1-\kappa }.)\) to \(\Lambda \), by setting \(f_{N} (x) = f_\ell (N^{1-\kappa }x)\), if \(|x| \le \ell \) and \(f_{N} (x) = 1\) for \(x \in \Lambda \), with \(|x| > \ell \). As a consequence,
where \(\chi _\ell \) denotes the characteristic function of the ball of radius \(\ell \). The Fourier coefficients of the function \(f_{N}\) are given by
for all \(p \in \Lambda ^*\). Next, we define \(w_\ell (x) = 1- f_\ell (x)\) for \(|x| \le N^{1-\kappa } \ell \) and \(w_\ell (x) = 0\) for all \(|x| > N^{1-\kappa } \ell \). Its rescaled version \(w_{N} : \Lambda \rightarrow {\mathbb {R}}\) is defined through \(w_{N} (x) = w_{\ell } (N^{1-\kappa }x)\) if \(|x| \le \ell \) and \(w_{N} (x) = 0\) if \(x \in \Lambda \) with \(|x| > \ell \). The Fourier coefficients of \(w_{N}\) are given by
where
denotes the Fourier transform of the (compactly supported) function \(w_\ell \). We find \({\widehat{f}}_{N} (p) = \delta _{p,0} - N^{3\kappa -3} {\widehat{w}}_\ell (p/N^{1-\kappa })\). From (3.2), we obtain
The next lemma summarizes important properties of the functions \(w_\ell \) and \( f_\ell \). Its proof can be found in [4, Appendix A] (replacing \(N\in {\mathbb {N}}\) by \(N^{1-\kappa }\) and noting that still \( N^{1-\kappa }\ell \gg 1\) for \(N\in {\mathbb {N}}\) sufficiently large and fixed \( \ell \in (0;1/2) \)).
Lemma 3.1
Let \(V \in L^3 ({\mathbb {R}}^3)\) be nonnegative, compactly supported and spherically symmetric. Fix \(\ell > 0\) and let \(f_\ell \) denote the solution of (3.1). For \(N\in {\mathbb {N}}\) large enough, the following properties hold true.
-
(i)
We have
$$\begin{aligned} \lambda _\ell = \frac{3\mathfrak {a}_0 }{ N^{3-3\kappa }\ell ^3} \left( 1 +{\mathcal {O}} \big (\mathfrak {a}_0 / \ell N^{1-\kappa }\big ) \right) . \end{aligned}$$(3.5) -
(ii)
We have \(0\le f_\ell , w_\ell \le 1\). Moreover there exists a constant \(C > 0\) such that
$$\begin{aligned} \left| \int V(x) f_\ell (x) dx - 8\pi \mathfrak {a}_0 \right| \le \frac{C \mathfrak {a}_0^2}{\ell N^{1-\kappa }}. \end{aligned}$$(3.6) -
(iii)
There exists a constant \(C>0 \) such that
$$\begin{aligned} w_\ell (x)\le \frac{C}{|x|+1} \quad \text { and }\quad |\nabla w_\ell (x)|\le \frac{C }{x^2+1}. \end{aligned}$$(3.7)for all \(x \in {\mathbb {R}}^3\) and all \(N\in {\mathbb {N}}\) large enough.
-
(iv)
There exists a constant \(C > 0\) such that
$$\begin{aligned} |{\widehat{w}}_{N} (p)| \le \frac{C}{N^{1-\kappa } p^2} \end{aligned}$$for all \(p \in {\mathbb {R}}^3\) and all \(N\in {\mathbb {N}}\) large enough (such that \(N^{1-\kappa } \ge \ell ^{-1}\)).
We define \(\eta : \Lambda ^* \rightarrow {\mathbb {R}}\) through
In position space, this means that for \( x\in \Lambda \), we have
so that we have in particular the \(L^\infty \)-bound
Lemma 3.1 also implies
for all \(p \in \Lambda _+^*=2\pi {\mathbb {Z}}^3 \backslash \{0\}\), and for some constant \(C>0\) independent of \(N\in {\mathbb {N}}\) (for \(N\in {\mathbb {N}}\) large enough). From (3.4), we find the relation
or equivalently, expressing the r.h.s. through the coefficients \(\eta _p\),
In our analysis, it is useful to restrict \(\eta \) to high momenta. To this end, let \(\alpha > 0\) and
We define \( \eta _H\in \ell ^2(\Lambda _+^*)\) by
Equation (3.11) implies that
and we assume from now on that \(\alpha > 2\kappa \) such that in particular
Notice, on the other hand, that the \(H^1\)-norm of \(\eta \) and \(\eta _{H}\) diverge, as \(N \rightarrow \infty \). From (3.9) and Lemma 3.1, part iii), we find
for all \(N \in {\mathbb {N}}\) large enough. We will mostly use the coefficients \(\eta _p\) with \(p\ne 0\). Sometimes, however, it will be useful to have an estimate on \(\eta _0\) (because Eq. (3.13) involves \(\eta _0\)). From Lemma 3.1, part iii), we obtain
It will also be useful to have bounds for the function \(\check{\eta }_H : \Lambda \rightarrow {\mathbb {R}}\), having Fourier coefficients \(\eta _H (p)\) as defined in (3.15). Writing \(\eta _H (p) = \eta _p - \eta _p \chi (|p| \le N^{\alpha })\), we obtain
so that
for all \(x \in \Lambda \), if \(N \in {\mathbb {N}}\) is large enough.
With the coefficients (3.15), we define the antisymmetric operator
and the generalized Bogoliubov transformation \(e^B : {\mathcal {F}}_+^{\le N} \rightarrow {\mathcal {F}}_+^{\le N}\). A first important observation is that conjugation with this unitary operator does not change the number of particles by too much. The proof of the following Lemma can be found in [7, Lemma 3.1] (a similar result has been previously established in [22]).
Lemma 3.2
Assume B is defined as in (3.21), with the coefficients \(\eta _p\) as in (3.8), satisfying (3.17). For every \(n \in {\mathbb {N}}\), there exists a constant \(C > 0\) such that
as an operator inequality on \({\mathcal {F}}_+^{\le N}\). (The constant depends only on \(\Vert \eta _H \Vert \) and on \(n \in {\mathbb {N}}\).)
With the generalized Bogoliubov transformation \(e^B\), we can now define the renormalized excitation Hamiltonian \({\mathcal {G}}_{N} : {\mathcal {F}}^{\le N}_+ \rightarrow {\mathcal {F}}^{\le N}_+\) by setting
In the next propositions, we collect important properties of \({\mathcal {G}}_{N}\). Recall the notation \({\mathcal {H}}_N = {\mathcal {K}}+ {\mathcal {V}}_N\), introduced in (1.9).
Proposition 3.3
Let \(V\in L^3({\mathbb {R}}^3)\) be compactly supported, pointwise nonnegative and spherically symmetric. Let \({\mathcal {G}}_N\) be defined as in (3.23). Assume that the exponent \(\alpha \) introduced in (3.14) is such that
Then,
and there exists \(C>0\) such that, for all \(\delta > 0\) and all \(N \in {\mathbb {N}}\) large enough, we have
and the improved lower bound
Furthermore, for \( \beta >0\), denote by \( {\mathcal {G}}_N^{\text {eff}}\) the excitation Hamiltonian
Then, there exists \(C > 0\) such that \({\mathcal {E}}_{{\mathcal {G}}_{N}} = {\mathcal {G}}_{N} - {\mathcal {G}}^{\text {eff}}_{N}\) is bounded by
for all \(N\in {\mathbb {N}}\) sufficiently large.
Furthermore, there exists a constant \(C > 0\) such that
for all \(\alpha \ge \gamma >0\), \(c>0\) fixed (independent of \(N\in {\mathbb {N}}\)) and \(N\in {\mathbb {N}}\) large enough.
Finally, for every \(k\in {\mathbb {N}}\), there exists a constant \(C>0\) such that
The proof of Proposition 3.3 is similar to the proof of [4, Prop. 4.2] and [3, Prop. 3.2], with the appropriate modifications dictated by the different scaling of the interaction. The main novelty in Proposition 3.3 is the bound (3.30) involving commutators of the restricted number of particles operator \({\mathcal {N}}_{\ge c N^\gamma }\). This can be obtained similarly to the bounds for \({\mathcal {E}}_{{\mathcal {G}}_N}\) and for \(i [ {\mathcal {N}}_+ , {\mathcal {G}}_N ]\), because we have a full expansion of the operator \({\mathcal {G}}_N\) in a sum of terms whose commutators with \({\mathcal {N}}_+\) and with \({\mathcal {N}}_{\ge cN^\gamma }\) retains essentially the same form. In the version of this paper that is posted on the arXiv, we give a complete proof of Proposition 3.3 in “Appendix A”, adapting the arguments of [4, Prop. 4.2], [3, Prop. 3.2].
