1 Introduction
In this manuscript we study a dilute Bose gas consisting of N quantum particles subject to Bose-Einstein statistics, in which the individual particles interact with each other via a three-body potential

defined in terms of a given bounded and non-negative function
$V:\mathbb {R}^{3}\times \mathbb {R}^3\longrightarrow \mathbb {R}$
with compact support. The quantum gas is then described by the self-adjoint operator

acting on the space of permutation-symmetric functions
$L^2_{\mathrm {sym}} \left (\Lambda ^N\right )$
, where
$\Lambda :=\left [-\frac {1}{2},\frac {1}{2}\right ]^3$
is the three-dimensional periodic torus, that is,
$\sum _{1\leq k\leq N}\Delta _{x_k}$
is defined as the closure of the Laplace operator acting on permutations symmetric and periodic
$C^2$
functions and
$x-y$
as well as
$x-z$
in Eq. (1) refer to the distance on the torus. We further assume that
$V_N$
defined in Eq. (1) is permutation symmetric in order to assure that
$H_N$
preserves permutation symmetry. The particular scaling in Eq. (1) with the number of particles N is referred to as the Gross-Pitaevskii regime and yields a short-range, but strong, interaction on the scale
$r=\frac {1}{\sqrt {N}}$
. This especially means that we are dealing with a dilute gas taking up a volume of the order
$Nr^3=\frac {1}{\sqrt {N}}$
. Due to the physical relevance of three-body interactions, which are for example responsible for 2% of the binding energy of liquid
$\mathrm {He}^4$
[Reference Murphy and Barker21] and 14% for water [Reference Mas, Bukowski and Szalewicz20], Dilute Bose gases with three-particle interactions have been studied extensively in [Reference Nam, Ricaud and Triay23, Reference Nam, Ricaud and Triay24, Reference Nam, Ricaud and Triay25, Reference Visconti26, Reference Junge and Visconti15], where the leading-order asymptotics of the ground state energy in the limit
$N\rightarrow \infty $
has been established as well as Bose-Einstein condensation (BEC) in the Gross-Pitaevskii regime. Here (BEC) refers to the observation that most of the particles occupy the state with zero momentum. Following this body of work, we will focus for the sake of simplicity on gases without two-body interactions and a repulsive three-body interaction, which is precisely the setting of [Reference Nam, Ricaud and Triay25, Conjecture 5].
In the Gross-Pitaevskii regime, the leading-order term in the asymptotics of the ground state energy has been derived in [Reference Nam, Ricaud and Triay23]

which is proportional to the number of particles N with a rather explicit constant
$b_{\mathcal {M}}(V)$
. Applying naive first-order perturbation theory, with
$-\sum _{1\leq k\leq N} \Delta _{x_k}$
as the unperturbed operator, would suggest the value
$\widehat {V}(0)$
for the constant
$b_{\mathcal {M}}(V)$
. It is, however, due to the singular nature of the scaling in Eq. (1) that we cannot ignore the presence of three particle correlations leading to a renormalized constant
$b_{\mathcal {M}}(V)<\widehat {V}(0)$
. In the following we will address a conjecture in [Reference Nam, Ricaud and Triay25], which claims that the subleading term in the asymptotic expansion of
$E_N$
is proportional to
$\sqrt {N}$
, see our main Theorem 1. The contributions to the ground state energy
$E_N$
of the order
$\sqrt {N}$
arise based on two-particle, three-particle, and four-particle correlations in the ground state. As a byproduct from the proof of Theorem 1, we obtain in addition that the ground state
$\Psi ^{\mathrm {GS}}_N$
of the operator
$H_N$
satisfies (BEC) with a rate
$\frac {1}{\sqrt {N}}$
, that is, we show that the ratio of particles outside the state with zero momentum compared to the total number of particles N is of the order
$O_{N\rightarrow \infty } \left (\frac {1}{\sqrt {N}}\right )$
. This is an improvement of the (BEC) result in [Reference Nam, Ricaud and Triay23], where the authors showed that the ratio is of the order
$o_{N\rightarrow \infty } \left (1\right )$
.
It is worth pointing out that much more is known for Bose gases with two-particle interactions, where the expansion of the ground state energy to second order is well known in the Gross-Pitaevskii regime, the thermodynamic limit, and interpolating regimes, see, for example, [Reference Boccato, Brennecke, Cenatiempo and Schlein3, Reference Brennecke, Caporaletti and Schlein7, Reference Brooks8, Reference Fournais and Solovej10, Reference Fournais and Solovej11, Reference Hainzl, Schlein and Triay13, Reference Nam, Napiorkowski, Ricaud and Triay22]. Furthermore, (BEC) is well known for the Gross-Pitaevskii regime and regular enough interpolating regimes, even with an (optimal) rate, see, for example, [Reference Adhikari, Brennecke and Schlein1, Reference Boccato, Brennecke, Cenatiempo and Schlein2, Reference Boccato, Brennecke, Cenatiempo and Schlein4, Reference Boccato and Seiringer5, Reference Brennecke, Brooks, Caraci and Oldenburg6, Reference Fournais9, Reference Haberberger, Hainzl, Nam, Seiringer and Triay12, Reference Lieb and Seiringer18], and the subleading term in the expansion of the ground state energy is known to be of the order
$O_{N\rightarrow \infty }(1)$
. This resolution of the energy is sharp enough to see the spectral gap, which is of the order
$O_{N\rightarrow \infty }(1)$
as well. For a Bose gas with three-particle interaction in the Gross-Pitaevskii regime we expect the spectral gap to be of the magnitude
$O_{N\rightarrow \infty }(1)$
, see the conjecture in [Reference Nam, Ricaud and Triay25]; however, the second-order expansion of the energy only allows for a resolution of the order
$O_{N\rightarrow \infty } \left (\sqrt {N}\right )$
, which is not sharp enough to see the spectral gap.
As it is not the goal of this manuscript to optimize the regularity of V, we will assume
$V\in C^\infty (\mathbb {R}^6)$
for the sake of convenience (although assuming, e.g.,
$V\in H^9(\mathbb {R}^6)$
would certainly be sufficient).
The correct constant
$b_{\mathcal {M}}(V)$
in the energy asymptotics Eq. (3) can be derived formally by making a translation-invariant ansatz for the correlation structure
$\varphi (x-u,y-u)$
between three particles at positions x, y, and u, where
$\varphi :\mathbb {R}^6\longrightarrow \mathbb {R}$
. Utilizing the matrix

and the modified Laplace operator
$\Delta _{\mathcal {M}}:=\left (\mathcal {M}\nabla _{\mathbb {R}^{3}\times \mathbb {R}^{3}}\right )^2$
, let us first express the action of the Laplace operator in relative coordinates as

For three particles, the energy of the trial state
$\Phi _\varphi (x-u,y-u):=1-\varphi (x-u,y-u)$
is then given by

Optimizing in
$\varphi $
leads to the definition

where
$\dot {H}^1(\mathbb {R}^d)$
refers to the space of functions
$g:\mathbb {R}^d\longrightarrow \mathbb {C}$
vanishing at infinity with
$|\nabla g|\in L^2 \left (\mathbb {R}^d\right )$
, see [Reference Lieb and Loss17, Section 8.3] where the notion
$D^1(\mathbb R^d)$
is used instead. It has been verified in [Reference Nam, Ricaud and Triay23] that a unique minimizer
$\omega $
to the variational problem in Eq. (4) exists satisfying the associated Euler–Lagrange equation

and the (modified) scattering length
$b_{\mathcal {M}}(V)$
describes the leading-order asymptotics of the ground state energy correctly, see Eq. (3). Notably, the solution
$\omega $
can formally be interpreted as a second-order correction to the condensate wavefunction
$\Psi \equiv 1$
, taking
$-2\Delta _{\mathcal {M}}+V$
, acting on functions vanishing at infinity, as the unperturbed operator and V, acting on the condensate
$\Psi \equiv 1$
, as the perturbation. Our main Theorem 1 confirms that the next term in the energy asymptotics in Eq. (3) is of the order
$O_{N\rightarrow \infty } \left (\sqrt {N}\right )$
due to contributions from the three-particle correlation
$\omega $
, as well as from two-particle and four-particle correlations.
In order to quantify the impact of two-particle correlations, we make a translation-invariant ansatz
$\xi (x - u)$
with
$\xi :\mathbb {R}^3\longrightarrow \mathbb {R}$
. Since
$(\Delta _x+\Delta _u)\xi (x-u)=(2\Delta \xi )(x-u)$
, we identify the kinetic energy of
$\xi $
as

Furthermore, the interaction energy of the wavefunction
$\Phi _\omega $
introduced above Eq. (4) with the state
$\Phi (x-u,y-u):=\xi (x-u)$
, which describes two correlated particles at position
$(x,u)\in \mathbb R^6$
and a particle in the condensate at position
$y\in \mathbb R^3$
, reads

where we have introduced the effective two-particle interaction

Adding up kinetic and interaction energy, and optimizing in
$\xi $
, immediately gives rise to the energy correction
$-\mu (V)$
with the proportionality constant
$\mu (V)$
defined as

Proceeding with the four-particle correlations, let us again commit to a translation-invariant ansatz
$\eta (x - u,y - u,z - u)$
with
$\eta :\mathbb {R}^9\longrightarrow \mathbb {R}$
. Defining
$\mathbb {V}:\mathbb {R}^9\longrightarrow \mathbb {R}$
and the matrix
$\mathcal {M}_*$
as

we can identify the action of the Laplace operator as

and the action of the potential on
$\eta $
as

Adding up the kinetic and potential energy of the four-particle correlation
$\eta $
, therefore yields

Moreover, the interaction energy
$V_*(x,y,z):=V(x,y)$
of
$\eta $
with the state

which describes three correlated particles at position
$(y,z,u)\in \mathbb R^9$
and a particle in the condensate at position
$x\in \mathbb R^3$
, reads

where we have introduced the function
$f(x,y,z):=V(x,y)\omega (y,z)$
. Combining kinetic, potential and interaction energy, yields

where we have introduced the functional

Note that
$\frac {f}{\mathbb {V}}$
is well defined and bounded on the support of
$\mathbb {V}$
, due to the sign of
$V\geq 0$
. Consequently, the corresponding energy correction is given by
$-\sigma (V)$
with

Finally, we observe in the presence of an additional particle at position z further interaction terms between
$\omega $
and itself. Utilizing again the potential
$V_*(x,y,z):=V(x,z)$
, and defining the states
$\Phi _1(x-u,y-u,z-u): =\omega (y-u,z-u)$
and
$\Phi _2(x-u,y-u,z-u): =\omega (x-u,z-u)$
, there are two relevant terms

giving rise to the energy correction
$\gamma (V)$

It is the content of our main Theorem 1, that
$\gamma (V),\mu (V)$
, and
$\sigma (V)$
describe the second-order correction to the leading-order asymptotics of the ground state energy
$E_N$
in Eq. (3), which is of the order
$O_{N\rightarrow \infty } \left (\sqrt {N}\right )$
. The mathematically precise implementation of the correlation structures
$\omega ,\xi $
, and
$\eta $
will be based on modified creation and annihilation operators, see, for example, [Reference Brooks8], and generalized (unitary) Bogoliubov transformations, see, for example, [Reference Boccato, Brennecke, Cenatiempo and Schlein2, Reference Boccato, Brennecke, Cenatiempo and Schlein3]. Furthermore, we show that the order
$O_{N\rightarrow \infty } \left (\sqrt {N}\right )$
term comes with a nonzero prefactor for a large class of potentials V.
Theorem 1. Let
$V\in C^\infty \left (\mathbb {R}^6\right )$
be a bounded and non-negative function with compact support, such that the function
$V_N$
defined in Eq. (1) is permutation symmetric. Furthermore, let
$\gamma (V),\mu (V)$
, and
$\sigma (V)\in \mathbb {R}$
be as in Eq. (10), Eq. (5) and Eq. (9) respectively, and let
$b_{\mathcal {M}}(V)$
be as in Eq. (4). Then the ground state energy
$E_N:=\inf \sigma (H_N)$
satisfies

Furthermore, there exists a
$\lambda (V)>0$
, such that for all
$0<\lambda \leq \lambda (V)$

Remark 1. While Theorem 1 concerns Bose gases in the ultra-dilute Gross-Pitaevskii regime occupying a volume of the order
$\frac {1}{\sqrt {N}}$
, the leading-order behavior of the ground state energy per unit volume
$e(\rho )$
is known in the thermodynamic regime as well as a function of the density
$\rho $
, see [Reference Nam, Ricaud and Triay24], and given in analogy to the leading-order asymptotics in Eq. (3) by

It is remarkable that the coefficients
$ \gamma (V)$
,
$\mu (V) $
, and
$\sigma (V)$
from Theorem 1 are defined in terms of variational problems on the unconfined space
$\mathbb {R}^{3d}$
and do not depend on the boundary conditions of the box
$\Lambda ^d$
. Substituting
$\rho $
with
$\frac {1}{\sqrt {N}}$
in Theorem 1 we therefore expect the second-order expansion of
$e(\rho )$
, as
$\rho \rightarrow 0$
, to be given by

This would be in contrast with the second-order expansion of a Bose gas with two-body interactions, where in the celebrated Lee-Huang-Yang formula, see, for example, [Reference Lee, Huang and Yang14, Reference Fournais and Solovej10] and [Reference Boccato, Brennecke, Cenatiempo and Schlein3] specifically for the periodic torus
$\Lambda $
, a summation of Fourier coefficients in the Pontryagin dual
$\left (2\pi \mathbb {Z}\right )^3$
of the locally compact group
$\Lambda $
appears. It is, however, expected that there is a corresponding Lee-Huang-Yang term for gases with three-body interactions, which should appear in a third-order expansion of the energy as a term of the order
$O_{N\rightarrow \infty }(1)$
.
Proof strategy of Theorem 1. Following the ideas in [Reference Nam, Ricaud and Triay23], respectively [Reference Boccato, Brennecke, Cenatiempo and Schlein2, Reference Boccato, Brennecke, Cenatiempo and Schlein3, Reference Fournais and Solovej10, Reference Fournais and Solovej11], which have been developed in the context of Bose gases with two-body interactions, we are going to unveil the correlation structure of the ground state with the help of a suitable coordinate transformation. Based on the strategy presented in [Reference Brooks8], our initial coordinate transformation will be of algebraic nature, that is, we introduce a new set of operators and observe that the many-body operator
$H_N$
is almost diagonal in these new variables. The algebraic approach immediately allows us to find satisfactory lower bounds on the ground state energy
$E_N$
. Furthermore, we show that this coordinate transformation can be realized in terms of a unitary map, at least in an approximate sense, which yields the corresponding upper bound on
$E_N$
.
In order to find a suitable transformation bringing
$H_N$
into a diagonal form, we observe that collisions between at most three particles will occur much more frequently compared to collisions between four or more particles, as we are in the dilute regime where the gas occupies only a volume of the magnitude
$\frac {1}{\sqrt {N}}$
. Consequently, we first look for a diagonalization of a gas with only three particles
$N=3$
, which will involve the three-particle correlation structure
$\omega $
, and subsequently lift it to a diagonalization of the full many-body problem. As it turns out, including the three-particle correlation structure is enough to identify the leading-order behavior of the ground state energy. To be more precise, utilizing the a priori information in Eq. (14), we are able to show at this point

