Skip to main content
Springer logoLink to Springer
. 2016 Nov 29;107(3):533–552. doi: 10.1007/s11005-016-0915-x

Triviality of a model of particles with point interactions in the thermodynamic limit

Thomas Moser 1,, Robert Seiringer 1
PMCID: PMC7089695  PMID: 32226208

Abstract

We consider a model of fermions interacting via point interactions, defined via a certain weighted Dirichlet form. While for two particles the interaction corresponds to infinite scattering length, the presence of further particles effectively decreases the interaction strength. We show that the model becomes trivial in the thermodynamic limit, in the sense that the free energy density at any given particle density and temperature agrees with the corresponding expression for non-interacting particles.

Keywords: Point interactions, Scattering length, Fermi gas, Thermodynamic limit

Introduction

Due to their relevance for cold-atom physics [27], quantum-mechanical models of particles with zero-range interactions have recently received a lot of attention. Of particular interest is the unitary limit of infinite scattering length, where one has scale invariance due to the lack of any intrinsic length scale (see, e.g., [3, 4, 11, 12, 25]). Despite some effort [57, 9, 21], it remains an open problem to establish the existence of a many-particle model with two-body point interactions. Such a model is known to be unstable in the case of bosons (a fact known as Thomas effect [3, 5, 24], closely related to the Efimov effect [8, 22, 26]) and hence can only exist for fermionic particles. In contrast, the two-body problem is completely understood and point interactions can be characterized via self-adjoint extensions of the Laplacian on R3\{0} (see [1] for details). These self-adjoint extensions can be interpreted as corresponding to an attractive point interaction, parametrized by the scattering length a, with interaction strength increasing with 1/a. For non-positive scattering length, a0, the attraction is too weak to support bound states, while there exists a negative energy bound state for a>0.

In the case of non-positive scattering length, a0, corresponding to the absence of two-body bound states, point interactions can alternatively be defined via the quadratic form

R31|x|-1a2f(x)2dxonL2(R3,(|x|-1-a-1)2dx) 1.1

The unitary limit corresponds to a-1=0. Recall that the scattering length is defined (see, e.g., [14, Appendix C]) via the asymptotic behavior of the solution to the zero-energy scattering equation, which in this case is simply equal to |x|-1-a-1, corresponding to f1. To see that (1.1) corresponds to a point interaction at the origin, note that an integration by parts shows that

|x|ϵ1|x|-1a2f(x)2dx=|x|ϵ1|x|-1af(x)2dx-|x|=ϵ1|x|-1a1|x|2|f(x)|2dω 1.2

for any ϵ>0. The last term vanishes as ϵ0 if f vanishes faster than |x|1/2 at the origin.

We consider here a many-body generalization of (1.1), which was introduced in [2]. It has the advantage of being manifestly well defined, via a non-negative Dirichlet form. As already noted above, in general it is notoriously hard to define many-body systems with point interactions, see [57, 9, 21], due to the inherent instability problems. The model under consideration here was studied in [10], where it was shown to satisfy a Lieb–Thirring inequality, i.e., the energy can be bounded from below by a semiclassical expression of the form Cρ(x)5/3dx, with ρ the particle density and C a positive constant. Up to the value of C, this is the same as the inequality for non-interacting fermions used by Lieb and Thirring [15, 16] in their proof of stability of matter. (For other recent work on Lieb–Thirring inequalities for interacting particles, see [1720].)

The model considered here has the disadvantage that the interaction is not purely two-body, however. In fact, it is a full many-body interaction, its strength depends on the position of all the particles and is weakened due to their presence. We shall show here that the effects of the interaction actually disappear in the thermodynamic limit, and the thermodynamic free energy density agrees with the one for non-interacting fermions.

In the next section, we shall introduce the model and explain our main results. The rest of the paper is devoted to their proof.