4 Cubic Renormalization
From Eq. (3.28), we observe that the cubic terms in \({\mathcal {G}}^\text {eff}_N\) still depend on the original interaction, which decays slowly in momentum (in contrast to the quadratic terms in the second line of (3.28), where the sum is now restricted to \(P_H^c = \{ p \in \Lambda ^*_+ : |p| < N^\alpha \}\)).
To renormalize the cubic terms in (3.28), we are going to conjugate \({\mathcal {G}}^\text {eff}_N\) with a unitary operator \(e^A\), where the antisymmetric operator \(A:{\mathcal {F}}_+^{\le N}\rightarrow {\mathcal {F}}_+^{\le N}\) is defined by
The high-momentum set \(P_H = \{ p \in \Lambda ^*_+ : |p| \ge N^\alpha \}\) is as in (3.14). The low-momentum set \(P_L\) is defined by
with exponent \(\beta >0\), that will be chosen as in (3.28).
Using the unitary operator \(e^A\), we define \({\mathcal {J}}_N:{\mathcal {F}}_+^{\le N}\rightarrow {\mathcal {F}}_+^{\le N}\) by
Observe here that we only conjugate the main part \({\mathcal {G}}_N^\text {eff}\) of the renormalized excitation Hamiltonian \({\mathcal {G}}_N\); this makes the analysis a bit simpler (the difference \({\mathcal {G}}_N - {\mathcal {G}}_N^\text {eff}\) is small and can be estimated before applying the cubic conjugation).
The next proposition summarizes important properties of \( {\mathcal {J}}_N\); it can be shown very similarly to [4, Prop. 5.2], of course with the appropriate changes of the scaling of the interaction. In the version of this paper that is posted on the arXiv, we give a complete proof of Proposition 4.1 in “Appendix B”, adapting the arguments of [4, Prop. 5.2].
Proposition 4.1
Suppose the exponents \(\alpha \) and \(\beta \) are such that
Let \( {\mathcal {J}}_N\) be defined as in (4.3), let
and set \(\mu = \max (3\alpha /2 +2\kappa -1, 3\kappa /2-\beta )\) (\(\mu < 0\) follows from (4.4)). Then, there exists a constant \(C > 0\) such that the self-adjoint operator \({\mathcal {E}}_{{\mathcal {J}}_{N}} = {\mathcal {J}}_N - {\mathcal {J}}_N^{\text {eff} }\) satisfies the operator inequality
in \( {\mathcal {F}}_+^{\le N}\) for all \(N\in {\mathbb {N}}\) sufficiently large.
The bounds for \({\mathcal {J}}_N\) given in Proposition 4.1 are still not enough to show Theorem 1.1. As we will discuss in the next section, the main problem is the quartic interaction term, contained in \({\mathcal {H}}_N\), which still depends on the singular interaction potential (in all other terms on the r.h.s. of (4.5), the singular potential has been replaced by the regular mean-field type potential, with Fourier transform \(8\pi \mathfrak {a}_0 N^\kappa \mathbf{1 }_{P_H^c} (p)\), supported on momenta \(|p| < N^\alpha \)). To renormalize the quartic interaction, we will have to conjugate \({\mathcal {J}}^\text {eff}_N\) with yet another unitary operator, this time quartic in creation and annihilation operators. This last conjugation (which will be performed in the next section) will produce error terms. These errors will controlled in terms of the observables \({\mathcal {N}}_+\), \({\mathcal {K}}\) and \({\mathcal {V}}_N\) (as in (4.6)) but also, as we stressed at the end of Sect. 1, in terms of observables having the form \({\mathcal {N}}_{\ge N^\gamma }\) (the number of excitations having momentum larger or equal to \(N^\gamma \)), \({\mathcal {N}}^2_{\ge N^\gamma }\), \({\mathcal {N}}_{\ge N^\gamma }^3\), \({\mathcal {K}}_{\le N^\gamma }\) (the kinetic energy of excitations with momentum below \(N^\gamma \)), \({\mathcal {K}}_L {\mathcal {N}}_{\ge N^\gamma }\). For this reason, we need to control the action of \(e^A\) on all these observables.
First of all, we bound the action of the cubic phase on the restricted number of particles operators \({\mathcal {N}}_{\ge \theta } = \sum _{p \in \Lambda ^*_+ : |p| \ge \theta } a_p^* a_p\). We will make use of the pull-through formula \(a_p {\mathcal {N}}_{\ge \theta } = ({\mathcal {N}}_{\ge \theta } + \mathbf{1 }_{[\theta , \infty )} (p)) a_p\), which in particular implies that
Lemma 4.2
Assume the exponents \(\alpha , \beta \) satisfy (4.4) (in fact, here it is enough to assume that \(\alpha > 2\kappa \)). Let \( k\in {\mathbb {N}}_0\), \(m=0,1,2\), \(0 < \gamma \le \alpha \), \(c \ge 0\) (and \(c < 1\) if \(\gamma =\alpha \)). Then, there exists a constant \(C>0\) such that the operator inequalities
for all \(s\in [-1;1] \) and all \(N\in {\mathbb {N}}\).
Proof
The case \(m=0\) follows from \(m=1\). We start therefore with the case \(m=1\). For \(\xi \in {\mathcal {F}}_+^{\le N}\), we define the function \( \varphi _\xi :{\mathbb {R}}\rightarrow {\mathbb {R}}\) by
which has derivative
where \(A_1\) as in (4.1). By the assumptions on \(\gamma \) and c, we have \(N^\alpha \ge N^{\alpha }-N^{\beta } \ge cN^{\gamma }\) for \(N\in {\mathbb {N}}\) large enough. This implies in particular that
for \( r\in P_H\) and \(p\in P_L\), by (2.1) and (2.10). We then obtain
as well as
for some function \( \Theta : {\mathbb {N}}\rightarrow (0;1)\) by the mean value theorem. Using the pull-through formula \( {\mathcal {N}}_+ a^*_p = a^*_p ({\mathcal {N}}_++1)\) and Cauchy–Schwarz, we estimate
With the operator inequality \({\mathcal {N}}_{\ge cN^{\gamma } } \ge {\mathcal {N}}_{\ge N^{\alpha }} \) and with (4.7), we find that
The same arguments show that
Finally, we have that
Recalling (4.9), (4.10) and that \( \alpha \ge 2\kappa \), the bounds (4.12) to (4.14) show that
Since the bounds are independent of \(\xi \in {\mathcal {F}}_+^{\le N}\) and the same bounds hold true replacing A by \(-A\) in the definition of \( \varphi _\xi \), the first inequality in (4.8) follows by Gronwall’s Lemma.
To prove (4.8) with \(m=2\), we proceed similarly. Given \(\xi \in {\mathcal {F}}_+^{\le N}\), we define the function \( \psi _\xi :{\mathbb {R}}\rightarrow {\mathbb {R}}\) by
Its derivative is equal to
Comparing the contribution containing the double commutator in the last line on the r.h.s. of the last equation with (4.10) and using once again that \(N^\alpha \ge N^{\alpha }-N^{\beta } \ge cN^{\gamma }\) for \(N\in {\mathbb {N}}\) large enough, we observe that
Hence, the bounds (4.12) and (4.13) prove that
To bound the second contribution on the r.h.s. in (4.15), we recall (4.10) and we estimate
Finally, the last contribution in (4.15) can be bounded as in (4.14), using (4.11). We have
where, in the last step, we used that \( {\mathcal {N}}_{\ge cN^{\gamma }}\le {\mathcal {N}}_+ \). In conclusion, we have proved that
Since the bounds are independent of \(\xi \in {\mathcal {F}}_+^{\le N}\) and the same bounds hold true replacing \(-A\) by A in the definition \(\psi _\xi \), Gronwall’s lemma implies the last inequality in (4.8). \(\square \)
We denote the kinetic energy restricted to low momenta by
We will need the following estimates for the growth of the restricted kinetic energy.
Lemma 4.3
Assume the exponents \(\alpha ,\beta \) satisfy (4.4) (here we only need \(\alpha \ge 2\kappa \) and \(\alpha > \beta \)). Let \(0 < \gamma _1, \gamma _2 \le \alpha \), and \(c_1, c_2 \ge 0\) (and also \(c_j < 1\), if \(\gamma _j = \alpha \), for \(j=1,2\)). Then, there exists a constant \(C>0\) such that the operator inequalities
for all \(s\in [-1;1] \) and all \(N\in {\mathbb {N}}\) sufficiently large.
Proof
Like the previous Lemma 4.2, this is an application of Gronwall’s lemma. Let us start to prove the first inequality in (4.18). Fix \(\xi \in {\mathcal {F}}_+^{\le N}\) and define \(\varphi _\xi :{\mathbb {R}}\rightarrow {\mathbb {R}} \) by \(\varphi _\xi (s) = \langle \xi , e^{-sA} {\mathcal {K}}_{\le c_1 N^{\gamma _1}} e^{sA}\xi \rangle \) such that
We notice first that
if \( r\in P_H \) and \(p\in P_L\), because \( |r|, |p+r| \ge N^{\alpha }- N^{\beta } > c_1 N^{\gamma _1}\) for all \(N\in {\mathbb {N}}\).