We want to emphasize that the proof of Eq. (12) depends on our ability to neglect collisions between four or more particles, and we note that the correlation structure involves mostly particles outside the state with zero-momentum. It is therefore crucial to have strong a priori information regarding the number of particles outside the state with zero momentum, which we will refer to as excited particles. In the language of second quantization, the number of excited particles can naturally be expressed as

where N is the total number of particles,
$\mathcal {N}_0$
counts the number of particles with zero momentum and
$a_0$
is the annihilation operator corresponding to the zero-momentum state, see also Section 2 for a more comprehensive introduction. The following result, which has been verified in [Reference Nam, Ricaud and Triay23], tells us that the number of excited particles is indeed small compared to the total number of particles N, that is,

for any sequence of states
$\Psi _N$
satisfying
$\langle \Psi _N , H_N \Psi _N\rangle =E_N+O_{N\rightarrow \infty } \left (1\right )$
. Notably, the results in [Reference Nam, Ricaud and Triay23] concern particles in
$\mathbb R^3$
subject to a confining external potential, which can be generalized to our setting on the periodic torus without significant modifications as is explained in [Reference Nam, Ricaud and Triay25, Eq. (19)]. Using the a priori information in Eq. (14) then allows us to identify the leading-order asymptotics of the ground state energy
$E_N$
in Eq. (12). In addition, we obtain at this point an improved (BEC) result

for any sequence of states
$\Psi _N$
satisfying
$\langle \Psi _N , H_N \Psi _N\rangle =E_N+O_{N\rightarrow \infty } \left (N^{\frac {1}{4}}\right )$
, see also the subsequent Theorem 2 where we further improve this result up to a rate of the order
$N^{-\frac {3}{4}}$
, which we believe to be of independent interest. Based on the observation that our constructed unitary maps create an order
$O_{N\rightarrow \infty }(1)$
amount of excited particles, we conjecture that the optimal rate of condensation is of the magnitude
$\frac {1}{N}$
.
Finally, we use an additional coordinate transformation, which implements the two-particle and four-particle correlation structure
$\xi $
and
$\eta $
, together with the improved control on the number of excited particles in Eq. (15), in order to identify the coefficient C in front of the
$\sqrt {N}$
term in the energy asymptotics

Notably, collisions between four particles do contribute to the subleading term in the energy expansion in Eq. (11); however, in analogy to Eq. (12) we can dismiss collisions between five or more particles.
Theorem 2. Let V satisfy the assumptions of Theorem 1 and let
$\Psi _N$
be a sequence of elements in
$L^2_{\mathrm {sym}} \left (\Lambda ^N\right )$
satisfying
$\|\Psi _N\|=1$
and

for some constant
$D>0$
. Furthermore let
$\mathcal {N}$
be the operator counting the number of excitations introduced in Eq. (13). Then there exists a constant
$C>0$
, such that

Outline. In Subsection 1.1 we are first deriving the three-particle correlation structure for a model where the total number of particles is
$N=3$
. Following the strategy proposed in [Reference Brooks8], we are implementing in a systematic way the correlation structures from Subsection 1.1 for gases with many particles
$N\gg 1$
in Section 2. Using Bose-Einstein condensation of the ground state as an input, this allows us to immediately recover the leading-order behavior of
$E_N$
as a lower bound and, in the subsequent Section 3, also as an upper bound. Furthermore, we obtain at this point an improved version of (BEC) with a rate. In Section 4, we are going to describe the two-particle and four-particle correlation structure, which gives rise to the correction
$\mu (V)$
and the correction
$\sigma (V)$
defined in Eq. (5) and Eq. (9) respectively. It is the purpose of Subsection 4.2 to verify the lower bound in our main Theorem 1, wherein we use the improved (BEC) result, and the purpose of Section 5 to verify the corresponding upper bound. In the following Section 6, we can then provide (BEC) with a rate of the order
$N^{-\frac {3}{4}}$
, which concludes the proof of Theorem 2. The sign of
$\gamma (\lambda V) - \mu (\lambda V)- \sigma (\lambda V)$
is established in Section 7 for small
$\lambda>0$
, alongside other useful properties of the scattering solutions that describe the correlation structure. Finally Appendix A contains a collection of operator inequalities.
1.1 The three-body Problem
While naive first-order perturbation theory would tell us that the ground state energy
$E_N$
is to leading order given by the energy of the uncorrelated wavefunction
$\Gamma _0(x_1,\dots ,x_N):=1$

it is due to the presence of correlations in the ground state of the operator
$H_N$
, that the leading-order coefficient
$b_{\mathcal {M}}(V)$
in the energy asymptotics of
$E_N$
in Eq. (3) satisfies

In order to quantify the correlation energy
$\int _{\mathbb {R}^3}\int _{\mathbb {R}^3}V(x,y)\, \mathrm {d}x\mathrm {d}y-b_{\mathcal {M}}(V)$
, we are going to follow the frame work developed in [Reference Brooks8], and investigate first the corresponding three-particle operator
$H_{(3)}:=-\Delta _{3}+V_N$
acting on
$L^2(\Lambda ^3)$
, where
$\Delta _3:=\Delta _{x_1}+\Delta _{x_2}+\Delta _{x_3}$
, before we study the many particle operator
$H_N$
defined in Eq. (2). It will be our goal to find a transformation

that removes correlations between states with low momenta and states with high momenta, that is, we want to bring
$H_{(3)}$
into a block-diagonal form, which allows us to extract the correlation energy. It is content of Section 2 to lift the block-diagonalization from the three-particle problem, described by the transformation T, to a block-diagonalization of the many-particle operator
$H_N$
, which will allow us to identify the correlation energy for the many-particle problem.
Let us first specify the set of low momenta as either the set where all three particles occupy the zero-momentum state

or the set where at most one of the three particles is allowed to have non-zero momentum

where
$0\leq K<\infty $
is a parameter that we are going to specify later. For the purpose of extracting the correlation energy, it is enough to consider
$K:=0$
, however, for technical reasons it is going to be convenient later to consider positive values
$K>0$
as well. Having the set
$\mathcal {L}_K$
at hand, we can define the projection
$\pi _K$
onto states with low momenta as

where
$u_k(x):=e^{ikx}$
for
$k\in (2\pi \mathbb {Z})^3$
and
$u_{k_1}u_{k_2}u_{k_3}$
has to be understood as the function
$u_{k_1}(x_1)u_{k_2}(x_2)u_{k_3}(x_3)$
. Let us furthermore introduce the projection Q acting on
$L^2(\Lambda )$
as

Following the strategy in [Reference Brooks8], let R be the pseudoinverse of the operator
$Q^{\otimes 3}(-\Delta _{3}+V_N)Q^{\otimes 3}$
, that is, using the function
$h(t):=\frac {1}{t}$
for
$t\neq 0$
and
$h(0):=0$
the operator R is given as

and let us define the Feshbach-Schur like transformation

Note that T would be a proper Feshbach-Schur map, in case we would exchange
$Q^{\otimes 3}$
with the projection
$1-\pi _K$
; however, we prefer to work with
$Q^{\otimes 3}$
for technical reasons. Using the notation
$\{A+\mathrm {H.c.}\}:=A+A^*$
and the observation
$T^{-1}=1-RV_N \pi _K$
, yields

and by taking the adjoint we furthermore obtain

Combining both equations yields

where we have used in the final identity Eq. (20) the decomposition

Defining the almost block-diagonal renormalized potential
$\widetilde {V}_N$
as

and multiplying Eq. (20) from the left by
$T^\dagger $
and from the right by T, therefore yields the algebraic identity

The presence of
$ \left \{\left (1-\pi _K- Q^{\otimes 3} \right )\left (V_N -V_N R V_N \right ) \pi _K +\mathrm {H.c.}\right \}$
in
$\widetilde {V}_N$
, which are the terms that violate the block-diagonal structure, is due to the usage of
$Q^{\otimes 3}$
instead of
$1-\pi _K$
; however, it turns out that these terms do not contribute to the correlation energy to leading order. One therefore expects to read off the leading-order coefficient
$b_{\mathcal {M}}(V)$
in the asymptotic expansion of the ground state energy
$E_N$
in Eq. (3) from the matrix entries of the renormalized potential

As we are going to verify in Lemma 16, we have indeed the asymptotic result

2 First-order lower bound
It is the goal of this Section to bring, in analogy to Eq. (22), the many-particle operator
$H_N$
in an approximate block-diagonal form, which allows us to obtain an asymptotically correct lower bound on the ground state energy
$E_N$
in Corollary 2. First, we are going to rewrite the operator
$H_N$
defined in Eq. (2) in the language of second quantization. For this purpose let
$a_k^\dagger $
denote the operator that creates a particle in the mode
$u_k$
, that is, for
$\Phi _n\in L^2_{\mathrm {sym}} \left (\Lambda ^n\right )$
we define
$a^\dagger _k\Phi _n\in L^2_{\mathrm {sym}} \left (\Lambda ^{n+1}\right )$
as

where
$\Xi _{n+1}$
is the orthogonal projection onto
$L^2_{\mathrm {sym}} \left (\Lambda ^{n+1}\right )\subseteq L^2 \left (\Lambda ^{n+1}\right )$
, and we write
$a_k$
for its adjoint, which annihilates a particle in the mode
$u_k(x):=e^{ik\cdot x}$
. With creation and annihilation operators at hand, we can write

where
$\left (V_N\right )_{ijk,\ell m n}$
are the matrix elements of
$V_N$
w.r.t. to the basis
$u_i u_j u_k$
defined below Eq. (17). If not indicated otherwise, we will always assume that indices run in the set
$(2\pi \mathbb {Z})^3$
, which we will usually neglect in our notation, and we write
$k\neq 0$
in case the index runs in the set
$(2\pi \mathbb {Z})^3\setminus \{0\}$
. Note that the operator on the right hand side of Eq. (23) is naturally defined on the full Fock space

while the left-hand side is only defined on
$L^2_{\mathrm {sym}} \left (\Lambda ^N\right )\subseteq \mathcal {F} \left (L^2 \left (\Lambda \right )\right )$
, and therefore Eq. (23) has to be understood as being restricted to the subspace
$L^2_{\mathrm {sym}} \left (\Lambda ^N\right )$
. Furthermore, we observe that
$V_N$
is a translation-invariant multiplication operator, and therefore the matrix elements of
$V_N$
satisfy
$\left (V_N\right )_{ijk,\ell m n}=0$
in case
$i+j+k\neq \ell +m+n$
and otherwise

Following the strategy proposed in [Reference Brooks8], we are going to introduce a many-particle counterpart to the three particle map T defined in Eq. (19), which is realized by the set of operators


Here denotes the matrix elements of T. The following Lemma 1 is the many-particle counterpart to Eq. (22), in the sense that it provides an (approximate) block-diagonal representation of the operator
$H_N$
in terms of the new variables
$c_k$
and
$\psi _{ijk}$
.
Lemma 1. Let
$\widetilde {V}_N$
be the operator defined in Eq. (21). Then we have

where the residual term
$\mathcal {E}$
is defined as

Proof. Using the permutation symmetry of T, we first identify
$\sum _{k}|k|^2 (c_k-a_k)^\dagger (c_k-a_k)$
as

where
$\Delta _3$
is the Laplace operator on
$L^2(\Lambda )^{\otimes 3}$
. Similarly

Since

we obtain

We observe that
$T^\dagger \left (-\Delta _3+\widetilde {V}_N\right )T + \Delta _3=V_N$
by Eq. (22), which concludes the proof by the representation of
$H_N$
in second quantization, see Eq. (23).
Making use of the sign
$ (1-\pi _K) V_N (1-\pi _K) \geq 0$
, we immediately obtain that

Therefore Lemma 1 allows us to bound
$H_N$
from below by

where we have used the fact that
$\psi _{ijk}=a_i a_j a_k$
in case one of the indices is zero, which is a direct consequence of the observation that
$(T-1)_{ijk,\ell mn}=0$
in case one of the indices in
$\{i,j,k\}$
is zero. Note that we can write

with A and B defined as

Using the sets
$\mathcal {L}^{(z)}:=\{(z,0,0),(0,z,0),(0,0,z)\}$
, let us first analyze the term involving A

Together with the definition of the coefficients

we can write

and utilizing the permutation symmetry of
$V_N -V_N R V_N$
we furthermore obtain for
$z\neq 0$

Therefore,

To keep the notation light, we do not explicitly indicate the N dependence of
$\lambda _{k,\ell }$
. Similarly

Putting together Eq. (28), Eq. (30) and Eq. (31) yields

where we define the operator
$\mathbb {Q}_K$
and the error term
$\mathcal {E}'$
as


Let us furthermore introduce the particle number operator

which counts the number of excited particles, that is, the number of particles with momentum
$k\neq 0$
. Since we have the operator identity
$\sum _{k}a_k^\dagger a_k=N$
on the Hilbert space
$L^2_{\mathrm {sym}}(\Lambda ^N)\subseteq \mathcal {F} \left (L^2 \left (\Lambda \right )\right )$
, we observe that
$a_0^\dagger a_0=N-\mathcal {N}$
, see also Eq. (13), that is, the number of particles with momentum
$k=0$
is given by the difference between the total number of particles N and the number of excited particles.
In order to control the terms arising in Eq. (32), it is imperative to understand the asymptotic behavior of the coefficients
$\lambda _{k,\ell }$
and the matrix entries
$T_{ijk,\ell m n}$
defined below Eq. (26). Since we want to focus our attention on the many-body analysis, we will postpone our study of the scattering coefficients to Section 7. For the convenience of the reader, we are going to state the relevant results, which are proven in Lemma 15 and Lemma 16 respectively,




Furthermore, we need the following result, which is verified in Lemma A2, see Appendix A, and which allows us to compare the new operators
$c_k$
with the annihilation operators
$a_k$
,


for
$0\leq \tau \leq 1$
,
$0\leq \sigma <\frac {1}{2}$
and integers
$s\geq 0$
. Utilizing Eq. (35)-(40), the following Lemma 2, Lemma 3 and Lemma 4, provide relevant bounds on the various terms appearing in Eq. (32), which will be instrumental in order to establish that the ground state energy
$E_N$
of
$H_N$
is, to leading order, bounded from below by
$\frac {1}{6}b_{\mathcal {M}}(V) N$
, see Corollary 2. In our first Lemma 2 we provide a lower bound on

for K large enough, which is an operator that is at most quadratic in the variables
$a_k$
and
$a_k^\dagger $
for
$k\neq 0$
.
Lemma 2. Let
$b_{\mathcal {M}}(V)$
be the modified scattering length defined in Eq. (4). Then there exists for all
$\tau ,\alpha>0$
, a constant
$K_0(\tau ,\alpha )$
and for all
$K\geq K_0(\tau ,\alpha )$
a constant
$C_K>0$
, such that

for
$K\geq K_0(\tau ,\alpha )$
.
Proof. First of all we observe that we can write

for a suitable
$D>0$
. Defining

we therefore obtain in combination with Eq. (37) and Eq. (38)

for a suitable constant
$D>0$
. Since

and since we have

we obtain

for a suitable, K-dependent, constant
$C_K$
. Finally, we choose
$K_0(\tau ,\alpha )$
large enough such that
$\frac {1}{2}b_{\mathcal {M}}(V)K_0(\tau ,\alpha )^{-2\tau }\leq \alpha $
, and therefore we have for all
$K\geq K_0(\tau ,\alpha )$