Model and main results

For N2, x=(x1,,xN)R3N, let g:R3NR denote the function

g(x)=1i<jN1|xi-xj|. 2.1

We consider fermions with q1 internal (spin) states, described by wave functions in the subspace AqNL2((R3×{1,,q})N,g(x)2dx) of functions that are totally antisymmetric with respect to permutations of the variables yi=(xi,σi), where xiR3 and σi{1,,q}. For ψAqN, our model is defined via the quadratic form

Eg(ψ)=i=1NR3Ng(x)2|iψ(y)|2dy 2.2

where i stands for the gradient with respect to xiR3, and we introduced the shorthand notation dy=σdx with σ=(σ1,,σN). Since g is a harmonic function away from the planes {xi=xj} of particle intersection, an integration by parts as in (1.2) shows that (2.2) corresponds to a model of point interactions, as Eg(ψ)=i=1N|igψ|2 in case ψ has compact support away from these planes. More generally, Eg(ψ)=i=1N|igψ|2 holds if ψ vanishes faster than the square root of the distance to the planes of intersection, which is in particular the case for smooth and completely antisymmetric functions of the spatial variables. In other words, the model is trivial for q=1.

For N particles in a cubic box [0,L]3R3, the free energy at temperature T=β-1>0 is defined as usual as:

Fg=-Tlntre-βHg 2.3

where Hg denotes the operator defined by the quadratic form (2.2), restricted to functions in AqNH1(R3N;g(x)2dx) with support in ([0,L]3)N. The latter restriction corresponds to choosing Dirichlet boundary conditions on the boundary of the cube [0,L]3. Alternatively, one can use the variational principle [13, Lemma 14.1] to write the free energy as:

Fg(β,N,L)=-Tlnsup{ψk}ψi|ψjg=δijke-βEg(ψk) 2.4

where ·|·g denotes the inner product on L2((R3×{1,,q})N,g(x)2dx),

ψi|ψjg=R3Ng2(x)ψi(y)¯ψj(y)dy, 2.5

and the supremum is over all finite sets of orthonormal functions in AqN with support in ([0,L]3)N. We are interested in the thermodynamic limit

fg(β,ρ)=limNρNFg(β,N,(N/ρ)1/3) 2.6

where ρ>0 denotes the particle density.

In the non-interacting case corresponding to taking g1, the free energy density can be evaluated explicitly, and is given by [23]

f(β,ρ)=supμRμρ-qT(2π)3R3ln1+e-β(p2-μ)dp 2.7

Our main result shows that the two functions, fg and f, are actually identical.

Theorem 2.1

For any β>0 and ρ>0, and any q1,

fg(β,ρ)=f(β,ρ) 2.8

We shall actually prove a stronger result below, namely a lower bound on Fg(β,N,L) for finite N which agrees with the corresponding expression for non-interacting particles, F(β,N,L), to leading order in N, with explicit bounds on the correction term. Note that the corresponding upper bound is trivial, since for functions ϕC0((R3×{1,,q})N)

Eg(ϕ/g)=i=1N|iϕ(y)|2dy 2.9

and hence Fg(β,N,L)F(β,N,L). Moreover, as already noted above one has Fg(β,N,L)=F(β,N,L) for q=1, since functions in A1N vanish whenever xi=xj for some ij. Hence, it suffices to consider the case q2.

Theorem 2.1 also holds true for the ground state energy, i.e., β=, where f(,ρ)=35(6π2/q)2/3ρ5/3. The proof of the equality (2.8) in this case is actually substantially easier, as the analysis of the entropy in Sect. 6 is not needed.

Intuitively, the result in Theorem 2.1 can be explained via a comparison of (2.2) with (1.1). Effectively, the scattering process between two particles, i and j, say, corresponds to a non-zero scattering length of the form

-1aeff={k,l}{i,j}1|xk-xl|. 2.10

In the limit of large particle number, the sum of these other terms diverges, corresponding to an effective scattering length zero, i.e., no interactions.

A minor modification of the proof shows that Theorem 2.1 also holds for a model where the function 1/|x| in (2.1) is replaced by 1/|x|-1/a for a0, corresponding to a two-body interaction with negative scattering length a. This only increases the effective scattering length aeff.

From Theorem 2.1, we conclude that the model (2.2) is not suitable to describe a gas of fermions with point interactions, as it becomes trivial in the thermodynamic limit. No non-trivial models that are proven to be stable for arbitrary particle number exist to this date, however. Such non-trivial models are not expected to be given by a Dirichlet form of the type (2.2), since such forms are naturally well defined even in the bosonic case, where point-interaction models are known to become unstable due to the Thomas effect [3, 5, 8, 22, 24, 26].

In the remainder of this paper, we shall give the proof of Theorem 2.1. We start with a short outline of the main steps in the next section.