Using the commutation relations (2.1), we then compute
With (4.19) and \(|p|\le N^{\beta }\) for \(p\in P_L\), we then find that
Finally, using Lemma 4.2 (with \( c=\frac{1}{2}\), \( \gamma =\alpha \) and \(N\in {\mathbb {N}}\) sufficiently large), we conclude
This proves the first inequality in (4.18), by Gronwall’s lemma.
Next, let us prove the second inequality in (4.18). We define \( \psi _\xi :{\mathbb {R}}\rightarrow {\mathbb {R}}\) by
and we compute
First, we proceed as in (4.20) and obtain with (4.7) that
Equation (4.21) and Lemma 4.2 then imply
Next, we recall the identity in (4.10) and that
whenever \( r\in P_H, p \in P_L\) and \(N\in {\mathbb {N}}\), by assumption on \( c_1\) and \(\gamma _1\). We then estimate
Hence, putting (4.22) and (4.23) together, we have proved that
This implies the second bound in (4.18), by Gronwall’s lemma. \(\square \)
Next, we seek a bound for the growth of the potential energy operator. To this end, we first compute the commutator of \({\mathcal {V}}_N\) with the antisymmetric operator A. We introduce here the shorthand notation for the low-momentum part of the kinetic energy
Proposition 4.4
Assume the exponents \(\alpha ,\beta \) satisfy (4.4). There exists a constant \(C>0\) such that
where the self-adjoint operator \({\mathcal {E}}_{[{\mathcal {V}}_N,A]} \) satisfies
for all \(\delta >0\) and for all \(N\in {\mathbb {N}}\) sufficiently large.
Proof
From (4.1), we have
Following [4, Prop. 8.1], we find
where
Here and in the following, the notation \( \sum ^*\) indicates that we only sum over those momenta for which the arguments of the creation and annihilation operators are nonzero. The first term on the r.h.s. of (4.27) appears explicitly in (4.25), so let us estimate next the size of the operators \( \Theta _{1}\) to \(\Theta _{4}\), defined in (4.28). The bounds can be obtained similarly as in the proof of [4, Prop. 8.1].
Consider first \(\Theta _{1}\). For \(\xi \in {\mathcal {F}}_+^{\le N}\), we switch to position space and find
The term \(\Theta _{2}\) on the r.h.s. of (4.28) can be controlled by
In the last step, we used (4.7) to estimate
for any \(\xi \in {\mathcal {F}}_+^{\le N}\). The contributions \(\Theta _{3}\) and \(\Theta _{4}\) can be bounded similarly. We find
as well as
Summarizing (using \(\alpha > 3\beta + 2 \kappa \)) we proved that
for any \(\delta >0\). Setting \({\mathcal {E}}_{[{\mathcal {V}}_N,A]} = \sum _{i=1}^4(\Theta _{i} +\text {h.c.}) \), this proves the claim. \(\square \)
From Proposition 4.4, we immediately get a bound for the action of \(e^A\) on \({\mathcal {V}}_N\).
Corollary 4.5
Assume the exponents \(\alpha ,\beta \) satisfy (4.4). Then, there exists a constant \(C>0\) such that
for all \(s\in [-1;1]\) and \(N\in {\mathbb {N}}\) large enough.
Proof
We apply Gronwall’s lemma. Given \(\xi \in {\mathcal {F}}_+^{\le N}\), we define \(\varphi _\xi (s) = \langle \xi , e^{-sA }{\mathcal {V}}_N e^{sA}\xi \rangle \) and compute its derivative s.t.
Hence, we can apply (4.25) and estimate
Here, we used (3.10), which shows that \(\Vert \check{\eta }\Vert _\infty \le CN \). Using Lemma 4.2, this simplifies to
Together with (4.25), the bound (4.26) (choosing \(\delta =1\)) and an application of Lemma 4.2 as well as of Lemma 4.3, the claim follows from Gronwall’s lemma.
\(\square \)
5 Quartic Renormalization
To explain why the bounds for \({\mathcal {J}}_N\) obtained in Prop. 4.1 are not enough to show Theorem 1.1, we introduce, for \( r\in \Lambda _+^*\), the operators
We denote the adjoints of \( c^*_r\) and \(e^*_r\) by \(c_r\) and \(e_r\), respectively. Notice in particular that \( e^*_r = e_{-r}\) for all \(r\in \Lambda _+^*\). A straightforward computation shows that
Together with (4.5), this suggests to bound the Hamiltonian \( {\mathcal {J}}_N\) from below by completing the square in the operators \(g_r^* := b^*_r + c^*_r + e^*_r\) and \(g_r := b_r + c_r + e_r\), for \( r\in P_H^c\subset \Lambda _+^*\). A better look at (4.5) reveals, however, that several terms that are needed to complete the square are still hidden in the energy \({\mathcal {H}}_N\). Since these terms are not small, we need to extract them from \({\mathcal {H}}_N\) by conjugation with a unitary operator \(e^{D}\), with
Since \([D,{\mathcal {N}}_+] = 0\), we have the identity
for all \(k \in {\mathbb {N}}\).
Using \(e^D\), we define the final excitation Hamiltonian
The next proposition provides an important lower bound for \({\mathcal {M}}_N\). Its proof is given in Sect. 7.
Proposition 5.1
Suppose the exponents \(\alpha \) (in the definition of the set \(P_H\) in (3.14)) and \(\beta \) (in the definition of the set \(P_L\) in (4.2)) are such that
Set \( \gamma = \min (\alpha , 1-\alpha -\kappa )\) (\(\gamma > 0\) from (5.6)) and let \(m_0\in {\mathbb {R}}\) be s.t. \( m_0\beta =\alpha \). Let \(V\in L^3({\mathbb {R}}^3)\) be compactly supported, pointwise nonnegative and spherically symmetric. Then, \( {\mathcal {M}}_N\), as defined as in (5.5), is bounded from below by
for a self-adjoint operator \( {\mathcal {E}}_{{\mathcal {M}}_N}\) satisfying
for all \(N\in {\mathbb {N}}\) sufficiently large.
6 Proof of Theorem 1.1
For \(\varepsilon >0 \) sufficiently small, we define
The choice \(\kappa < 1/43\) guarantees, if \(\varepsilon > 0\) is small enough, that all conditions in (5.6) (and thus also in (3.24) and (4.4)) are satisfied.
From (3.25) and (3.26), we obtain the upper bound
for the ground state energy of \(H_N\). From (3.25) and (3.27), on the other hand, we obtain
With (6.2) and setting \({\mathcal {G}}'_N = {\mathcal {G}}_N - E_N\), we deduce that
Next, we prove (1.5). From (3.29) and (6.3) we arrive at
Writing \({\mathcal {G}}_\text {eff} = e^A {\mathcal {J}}_N e^{-A}\) and recalling that \(\kappa < 1/43\) (and that \(\varepsilon > 0\) is small enough), Prop. 4.1 and (6.3) imply that
Inserting \({\mathcal {J}}_\text {eff} = e^D {\mathcal {M}}_N e^{-D}\) and applying Prop. 5.1, we obtain
With \({\mathcal {K}}\ge (2\pi )^2 {\mathcal {N}}_+\) and Lemma 4.2 (with \(m=0\) and \(k=1\)) we have
for a constant \(c > 0\) small enough (but independent of N). If N is large enough, we conclude (using also the upper bound (6.2)), that
To bound the error term \(e^A e^D {\mathcal {E}}_{{\mathcal {M}}_N} e^{-D} e^{-A}\), we need (according to (5.8)) to control observables of the form \(N^{-1} {\mathcal {K}}{\mathcal {N}}_{\ge c N^\gamma }\). To this end, we observe, first of all, that, by Cauchy–Schwarz and by (6.3),
Choosing \(\delta >0\) sufficiently small, we thus have
We write
Using (6.3) (similarly as we did in (6.7)) and \({\mathcal {N}}_{\ge cN^\gamma } \le N\), \({\mathcal {N}}_{\ge cN^{\gamma }}\le CN^{-2\gamma }{\mathcal {K}}\), we can bound the expectation of the first term on the r.h.s. of the last equation, for an arbitrary \(\xi \in {\mathcal {F}}_+^{\le N}\), by
On the other hand, to estimate the commutator term in Eq. (6.9), we notice that \( {\mathcal {A}}:= ({\mathcal {H}}_N+1)^{-1/2} i [{\mathcal {G}}_N',{\mathcal {N}}_{\ge cN^{\gamma }} ] ({\mathcal {H}}_N+1)^{-1/2} \) is a bounded, self-adjoint operator with \( \Vert {\mathcal {A}}\Vert \le CN^{\kappa +\alpha /2-\gamma } + CN^{\kappa +\gamma /2}\), by (3.30). Setting \( \mu = \max (\alpha , 3\gamma )\), this implies, with (6.3),
for all \(\xi \in {\mathcal {F}}_+^{\le N}\). Plugging (6.10) and (6.11) into (6.9), we find that, for sufficiently small \(\delta > 0\),
Inserting into (6.8) and choosing \(\delta > 0\) small enough, we obtain
Applying (6.13) to the r.h.s. of (5.8) we find, using also (6.3), (6.1), and the choice \(\kappa < 1/43\),
Inserting the last equation into (6.6) and using (6.2), we conclude that for N large enough,
For \(\psi _N \in L^2_s (\Lambda ^N)\) with \(\Vert \psi _N \Vert = 1\) and \(\langle \psi _N, (H_N - E_N)^2 \psi _N \rangle \le \zeta ^2\), the corresponding excitation vector \(\xi _N = e^{B} U_N \psi _N\) is such that \(\langle \xi _N, {\mathcal {G}}^{'2}_N \xi _N \rangle \le \zeta ^2\) and thus
which proves (1.5), using Lemma 3.2. From (6.3), we obtain also
an estimate that will be needed to arrive at (1.6).