In the Lemma 3, we will provide estimates on the residual term
$\mathcal {E}$
defined in Lemma 1, which will allow us to compare the size of
$\mathcal {E}$
with the kinetic energy
$ \sum _{k}|k|^2 c_k^\dagger c_k$
in the variables
$c_k$
. Before we come to Lemma 3, we need the following Corollary 1, which is a consequence of Eq. (40) and allows us to estimate monomials in the operators
$a_k$
and
$a_k^\dagger $
by the kinetic energy in the variables
$c_k$
and powers of the particle number operator
$\mathcal {N}$
.
Corollary 1. Let
$\mathcal {K}_{\tau ,t}:=\sum _{i=1}^t (-\Delta _{x_i})^\tau $
. Given
$0\leq \tau \leq 1$
, and integers
$s,t\geq 1$
and
$\alpha ,\beta \geq 0$
, there exist
$\delta>0$
and
$C>0$
, such that for
$\epsilon>0$

where
$G:\left (\mathrm {ran}Q\right )^{\otimes s}\longrightarrow \left (\mathrm {ran}Q\right )^{\otimes t}$
and
$X:\mathcal {F}\left (L^2(\Lambda )\right )\longrightarrow \mathcal {F}\left (L^2(\Lambda )\right )$
. In case
$s=0$

Proof. Let us define for
$s,t\geq 1$
the operator-valued vector and operator-valued matrix

which allow us to represent

Using the fact that
$\|\Upsilon \|\leq \left \|\mathcal {K}_{\tau ,t}^{-\frac {1}{2}}G\, \mathcal {K}_{\tau ,s}^{-\frac {1}{2}}\right \|\big \| \mathcal {N}^{-\frac {\beta }{2}}X\mathcal {N}^{-\frac {\alpha }{2}}\big \|$
, we obtain by Cauchy-Schwarz

Since
$\mathcal {N}\leq N$
we furthermore have by Eq. (40)

With Corollary 1 at hand, we can verify the subsequent Lemma 3.
Lemma 3. For
$K\geq 0$
, there exists a constant
$C_K>0$
, such that

Proof. Let us denote with
$\mathcal {I}\subseteq (2\pi \mathbb {Z})^{3\times 3}$
the index set

Then we define for
$I=(I_1,I_2,I_3),I'=(I^{\prime }_1,I^{\prime }_2,I^{\prime }_3)\in \mathcal {I}$
the operator
$K^{(I,I')}$
acting on
$L^2(\Lambda )$
and the operator
$G^{(I,I')}$
acting on
$L^2(\Lambda )^{\otimes 2}$
as

as well as
$\mathcal {K}_{\tau ,2}:=(-\Delta _{x_1})^{\tau }+(-\Delta _{x_2})^{\tau }$
acting on
$L^2(\Lambda )^{\otimes 2}$
. Then we can write
$\mathcal {E}$
as

By the weighted Schur test, the operator norm of
$\mathcal {K}_{\tau ,2}^{-\frac {1}{2}} G^{(I,I')}\mathcal {K}_{\tau ,2}^{-\frac {1}{2}}$
is bounded by

where we define
$\alpha ^{(I,I')}:=\sup _{i'j'}\sum _{ij}\frac {|G^{(I,I')}_{ij,i' j'}|}{|i|^{2\tau }+|j|^{2\tau }}$
. Let us furthermore introduce
$s:=I_1+I_2+I_3$
and
$s':=I^{\prime }_1+I^{\prime }_2+I^{\prime }_3$
. Making use of Eq. (36), we obtain for the concrete choice
$\tau :=\frac {2}{3}$

Consequently
$\|\mathcal {K}_{\tau ,2}^{-\frac {1}{2}} G^{(I,I')}\mathcal {K}_{\tau ,2}^{-\frac {1}{2}}\|\lesssim N^{-4}$
. Furthermore, the operator

satisfies
$\|X^{(I,I')}\|\leq N^3$
. Therefore we obtain by Corollary 1

Again by Lemma 15 we have
$\|K^{(I,I')}\|\lesssim N^{-\frac {7}{2}}$
, which concludes the proof by Corollary 1, together with the observation that the set
$\mathcal {I}$
in the definition of
$\mathcal {E}$
in Eq. (42) is finite.
The next Lemma 4 will give us sufficient bounds on the error term
$\mathcal {E}'$
defined in Eq. (34), which will be responsible for the appearance of an order
$O_{N\rightarrow \infty } \left (\sqrt {N}\right )$
error in the main results of this Section Theorem 3 and Corollary 2.
Lemma 4. There exist constants
$C,C_K>0$
such that for
$K\leq \sqrt {N}$
, where K is as in the definition of
$\pi _K$
below Eq. (17), and
$\epsilon>0$


Furthermore, we have

Proof. Given
$m\in \{0,1\}$
, let us define the operator
$X:=a_0^\dagger a_0^3\frac {\mathcal {N}^m}{N^m}$
and the coefficients

and observe that by Eq. (36) there exists a constant
$C>0$
such that

where we have assumed w.l.o.g. that
$|n|\leq \sqrt {N}$
, since
$\Lambda ^{(n)}_{\ell ,k}=0$
in case
$|n|>K$
and
$K\leq \sqrt {N}$
. In order to verify Eq. (43), respectively Eq. (45), let us write

Regarding the first term in Eq. (47), note that we have for
$\epsilon>0$
the estimate

Using
$|\lambda _{k,\ell }|\lesssim N^{-2}(1+\frac {|\ell |^2}{N})^{-1}$
, see Eq. (35), we have
$\sum _{|\ell |>K}\frac {|\lambda _{0,\ell }|^2}{\ell ^2}\lesssim N^{-\frac {7}{2}}$
and
$\frac {|\lambda _{0,\ell }|^2}{\ell ^2}\leq \frac {1 }{N^4 (K^2+1)}$
for
$|\ell |>K$
, and therefore we obtain for such K

where we have used
$X^\dagger X\lesssim N^2 \frac {\mathcal {N}^{2m}}{N^{2m}}$
and
$[X,\mathcal {N}]=0$
. Regarding the second term in Eq. (47), let us use
$\sum _{|\ell |>K}|\ell |^{-\frac {1}{2}}|N^4\lambda _{0,\ell }|^2 a_{\ell }^\dagger a_{\ell }\lesssim \frac {\mathcal {N}}{\sqrt {K+1}}$
as well as the fact that

by Eq. (39), to estimate for
$\kappa>0$

Regarding the final term in Eq. (47) we have that
$\sum _{\ell }|\Lambda ^{(n)}_{\ell ,0}|\lesssim N^{-\frac {3}{2}}$
by Eq. (46), and hence

For
$m=0$
, the choice
$\kappa :=\epsilon $
yields Eq. (43) and for
$m=1$
the choice
$\kappa :=\sqrt {N}$
and
$\epsilon :=\frac {1}{\sqrt {N}}$
yields Eq. (45).
Regarding the proof of Eq. (44), let us define the operators
$d_{k,\ell }:=\lambda _{k,\ell } N^{\frac {3}{2}} a_{k-\ell }^\dagger a_k$
and write
$a^\dagger _\ell d_{k,\ell }=c_\ell ^\dagger d_{k,\ell }+d_{k,\ell } (c_\ell -a_\ell )^\dagger +[(c_\ell -a_\ell )^\dagger ,d_{k,\ell }]$
. We compute

where we define the coefficients

Using again Eq. (46) and
$|\lambda _{k,\ell }|\lesssim N^{-2}(1+\frac {|\ell |^2}{N})^{-1}$
, we immediately obtain
$|\mu ^{(1)}_k|\lesssim \frac {1}{\sqrt {N}}$
,
$|\mu ^{(2)}_{k,n}|\lesssim \frac {1}{N}$
,
$|\mu ^{(3)}_{k,i,\ell }|\lesssim N^{-\frac {3}{2}}$
and
$\sum _{i}|\mu ^{(3)}_{k,i,\ell }|\lesssim \frac {1}{N}$
, and therefore by Cauchy-Schwarz

Consequently,

Similar to the proof of Eq. (43), we observe that
$\frac {1}{K^3}\sum _{0<|k|\leq K,\ell }d_{k,\ell } d_{k,\ell }^\dagger \lesssim \frac {\mathcal {N}}{N}\mathcal {N}$
and, using Eq. (39),

Having Lemma 2, Lemma 3 and Lemma 4 at hand, we can use the lower bound in Eq. (32) in order to derive the following Theorem 3, which provides strong lower bounds on the quantity . Note, however, that Theorem 3 is only applicable for states
$\Psi $
which satisfy the strong condition that
$\Psi $
is in the spectral subspace
$\mathcal {N}\leq \epsilon N$
, that is,

where the orthogonal projection is defined by means of functional calculus. In the following we will refer to the property in Eq. (48) as (BEC) in the spectral sense
. Furthermore, we refer to a Hilbert space element
$\Psi $
as a state, in case it satisfies
$\|\Psi \|=1$
.
Theorem 3. There exist constants
$\delta ,C,N_0>0$
and
$\epsilon>0$
, such that

for any state
$\Psi $
satisfying
and
$N\geq N_0$
, where
$\mathcal {N}:=\sum _{k\neq 0}a_k^\dagger a_k$
.
Proof. By Eq. (32) together with the estimates in Lemma 2, Lemma 3 and Lemma 4 we have for
$\alpha ,\tau ,\epsilon '>0$
and
$K\geq K_0(\alpha ,\tau )$

where
$C,C_{K,\epsilon '}$
and
$K_0(\alpha ,\tau )$
are suitable constants. In the following let
$\Psi $
be a state satisfying
and define
$\Psi _k:=c_k \Psi $
. By the definition of
$c_k$
it is clear that
, and therefore

In a similar fashion we have

Furthermore, note that for a suitable constant
$D_1>0$

by Eq. (40) for
$\tau <\frac {1}{2}$
, and by Eq. (39) we have

for a suitable constant
$R>0$
. Using
with
$\epsilon $
small enough such that
$R\epsilon <1$
, we therefore obtain

for a suitable constant
$D_2>0$
. Combining Eq. (49)-(53), therefore yields for suitable constants D and
$D_{K,\epsilon ',\alpha }$
, and
$0<\epsilon <\frac {1}{R}$
,

We can now make our choice of parameters concrete. First we take choose
$\tau $
such that
$0<\tau < \frac {1}{2}$
and
$\alpha ,\epsilon '>0$
small enough, such that
$D\left (\epsilon '+\alpha \right )<\frac {1}{2}$
, and then we take
$K\geq K_0(\alpha ,\tau )$
large enough, such that

Finally, we take
$0<\epsilon <\frac {1}{R}$
small enough and N large enough, such that

It has been verified in [Reference Nam, Ricaud and Triay23], for the more general setting of particles being confined by an additional external potential, that any approximate ground state
$\Psi _N$
of the operator
$H_N$
satisfies complete Bose-Einstein condensation

Adapting the localization procedure presented in [Reference Lieb and Solovej19, Theorem A.1] in the form stated in [Reference Lewin, Nam, Serfaty and Solovej16, Proposition 6.1] for the following Lemma 5 allows us to lift Bose-Einstein condensation in the sense of Eq. (54), to Bose-Einstein condensation in the spectral sense, which is a crucial assumption of the previous Theorem 3.
Lemma 5. Let
$\Psi $
satisfy
with
$\delta \leq N$
. Then there exists a constant
$C>0$
, such that there exists for all
$1\leq M\leq N$
states
$\Phi $
satisfying
and

Furthermore, there exists a state
$\widetilde {\Phi }$
such that
and

Proof. In the following let
$f,g:\mathbb {R}\rightarrow [0,1]$
be smooth functions satisfying
$f^2+g^2=1$
and
$f(x)=1$
for
$x\leq \frac {1}{2}$
, as well as
$f(x)=0$
for
$x\geq 1$
, and let us define
$ m:= \|f \left (\frac {\mathcal {N}}{M}\right )\Psi \|^2$
and

Note that
$\|\Phi \|=\|\widetilde \Phi \|=1$
,
$0\leq m \leq 1$
and clearly we have
and
. Making use of the algebraic identity

with the residual term

we obtain

In order to estimate , let
$\pi ^0$
denote the projection onto the constant function in
$L^2(\Lambda )$
and
$\pi ^1:=1-\pi ^0$
. Then we can rewrite
$\mathcal {E}$
as

with

and
$\#_I$
counting the number of indices in I that are equal to
$1$
. Using
$0\leq X_{I,J}\leq X$
, where

we obtain by the Cauchy-Schwarz inequality

In the following we want to show that for any
$I\in \{0,1\}^3$
, the
$\Psi $
-expectation value of the corresponding term appearing in the sum on the right side of Eq. (57) is of the order
$\sqrt {N}M+N$
. For
$I=(0,0,0)$
we have

Similarly,

in the case
$I=(0,0,1)$
. Regarding the case
$I=(0,1,1)$
, let us first observe that we have the upper bound

with the constant
$C_N$
being defined as

Due to our regularity assumptions on V we have
$\big |V \big (N^{-\frac {1}{2}}t\big )\big |\lesssim \frac {1}{1+N^{-1}|t|^2}$
and therefore

where we have used the Hardy-Littlewood inequality in the first estimate of Eq. (59). Since

we obtain by Eq. (58)

where we have used the assumption
$\delta \leq N$
and the upper bound on
$E_N$
derived in Theorem 4. The only distinguished case left is
$I=(1,1,1)$
. We start its analysis by defining

which allows us to estimate, using the Cauchy-Schwarz inequality,

From the previous cases we know that

and therefore . Summarizing what we have so far yields the inequality

Using and the simple observation that
immediately yields Eq. (55), and using
we obtain for a suitable constant
$C>0$

In order to derive Eq. (56), we note that
$\mathcal {N}=f \left (\frac {\mathcal {N}}{M}\right )\mathcal {N} f \left (\frac {\mathcal {N}}{M}\right )+g \left (\frac {\mathcal {N}}{M}\right ) \mathcal {N} g \left (\frac {\mathcal {N}}{M}\right )$
and therefore

Before we come to the lower bound on the ground state energy
$E_N$
in the main result of this Section Corollary 2, let us first state the corresponding upper bound in the subsequent Theorem 4, which has essentially been verified in [Reference Nam, Ricaud and Triay23]. To be precise, it has been shown in [Reference Nam, Ricaud and Triay23] that
$E_N\leq \frac {1}{6}b_{\mathcal {M}}(V)N+CN^{\frac {2}{3}}$
, and, as is explained in [Reference Nam, Ricaud and Triay25], the method in [Reference Nam, Ricaud and Triay23] can be improved to yield
$E_N\leq \frac {1}{6}b_{\mathcal {M}}(V)N+CN^{\frac {1}{2}}$
as well. However, since the computations in the proof of Theorem 4 are relevant for the proof of the upper bound in Theorem 1, and for the sake of completeness, we are nevertheless going to verify Theorem 4 in detail in the subsequent Section 3.
Theorem 4. There exists a constant
$C>0$
such that the ground state energy
$E_N$
is bounded from above by
$E_N\leq \frac {1}{6}b_{\mathcal {M}}(V)N+C\sqrt {N}$
.
Using Bose-Einstein condensation in the spectral sense, Theorem 3 allows us to derive an asymptotically correct lower bound on the ground state energy in Corollary 2 with an error of the order
$\sqrt {N}$
, see Eq. (62). In this context we call
$\Psi _N$
an approximate ground state, in case
$\|\Psi _N\|=1$
and there exists a constant
$C>0$
such that

Note that the assumption in Eq. (60) is more restrictive compared to the one employed in [Reference Nam, Ricaud and Triay23], where the authors call
$\Psi _N$
an approximate ground state in case
$\|\Psi _N\|=1$
and