Evaluating (6.14) on a normalized ground state \(\xi _N\) of \({\mathcal {G}}_N\) and inserting the result in (6.4) we also deduce that
Together with the upper bound (6.2), this concludes the proof of (1.3).
We still have to show (1.6) for \(k > 0\). To this end, we will prove the stronger bound (1.8); Eq. (1.6) follows then immediately from \({\mathcal {N}}_+ \le {\mathcal {H}}_N\) and by Lemma 3.2. We denote by \(Q_\zeta \) the spectral subspace of \({\mathcal {G}}_N\) associated with energies below \(E_N + \zeta \). We use induction to show that for all \(k \in {\mathbb {N}}\), there exists a constant \(C > 0\) (depending on k) such that
for all \(k \in {\mathbb {N}}\). This proves (1.8) and thus, with the bound \( {\mathcal {N}}_+\le {\mathcal {H}}_N\) and with Lemma 3.2, also (1.6). The case \(k=0\) follows from (6.15). From now on, we assume (6.16) to hold true, and we prove the same bound, with k replaced by \((k+1)\) (and with a new constant C). To this end, we start by observing that combining (6.3) and (6.6),
Hence,
We estimate the first term on the r.h.s. by
By Cauchy–Schwarz, we find
With \(({\mathcal {N}}_+ +1)^{2(k+1)} \le ({\mathcal {N}}_+ +1)^{2k+1} ({\mathcal {H}}_N+1)\) and with the estimate
from (3.31) we obtain, using again Cauchy–Schwarz,
for every \(\xi \in Q_{\zeta }\). Hence, for any \(\delta > 0\), we have
To bound the contribution proportional to \(e^A e^D {\mathcal {E}}_{{\mathcal {M}}_N} e^{-D} e^{-A}\) on the r.h.s. of (6.17), we have to control, according to (6.8), terms of the form
For an arbitrary \(\xi \in Q_{\zeta }\), we can bound the expectation of \(\text {A}\) by Cauchy–Schwarz as
As for the term \(\text {B}\), we can write
From (6.18) and using (3.30) to estimate
we obtain for every \(\xi \in Q_{\zeta }\) that
Applying the bounds \( {\mathcal {N}}_+\le N\), \( {\mathcal {N}}_{\ge c N^\gamma }\le C N^{-2\gamma }{\mathcal {K}}\) and (6.3) yields on the one hand
for any \(\delta > 0\). Since \( 8\kappa +2\varepsilon -\gamma \le 1+\kappa /2-\gamma \) and \( \kappa +\gamma /2 \le 1+\kappa /2-\gamma \) for all \(\gamma \le \alpha \) if \( \kappa <1/43\), this implies with the choice \( \delta =\frac{1}{4} (N^{8\kappa +2\varepsilon -\gamma } + N^{ \kappa +\gamma /2} )^{-1} \) that
On the other hand, we can estimate
Expressing \({\mathcal {V}}_N\) in position space, we find, with \(\phi = {\mathcal {N}}_{\ge cN^\gamma }({\mathcal {N}}_+ +1)^{k+1} \xi \),
We have
where
is such that \(\Vert \check{\chi }_x \Vert = \Vert \chi \Vert \le C N^{3\gamma /2}\). Hence, we find
Inserting in (6.23), we find
From (6.22), we conclude that
for all \(\gamma \le \alpha = 14 \kappa + 4\varepsilon \), if \(\kappa < 1/43\). Using now similar arguments as before (6.21), we conclude that together with (6.21), we have
Combining this with (6.20), we arrive at
for all \(\xi \in Q_z\). With (6.8), we obtain
Applying this bound to (5.8) and recalling that \( \kappa <1/43\), we conclude that
Therefore, for any \(\delta > 0\), we find (if N is large enough)
From the last bound, (6.19) and (6.17), we obtain
for any \(\xi \in Q_{\zeta }\). Taking the supremum over all \(\xi \in Q_{\zeta }\), and choosing \(\delta > 0\) small enough, we arrive at
by the induction assumption. \(\square \)
7 Analysis of \( {\mathcal {M}}_N \)
This section is devoted to the proof of Proposition 5.1. In Sect. 7.1 we establish bounds on the growth of the number of excitations and of their energy with respect to the action of \(e^D\), with the quartic operator \(D = D_1 - D_1^*\) with
as defined in (5.3). In Sect. 7.2, we compute the different parts of the excitation Hamiltonian \({\mathcal {M}}_N\), introduced in (5.5). Finally, in Sect. 7.3, we conclude the proof of Proposition 5.1.
7.1 Growth of Number and Energy of Excitations
The first lemma of this section controls the growth of the number of excitations with high momentum.
Lemma 7.1
Assume the exponents \(\alpha , \beta \) satisfy (5.6). Let \(k\in {\mathbb {N}}_0\), \(m=1,2,3\), \(0< \gamma \le \alpha \) and \(c>0\) (\(c < 1\) if \(\gamma = \alpha \)). Then, there exists a constant \(C>0\) such that
for all \(s\in [-1;1] \) and all \(N\in {\mathbb {N}}\) large enough.
Proof
Since \([{\mathcal {N}}_+ , {\mathcal {N}}_{\ge cN^\gamma } ] = 0\) and \([ {\mathcal {N}}_+ , D] =0\), it is enough to prove the lemma for \(k=0\). We consider first \(m=1\). For \(\xi \in {\mathcal {F}}_+^{\le N}\), we define the function \( \varphi _\xi :{\mathbb {R}}\rightarrow {\mathbb {R}}\) by
so that differentiating yields
with \(D_1\) as in (7.1). By assumption, \(N^\alpha \ge N^{\alpha }-N^{\beta } \ge cN^{\gamma }\) for sufficiently large \(N\in {\mathbb {N}}\). This implies that
for \( r\in P_H\) and \(p, q\in P_L\), by (2.1) and (2.10). We then compute
and apply Cauchy–Schwarz to obtain
Since the bound is independent of \(\xi \in {\mathcal {F}}_+^{\le N}\) and it also holds true if we replace D by \(-D\) in the definition of \( \varphi _\xi \), this proves (7.2), for \(m=1\).
For \(m=3\), we define
with derivative
We have
The contribution of the first term on the r.h.s. of (7.6) can be controlled as in (7.5) (replacing \(e^{sD} \xi \) with \(({\mathcal {N}}_{\ge c N^{\gamma }}+1) e^{sD} \xi \)). With (7.4) and using again that \(N^\alpha \ge N^{\alpha }-N^{\beta } \ge cN^{\gamma }\), we obtain that
All these contributions can be controlled like those in (7.4). We conclude that
This proves (7.2) with \(m=3\). The case \(m=2\) follows by operator monotonicity of the function \(x \mapsto x^{2/3}\). \(\square \)
Next, we prove bounds for the growth of the low-momentum part of the kinetic energy, defined as in (4.17).
Lemma 7.2
Assume the exponents \(\alpha , \beta \) satisfy (5.6). Let \(0< \gamma _1, \gamma _2 \le \alpha \), \(c_1, c_2 \ge 0\) (and \(c_j \le 1\) if \(\gamma _j = \alpha \), for \(j=1,2\)). Then, there exists a constant \(C>0\) such that
for all \(s\in [-1;1] \) and all \(N\in {\mathbb {N}}\) sufficiently large.
Proof
Fix \(\xi \in {\mathcal {F}}_+^{\le N}\) and define \(\varphi _\xi :{\mathbb {R}}\rightarrow {\mathbb {R}} \) by \(\varphi _\xi (s) = \langle \xi , e^{-sD} {\mathcal {K}}_{\le c_1 N^{\gamma _1}} e^{sD}\xi \rangle \) such that
We notice that
if \( r\in P_H \) and \(p,q\in P_L\), because \( |r|, |p+r|, |q-r|\ge N^{\alpha }- N^{\beta } > c_1 N^{\gamma _1}\) for \(N\in {\mathbb {N}}\) large enough.
Using (2.1), we then compute
and, using that \(|p|\le N^{\beta }\) for \(p\in P_L\), we obtain with Cauchy–Schwarz
With Lemma 7.1 choosing \( c=\frac{1}{2}\) and \( \gamma =\alpha \), this implies for \(N\in {\mathbb {N}}\) large enough that
This proves the first inequality in (7.7), by Gronwall’s lemma and \( \alpha > 3\beta +2\kappa \ge 0\).
Next, let us prove the second inequality in (7.7). We define \( \psi _\xi :{\mathbb {R}}\rightarrow {\mathbb {R}}\) by
and we compute
First, we proceed as in (7.9) and obtain with (4.7) that
Here, we used in the last step that \( [a_{q-r}, {\mathcal {N}}_{\ge c_2N^{\gamma _2}}] = a_{q-r}\) for \(r\in P_H\), \(q\in P_L\) and that \( {\mathcal {N}}_{c_2 N^{\gamma _2}} \ge {\mathcal {N}}_{ N^{\alpha }-N^{\beta }}\) for \(N\in {\mathbb {N}}\) large enough. The last bound and Lemma 7.1 imply that
Next, we recall the identity (7.4) and that
whenever \( r\in P_H, p,q \in P_L\) and \(N\in {\mathbb {N}}\) is sufficiently large. We then obtain
Hence, putting (7.10) and (7.11) together, we have proved that
which implies the second bound in (7.7), by Gronwall’s lemma. \(\square \)
It will also be important to control the potential energy operator, restricted to low momenta. We define
Notice that \( {\mathcal {V}}_{N,L} = {\mathcal {V}}_{N,L}^*\) by symmetry of the momentum restrictions. To calculate \(e^D {\mathcal {V}}_{N,L} e^{-D}\), we will use the next lemma, which will also be useful in the next subsections.
Lemma 7.3
Assume the exponents \(\alpha , \beta \) satisfy (5.6). Let \( F = (F_p)_{p\in \Lambda _+^*}\in \ell ^\infty (\Lambda _+^*)\) and define
Then, there exists a constant \(C>0\) such that
for all \(s\in [-1;1] \), and for all \(N\in {\mathbb {N}}\) sufficiently large.
Proof
Given \( \xi \in {\mathcal {F}}_+^{\le N}\), we define \( \varphi _\xi :{\mathbb {R}}\rightarrow {\mathbb {R}}\) by
which has derivative
By assumption, we have \( \alpha > 3\beta + 2\kappa \) so that \( |r|, |v+r|, |w-r| \ge N^\alpha -N^\beta > N^\beta \) if \( r\in P_H\) and \( v,w\in P_L\), for sufficiently large \(N\in {\mathbb {N}}\). This implies in particular that
whenever \(q+u, p\in P_L \) and \( r\in P_H\), \( v,w\in P_L\). As a consequence, we find
With (4.7) and \( N^\alpha -N^\beta > \frac{1}{2} N^\alpha \) for \(N\in {\mathbb {N}}\) large enough, we can bound
and
Lemmas 7.1, 7.2 and the assumption \( \alpha > 3\beta + 2\kappa \ge 0\) implies
Hence, integrating the last equation from zero to \( s\in [-1; 1]\) proves the lemma.
\(\square \)
With \(\sup _{p\in \Lambda ^*}| N^{\kappa }{\widehat{V}}(p/N^{1-\kappa })|\le CN^\kappa \), we obtain immediately the following result.
Corollary 7.4
Assume the exponents \(\alpha , \beta \) satisfy (5.6). Then, there exists a constant \(C>0\) such that
for all \(s\in [-1;1] \), and for all \(N\in {\mathbb {N}}\) sufficiently large.
We also need rough bounds for the conjugation of the full potential energy operator \({\mathcal {V}}_N\). To this end, we will make use of the following estimate for the commutator of \({\mathcal {V}}_N\) with \(D = D_1 - D_1^*\), with \(D_1\) defined in (7.1).
Proposition 7.5
Assume the exponents \(\alpha , \beta \) satisfy (5.6). Then,
and there exists a constant \(C>0\) such that
for all \(\delta >0\) and for all \(N\in {\mathbb {N}}\) sufficiently large.
Proof
We have
To compute the commutator \( [ {\mathcal {V}}_N, D_1 ]\), we compute first of all that
Putting the terms in the first and last line on the r.h.s. into normal order, we obtain
where
The first term on the r.h.s. in (7.18) appears explicitly in (7.16). Hence, let us estimate the size of the operators \( \Phi _{1}\) to \(\Phi _{4}\), defined in (7.19).
Starting with \(\Phi _1\), we switch to position space and find
The term \(\Phi _2\) on the r.h.s. of (7.19) can be controlled by
Finally, the contributions \(\Phi _3\) and \(\Phi _{4}\) can be bounded as follows. We obtain
as well as
In conclusion, the previous bounds imply with the assumption (5.6) (in particular, since \( \alpha > 3\beta +2\kappa \) and \(3\beta -2 < 0\)) that
holds true in \( {\mathcal {F}}_+^{\le N}\) for any \(\delta >0\). This concludes the proof. \(\square \)
With Proposition 7.5, we obtain a bound for the growth of \({\mathcal {V}}_N\).
Corollary 7.6
Assume the exponents \(\alpha , \beta \) satisfy (5.6). Then, there exists a constant \(C>0\) such that the operator inequality
for all \( s\in [-1;1]\) and for all \(N\in {\mathbb {N}}\) sufficiently large.
Proof
We apply Gronwall’s lemma. Given a normalized vector \(\xi \in {\mathcal {F}}_+^{\le N}\), we define \(\varphi _\xi (s) = \langle \xi , e^{-sD }{\mathcal {V}}_N e^{sD}\xi \rangle \) and compute its derivative s.t.
Hence, we can apply (7.16) and estimate
Here, we used (3.10), which shows that \(\Vert \check{\eta }\Vert _\infty \le CN \). Using Corollary 7.4 (recalling that \( \alpha > 3\beta +2\kappa \) and \( 2\beta \le 1\)) and \( {\mathcal {N}}_{\ge \frac{1}{2} N^{\alpha } }\le N\) in \( {\mathcal {F}}_+^{\le N}\), this simplifies to
Together with (7.16), the bound (7.17) (choosing \(\delta =1\)) and an application of Lemma 7.1 and of Lemma 7.2, the claim follows now from Gronwall’s lemma.
\(\square \)
Finally, we need control for the growth of the full kinetic energy operator \({\mathcal {K}}\). To this end, we need to estimate its commutator with D.
Proposition 7.7
Assume the exponents \(\alpha , \beta \) satisfy (5.6). Let \(m_0 \in {\mathbb {R}}\) be such that \(m_0\beta = \alpha \) (from (5.6) it follows that \(3< m_0 < 5\)). Then,
where the self-adjoint operator \({\mathcal {E}}_{[{\mathcal {K}},D]} \) satisfies
for all \(\delta >0\) and for all \(N\in {\mathbb {N}}\) sufficiently large.
Proof
Using that \([{\mathcal {K}}, D] = [{\mathcal {K}}, D_1] +\text {h.c.}\), a straight forward computation shows that
where
Let us estimate the size of the operators \( \Sigma _{1}, \Sigma _{2}\) and \(\Sigma _{3}\). Using \( \big |({\widehat{V}}(./N^{1-\kappa })*{\widehat{f}}_{N})(r)\big |\le C\), we control the operator \( \Sigma _1\) by
By Cauchy–Schwarz, the first term on the r.h.s. of (7.27) can be controlled by
The second contribution on the r.h.s. of (7.27) can be bounded by
Similarly, we find that
In summary, the previous three bounds imply that
for some constant \(C>0\) and all \(\delta >0\).
Next, let us switch to \( \Sigma _2\) and \(\Sigma _3\), defined in (7.26). Since \(({\widehat{\chi }}_\ell * {\widehat{f}}_N) (r) = {\widehat{\chi }}_\ell (r) + N^{-1} \eta _r\), with
we find
This, together with Lemma 3.1(i), Cauchy–Schwarz and \(\alpha >3\beta +2\kappa \), implies that
Similarly, we obtain
where we used that \( |r|/|v+r|\le 2\) for \(r\in P_H\), \(v\in P_L\) and \(N\in {\mathbb {N}}\) large enough. Combining (7.30), (7.31) and (7.32) and defining \( {\mathcal {E}}_{[{\mathcal {K}},D]} = \sum _{i=1}^3(\Sigma _i+\text {h.c.})\) proves the claim. \(\square \)
Corollary 7.8
Assume the exponents \(\alpha , \beta \) satisfy (5.6). Let \( m_0\in {\mathbb {R}}\) be such that \(m_0 \beta =\alpha \) (\(3< m_0 < 5\) from (5.6)). Then, there exists a constant \(C>0\) such that
for all \( s\in [-1;1]\) and for all \(N\in {\mathbb {N}}\) sufficiently large.