The fact that Eq. (60) implies Eq. (61) follows immediately from the leading-order asymptotics in Eq. (3), which has been verified in [Reference Nam, Ricaud and Triay23], together with the trivial lower bound .
Corollary 2. The ground state
$\Psi _N^{\mathrm {GS}}$
of the operator
$H_N$
satisfies for a suitable
$C>0$

and we have the lower bound

Furthermore there exists a constant
$C>0$
and states
$\Phi _N$
, such that
$\Phi _N$
is an approximate ground state of
$H_N$
satisfying (BEC) in the spectral sense with rate
$\frac {1}{\sqrt {N}}$
, that is,

and we have the estimate on the kinetic energy
$\left \langle \Phi _N, \sum _{k}|k|^2 c_k^\dagger c_k \Phi _N \right \rangle \leq C\sqrt {N}$
.
Proof. From the results in [Reference Nam, Ricaud and Triay23] we know that the ground state
$\Psi ^{\mathrm {GS}}_N$
of
$H_N$
satisfies

Consequently, we know by Lemma 5 that there exist states
$\xi _N$
satisfying

and , where we choose
$\epsilon>0$
as in Theorem 3. By Theorem 3 and Theorem 4 we therefore obtain for a suitable constant
$C>0$

This immediately implies
$E_N\geq \frac {1}{6} b_{\mathcal {M}}(V)N-C\sqrt {N}$
for a suitable constant
$C>0$
and

as well as
$\left \langle \xi _N , \sum _{k}|k|^2 c_k^\dagger c_k \xi _N \right \rangle =O_{N\rightarrow \infty } \left (\sqrt {N}\right ) $
. Furthermore, there exist by Lemma 5 states
$\widetilde {\xi }_N$
satisfying
and

In the following we show by a contradiction argument, similar to the one employed in the proof of [Reference Boccato, Brennecke, Cenatiempo and Schlein4, Theorem 1.1], that

For this purpose let us assume that Eq. (67) is violated, that is, we assume that there exists a subsequence
$N_j$
and a constant
$C>0$
such that
. Let us complete this subsequence to a proper sequence by defining
$\xi ^{\prime }_N:=\widetilde \xi _N$
in case
$N=N_j$
for some j and
$\xi ^{\prime }_N:=\Psi ^{\mathrm {GS}}_N$
otherwise. Clearly
$\xi ^{\prime }_N$
is a sequence of approximate ground states, see Eq. (60), and as such
$\xi ^{\prime }_N$
satisfies complete (BEC) by the results in [Reference Nam, Ricaud and Triay23], that is,
. This is, however, a contradiction to

which concludes the proof of Eq. (67). Combining Eq. (66) and Eq. (67) yields

for a suitable constant
$C>0$
. Applying again Lemma 5 for the state
$\Psi ^{\mathrm {GS}}_N$
and
$M:=K \sqrt {N}$
, we obtain states
$\Phi _N$
satisfying
and

for a large enough
$C>0$
. Consequently
$\Phi _N$
is a sequence of approximate ground states for
$K>2C$
. Finally we notice that the states
$\Phi _N$
satisfy the chain of inequalities in Eq. (64) as well, and therefore

3 First-order upper bound
It is the goal of this section to introduce a trial state
$\Gamma $
, which simultaneously annihilates the variables
$c_k$
for
$k\neq 0$
and
$\psi _{\ell m n}$
in case
$(\ell ,m,n)\neq 0$
, at least in an approximate sense, which allows us to verify the upper bound on the ground state energy
$E_N$
in Theorem 4. For the rest of this Section we specify the parameter K introduced above the definition of
$\pi _K$
in Eq. (17) as
$K:=0$
, which especially means that with
$\eta _{ijk}:=(T-1)_{i j k, 0 0 0}$
we have


In order to find a suitable state
$\Gamma $
, let
$\eta _{ijk}:=(T-1)_{i j k, 0 0 0}$
and let us follow the strategy in [Reference Nam, Ricaud and Triay23], respectively in the case of Bose gases with two-particle interactions see, for example, [Reference Boccato, Brennecke, Cenatiempo and Schlein3, Reference Brennecke, Caporaletti and Schlein7, Reference Hainzl, Schlein and Triay13], by defining the generator

of a unitary group
$U_s:=e^{s\mathcal {G}^\dagger -s\mathcal {G}}$
and
$U:=U_1$
. The generator
$\mathcal {G}$
is chosen, such that

and in particular, as we show this section, the unitary U has the property
$U^{-1}c_k U\approx a_k$
and
$U^{-1}\psi _{ijk}U\approx a_i a_j a_k$
in a suitable sense. Denoting with

the constant function in
$L^2_{\mathrm {sym}} \left (\Lambda ^N\right )$
, that is,
$a_k \Gamma _0=0$
for
$k\neq 0$
, we observe that
$\Gamma :=U \Gamma _0$
is a suitable trial state for the (approximate) annihilation of
$c_k$
, given
$k\neq 0$
, and
$\psi _{\ell m n}$
, given
$(\ell ,m,n)\neq 0$
, where the action of the unitary U introduces a three-particle correlation structure on the completely uncorrelated wavefunction
$\Gamma _0$
. We note at this point, that the action of the unitary operator U only creates an
$O(1)$
amount of particles, in the sense that

as is proven in Appendix A, see Lemma A1.
For the purpose of verifying that
$U^{-1}\psi _{ijk}U$
is approximately identical to
$a_i a_j a_k$
, we first apply Duhamel’s formula, which yields

Furthermore, note that we can write

using the definition

Therefore we can identify the transformed operators
$U^{-1}\psi _{ijk}U$
as

where we have used Duhamel’s formula to express
$U^{-1}\eta _{ijk}a_0^3U-U_{-s}\eta _{ijk}a_0^3U_s$
. The following Lemma 6 demonstrates that we can treat the quantities
$\delta _1 \psi $
and
$\delta _2 \psi $
in Eq. (74) as error terms. In order to formulate Lemma 6, recall the set
$\mathcal {L}_0:=\{(0,0,0)\}$
from Eq. (16) and let us define

and the potential energy
$\mathcal {E}_{\mathcal {P}}$
of an operator-valued three particle vector
$\Theta _{i_1 i_2 i_3}$
as

To keep the notation light, we will occasionally write
$\mathcal {E}_{\mathcal {P}}(\Theta _{i_1 i_2 i_3})$
for
$\mathcal {E}_{\mathcal {P}}(\Theta )$
with dummy indices
$i_1 i_2 i_3$
.
Lemma 6. There exists a constant
$C>0$
, such that


Furthermore,
$\mathcal {E}_P \left (\big [a_{i_1} a_{i_2} a_{i_3} ,\mathcal {G}^\dagger \big ]\right )\leq C N^{-\frac {3}{2}}(\mathcal {N}+1)^5$
.
Proof. Let us define
$A_j$
as the set of all s such that
$(-s, j-s, 0)\in A$
and

Making use of the fact that
$V_N\geq 0$
, we obtain by the Cauchy-Schwarz inequality

By Lemma 15 and the fact that
$|(V_N)_{-s (j-s) 0,-t (j-t) 0}|\lesssim \frac {1}{N^2}$
, we have
$N^4\sum _j \alpha _j\lesssim C N^{\frac {1}{2}}$
, which concludes the proof of Eq. (76). In order to verify Eq. (77), let us first define

Then we obtain, using the sign
$V_N\geq 0$
and a Cauchy-Schwarz estimate,

Proceeding as in the proof of Eq. (76), we obtain

Regarding the term
$\mathcal {E}_{\mathcal {P}} \left (\widetilde {\delta }_{2,\mathrm {id}} \psi \right )$
, let us define

for
$(i_1,i_2,-j_3-j_4)\in A$
and
$(j_1,j_2,-i_3-i_4)\in A$
, and
$(G)_{i_1.. i_4, j_1..j_4}:=0$
otherwise. Then

with
$(G')_{i_1.. i_3, j_1..j_3}:=\sum _k G_{i_1.. i_3 k, j_1..j_3 k}$
and
$(G")_{i_1 i_2, j_1 j_2}:=\sum _{k_1 k_2} G_{i_1 i_2 k_1 k_2, j_1 j_2 k_1 k_2}$
. In the following let us study the term involving
$G"$
, which is responsible for the largest contribution. Since
$(V_N)_{i_1 i_2 (-k_1-k_2),j_1 j_2 (-k_1-k_2)}=(V_N)_{i_1 i_2 0,j_1 j_2 0}$
, see Eq. (24), we obtain

where we have used

see Lemma 15. Proceeding similarly for the other terms in Eq. (78), concludes the proof of Eq. (77). Regarding the bound on
$\mathcal {E}_P \left (\big [a_{i_1} a_{i_2} a_{i_3} ,\mathcal {G}^\dagger \big ]\right )$
, let us identify

where
$\mathbb {A}:=\frac {1}{6}\sum _{ijk}\eta _{ijk}a_i a_j a_k$
. Due to the sign
$V_N\geq 0$
and the permutations symmetry of
$V_N$
, as well as to the fact that there are
$6$
permutations of the set
$\{1,2,3\}$
, we can bound the operator
$ \mathcal {E}_P \left (\big [a_{i_1} a_{i_2} a_{i_3} ,\mathcal {G}^\dagger \big ]\right )$
from above by

In the following we focus on , the other terms can be treated in a similar fashion. By Lemma 15 we have
$\mathbb {A}^\dagger \mathbb {A}\lesssim N^{-3}(\mathcal {N}+1)^3$
, and therefore

where
$A^0$
contains all pairs
$(jk)$
such that
$(0jk)\in A$
.
As a consequence of Lemma 6, we obtain that the trial state
$\Gamma $
defined below Eq. (71) has a potential energy
$\mathcal {E}_{\mathcal {P}}(\psi )$
of the order
$O_{N\rightarrow \infty } \left (\sqrt {N}\right )$
, see the following Corollary 3.
Corollary 3. There exists a constant
$C>0$
, such that
$ \left \langle \Gamma , \mathcal {E}_{\mathcal {P}}(\psi +\delta _1 \psi )\Gamma \right \rangle \leq C$
and

Proof. Recall that we can express the transformed quantity
$U^{-1}\psi _{ijk} U$
by Eq. (74) as

where we have used Duhamel’s formula to express
$U^{-1}\eta _{ijk}a_0^3U-U_{-s}\eta _{ijk}a_0^3U_s$
. Using the sign
$V_N\geq 0$
and Lemma 6, we estimate using the Cauchy-Schwarz inequality

for suitable
$C,C'$
, where we utilize Eq. (72) in order to estimate
$U_{-s}(\mathcal {N}+1)^4 U_s$
. Similarly

follows from Lemma 6. Regarding the term in the last line of Eq. (74), we note that

follows from an analogous argument as we have seen in the proof of Lemma 6 and

by Eq. (36). Therefore

where we have used Eq. (72) again. Using
$a_i a_j a_k \Gamma _0=0$
in case
$(ijk)\in A$
, we obtain by Eq. (74) together with Eq. (79), Eq. (80) and Eq. (81) for a suitable constant C

Analogously we obtain
$ \left \langle \Gamma , \mathcal {E}_{\mathcal {P}}(\psi +\delta _1 \psi )\Gamma \right \rangle \leq C$
.
Regarding the variable
$c_k=a_k+[a_k,\mathcal {G}]$
from Eq. (68), let us apply Duahmel’s formula

where we have used
$\left [a_k,\mathcal {G}^\dagger \right ]=0$
for
$k\neq 0$
and
$[[a_k,\mathcal {G}],\mathcal {G}]=0$
, which follows from the observation that
$\eta _{ijk}=0$
in case one of the indices in
$\{i,j,k\}$
is zero. The following Lemma 7 provides useful estimates on the quantity
$\left [[a_k,\mathcal {G}],\mathcal {G}^\dagger \right ]$
. In order to formulate Lemma 7 let us define the kinetic energy of an operator-valued one-particle vector
$\Theta _k$
, written as
$\mathcal {E}_{\mathcal {K}}(\Theta )$
or
$\mathcal {E}_{\mathcal {K}}(\Theta _k)$
with k being a dummy index, as

Lemma 7. For
$m\geq 0$
there exists a constant
$C_m>0$
, such that

Proof. Let us write the double commutator as
$\left [[a_k,\mathcal {G}],\mathcal {G}^\dagger \right ]= (\delta _1 c)_k+(\delta _2 c)_k+(\delta _3 c)_k$
, where

By Eq. (36) it is clear that

and therefore

Using
$J_{p_1 p_2 p_3,p_1' p_2' p_3'}:=\sum _{q q' k}|k|^2\overline {\eta _{q' p_2' p_3'}\eta _{q p_1' k}}\eta _{q' p_1 k}\eta _{q p_2 p_3}$
and
$\widetilde {J}_{p_2 p_3,p_2' p_3'}:=\sum _{p_1}J_{p_1 p_2 p_3,p_1 p_2' p_3'}$

Utilizing the operator
$X_{jk,j'k'}:=\sum _q |k|\eta _{q j k}\overline {\eta _{q j' k'}}$
acting on
$L^2(\Lambda )^{\otimes 2}$
and the permutation operator
$(S\Psi )(x_1,x_2,x_3):=\Psi (x_2,x_1,x_3)$
acting on
$L^2(\Lambda )^{\otimes 3}$
, we can write

and
$\widetilde {J}=X^\dagger X$
. Consequently

By Eq. (36) we have
$\|\widetilde {J}\|\leq C N^{-\frac {15}{2}}$
for a suitable constant C. Consequently

Similarly one can show that
$\mathcal {E}_{\mathcal {K}} \left (\mathcal {N}^{m} \delta _3 c\right )\lesssim N^{-1}(\mathcal {N}+1)^{5+2m}$
, and therefore

As a consequence of Lemma 7, we obtain that the trial state
$\Gamma $
defined below Eq. (71) has a kinetic energy
$\mathcal {E}_{\mathcal {K}}(c)$
of the order
$O_{N\rightarrow \infty } \left (1\right )$
in the subsequent Corollary 4. Since in the residual term
$\mathcal {E}$
defined in Lemma 1 the term
$\mathcal {E}_{\mathcal {K}} \left (\sqrt {\mathcal {N}}c\right )\leq \frac {1}{2}\mathcal {E}_{\mathcal {K}} \left (c\right )+\frac {1}{2}\mathcal {E}_{\mathcal {K}} \left (\mathcal {N} c\right )$
appears, it will be convenient to estimate the expectation value in the state
$\Gamma $
of
$\mathcal {E}_{\mathcal {K}} \left (\mathcal {N}^m c\right )$
for
$m\geq 1$
as well.
Corollary 4. Let
$\Gamma $
be the state defined below Eq. (71) and
$m\geq 0$
. Then there exists a constant
$C>0$
, such that
$\left \langle \Gamma , \mathcal {E}_{\mathcal {K}} \left (\mathcal {N}^m c\right )\Gamma \right \rangle \leq \frac {C_m}{N}$
.
Proof. By Eq. (72) we have

and hence

where we have used that for operators
$f_k$
and
$A,B$
satisfying
$A^\dagger A\leq C B^\dagger B$
we have

Proceeding as in the proof of Corollary 3, we obtain by Eq. (82) and Lemma 7

where we have made use of Eq. (36) again. Using
$a_k\Gamma _0=0$
for
$k\neq 0$
, therefore yields

Having Corollary 3 and Corollary 4 at hand, we are in a position to verify the upper bound on the ground state energy
$E_N$
in Theorem 4.
Proof of Theorem 4.
Let
$A:=(2\pi \mathbb {Z})^9 \setminus \{(0,0,0)\}$
, and let
$\Gamma $
be the state defined below Eq. (71). Using Eq. (27) and Eq. (21), and the fact that
$\left (\widetilde {V}_N\right )_{i j k,\ell m n}=\left (V_N\right )_{i j k,\ell m n}$
for index triples
$(ijk),(\ell m n)\in A$
, we obtain