Proof
Given \( \xi \in {\mathcal {F}}_+^{\le N}\), we define \( \varphi _\xi (s) = \langle \xi , e^{-sD} {\mathcal {K}}e^{sD}\xi \rangle \). Differentiation yields
s.t., to bound the derivative of \( \varphi _\xi \), we can apply Proposition 7.7. Arguing exactly as in (7.22), we obtain with \( \sup _{x\in \Lambda } |f_N(x)|\le 1\) the operator inequality
Now, the claim follows from the bound (7.24) (choosing \(\delta =1\)), the previous bound and an application of Corollaries 7.6, 7.4, Lemmas 7.1, 7.2 and the operator bound \( {\mathcal {N}}_{\ge \frac{1}{2} N^\alpha }\le 4 N^{-2\alpha }{\mathcal {K}}\), by Gronwall’s Lemma. \(\square \)
7.2 Action of Quartic Renormalization on Excitation Hamiltonian
We compute now the main contributions to \({\mathcal {M}}_N= e^{-D} {\mathcal {J}}_{N}^\text {eff} e^{D}\). From (4.5) and recalling that \([{\mathcal {N}}_+ , D] = 0\), we can decompose
where the operators \( {\mathcal {M}}_{N}^{(i)}, i= 2,3,4,\) are defined by
7.2.1 Analysis of \( {\mathcal {M}}_N^{(2)}\)
In this section, we determine the main contributions to \( {\mathcal {M}}_N^{(2)}\), defined in (7.35) by
The main result of this section is the following proposition.
Proposition 7.9
Assume the exponents \(\alpha , \beta \) satisfy (5.6). Then
and there exists a constant \(C>0\) such that
for all \(N\in {\mathbb {N}}\) sufficiently large.
Proof
We start with the identity
and a straight-forward computation shows that
As a consequence, we find that
where
Here, \( \chi _{\{ p \in S\}}\) denotes as usual the characteristic function for the set \(S\subset \Lambda _+^*\), evaluated at \( p\in \Lambda _+^*\). Let us briefly explain how to bound the different contributions \( \text {V}_1\) to \(\text {V}_5\), defined in (7.41). Using Cauchy–Schwarz, the first two contributions are bounded by
where, for \( \text {V}_2\), we used that \(v+r\in P_H^c\) implies that \( |r|\le N^{\alpha }+N^{\beta }\) and furthermore that \( \sum _{N^\alpha \le |r|\le N^\alpha +N^\beta }|\eta _r|\le N^{\kappa +\beta }\). The contributions \( \text {V}_3\) to \(\text {V}_5\), on the other hand, can be controlled by
for any \( \xi \in {\mathcal {F}}_+^{\le N}\). In conclusion (since \(2\kappa + 3\beta -\alpha /2-1 < \kappa \) from (5.6)), we have proved that
Now, applying this bound together with (7.40), Lemmas 4.2, 4.3, 7.1, 7.2 and the operator inequality \( {\mathcal {N}}_{\ge \frac{1}{2} N^\alpha }\le 4 N^{-2\alpha }{\mathcal {K}}\) proves the claim. \(\square \)
7.2.2 Analysis of \( {\mathcal {M}}_N^{(3)}\)
In this section, we determine the main contributions to \( {\mathcal {M}}_N^{(3)}\), defined in (7.35) by
Proposition 7.10
Assume the exponents \(\alpha , \beta \) satisfy (5.6). Then, we have that
and there exists a constant \(C>0\) such that
for all \(N\in {\mathbb {N}}\) sufficiently large.
Proof
Let us define the operator \( \text {Y}:{\mathcal {F}}_+^{\le N}\rightarrow {\mathcal {F}}_+^{\le N} \) by
so that \( {\mathcal {M}}_N^{(3)} = e^{-D}\text {Y}e^D\). We recall the definition (7.1) and observe that
This implies that it is enough to control the commutator \( [\text {Y}, D_1]\) after conjugation with \( e^{tD}\), for any \(t\in [-1;1] \). Note that, if \( p\in P_H^c, q\in P_L, r\in P_H\) and \(v,w \in P_L\), we have \( |v+r|\ge N^\alpha -N^\beta>\frac{1}{2} N^\alpha >N^\beta \) s.t. \([a^*_{-p}a_q, a^*_{v+r}a^*_{w-r}] = 0 \), for \(N\in {\mathbb {N}}\) large enough. Then, a lengthy but straightforward calculation shows that
and
As a consequence, we conclude that
where
Let us explain how to control the operators \( \Psi _1\) to \(\Psi _6\), defined in (7.48). We start with \(\Psi _1\). Given \( \xi \in {\mathcal {F}}_+^{\le N}\), we find that
The contribution \( \Psi _2\) can be bounded by
Notice here, that we used that \( |r|\le N^\alpha +N^\beta \) if \( r+v\in P_H^c\) and \(v\in P_L\). Next, we apply as usual Cauchy–Schwarz to estimate the terms \( \Psi _3\) to \(\Psi _5 \) by
for all \( \alpha >3\beta +2\kappa \). Finally, the term \( \Psi _6\) can be controlled by
In conclusion, the previous estimates show that
so that, together with (7.46) and (7.47), an application of the Lemmas 4.2, 4.3, 7.1, 7.2 and the operator bound \( {\mathcal {N}}_{\ge \frac{1}{2} N^\alpha }\le 4N^{-2\alpha }{\mathcal {K}}\) proves the claim. \(\square \)
7.2.3 Analysis of \( {\mathcal {M}}_N^{(4)}\)
In this section, we determine the main contributions to \( {\mathcal {M}}_N^{(4)} = e^{-D}{\mathcal {H}}_N e^D\), defined in (7.35). To this end, we start with the observation that
with \(D_1\) defined in (7.1). By Propositions 7.5 and 7.7, this implies that
where we used that \( {\widehat{V}}(\cdot /N^{1-\kappa })*({\widehat{f}}_N-\eta /N)(r) = {\widehat{V}}(\cdot /N^{1-\kappa })(r)\) for all \(r\in \Lambda _+^*\). Moreover, the operators \( {\mathcal {E}}_{[{\mathcal {V}}_N,D]}\) and \( {\mathcal {E}}_{[{\mathcal {K}}, D]}\) are explicitly given by
where we recall the definitions (7.19) and (7.26). Let us analyze the different contributions in (7.50), separately. We start with the second term on the r.h.s. of (7.50).
Proposition 7.11
Assume the exponents \(\alpha , \beta \) satisfy (5.6). Then, we have
and there exists a constant \(C>0\) s.t. \( {\mathcal {E}}_1(s) \) and \( {\mathcal {E}}_2(s)\) satisfy
for all \( \delta >0\), \( s\in [-1;1]\) and for all \(N\in {\mathbb {N}}\) sufficiently large.
Proof
For definiteness, let us denote by \( \text {W}:{\mathcal {F}}_+^{\le N}\rightarrow {\mathcal {F}}_+^{\le N} \) the operator
and consider the identity
Now, observe that
for all \( p,q \in P_L\) and \( r\in P_H\), \(v,w\in P_L\) and \(N\in {\mathbb {N}}\) sufficiently large. Then, proceeding as in the proof of Proposition 7.5, we obtain
and
Combining the last two identities and putting non-normally ordered contributions into normal order, we find that
where
Let us briefly explain how to control the operators \( \zeta _{1}\) to \(\zeta _{6}\), defined in (7.2.3).
Noting that \( v+u\in P_L \) implies \( |u|\le 2N^\beta \) whenever \(v\in P_L\), the first two contributions \( \zeta _1\) and \(\zeta _2\) in (7.2.3) can be controlled by
By switching to position space, the term \(\zeta _3\) can be bounded by
We proceed similarly as above for the terms \( \zeta _4\) and \(\zeta _5\) which yields
where, for \( \zeta _5\), we used that \( v+r+u\in P_L\) implies that \( |u|\ge \frac{3}{4} N^\alpha \), and thus \(|q+u|\ge \frac{1}{2} N^\alpha \), whenever \( v,q\in P_L \), \(r\in P_H\) and \(N\in {\mathbb {N}}\) sufficiently large (otherwise \( |v+r+u|\ge \frac{1}{4}N^{\alpha }-N^\beta > N^\beta \) for large enough \(N\in {\mathbb {N}}\)). Finally, \( \zeta _6\) can be controlled by
In summary, the previous estimates show that
for all \( \delta >0\). On the other hand, by Lemma 7.3, we also know that
Now, going back to (7.55), the bounds (7.62) and (7.63) imply that
where the self-adjoint operators \( {\mathcal {E}}_1(s)\) and \({\mathcal {E}}_2(s)\) are bounded by
as well as
for all \( \delta >0\) and uniformly in \( s\in [-1;1]\). Defining \( {\mathcal {E}}_2(s) = {\mathcal {E}}_2(s, N^{-\beta -\kappa })\), this concludes the proof. \(\square \)
Equipped with Proposition 7.11, we go back to (7.50) and conclude that
for all \( \alpha \ge 3\beta +2\kappa \ge 0\) with \( \alpha +\beta +2\kappa -1<0\), \(0\le \kappa <\beta \) and \(N\in {\mathbb {N}}\) large enough.
Next, let us analyse the error terms related to \( {\mathcal {E}}_2(s) \) and \( {\mathcal {E}}_{[{\mathcal {V}}_N, D]}\) further. The bounds (7.53) and (7.21) (with \(\delta = c N^{-\beta -\kappa }\) for a sufficiently small \(c>0\); this choice guarantees that we can extract the term \({\mathcal {V}}_{N,L}\) in (7.66), with an error that can be absorbed in \({\mathcal {K}}\)) imply, together with Lemmas 7.1, 7.2, Corollaries 7.4 and 7.6 and with the assumption (5.6) on the exponents \(\alpha , \beta \), that
for all \(N\in {\mathbb {N}}\) large enough and for an arbitrarily small constant \({\widetilde{C}} > 0\). With Corollary 7.4 and (7.65), we conclude that
where the error \( {\mathcal {E}}_{{\mathcal {M}}_N}^{(41)}\) is such that
Applying Lemmas 4.2, 4.3 and Corollary 4.5, we deduce with the operator inequality \( {\mathcal {N}}_{\ge \frac{1}{2} N^\alpha }\le 4N^{-2\alpha }{\mathcal {K}}\) that
for all \(N\in {\mathbb {N}}\) large enough.
Now, we switch to the contribution containing the operator \( {\mathcal {E}}_{ [{\mathcal {K}}, D] }\) on the r.h.s. of the lower bound (7.66). We recall once again that
where the operators \( \Sigma _1, \Sigma _2\) and \(\Sigma _3\) were defined in (7.26). It turns out that \( \Sigma _2\) and \(\Sigma _3\) are negligible errors while \( \Sigma _1\) still contains an important contribution of leading order. We start with the analysis of the contribution related to \( \Sigma _1\).
Proposition 7.12
Assume the exponents \(\alpha , \beta \) satisfy (5.6). Then, we have that
and there exists a constant \(C>0\) such that
for all \( s\in [-1;1]\) and for all \(N\in {\mathbb {N}}\) sufficiently large.
Proof
We proceed as in Proposition 7.11 and recall \( \Sigma _1:{\mathcal {F}}_+^{\le N}\rightarrow {\mathcal {F}}_+^{\le N}\) to be
We then have
Similarly as in (7.58) and (7.2.3), we find that
where
The operators \( \Gamma _1\) to \(\Gamma _6\) can be bounded similarly as in the proof of Proposition 7.11. Let us start with \( \Gamma _1\). Applying as usual Cauchy-Schwarz implies that
where we used that \(v+u+r \in P_L \) implies \( |u|\ge N^{\alpha }-3N^\beta \) and \( |r|\le N^{\alpha }+3N^\beta \) whenever \( u\in P_H^c, r\in P_H\) and \( v\in P_L\) (otherwise \( |u+r+v| \ge |r| - |u|-N^\beta \ge 2N^\beta > N^\beta \) if either \( |u|\le N^{\alpha }-3N^\beta \) or \( |r|\ge N^{\alpha }+3N^\beta \), in contradiction to \( u+r+v\in P_L\)) for \(N\in {\mathbb {N}}\) sufficiently large. Notice in addition that \( \sum _{ N^\alpha -3N^\beta \le |u|\le N^\alpha } \le C N^{2\alpha +\beta }\).
The term \( \Gamma _2\) can be estimated exactly as the term \( \zeta _2\) in (7.60), that is
The contribution \(\Gamma _3\) can be controlled by
The terms \( \Gamma _4\) and \(\Gamma _5\) can be bounded exactly as in (7.61). We find
Finally, the last contribution \( \Gamma _6\) is bounded by
In conclusion, the above estimates imply that
for all \( \alpha >3\beta +2\kappa \ge 0\) and for all \(N\in {\mathbb {N}}\) sufficiently large. Combining this estimate with the identites (7.70) and (7.71), and applying Lemmas 4.2, 4.3, 7.1 as well as Lemma 7.2 together with the operator inequality \( {\mathcal {N}}_{\ge \frac{1}{2} N^\alpha }\le 4 N^{-2\alpha }{\mathcal {K}}\) proves the proposition. \(\square \)
Applying Proposition 7.12 to the lower bound (7.66) and defining \( {\mathcal {E}}_{{\mathcal {M}}_N}^{(42)} = \int _0^1 ds\; {\mathcal {E}}_3(s)\) with \( {\mathcal {E}}(s)\) from Proposition 7.12, we conclude that
where \( {\mathcal {E}}_{{\mathcal {M}}_N}^{(41)}\) satisfies the lower bound (7.67), \({\mathcal {E}}_{{\mathcal {M}}_N}^{(42)}\) satisfies the bound (7.69) and where the operators \(\Sigma _2 \) and \(\Sigma _3\) were defined in (7.26).
Let us finally estimate the size of the error in the last line of (7.72), involving the two operators \(\Sigma _2 \) and \(\Sigma _3\). Using the estimate (7.31) together with Lemmas 4.2, 4.3, 7.1 and 7.2, we find for \( {\mathcal {E}}_{{\mathcal {M}}_N}^{(43)} = \int _0^1ds\; e^{-sD} \big ( \Sigma _2+\text {h.c.}\big ) e^{sD}\)
Finally, consider the operator \( {\mathcal {E}}_{{\mathcal {M}}_N}^{(44)} = \int _0^1ds\; e^{-sD} \big ( \Sigma _3+\text {h.c.}\big ) e^{sD}\), with \(\Sigma _3\) defined in (7.26). Let \( m_0\in {\mathbb {R}}\) be such that \( m_0\beta =\alpha \) (in particular, \( \lfloor m_0 \rfloor \ge 3\)). Here, we use the bound (7.32) to find first of all that
for any \( \xi \in {\mathcal {F}}_+^{\le N}\) with \( \Vert \xi \Vert =1\). Notice that we applied once again Lemmas 7.1 and 7.2 in the second factor. With Corollary 7.8, the first factor is bounded by
for all exponents \(\alpha , \beta \) satisfying (5.6) and \(N\in {\mathbb {N}}\) sufficiently large. It follows that
where
with an arbitrarily small constant \({\widetilde{C}} >0\) and where after an additional application of Lemmas 4.2, 4.3, 7.1 and 7.2 together with the operator bound \({\mathcal {N}}_{\ge \Theta }\le \Theta ^{-2}{\mathcal {K}}\), the error \( {\mathcal {E}}_{{\mathcal {M}}_N}^{(443)}\) is such that
for all exponents \(\alpha , \beta \) satisfying (5.6) and \(N\in {\mathbb {N}}\) sufficiently large.
Choosing \({{\widetilde{C}}}>0\) sufficiently large (but independently of \(N\in {\mathbb {N}}\)) and arguing as right before (7.66), we deduce that
for all \(\alpha , \beta \) satisfying (5.6) and \(N\in {\mathbb {N}}\) sufficiently large. This follows through another application of Corollaries 4.5, 7.4 and 7.6, together with Lemmas 4.2, 4.3, 7.1 and 7.2. We summarize these bounds in the following corollary.
Corollary 7.13
Let \( m_0\in {\mathbb {R}}\) be such that \( m_0\beta =\alpha \) and let \( {\mathcal {M}}_N^{(4)}\) be defined as in (7.35). For every \({\widetilde{C}} > 0\), there exists a constant \(C>0\) such that
where
for all exponents \(\alpha ,\beta \) satisfying (5.6) and for all \(N\in {\mathbb {N}}\) sufficiently large.
Proof
The proof follows from defining \( {\mathcal {E}}_{{\mathcal {M}}_N}^{(4)} = \sum _{j=1}^3{\mathcal {E}}_{{\mathcal {M}}_N}^{(4j)} + \sum _{j=1}^3{\mathcal {E}}_{{\mathcal {M}}_N}^{(44j)}\) and combining (7.67), (7.72), (7.69), (7.73), (7.74), (7.75), (7.77), (7.76) and the operator bound \( {\mathcal {N}}_+\le (2\pi )^{-2}{\mathcal {K}}\) in \( {\mathcal {F}}_+^{\le N}\). \(\square \)
7.3 Proof of Proposition 5.1
Recall from (7.34) the decomposition
Collecting the results of Propositions 7.9, 7.10 and Corollary 7.13, we deduce that
where \({\mathcal {E}}'_{{\mathcal {M}}_N}\) satisfies the lower bound
for all \(N\in {\mathbb {N}}\) sufficiently large.