By a symmetry argument it is clear that . Applying Corollary 3 as well as Corollary 4, with
$m=0$
, yields for suitable constants
$C>0$

Furthermore, observe that
$N^3\lambda _{0,0}\leq \frac {1}{6}b_{\mathcal {M}}(V)N +C'$
by Eq. (37) for a suitable
$C'$
. In order to estimate the final term
, note that we have by Eq. (72) for
$m\in \mathbb {N}$

Using Lemma 3 together with the estimate from Corollary 4 for
$m=0$
and
$m=1$
, we therefore obtain
.
4 Refined correlation structure
Utilizing the set of operators defined in Eq. (25) and Eq. (26), we were able to identify the ground state energy
$E_N$
up to errors of the magnitude
$O_{N\rightarrow \infty }(\sqrt {N})$
in the previous Sections 2 and 3. It is the purpose of this Section to obtain a higher resolution of the energy, which especially captures the subleading term proportional to
$\sqrt {N}$
in the asymptotic expansion of
$E_N$
, using a more refined correlation structure compared to the one introduced in Subsection 1.1. On a technical level, the new correlation structure is implemented by the new set of operators
$d_k$
and
$\xi _{ijk}$
defined below in Eq. (89) and Eq. (90), which constitute a refined version of the operators
$c_k$
and
$\psi _{ijk}$
respectively. Writing the operator
$H_N$
in terms of
$d_k$
and
$\xi _{ijk}$
will then allow us to verify the lower bound from Theorem 1 in Subsection 4.2 and the corresponding upper bound in Section 5.
The approach presented in Sections 2 and 3 fails to capture the correct term of order
$\sqrt {N}$
for two reasons: (I) The following expression appearing in Eq. (32)

is expected to lower the ground state energy by an amount proportional to
$\sqrt {N}$
, to be precise naive perturbation theory suggests that the term in Eq. (85) is giving rise to an energy correction proportional to

see Eq. (35), and therefore consistent with our estimate in Lemma 4. (II) In the pursue of an upper bound on
$E_N$
we expressed the unitary conjugated variables
$U^{-1} \psi _{ijk} U$
as a sum of
$a_i a_j a_k$
and various error terms according to Eq. (74) as

While most of the terms appearing in Eq. (87) give a contribution of the order
$o_{N\rightarrow } \left (\sqrt N \right )$
, the term
$\delta _1 \psi $
is expected to increase the ground state energy by an amount proportional to
$\sqrt {N}$
, which is consistent with our estimate in Eq. (76). In order to extract the energy shift due to the expression in Eq. (85), we follow the strategy in Subsection 1.1 and introduce an additional two-particle correlation structure via a map acting on the two-particle space

in Eq. (88), which will give rise to the negative energy correction
$-\mu (V)\sqrt {N}$
from Theorem 1. To be precise, we define the map
$T_2$
via its matrix elements as

for
$|\ell |>K$
and
$(T_2-1)_{jk,mn}:=0$
otherwise, where
$\lambda _{k,\ell }$
is defined below Eq. (32). Regarding the energy shift associated with
$\delta _1 \psi $
, it is a natural idea to include this term in the definition of our new operators
$\xi _{ijk}$
, giving rise to the positive energy correction
$\gamma (V)\sqrt {N}$
from Theorem 1. However, a computation in Eq. (93) demonstrates that the presence of
$\delta _1 \psi $
produces new four-particle correlation terms of the form

with coefficients
$\theta _{uijk}$
proportional to
$N^2\sum _{mn}(V_N)_{ijk,0mn}\eta _{mnu}$
, which behave like

for momenta of the order
$\sqrt {N}$
, and decay fast for higher momenta. Therefore, the four-particle correlation terms are expected to lower the ground state energy, similar to Eq. (86), by an amount of the order

Again we extract the correlation energy by introducing a map, acting this time on the four-particle space

in Eq. (95), which gives rise to the negative energy correction
$-\sigma (V)\sqrt {N}$
from Theorem 1.
In the following let
$T:L^2(\Lambda ^3)\longrightarrow L^2(\Lambda ^3)$
be the map constructed in Eq. (19), and for now let us think of
$T_2:L^2(\Lambda ^2)\longrightarrow L^2(\Lambda ^2)$
and
$T_4:L^2(\Lambda ^4)\longrightarrow L^2(\Lambda ^4)$
as generic bounded permutation-symmetric operators modeling the two-particle and four-particle correlation structure respectively. Following the approach in Section 2, we are implementing many-particle counterparts to the transformations T,
$T_2$
, and
$T_4$
as


Note that
$T_2$
is not included in the definition of
$\xi _{ijk}$
, as it would only give contributions of the order
$O_{N\rightarrow \infty }(1)$
. Using the Laplace operator
$\Delta _s$
acting on the space
$L^2(\Lambda )^{\otimes s}$
and the coefficients

let us furthermore define the operators
$X_2: =T_2^\dagger (-\Delta _2)T_2+\Delta _2$
and

A straightforward computation, similar to the one in Eq. (27), reveals that up to excess terms involving
$X_2$
,
$X_4$
and an error term
$\widetilde {\mathcal {E}}$
, we can write the operator
$H_N$
as a sum of squares in the variables
$d_k$
and
$\xi _{ijk}$
according to

where the error
$\widetilde {\mathcal {E}}$
contains all the non-fully contracted products appearing in the squares

In this context we define the fully contracted part of a product of monomials

as
$C_{j_1\dots j_{t},i^{\prime }_1\dots i^{\prime }_{r'}} a_{i_1}^\dagger \dots a_{i_r}^\dagger a_{j^{\prime }_1}\dots a_{j^{\prime }_{t'}}$
with
$C_{j_1\dots j_{t},i^{\prime }_1\dots i^{\prime }_{r'}}$
being the expectation of
$a_{j_1}\dots a_{j_t}a_{i^{\prime }_1}^\dagger \dots a_{i_{r'}}^\dagger $
in the vacuum. For a term by term definition of
$\widetilde {\mathcal {E}}$
see Eq. (111) in Subsection 4.1.
In the following we want to choose
$T_4$
, such that the term
$4( \widetilde {V}_N\otimes 1)\chi $
is cancelled in the expression
$\{..\}$
from Eq. (92), at least after symmetrization and projection onto the range of
$Q^{\otimes 4}$
, that is, we define

where
$\Pi _{\mathrm {sym}}$
is the orthogonal projection onto the subspace
$L^2_{\mathrm {sym}}(\Lambda ^4)\subseteq L^2(\Lambda ^4)$
and
$R_4$
is the pseudoinverse of

In order to obtain an improved representation of the operator
$X_4$
defined in Eq. (92), let us introduce the constants


which allow us to write
$\gamma _N-\sigma _N=\frac {N^4}{24}(X_4)_{0000,0000}$
. Furthermore, we define the three-particle state
$\Theta $
as

for
$\{i,j,k\}$
all different from zero and
$ (\Theta )_{ijk}:=0$
otherwise. According to the definition of
$T_4$
we have
$Q^{\otimes 4}\Pi _{\mathrm {sym}}X_4\Pi _{\mathrm {sym}} P^{\otimes 4}=0$
, and therefore

In order to understand the size of the term in Eq. (100) better, we are going to rewrite it in terms of the variables
$\psi _{ijk}$
defined in Eq. (26), respectively the variables

defined in Eq. (26) for the concrete choice
$K:=0$
, see Eq. (69), with the corresponding operator
$T_{K=0}:=1+RV_N\pi _0$
, see Eq. (17). Note that

and therefore

In order to address the correlation term in Eq. (85), we use the concrete choice for
$T_2$
from Eq. (88). With

this choice for a transformation
$T_2$
yields

Summarizing what we have so far allows us to write the operator
$H_N$
in terms of the new variables
$d_k$
and
$\xi _{ijk}$
as

Defining the error term

we obtain as a consequence of Eq. (103) the following Corollary 5.
Corollary 5. Let
$d_k$
and
$\xi _{ijk}$
be as in Eq. (89) and Eq. (90), with
$T_2$
defined in Eq. (88) and
$T_4$
defined in Eq. (95),
$\gamma _N,\sigma _N$
, and
$\mu _N$
as in Eq. (98), Eq. (97) and Eq. (102), and let
$\mathcal {E}_*$
be as in Eq. (104),
$\Theta $
as in Eq. (99) and
$\widetilde {\psi }_{ijk}$
as in Eq. (101). Furthermore, recall the definition of
$\lambda _{0,0}$
in Eq. (29) and
$\mathbb {Q}_K$
in Eq. (33). Then

Making use of the notation
$\mathcal {E}_{\mathcal {K}}(d)$
from Eq. (83) and
$\mathcal {E}_{\mathcal {P}}(\xi )$
from Eq. (75), we obtain in the case
$K=0$
the identity

Proof. Using Eq. (103) and the definition of
$\widetilde {V}_N$
in Eq. (21), as well as the identities in Eq. (30) and Eq. (31), we obtain

Since
$\mathbb {Q}_0=0$
and

4.1 Analysis of the error terms
In the following we are providing an explicit representation of the error term
$\widetilde {\mathcal {E}}$
introduced in Eq. (93), which we subsequently use in Lemma 8 in order to control
$\widetilde {\mathcal {E}}$
. For this purpose, we are going to utilize the following estimates on the matrix elements of
$T,T_2$
and
$T_4$



which are verified in Lemma 15 and Lemma 18 respectively. Furthermore, it is useful to introduce the two-particle state
$(\varphi _2^0)_{jk}:=N(T_2-1)_{jk,00}$
, the three-particle states
$(\varphi ^0_3)_{ijk}:=\frac {N^{\frac {3}{2}}}{2}(T-1)_{ijk,0 0 0}$
and for
$m\in (2\pi \mathbb {Z})^3\setminus \{0\}$

and the four particle state
$(\varphi ^0_4)_{uijk}:=\frac {N^2}{6} (T_4-1)_{uijk,0000}$
as well as

Additionally, let us introduce for
$\varphi \in L^2(\Lambda ^s)$
and
$\psi \in L^2(\Lambda ^t)$
with
$s,t\geq 0$
, and
$\ell \leq \min \{s,t\}$
, the operator

acting on
$L^2(\Lambda ^{t-\ell })\longrightarrow L^2(\Lambda ^{s-\ell })$
. In coordinates, the operator is given by

Finally let
$\widetilde {G}:=\mathrm {Tr}_{1\rightarrow 3} \left [\widetilde {V}_N\otimes 1 \varphi _4 \varphi _4^\dagger \right ]$
. With this at hand we can write

where
$C_{s,t,\ell }:=\frac {(s-1)! (t-1)!}{(\ell -1)! (s-\ell )! (t-\ell )!}$
and the set of allowed configurations
$(s,t,\ell ,m,n)\in \mathcal {S}$
is defined by the rules
$\ell \leq \min \{s,t\}$
and
$\ell <\max \{s,t\}$
, where
$2\leq s,t\leq 4$
and
$|n|,|m|\leq K$
with
$m=0$
, respectively
$n=0$
, in case
$s\neq 3$
, respectively
$t\neq 3$
. Note that the criterion
$\ell <\max \{s,t\}$
makes sure that we only include non-fully contracted parts of the first product in Eq. (94) and
$\widetilde {G}$
is the kernel associated with the non-fully contracted part of the second product in Eq. (94). Furthermore,
$C_{s,t,\ell }$
counts the various different ways to have an
$\ell -1$
fold contraction between
$s-1$
many annihilation operators and
$t-1$
many creation operators, that is,
$C_{s,t,\ell }$
counts the number of partially defined injective functions

with
$\# \mathrm {dom}f=\ell -1$
. Let us illustrate the derivation of Eq. (94), looking at the term

from Eq. (94). According to the definition of
$d_k$
in Eq. (89), the term
$d_k-a_k$
decomposes into three terms, and in the following we focus only on the last one involving
$T_4-1$

Notably, the final term in Eq. (112) is a fully contracted contribution and therefore part of the operator
$X_4$
defined in Eq. (92) and not of the error term
$ \widetilde {\mathcal {E}} $
.
Estimating the various terms appearing in Eq. (111) individually allows us to prove the following Lemma 8.
Lemma 8. There exists a constant
$C>0$
and a function
$\epsilon :[0,\infty )\longrightarrow (0,C)$
satisfying
${\lim _{K\rightarrow \infty }\epsilon (K)=0}$
, such that we have for K as in the definition of
$\pi _K$
below Eq. (17)

Proof. In the following let
$\tau \leq \frac {1}{2}$
. Using the fact that
$\|\frac {a_m^\dagger (a_0^\dagger )^{s-1} a_0^{t-1}a_n}{N^{ \frac {s+t}{2} }}\|\leq 1$
, there exists by Corollary 1 a constant
$C>0$
such that for
$\delta>0$
and
$s,t\geq \ell +1$

For
$\tau =\sigma =0$
, we have the improved bound on the left-hand side of Eq. (113)

In the case that either s or t is equal to
$\ell $
, for example,
$t=\ell $
we obtain by Corollary 1

In order to obtain good estimates on the operator norms
$\left \|\mathcal {K}_{\sigma ,s-\ell }^{-\frac {1}{2}} G_{\ell }(\varphi _s^m,\varphi _t^n)\mathcal {K}_{\tau ,t-\ell }^{-\frac {1}{2}}\right \|$
, observe that we obtain by a Cauchy-Schwarz argument

that is, it is enough to control the norm of the symmetric ones. In the following we choose
$\sigma =\tau =\frac {1}{2}$
, except for the case
$s=t=2$
where we choose
$\sigma =\tau =0$
. By Eq. (107) and Eq. (109) we obtain for a suitable
$C>0$

Furthermore by Eq. (108),
$\left \| G_{1}(\varphi _2^0,\varphi _2^0)\right \|\leq \epsilon $
in case K is large enough and
$\left \| G_{2}(\varphi _2^0,\varphi _2^0)\right \|\leq C \sqrt {N}$
, as well as
$\left \| \widetilde {G}\right \|\leq C N^{-1}$
by Eq. (109). Choosing
$\delta :=N^{\frac {t-s}{4}}$
, and combining the estimates on the operator norms with Eq. (113), respectively Eq. (114), and Eq. (115) concludes the proof by Eq. (111).
Following the ideas in the proof of Lemma 8, we can furthermore compare the operator
$\sum _{k}|k|^{2\tau }a_k^\dagger a_k$
with the corresponding operator
$\sum _{k}|k|^{2\tau }d_k^\dagger d_k$
in the variables
$d_k$
defined in Eq. (89). This is the content of the subsequent Lemma 9.
Lemma 9. Let
$0\leq \tau < \frac {1}{4}$
. Then there exists
$K_0,C>0$
such that for
$K\geq K_0$
, with K as in the definition of
$\pi _K$
below Eq. (17),

Proof. By Eq. (40), there exists a constant
$C>0$
such that

Furthermore, we have by Cauchy-Schwarz the estimate

Similar to the definition of
$G_\ell $
in Eq. (110) let us introduce

A similar computation as in Eq. (111) together with a Cauchy-Schwarz estimate yields

for a suitable constant
$C>0$
. Utilizing the estimates in Eq. (108) and Eq. (109) we obtain that
$|G^{\prime }_2|\lesssim 1$
,
$\|G^{\prime }_2\|\lesssim \frac {1}{K^2}$
,
$|G^{\prime \prime }_4|\lesssim N^{\tau -\frac {1}{2}}\leq 1$
and
$\|G^{\prime \prime }_\ell \|\lesssim N^{\tau -2}\leq N^{-\frac {3}{2}}$
for
$\ell \leq 3$
. Consequently there exists a
$C>0$
such that