We combine next the terms on the third, fourth and fifth lines in (7.3). We first notice that
Arguing in the same way for the contribution on the fifth line in (7.3), using that \( ({\widehat{f}}_N-\eta /N)(p)=\delta _{p,0}\) for all \(p\in \Lambda _+^*\), and using that \( v\in P_L\) and \( v+r\in P_L \) implies in particular that \( r\in P_H^c\), we therefore obtain that
Now, notice furthermore that
such that, after switching to position space, the pointwise positivity \(V\ge 0\) implies
Here, we used that \( \Lambda _+^* = P_L \cup P_L^c \) and we denote by \( \check{\chi }_{S}\) the distribution which has Fourier transform \( \chi _S\), the characteristic function of the set \(S\subset \Lambda _+^*\).
Combining (7.3), (7.3), (7.3) and (7.84), it follows that
Using Lemma 3.1, part ii), we have \(\big ({\widehat{V}}(./N^{1-\kappa })*{\widehat{f}}_N\big )(0) = 8\pi {\mathfrak {a}}_0 + {\mathcal {O}}(N^{\kappa -1}) \). This implies
where, by (7.81) and Lemmas 4.2 and 7.1,
Similarly, for \( r\in P_H^c\), we know that
Therefore, proceeding exactly as between (7.27) and (7.30), with \(\big ({\widehat{V}}(./N^{1-\kappa })*{\widehat{f}}_N\big )(r)\) replaced by \(\big | \big ({\widehat{V}}(/N^{1-\kappa })*{\widehat{f}}_N\big )(r) - 8\pi {\mathfrak {a}}_0 \big | \), we deduce that
with \({\mathcal {E}}'''_{{\mathcal {M}}_N}\) satisfying the same bound (7.87) as \({\mathcal {E}}''_{{\mathcal {M}}_N}\). Here we used Lemmas 4.2, 4.3, 7.1 and 7.2, as well as the assumption (5.6).
Finally, recalling the definition (5.1) and the identity (5.2), we find
To express also the first term in the third line of (7.89) in terms of the modified creation and annihilation fields defined in (5.1), we first observe that
Then, for a fixed \( r\in P_H^c\), we have that
where
In particular, the union \( \bigcup _{j=1}^7S_j\) is a disjoint union. As a consequence, we find that
Inserting in (7.88), we obtain
with
Let us now estimate the remaining terms on the last line of (7.90). For \(\xi \in {\mathcal {F}}_+^{\le N}\), we have
and
Similarly, we can bound
Thus, choosing the constant \({\widetilde{C}} > 0\) small enough and applying Lemmas 7.2, 4.3 and 4.2 to the r.h.s. of (7.91) and to the second term on the r.h.s. of (7.92), we conclude that
where \( {\mathcal {E}}''''_{{\mathcal {M}}_N}\) is such that
We introduce the operators
With the algebraic identity
we conclude that
Since
we obtain that
A straightforward computation then shows that
Thus
where \({\mathcal {E}}_{{\mathcal {M}}_N}\) satisfies
This concludes the proof of Proposition 5.1. \(\square \)
Notes
Going through the proof of [18, Theorem 5.1], one can observe that the authors actually show that \(1 - \langle \varphi _0 , \gamma _N \varphi _0 \rangle \le C N^{-2/17}\).
For \(\kappa > 0\), the rate (1.6) is not expected to be optimal. Bogoliubov theory predicts that the number of excitations of the Bose–Einstein condensate in a Bose gas with density \(\rho \) is of the order \(N \rho ^{1/2}\); see [5]. In our regime, this corresponds to \(N^{3\kappa /2}\) excitations.
Observe that the renormalized potential with Fourier transform \(8\pi \mathfrak {a}_0N^{-1+\kappa } \mathbf{1 } (|p| < N^\alpha )\) that emerges in our rigorous analysis after a series of unitary transformations is reminiscent of the interaction that appears through an ad hoc substitution in the pseudo-potential method of [12, 13].
References
Balaban, T., Feldman, J., Knörrer, H., Trubowitz, E.: Complex bosonic many-body models: overview of the small field parabolic flow. Ann. Henri Poincaré 18, 2873–2903 (2017)
Boccato, C., Brennecke, C., Cenatiempo, S., Schlein, B.: Complete Bose–Einstein condensation in the Gross–Pitaevskii regime. Commun. Math. Phys. 359(3), 975–1026 (2018)
Boccato, C., Brennecke, C., Cenatiempo, S., Schlein, B.: Bogoliubov theory in the Gross–Pitaevskii limit. Acta Math. 222(2), 219–335 (2019)
Boccato, C., Brennecke, C., Cenatiempo, S., Schlein, B.: Optimal Rate for Bose-Einstein Condensation in the Gross-Pitaevskii Regime. Commun. Math. Phys. (2019). https://doi.org/10.1007/s00220-019-03555-9. arXiv:1812.03086
Bogoliubov, N.N.: On the theory of superfluidity. Izv. Akad. Nauk. USSR 11 (1947), 77. Engl. Transl. J. Phys. (USSR) 11 (1947), 23
Brennecke, C., Caporaletti, M., Schlein, B.: Excitation spectrum for Bose gases beyond the Gross–Pitaevskii regime. In preparation
Brennecke, C., Schlein, B.: Gross–Pitaevskii dynamics for Bose–Einstein condensates. Anal. PDE 12(6), 1513–1596 (2019)
Brietzke, B., Fournais, S., Solovej, J.P.: A simple 2nd order lower bound to the energy of dilute bose gases. Commun. Math. Phys. 376, 323–351 (2020)
Fournais, S.: Length scales for BEC in the dilute Bose gas. arXiv:2011.00309
Fournais, S., Solovej, J.P.: The energy of dilute Bose gases. Ann. Math. 192(3), 893–976 (2020)
Hainzl, C.: Another proof of BEC in the GP-limit. arXiv:2011.09450
Huang, K., Yang, C.N.: Quantum-mechanical many-body problem with hard-sphere interaction. Phys. Rev. 105(3), 757–767 (1957)
Lee, T.D., Huang, K., Yang, C.N.: Eigenvalues and eigenfunctions of a bose system of hard spheres and its low-temperature properties. Phys. Rev. 106(6), 1135–1145 (1957)
Lee, T.D., Yang, C.N.: Many-body problem in quantum mechanics and quantum statistical mechanics. Phys. Rev. 105, 1119–1120 (1957)
Lewin, M., Nam, P.T., Serfaty, S., Solovej, J.P.: Bogoliubov spectrum of interacting bose gases. Commun. Pure Appl. Math. 68(3), 413–471 (2014)
Lieb, E.H., Seiringer, R.: Proof of Bose–Einstein condensation for dilute trapped gases. Phys. Rev. Lett. 88, 170409 (2002)
Lieb, E.H., Seiringer, R.: Derivation of the Gross–Pitaevskii equation for rotating Bose gases. Commun. Math. Phys. 264(2), 505–537 (2006)
Lieb, E.H., Seiringer, R., Solovej, J.P., Yngvason, J.: The mathematics of the bose gas and its condensation. Oberwolfach Seminars. Birkhäuser Verlag, Series (2005)
Lieb, E.H., Seiringer, R., Yngvason, J.: Bosons in a trap: a rigorous derivation of the Gross–Pitaevskii energy functional. Phys. Rev. A 61, 043602 (2000)
Nam, P.T., Napiórkowski, M., Ricaud, J., Triay, A.: Optimal rate of condensation for trapped bosons in the Gross–Pitaevskii regime. arXiv:2001.04364
Nam, P.T., Rougerie, N., Seiringer, R.: Ground states of large bosonic systems: the Gross–Pitaevskii limit revisited. Anal. PDE. 9(2), 459–485 (2016)
Seiringer, R.: The excitation spectrum for weakly interacting bosons. Commun. Math. Phys. 306, 565–578 (2011)
Acknowledgements
We would like to thank C. Boccato and S. Cenatiempo for many helpful discussions with regards to the quartic renormalization. B. Schlein gratefully acknowledges partial support from the NCCR SwissMAP, from the Swiss National Science Foundation through the Grant “Dynamical and energetic properties of Bose–Einstein condensates” and from the European Research Council through the ERC-AdG CLaQS.
Funding
Open access funding provided by University of Zurich. Funding was provided by H2020 European Research Council (Grant No. ERC - ADG 2018 - Project 834782 CLaQS), Schweizerischer Nationalfonds zur Förderung der Wissenschaftlichen Forschung (Grant Nos. NCCR SwissMAP and Projekt 200020_172623).
Author information
Authors and Affiliations
Corresponding author
Additional information
Communicated by Vieri Mastropietro.
Publisher's Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Adhikari, A., Brennecke, C. & Schlein, B. Bose–Einstein Condensation Beyond the Gross–Pitaevskii Regime. Ann. Henri Poincaré 22, 1163–1233 (2021). https://doi.org/10.1007/s00023-020-01004-1
Received:
Accepted:
Published:
Issue Date:
DOI: https://doi.org/10.1007/s00023-020-01004-1