Using
$\mathcal {N}\leq \sum _{k}|k|^{2\tau }a_k^\dagger a_k$
and
$\sum _{k}|k|^{2\tau }d_k^\dagger d_k\leq \sum _{k}|k|^{2}d_k^\dagger d_k$
we therefore obtain

Choosing K large enough such that
$\frac {C}{K^2}<1$
concludes the proof.
Before we come to the proof of the lower bound in Theorem 1 in the following Subsection 4.2, we are going to derive sufficient estimates on

in the following Lemma 10. In order to verify Lemma 10, we require the estimate

for
$ \tilde {c}_k:=a_k +\frac {1}{2}\sum _{ij} (T-1)_{ijk,000}a_i^\dagger a_j^\dagger a_0^3$
, which is verified in Appendix A, see Lemma A3.
Lemma 10. Let
$0\leq \gamma <\frac {1}{4}$
. Then there exists a
$C>0$
such that

Proof. Let us define for
$ijk\neq 0$

where
$\chi $
is defined in Eq. (91). Then we have the decomposition

Note that
$\zeta _k= \frac {1}{24}\Big ( \Pi _{\mathrm {sym}} 4( V_N\otimes 1)(T_4-1+\chi )\Big )_{0 0 k (-k),0 0 0 0}$
for
$|k|>K$
. Using the regularity of V and the bounds derived in Eq. (107) and Eq. (109), we observe that we have
$|N^3 \zeta _k|\lesssim N^{-\frac {1}{2}}\left (1+\frac {|k|^2}{N}\right )^{-1}$
, and therefore

where we have used Eq. (40). Choosing
$\epsilon $
of the order
$N^{-\frac {1}{4}}$
and
$\tau =\frac {1}{2}$
concludes the analysis of the second term in Eq. (118). Regarding the first term, we use the definition of
$\tilde {c}_k$
below Eq. (116), in order to identify
$\sum _{ijk}\overline {\zeta _{ijk}} (a_0^\dagger )^4 \widetilde {\psi }_{ijk}a_0$
as

In the following we are going to verify that the most significant term
$\sum _{ijk}\overline {\zeta _{ijk}}(a_0^\dagger )^4 a_i a_j \tilde {c}_k a_0 $
in Eq. (119) satisfies the desired bound. By Cauchy-Schwarz, we have for
$\epsilon>0$

with
$G^{(0)}:=N^5\sum _{ijk}\frac {|\zeta _{ijk}|^2}{|k|^{2}}$
,
$G^{(1)}_{i,i'}:=N^5\delta _{i,i'}\sum _{jk}\frac {|\zeta _{ijk}|^2+\overline {\zeta _{ijk}}\zeta _{jik}}{|k|^{2}}$
and
$G^{(2)}_{ij,i' j'}:=N^5\sum _k \frac {\zeta _{ijk}\zeta _{i' j' k}}{|k|^{2}}$
, and
$X:=N^{-5}a_0^{4\dagger }a_0 a_0^\dagger a_0^4$
. Using again the regularity of V and Eq. (107), as well as the bounds on
$T_4$
in Eq. (109), yields

and therefore
$|G^{(0)}|\lesssim 1 $
and
$\|G^{(1)}\|\lesssim N^{-\frac {3}{2}}$
. The choice
$\epsilon :=N{-\frac {1}{4}}$
then yields

Finally
$\|G^{(2)}\|\leq N^{-\frac {3}{2}}$
, and therefore

This concludes the proof together with Eq. (116).
4.2 Proof of the lower bound in Theorem 1
In this subsection, we are going to verify the lower bound in Theorem 1 making use of the sequence of states
$\Phi _N$
constructed in Corollary 2, which simultaneously satisfies

Starting point for our investigations is then the lower bound

see Eq. (105). As is proven in Section 7, the coefficients
$\gamma _N,\mu _N$
and
$\sigma _N$
converge to the corresponding constants
$\gamma (V)$
,
$\mu (V)$
and
$\sigma (V)$
introduced in Eq. (10), Eq. (5) and Eq. (9)



see Lemma 17. Given
$\epsilon>0$
, assume that K is large enough such that the function
$\epsilon (K)$
from Lemma 8 satisfies
$\epsilon (K)\leq \epsilon $
. Making use of the fact that

we immediately obtain for C and
$\widetilde {C}$
large enough

Similarly we obtain by Lemma 10 and Lemma 4 for suitable
$C,\widetilde {C}>0$

By Lemma 2 we furthermore obtain for
$\tau ,\epsilon>0$
and K large enough, and a suitable
$C>0$
,

Moreover we note that we have by Eq. (120)-(122)

and similarly . Finally by Lemma 9

Choosing
$\tau <\frac {1}{4}$
and
$\epsilon <\frac {1}{2C}$
concludes the proof, since

5 Second-order upper bound
It is the goal of this Section to introduce a trial state
$\Phi $
, which simultaneously annihilates the variables
$d_k$
for
$k\neq 0$
and
$\xi _{\ell m n}$
in case
$(\ell ,m ,m)\neq 0$
, at least in an approximate sense. We are then going to use this trial state
$\Phi $
to verify the upper bound in Theorem 1. For the rest of this Section we specify the parameter K introduced above the definition of
$\pi _K$
in Eq. (17) as
$K:=0$
. In order to find
$\Phi $
, we define
$\alpha _{jk}:=(T_2-1)_{jk,00}$
and
$\beta _{u i j k}:=(T_4-1)_{uijk,0000}$
, and the generator

of a unitary group
$W_s:=e^{s(\mathcal {G}_2+\mathcal {G}_4)^\dagger - s (\mathcal {G}_2+\mathcal {G}_4)}$
and
$W:=W_1$
. We note at this point, that the action of the unitary operator W only creates an
$O(1)$
amount of particles, in the sense that

as is proven in Appendix A, see Lemma A1. Applying Duhamel’s formula, we can express
$W^{-1}a_{i_1} a_{i_2} a_{i_3} W$
as

Furthermore, note that we can write

where we define the error term

Therefore we can write the transformed operators
$W^{-1}\xi _{i_1 i_2 i_3}W$
as

Recall the definition of
$\mathcal {E}_{\mathcal {P}}$
defined in Eq. (75). The following Lemma 11 provides sufficient bounds on the various error terms appearing in Eq. (128).
Lemma 11. There exists a constant
$C>0$
, such that

Furthermore, we have
$\mathcal {E}_{\mathcal {P}} \left ([a_{i_1} a_{i_2} a_{i_3},\mathcal {G}_2^\dagger + \mathcal {G}_4^\dagger - \mathcal {G}_2]\right )\leq C (\mathcal {N}+1)^6$
and


Proof. Let us define

where we have used Eq. (109) to estimate
$C_N$
. Applying Cauchy-Schwarz yields

Using the fact that
$C_N(a_0^\dagger )^4a_0^4 \left (\sum _{i_1} a_{i_1}^\dagger a_{i_1}\right )\leq C_N N^5\lesssim 1$
, we observe that the quantity in Eq. (132) is bounded by the right-hand side of Eq. (129). Let us furthermore define

In the following we want to distinguish between the cases
$A':=\{(i_1 i_2 i_3)\in A: i_1,i_2\neq 0\}$
and
$A":=A\setminus A'$
, leading to the definition

where we have again used Eq. (109). Applying Cauchy-Schwarz leads to the estimate

which is bounded by the right-hand side of Eq. (129). Finally we use that
$V_N$
is permutation-symmetric and non-negative, and therefore the left-hand side of Eq. (129) is bounded by

Regarding the term
$\mathcal {E}_{\mathcal {P}} \left ([a_{i_1} a_{i_2} a_{i_3},\mathcal {G}_2^\dagger + \mathcal {G}_4^\dagger - \mathcal {G}_2]\right )$
, let us analyze the term involving the commutator with
$\mathcal {G}_2$
, the terms involving
$\mathcal {G}_2^\dagger + \mathcal {G}_4^\dagger $
can be analyzed in a similar fashion as has been done in Lemma 6. We compute

In order to analyze the first term on the right-hand side of Eq. (133), let us define
$D_N:=\sup _{i_1}\sum _{i_2 i_3, i_1' i_2' i_3'}\left |(V_N)_{i_1 i_2 i_3, i_1' i_2' i_3'} \alpha _{i_2 i_3}\alpha _{i_2' i_3'}\right |$
and note that
$D_N\lesssim N^{-3}$
by Eq. (108). Hence

Regarding the second term on the right-hand side of Eq. (133), we use again the split
$A=A'\cup A"$
and define

where we have used Eq. (108). Consequently

and therefore
$\mathcal {E}_{\mathcal {P}} \left ([a_{i_1} a_{i_2} a_{i_3},\mathcal {G}_2]\right )\leq 12 \mathcal {E}_{\mathcal {P}} \left (\alpha _{i_2 i_3}a_{i_1} a_0^2\right )+ 12 \mathcal {E}_{\mathcal {P}} \left (\sum _u \alpha _{u i_3}a_u^\dagger a_{i_2} a_{i_3}\right )\lesssim (\mathcal {N}+1)^3$
. The inequalities in Eq. (130) and Eq. (131) can be verified similarly.
With Lemma 11 at hand, we show in the subsequent Corollary 6 that after conjugation with the unitary W, the potential energy of the operators
$\xi _{ijk}$
is comparable to the potential energy of
$(\psi +\delta _1)_{ijk}$
.
Corollary 6. There exists a constant
$C>0$
, such that

Proof. Using the sign
$V_N\geq 0$
, we obtain by the Cauchy-Schwarz inequality and the representation of
$W^{-1}\xi _{i_1 i_2 i_3} W$
in Eq. (128) the estimate

where we have first used Lemma 11 and subsequently Eq. (125) in the last line.
Regarding the variable
$d_k$
, Duhamel’s formula yields for
$k\neq 0$

Recall the kinetic energy of an operator-valued one-particle vector
$\Theta _k$
defined in Eq. (83). Then the following Lemma 12 provides sufficient bounds on the various error terms appearing in Eq. (134).
Lemma 12. There exists a constant
$C>0$
, such that for
$m\in \mathbb {N}$

Proof. Let us compute as an example for
$k\neq 0$

and let us focus on the term

Defining

where we have used Eq. (109), we obtain

The other estimates in Lemma 12 can be verified similarly.
Similar to Corollary 6, we show in the following Corollary 7 that, after conjugation with the unitary W, the kinetic energy of the operators
$ d_k$
is comparable to the one of
$c_k$
.
Corollary 7. There exists a constant
$C>0$
, such that for
$m\in \mathbb {N}$

Proof. By Eq. (125) we have
$(W^{-1}\mathcal {N}^{m}W)^* W^{-1}\mathcal {N}^{m}W=W^{-1}\mathcal {N}^{2m}W\lesssim \mathcal {N}^{2m}+1$
, hence

Following the ideas in Corollary 6, we estimate using Eq. (134)

where we have first used Lemma 12 and subsequently Eq. (125) in the last line.
Before we come to the proof of the upper bound in Theorem 1, we are showing in the following Lemma 13 that even without a unitary conjugation, the kinetic energy of
$c_k$
is comparable with the one of
$d_k$
. The price for dropping the unitary W is that we obtain an order
$O_{N\rightarrow \infty } \left (\sqrt {N}\right )$
prefactor in front of the excess term
$(\mathcal {N}+3)^{3+2m}$
, instead of a prefactor of the order
$O_{N\rightarrow \infty } \left (1\right )$
.
Lemma 13. There exists a constant
$C>0$
, such that for
$m\in \mathbb {N}$

Proof. Note that we can write
$c_k$
as

and therefore

Defining the constant
$C_N:=\sum _{jk}|k|^2 |\alpha _{jk}|^2\lesssim N^{-\frac {3}{2}}$
, which follows from Eq. (108), we obtain

Similarly we obtain

Proof of the upper bound in Theorem 1.
Let us define the trial state
$\Phi :=W\Gamma $
, where
$\Gamma $
is the state defined below Eq. (71), and recall the representation of
$H_N$
in Eq. (106)

By Eq. (125) and the fact that
$\pm \left (N^{-m}(a_0^\dagger )^m a_0^m-1\right )\lesssim N^{-1}\mathcal {N}$
, we obtain that

see Eq. (84) for the last estimate. Making use of Lemma 8, Lemma 4 and Lemma 10 yields

Observe that and furthermore we have by Lemma 13 and Corollary 7

where we used Corollary 4 in the last estimate. Putting together what we have so far, and utilizing Eq. (120)-(122) again, yields

Using Corollary 6, Corollary 7, Corollary 3 and Corollary 4, we further have

6 Proof of Theorem 2
In the following, we want to verify Theorem 2, claiming that any sequence of states
$\Psi _N$
with

satisfies complete Bose-Einstein condensation with a rate of the order
$N^{-\frac {3}{4}}$
. Together with Theorem 1, the statement follows immediately once we can show that

for some constant C. In order to prove Eq. (135), we observe that by the results in [Reference Nam, Ricaud and Triay23, Section 7], see also the comment below [Reference Nam, Ricaud and Triay25, Theorem 4], the modified operators
$H_N(2\alpha )$
satisfy the asymptotic identity

where the modified Gross-Pitaevskii functional is defined as

using the projection
$P=1-Q$
, with Q being introduced in Eq. (18). Note that in the notation of [Reference Nam, Ricaud and Triay25] the operator
$H_N(2\alpha )$
reads

Furthermore, for
$\alpha <2\pi ^2$
, we have that

and by Hölder’s inequality we have

for any
$u\in L^2(\Lambda )$
with
$\|u\|=1$
, leading to the lower bound

Therefore, the ground state
$\Psi _{N,\alpha }$
of
$H_N(\alpha )$
satisfies

and by Theorem 1 we obtain the matching upper bound

As a consequence, the states
$\Psi _{N,\alpha }$
satisfy complete Bose-Einstein condensation

and we can proceed exactly as in Corollary 2, as long as the additional condition
$\alpha <\delta $
holds with
$\delta $
being the constant in Eq. (64). In particular, there exist states
$\Phi _{N,\alpha }$
such that

and we have the estimate on the kinetic energy
$\left \langle \Phi _{N,\alpha }, \sum _{k}|k|^2 c_k^\dagger c_k \Phi _{N,\alpha } \right \rangle \leq C\sqrt {N}$
. Note that the localization results in Lemma 5 hold without any modification for the operator
$H_N(\alpha )$
, since
$\mathcal {N}$
commutes with the localization functions
. Following Subsection 4.2, we therefore arrive at the lower bound

Using again Eq. (123), for
$\tau <\frac {1}{4}$
, we obtain for a large enough constant C

Choosing
$\alpha $
and
$\epsilon $
small enough such that
$2\epsilon +\alpha <\frac {1}{C}$
concludes the proof of Eq. (135).
7 Analysis of the scattering coefficients
This Section is devoted to the study of the variational problems in the definition of
$b_{\mathcal {M}}(V)$
in Eq. (4) and the definition of
$\sigma (V)$
in Eq. (9) as well as the study of their corresponding minimizers
$\omega $
and
$\eta $
. Especially we want to compare
$\gamma (V),\mu (V)$
, and
$\sigma (V)$
defined in Eq. (10), Eq. (5) and Eq. (9) with
$\gamma _N,\mu _N$
, and
$\sigma _N$
defined in Eq. (97), Eq. (102) and Eq. (98), see Lemma 17. The proof will be based on the observation that the N-dependent quantities can be seen as a counterpart on the three-dimensional torus
$\Lambda $
to the N-independent quantities defined in terms of variational problems on the full space
$\mathbb {R}^3$
. Similarly we will compare in Lemma 16 the modified scattering length
$b_{\mathcal {M}}(V)$
, which can be expressed in terms of the minimizer
$\omega $
as

see [Reference Nam, Ricaud and Triay23], with its counterpart on the torus
$\Lambda $
defined in Eq. (29) as

The proof of Lemma 16 is based on the observation that
$(1-\omega )V=V-\omega V$
is the full space counterpart to the renormalized potential
$V_N -V_N R V_N$
.
In the following Lemma 14 we want to derive properties of
$\mathcal {Q}$
defined in Eq. (8) as

and its minimizers. For this purpose it will be useful to introduce for a given cut-off parameter
$\ell $
and a smooth function
$\chi :\mathbb {R}^6\longrightarrow \mathbb {R}$
function with
$\chi (x)=1$
for
$|x|_\infty \leq \frac {1}{3}$
and
$\chi (x)=0$
for
$|x|_\infty>\frac {1}{2}$
, the modified function

Furthermore, we define the corresponding functional, acting on
$\dot {H}^1(\mathbb {R}^9)$
, as

and
$\sigma _\ell (V):=\mathcal {Q}_\ell (0)-\inf _{\varphi \in \dot {H}^1(\mathbb {R}^9)}\mathcal {Q}_\ell (\varphi )$
.
Lemma 14. There exists a unique minimizer
$\eta $
of the functional
$\mathcal {Q}$
in
$\dot {H}^1(\mathbb {R}^9)$
, and
$\eta $
satisfies the point-wise bounds
$0\leq \eta \leq \frac {1}{-2\Delta _{\mathcal {M}_*}}f$
and
$\sigma (V)=\int _{\mathbb {R}^9}f(x)\eta (x)\mathrm {d}x$
, as well as

in the sense of distributions. Furthermore,
$\mathcal {Q}_\ell $
has a unique minimizer
$\eta _\ell $
, and
$\eta _\ell $
satisfyies
$0\leq \eta _\ell \leq \frac {1}{-2\Delta _{\mathcal {M}_*}}f_\ell $
and
$\sigma _\ell (V)=\int _{\mathbb {R}^9}f_\ell (x)\eta _\ell (x)\mathrm {d}x$
, as well as
$( -2\Delta _{\mathcal {M}_*}+\mathbb {V})\eta _\ell =f_\ell $
and

Proof. Following the proof of [Reference Nam, Ricaud and Triay23], we observe that since
$\mathcal {Q}(0)<\infty $
, there exists a minimizing sequence
$\varphi _n\in \dot {H}^1(\mathbb R^9)$
for
$\mathcal {Q}$
with
$\sup _{n}\|\nabla \varphi _n\|<\infty $
and
$\sup _{n}\left \|\sqrt {\mathbb {V}}\left (\frac {f}{\mathbb {V}}-\varphi _n\right )\right \|<\infty $
, and therefore there exists by Banach-Alaoglu a subsequence
$\varphi _n$
and elements
$X,Y\in L^2(\mathbb R^9)$
such that
$\nabla \varphi _n\rightharpoonup X$
and
$\sqrt {\mathbb {V}}\left (\frac {f}{\mathbb {V}}-\varphi _n\right )\rightharpoonup Y$
converge weakly in
$L^2(\mathbb R^9)$
. By [Reference Lieb and Loss17, Theorem 8.6], we obtain that there exists an element
$\eta \in \dot {H}^1(\mathbb R^9)$
such that
$X=\nabla \eta $
and
$\varphi _n|_{A}$
converges (strongly) to
$\eta |_{A}$
in
$L^2(A)$
for any set
$A\subseteq \mathbb R^9$
of finite measure. Since
$\sqrt {\mathbb {V}}$
is a bounded function, we further have the convergence of
$\sqrt {\mathbb {V}}\left (\frac {f}{\mathbb {V}}-\varphi _n\right )|_A$
to
$\sqrt {\mathbb {V}}\left (\frac {f}{\mathbb {V}}-\eta \right )|_A$
, and in particular
$Y=\sqrt {\mathbb {V}}\left (\frac {f}{\mathbb {V}}-\eta \right )$
. In summary we have

weakly in
$L^2(\mathbb R^9)$
, and therefore we observe that
$\eta $
is a minimizer of
$\mathcal {Q}$

Computing
$0=\frac {\mathrm {d}}{\mathrm {d}t}\mathcal {Q}(\eta +t\varphi )$
for
$\varphi \in C^\infty _{0}(\mathbb {R}^9)$
immediately gives in the sense of distributions

and computing
$0=\frac {\mathrm {d}}{\mathrm {d}t}\mathcal {Q}(\eta +t\eta )$
yields
$\sigma (V)=\int _{\mathbb {R}^9}f(x)\eta (x)\mathrm {d}x$
. Regarding the uniqueness, we note that
$\varphi \mapsto \|\mathcal {M}_* \nabla \varphi \|^2$
is strictly convex on
$\dot {H}^1(\mathbb {R}^9)$
, and therefore
$\mathcal {Q}$
is strictly convex too. Consequently the minimizer
$\eta $
is unique. Using that
$\frac {f}{\mathbb {V}}\geq 0$
, we have

and using furthermore the fact that
$\varphi \mapsto \int _{\mathbb R^9}|\nabla \varphi (x)|^2\mathrm {d}x$
is a Dirichlet form yields

Therefore,
$\mathcal {Q}(|\varphi |)\leq \mathcal {Q}(\varphi )$
for all
$\varphi \in \dot {H}^1(\mathbb {R}^9)$
and by the uniqueness of the minimizer we obtain
$\eta =|\eta |\geq 0$
. For the purpose of obtaining an upper bound on
$\eta $
, we observe that
$\frac {1}{\mathcal {M}_*\nabla }f\in L^2(\mathbb {R}^9)$
and define the functional

Since
$\widetilde {\mathcal {Q}}(\varphi )=\mathcal {Q}(\varphi )+\widetilde {\mathcal {Q}}(0)-\mathcal {Q}(0)$
, we observe that
$\eta $
is the unique minimizer of
$\widetilde {Q}$
. It is furthermore clear that

and utilizing again that
$\varphi \mapsto \int _{\mathbb R^9}|\nabla \varphi (x)|^2\mathrm {d}x$
is a Dirichlet form yields

In particular,
$\widetilde {\mathcal {Q}}(\min \{\varphi ,\frac {1}{-2\Delta _{\mathcal {M}_*}}f\})\leq \widetilde {\mathcal {Q}}(\varphi )$
, and therefore we obtain by the uniqueness of minimizer for
$\widetilde {\mathcal {Q}}$

The properties of
$\mathcal {Q}_\ell $
can be verified analogously.
In order to compare
$\sigma (V)$
with
$\sigma _\ell (V)$
, let us first verify the point-wise bounds

For this purpose we introduce the additional functionals
$\mathcal {Q}^{\prime }_\ell $
and
$\mathcal {Q}^{\prime \prime }_\ell $
as

By a straightforward computation, we observe that
$\mathcal {Q}^{\prime }_\ell (\varphi )=\mathcal {Q}_\ell (\varphi )+\mathcal {Q}^{\prime }_\ell (0)-\mathcal {Q}_\ell (0)$
and similarly
$\mathcal {Q}^{\prime \prime }_\ell (\varphi )=\mathcal {Q}(\varphi )+\mathcal {Q}^{\prime \prime }_\ell (0)-\mathcal {Q}(0)$
. Therefore
$\eta _\ell $
is the unique minimizer of
$\mathcal {Q}^{\prime }_\ell $
, and since
$f(x)\geq f_\ell (x)$
we further have

Consequently
$\eta _\ell \leq \eta $
. The second inequality in Eq. (136) follows analogously, utilizing that
$\eta $
is the unique minimizer of
$\mathcal {Q}^{\prime \prime }_\ell $
and that

Using the fact that , see [Reference Nam, Ricaud and Triay23], the fundamental solution

for the differential operator
$-2\Delta _{\mathcal {M}_*}$
and the observation
$\frac {1}{|\mathcal {M}_*^{-1} v|}\leq \frac {\|\mathcal {M}_*\|}{|v|}$
, we obtain

where we have used symmetric rearrangement. Since
$\int _{\mathbb R^9}\frac {1}{(1+|y|)^4 |y|^{7}}\mathrm {d}y<\infty $
is finite, we obtain that
$\frac {1}{-2\Delta _{\mathcal {M}_*}}(f-f_\ell )$
converges point-wise to zero and consequently
$\eta _\ell $
converges point-wise to
$\eta $
by Eq. (136). Using Fatou’s Lemma and
$f_\ell (x)\eta _\ell (x)\geq 0$
, as well as the fact that
$f_\ell $
converges point-wise to f, therefore yields

where we have used in the last inequality that
$f_\ell \eta _\ell \leq f \eta $
by Eq. (136).
Before we can compare the modified scattering length
$b_{\mathcal {M}}(V)$
with its counterpart on the torus in Lemma 16, we need the following auxiliary result Lemma 15.
Lemma 15. Recall the definition of the coefficients
$\lambda _{k,\ell }$
in Eq. (29) and the definition of T in Eq. (19). Then there exists a constant
$C>0$
such that
$|\lambda _{k,\ell }|\leq \frac {C}{N^2}\left (1+\frac {|\ell |^2}{N }\right )^{-1}$
and

Proof. In order to verify Eq. (137), we observe that for
$|\ell |\leq K$
and
$n\geq 2$

can be written as the sum of terms of the form

where the coefficients satisfy
$k_1+\dots + k_m + 2m +a +2b=n$
and either (I) that
$b=1$
, (II) that
$b=0$
and
$a=1$
or (III) that
$b=0$
,
$a=0$
and
$m\geq 1$
. We are going to verify Eq. (138) by induction, using the resolvent identity

We start with the case
$n=2$

and observe that the first term in Eq. (139) is of the type (III) and the second one is of type (I). For the inductive argument, let

be of type X with X being (I), (II) or (III), and let us compute


The terms in the first two lines, see Eq. (140), are clearly of type X again. Regarding the term in the third line, see Eq. (141), we obtain that
$\nabla T$
is type I in case T itself is type I. In case T is type III we obtain that
$\nabla T$
is type II, and finally in case T is type II, we have
$a=1$
and use Eq. (139) again


The right-hand side of Eq. (143) is clearly a sum of a type I and type III term, which concludes the inductive proof of Eq. (138).
In the following we are going to verify individually for the three cases (I)–(III) that the Fourier transform of the expression in Eq. (138) has an
$L^{\infty }$
bound of the order
$N^{-2}(\sqrt {N}+|\ell |)^{n-2}$
for
$n\geq 2$
, which immediately implies Eq. (137). Let us first of all state the useful bounds



for
$k\geq 0$
. Regarding the case (I), we obtain immediately by Eq. (144) and Eq. (146)

Since the case (II) is similar to the case (III), let us directly have a look at the case (III), where we use the fact that by Eq. (145)

to obtain

As a consequence of Eq. (144) and Eq. (146) we have

This yields the desired estimate for the term in Eq. (147), since
$\left \|\frac {1}{\nabla }Q^{\otimes 3} V_N e^{i \ell x} \right \|\lesssim \sqrt {N}^{-2}$
. The bounds on
$\lambda _{k,\ell }$
can be verified similarly.
In the following Lemma 16, we show that the renormalized potential
$N^2(V_N - V_N R V_N)$
converges to
$b_{\mathcal {M}}(V)\delta (x-y,x-z)$
in a suitable sense. The analogous result for Bose gases with two-particle interactions has been verified in [Reference Brooks8, Lemma 1].
Lemma 16. Let
$b_{\mathcal {M}}(V)$
be the modified scattering length introduced in Eq. (4). Then

Furthermore,
$(V_N - V_N R V_N)_{ijk,\ell m n}=0$
in case
$i+j+k\neq \ell +m+n$
and otherwise

for suitable constants
$C_{ijk,\ell m n}$
.
Proof. Let
$\omega $
be the unique minimizer to the variational problem in Eq. (4), which exists according to [Reference Nam, Ricaud and Triay23] and satisfies in the sense of distributions

Furthermore, let
$\chi $
be a smooth function with
$\chi (x)=1$
for
$|x|_\infty \leq \frac {1}{3}$
and
$\chi (x)=0$
for
$|x|_\infty>\frac {1}{2}$
, and let us denote for a function f the rescaled version with
$f^L(x):=f(Lx)$
. Then we define for
$n=(n_1,n_2,n_3)\in \left (2\pi \mathbb {Z}\right )^{3\times 3}$
and
$0<\ell <{\sqrt {N}}$

In the following we want to show that
$\psi _n$
is an approximation of
$RV_Ne^{i n_1 x}e^{i n_2 y}e^{i n_3 z}$
. For this purpose we observe that the function
$\psi _n$
satisfies the differential equation

where we define
$\xi _n:\mathbb {T}^2\longrightarrow \mathbb {R}$
as


In order to verify that
$\xi _n$
can be treated as an error term, we first note that we have

and utilizing the density
$\rho :=-2\Delta _{\mathcal {M}}\omega $
we obtain

Furthermore, we observe that we can write the Fourier transform of
$\xi _n$
as

Since
$\rho \in L^1(\mathbb {R}^6)$
, see [Reference Nam, Ricaud and Triay23], we have
$\widehat {\rho }\in L^\infty (\mathbb {R}^6)$
, and distinguishing between the cases
$|K|\lesssim \frac {{\sqrt {N}}}{\ell }$
and
$|K|\gg \frac {{\sqrt {N}}}{\ell }$
, yields the estimate

Using that
$\rho $
has compact support as a consequence of the scattering equation, we obtain that
${x\nabla \rho (x)\in L^1(\mathbb {R}^6)}$
and therefore

Since
$\widehat {\mathcal {M}^2\nabla \chi }$
is reflection antisymmetric, we furthermore have

Summarizing what we have so far, we can estimate the Fourier coefficients of
$\xi _0$
by

Proceeding similarly for general
$n\neq 0$
we observe the slightly weaker estimate

In the following let R denote the resolvent of the operator
$Q^{\otimes 3}(-\Delta +V_N)Q^{\otimes 3}$
on the torus, and note that we obtain as a consequence of the differential equation Eq. (150)

Using the fact that V has compact support, there exists a
$\ell _0>0$
such that
$\chi ^{\frac {1}{\ell }}(x)=1$
for
$x\in \mathrm {supp}(V)$
and
$\ell \geq \ell _0$
, and therefore we obtain for
$n=(n_1,n_2,n_3)$
and
$m=(m_1,m_2,m_3)$

where
$\overline {n}:=n_1+n_2+n_3$
and we have used that we can express the minimum in Eq. (4) according to [Reference Nam, Ricaud and Triay23] as

We observe that in the case
$m=0$
and
$n=0$
, we even have the exact identity

Consequently we obtain

Using Lemma 15 and the fact that
$(RV)_{ijk,000}=(T-1)_{ijk,000}=0$
in case
$i\neq -(j+k)$
, we can estimate

Again by Lemma 15 we have that

and following the proof of Eq. (153) we obtain that . Therefore

Choosing
$\ell $
of the order
$\sqrt {N}$
yields

For general
$n=(n_1, n_2 ,n_3)$
and
$m=(m_1 , m_2 ,m_3)$
with
$n_1+n_2+n_3=m_1+m_2+m_3$
the estimates in Eq. (155) and Eq. (157) yield in a similar fashion the desired estimate.
In Eq. (156) we saw that
$RV_N$
, an object defined on the torus
$\Lambda $
, is approximated by

which involves the corresponding object
$\omega $
defined on the full space. In the following Lemma 17 we make use of this correspondence again, to compare
$\gamma _N,\mu _N$
and
$\sigma _N$
with
$\gamma ,\mu $
and
$\sigma $
.
Lemma 17. Let
$\gamma _N,\mu _N$
and
$\sigma _N$
be as in Eq. (97), Eq. (102) and Eq. (98), and
$\sigma (V)$
,
$\mu (V)$
and
$\gamma (V)$
as in Eq. (10), Eq. (5) and Eq. (9). Then

and there exists a constant
$\lambda (V)>0$
such that for
$0<\lambda \leq \lambda (V)$

Furthermore,
$\sigma _N$
and
$\gamma _N$
are independent of the parameter K from the definition of
$\pi _K$
below Eq. (17), and the limit
$\mu (V)=\lim _N \frac {\mu _N}{\sqrt {N}}$
is independent of K as well.
Proof. In order to analyze
$\gamma _N$
, let us denote with
$L_i:L^2 \left (\Lambda ^4\right )\longrightarrow L^2 \left (\Lambda ^4\right )$
the linear map that exchanges the fist factor in the tensor product
$L^2 \left (\Lambda ^4\right )\cong L^2 \left (\Lambda \right )^{\otimes 4}$
with the i-th factor and observe that

Furthermore, recall the definition of
$\psi _0$
from Eq. (149) in the proof of Lemma 16 and define

where the second identity holds by a scaling argument for all
$0<\ell <N$
. We observe that by the permutation symmetry of
$V_N$
we have
$L_i V_N\otimes 1 L_i=V_N\otimes 1$
and therefore

Using
$L_i V_N\otimes 1 L_i=V_N\otimes 1$
again, together with the identity in Eq. (156) and the Cauchy-Schwarz inequality yields

Regarding the analysis of the term on the right side of Eq. (160), we observe that

satisfies
$\nabla ^k \rho \in L^1$
due to the regularity assumptions on V. Proceeding as in Eq. (153) we obtain the improved version of Eq. (154)

Similar to Eq. (138) we can write
$\nabla ^n R\xi _0 $
, where
$\xi _0$
is introduced in Eq. (151), as the sum of terms of the form

where the coefficients satisfy
$k_1+\dots + k_m + 2m +a +2b=n$
and either (I) that
$b=1$
, (II) that
$b=0$
and
$a=1$
or (III) that
$b=0$
,
$a=0$
and
$m\geq 1$
as well as
$k_m=0$
. In the following we are going to verify individually for the three cases (I)–(III) that the Fourier transform of the expression in Eq. (162) has an
$L^{\infty }$
bound of the order
$\frac {\sqrt {N}^{n}}{N^3 \ell ^2}$
for
$n\geq 4$
, and consequently

Regarding the case (I), we obtain using Eq. (161) and our regularity assumptions on V by a direct computation in Fourier space, for
$n\geq 4$

Since the case (II) is similar to the case (III), let us directly have a look at the case (III), where we use the fact that
$\|\sqrt {Q^{\otimes 3} V_N Q^{\otimes 3}}R\nabla \|\lesssim 1$
to obtain

Since we have
$\left \|\frac {1}{\nabla }Q^{\otimes 3}\xi _0\right \|\lesssim \frac {1}{N\ell ^2 }$
, we obtain together with Eq. (148) that the term in Eq. (164) is bounded by
$\frac {\sqrt {N}^{k_1+\dots + k_m+2(m-1)-2}}{N\ell ^2 }=\frac {\sqrt {N}^{n}}{N^3 \ell ^2 }$
, which concludes the proof of Eq. (163). Consequently

Using by Lemma 15 for
$i\neq j$
, we further have

By Eq. (160) we consequently obtain
$\left |\sqrt {N}^{-1}\gamma _N-\gamma ^{(\ell )}\right |\lesssim \ell ^{-\frac {3}{2}}$
for
$\ell \leq \sqrt {N}$
. Note that for
$\ell _1,\ell _2>0$
we can always pick an arbitrary
$N\geq \max \{\ell _1,\ell _2\}^2$
yielding

that is,
$\gamma ^{(\ell )}$
is convergent with rate
$\frac {1}{\ell ^{\frac {3}{2}} }$
, and by monotone convergence the limit is given by
$\gamma (V)$
.
In order to establish the convergence of
$\sigma _N$
, let us define
$f_{N,\ell }:=(V_N\otimes 1) 1\otimes \psi _0$
, where we keep track of the N and
$\ell $
dependence in our notation, and

for
$\ell <\sqrt {N}$
and let
$R_4$
be defined above Eq. (96). As a consequence of the operator inequality

we obtain by Eq. (165) and Eq. (166)

Using the identity Eq. (156), this immediately implies for
$\ell <\sqrt {N}$

To understand the dependence of
$\sigma _{N,\ell }$
on the parameter N, recall the function

and
$\eta _\ell :\mathbb {R}^9\longrightarrow \mathbb {R}$
from Lemma 14, which solves in the sense of distributions

By Lemma 14 we have the point-wise bound
$0 \leq \eta _\ell \leq \eta _\ell ^*$
with

In the following let us write
$x_1 g$
for the function
$x\mapsto x_1 g(x)$
. By Eq. (168) we obtain that
$\rho _{\ell }:=-2\Delta _{\mathcal {M}_*}\eta _\ell $
satisfies the (uniform in
$\ell $
) bounds


where we have used in the second estimates that
$f_\ell (x)$
is compactly supported in the variables
$x_1$
and
$x_2$
, and satisfies
$\sup _{x_1,x_2}f_\ell (x)\lesssim \frac {1}{1+|x_3|^4}$
, see the estimates on
$\omega $
in [Reference Nam, Ricaud and Triay23], and therefore
${\left \|(1+|x_1|) f_\ell \right \|_{L^1(\mathbb {R}^9)}\lesssim 1}$
, as well as the fact that
$x\mapsto \frac {1}{|x|^4}\mathbb {V}(x)\in L^1(\mathbb {R}^9)$
and hence

and similarly we obtain
$ \| \mathbb {V} \eta _\ell ^*\|_{L^1(\mathbb {R}^9)}\lesssim 1$
. Using Eq. (168), we obtain the analogue estimates on the derivatives of
$\rho _\ell $

Having
$\eta _\ell $
at hand, we use a smooth function
$\chi _*$
with
$\chi _*(x)=1$
for
$|x|_\infty \leq \frac {1}{2}$
and
$\chi _*(x)=0$
for
$|x|_\infty>\frac {2}{3}$
, in order to define

Notably, the state
$\Psi $
allows us to express

with
$\zeta :=[2\Delta _{\mathcal {M}_*},\chi _*] \eta _{\ell }^{\sqrt {N}}=2\Delta _{\mathcal {M}_*}(\chi _*)\eta _{\ell }^{\sqrt {N}}+4\mathcal {M}_*^2 \nabla (\chi _*)\nabla \eta _{\ell }^{\sqrt {N}}$
and

Proceeding as in the proof of Eq. (154), we have by Eq. (171)

Using the fact that
$\|\sqrt {Q^{\otimes 4}\mathbb {V}_N Q^{\otimes 4}}R_4\nabla \|\leq 1$
we furthermore obtain

Using furthermore Eq. (162), we can utilize Eq. (173) to improve this result to

In analogy to Eq. (161), one can show that
$\left |\widehat {\chi ^{\frac {\sqrt {N}}{\ell }}\omega ^{\sqrt {N}}} (k)\right |\lesssim N^{-2}\frac {1}{1+|k|^2} \left (1+\frac {|k|^2}{N}\right )^{-m}$
, and therefore we have
$\left |\widehat {f}_{N,\ell }(K)\right |\lesssim N^{-\frac {7}{2}}\left (1+\frac {|K|^2}{N}\right )^{-m}$
, which yields together with Eq. (174)

Furthermore, in analogy to Eq. (174), we have the estimate

Denoting with
$\mathbb {I}$
the set of all indices
$K=(k_1,\dots ,k_4)$
such that
$k_1+\dots + k_4=0$
and at least one of the indices satisfies
$k_\alpha =0$
, we obtain

where we used

Applying Eq. (172), Eq. (175) and Eq. (176) therefore yields for
$\ell <\frac {\sqrt {N}}{2}$

In combination with Eq. (167) and the fact that
$\sigma (V)=\lim _\ell \sigma _\ell (V)$
, see Lemma 14, we obtain that
$|\sigma _\ell (V)-\sigma (V)|\lesssim \frac {1}{\sqrt {\ell }}$
and conclude

To establish the convergence of
$\sqrt {N}^{-1}\mu _N$
, let us recall the effective potential

and let
$\theta $
solve
$-2\Delta \theta =V_{\mathrm {eff}}$
with
$\theta (x)\underset {|x|\rightarrow \infty }{\longrightarrow } 0$
. Then

Applying the techniques developed in this proof so far, yields furthermore

Finally, in order to establish Eq. (159) let us denote with
$\omega _\lambda $
the minimizer in Eq. (4) for the rescaled potential
$\lambda V$
, which satisfies
$0\leq \omega _\lambda \leq 1$
and
$\omega _\lambda (x,y)\leq \frac {\lambda C(V)}{1+|x|^4+|y|^4}$
, for a V dependent, constant
$C(V)>0$
. Consequently
$\lim _{\lambda \rightarrow 0}(1-\omega _\lambda )=1$
, and hence we obtain by dominated convergence

This concludes the proof, since
$\sigma (V)\geq 0$
and

Making use of Eq. (172) again, we can furthermore verify decay properties for the matrix entries of
$T_4-1$
in momentum space in the subsequent Lemma 18.
Lemma 18. Recall the definition of the linear map
$T_2$
in Eq. (88) and
$T_4$
in Eq. (95). Then there exists a constant
$C>0$
such that
,

Proof. For the purpose of verifying the bound on

with
$K=(uijk)$
, let us choose
$\ell :=\frac {\sqrt {N}}{3}$
and recall the elements
$\zeta $
and
$\Psi $
from Eq. (172), and the set
$\mathbb {I}$
above Eq. (176), in the proof of Lemma 17. With these elements at hand, we can write

In the proof of Lemma 17 we have established

Regarding the sum over
$\mathbb {I}$
we have

Regarding the final term in Eq. (177), we observe that we have the estimate

By Eq. (165) and Eq. (166) we know that
$\left \langle 1 \otimes \big \{RV_N - \psi \big \}, (V_N\otimes 1)1 \otimes \big \{RV_N - \psi \big \}\right \rangle \lesssim N^{-5}$
,

Finally we note that the bound on
$T_2$
is an immediate consequence of the regularity of V and the bounds on
$RV_N$
established in Lemma 15.
A Appendix A
In the following we establish comparability results between transformed and nontransformed quantities. The first result in this direction, Lemma A1, establishes that the unitarily transformed powers of the particle number operator
$\mathcal {N}$
, w.r.t. the transformations
$U_s$
and
$W_s$
, are again of the same order as the bare powers in
$\mathcal {N}$
.
Lemma A1. Let
$U_s$
be the unitary map defined below Eq. (70) and
$W_s$
the one defined below Eq. (124). Then there exists for all
$m\in \mathbb {N}$
constants
$C_m>0$
, such that


Proof. Let us recall the definition of the generator
$\mathcal {G}^\dagger - \mathcal {G}$
with

of the unitary group
$U_t$
from Eq. (70). As a consequence of the bounds on T from Lemma 15, we have

Together with
$0\leq (x+n+3)^k-(x+n)^k\leq C_{n,k} (x+3)^{k-1}$
for a suitable
$C_{n,k}>0$
, we obtain

Applying Duhamel’s formula then yields

Consequently Grönwall’s inequality gives us

for a suitable constant
$C>0$
, which concludes the proof of Eq. (A.1). The proof of Eq. (A.2) follows analogously from
$ \pm \left (\mathcal {G}_2+\mathcal {G}_2^\dagger \right )\lesssim \mathcal {N}+1$
and

where we have used Lemma 18 in order to control the coefficients of
$T_2$
and
$T_4$
.
In the subsequent Lemma A2 we are going to compare the kinetic energy
$ \sum _{k}|k|^{2\tau }a_k^\dagger a_k$
in the operators
$a_k$
with a fractional Laplace
$(-\Delta )^{\tau }$
, with the corresponding expression in the variables
$c_k$
.
Lemma A2. Let
$0 \leq \tau \leq 1$
and
$0\leq \sigma <\frac {1}{2}$
. Then
$\sum _{k}|k|^{2\sigma }(c_k-a_k)(c_k-a_k)^\dagger \lesssim \frac {1}{N}\mathcal {N}^2$
, and furthermore we have for integers
$s\geq 0$

Proof. Let us define
$\left (G^{(I,I')}_\tau \right )_{ij,i' j'}:=\frac {1}{4}\sum _{k}|k|^{2\tau } \overline {(T-1)_{i'j'k,I' }}(T-1)_{ijk,I}$
for

as well as for
$0\leq \gamma \leq 1$
the operator-valued vector and matrix

with
$\mathcal {K}_{ \gamma ,2}:=(-\Delta _x)^\gamma + (-\Delta _y)^\gamma $
. With these definitions at hand we obtain

For
$\gamma>\tau -\frac {1}{2}$
we have by the estimates from Lemma 15 that

Together with

on the N particle sector, we obtain
$\left (\Upsilon ^{(I,I')}_{\gamma ,\tau }+\mathrm {H.c.}\right ) \leq \frac {C}{N}\left (\mathcal {N}+1\right )^s$
for
$\gamma>\tau -\frac {1}{2}$
and a suitable constant C. Using Cauchy-Schwarz we therefore have

Applying this result for
$\tau ':=\sigma $
,
$\gamma ':=0$
and
$s':=0$
, yields the first claim of the Lemma

Concerning Eq. (A.3), we have

and furthermore we can express

with

Following the proof of the first part of the Lemma, we obtain for
$\gamma>\tau -\frac {1}{2}$

Using Lemma 15 again, yields
$|f|\lesssim N^{\tau -3}$
and
$|g_j|\lesssim N^{\max \{\tau -\frac {1}{2},0\}}-4$
, and consequently

Summarizing what we have so far we obtain for
$\gamma>\tau -\frac {1}{2}$

Choosing
$\gamma :=\max \{\tau -\frac {1}{3},0\}$
and iterating this equation at most two times with
$\tau ':=\max \{\tau -\frac {1}{3},0\}$
and
$\gamma ':=\max \{\gamma -\frac {1}{3},0\}$
, and using
$\mathcal {N}\leq N$
, yields the desired statement.
Similar to Lemma A2, the following Lemma A3 allows us to compare the operators

with the operators
$c_k$
.
Lemma A3. Then there exists a
$C>0$
, such that

Proof. Similar to Eq. (A.4), we can write

where
$f^{I,I'},g_k^{I,I'}, \widetilde {\Upsilon }^{(I,I')}_{1,1}, X^{I,I'}_0$
and
$X^{I,I'}_1$
are defined below Eq. (A.4) for the concrete choice
$s:=0$
and

Using
$I,I'\neq 0$
, we obtain the improved estimates
$\pm X^{I,I'}_0\lesssim N^2 \mathcal {N}$
and
$\pm X^{I,I'}_1\lesssim N^2 \mathcal {N}$
. Consequently

Furthermore,

where we have used Eq. (40) in the last estimate.
Acknowledgments
We would like to thank Marco Caporaletti and Benjamin Schlein for insightful discussions.
Competing interest
The authors have no competing interest to declare.
Financial Support
Funding from the ERC Advanced Grant ERC-AdG CLaQS, grant agreement n. 834782, is gratefully acknowledged.