Outline of the proof

In the first step in Sect. 4, we shall localize particles in small boxes. This part of the Dirichlet–Neumann bracketing technique is quite standard, but it does not directly allow us to reduce the problem to fewer particles, as the interactions depend on the location of all the particles, including the ones in different boxes. Still this step allows us to compare our model with the corresponding one for non-interacting fermions, by utilizing a suitable version of the Hardy inequality to quantify the effect of the deviation of the weight function g in (2.1) from being a constant. This analysis is done in Sect. 5. Note that the relevant constant to compare g with depends on the distribution of the particles in the various boxes, hence the importance of the first step. An important point in the analysis is a control on the particle number distribution, which is obtained in Proposition 5.4.

In Sect. 6, we shall give a rough bound on the entropy for large energy, which will allow us to conclude that to compute the free energy (2.4), it suffices to consider only states with energy ENlnN. We do this by applying the localization technique to very small boxes, with side length decreasing with energy, in order to have to consider effectively only the ground states in each small box.

In the low energy sector, corresponding to energies ENlnN, our bounds in Sect. 5 allow to make a direct comparison of our model with non-interacting fermions. This comparison is detailed in Sect. 7. For this purpose, we shall choose much larger boxes than in the previous step, very slowly increasing to infinity with N in order for finite size effects to vanish in the thermodynamic limit. Finally, Sect. 8 collects all the results in the previous sections to give the proof of Theorem 2.1.

Throughout the proof, we shall use the letter c for universal constants independent of all parameters, even though c might have different values at different occurrences. Similarly, we use cη for functions of η=βρ2/3 that are uniformly bounded for η>ε for any ε>0. Note that the free energy for noninteracting particles in (2.7) satisfies the scaling relation

f(β,ρ)=ρ5/3f(η,1),η=βρ2/3, 3.1

and η corresponds to the zero-temperature limit.

Particle localization in small boxes

Given an integer m2, we shall divide the cube [0,L]3 into M=m3 disjoint cubes of side length =L/m, denoted by {Bi}i=1M. To obtain a lower bound on Eg, we introduce Neumann boundary conditions on the boundary of each box Bi.

Specifically, given a vector n={n1,,nM} of nonnegative integers with j=1Mnj=N, let Bsym(n) denote the subset of [0,L]3N where exactly nj particles are in Bj, for all 1jM. More precisely, if

B(n)=B1n1××BMnM 4.1

and, for general AR3N and πSN (the permutation group of N elements)

π(A)={x:π-1(x)A},π(x):=(xπ(1),,xπ(N)) 4.2

we have

Bsym(n)=πSNπ(B(n)) 4.3

Then, clearly

1=nχBsym(n)(x) 4.4

for almost every x[0,L]3N. Correspondingly, one can write for any ψAqN supported in [0,L]3N

ψ(y)=nχBsym(n)(x)ψ(y)=:nψn(y). 4.5

Note that each ψn is a function in AqN with the property that it is non-zero only if exactly nj particles are in Bj for any 1jM. In particular, the functions appearing in the decomposition on the right side of (4.5) all have disjoint support.

Conversely, given a set of functions ψnAqN supported in Bsym(n), we can define ψAqN via (4.5). Hence, there is a one-to-one correspondence between functions in AqN and sets of functions ψn. We now redefine our energy functional Eg as:

Eg(ψ)=ni=1NBsym(n)g(x)2|iψn(y)|2dy 4.6

This coincides with the definition (2.2) in case ψH1((R3×{1,,q})N,g(x)2dx), but is more general since it allows for wave functions that are discontinuous at the boundaries of the Bj, effectively introducing Neumann boundary conditions there.

Note that with the definition (4.6) above, we have

Eg(ψ)=nEg(ψn)forψ=nψn 4.7

In particular, the corresponding operator is diagonal with respect to the direct sum decomposition of AqN into functions supported on Bsym(n), and hence the min–max principle implies the bound

sup{ψk}ψi|ψjg=δijke-βEg(ψk)nsup{ψkn}ψin|ψjng=δijke-βEg(ψkn) 4.8

where on the right side it is understood that each ψjn is supported in Bsym(n).

As a final step in this section, we want to simplify the problem by getting rid of the antisymmetry requirement for particles localized in different boxes. There exists a simple isometry between functions ψn in AqN and functions whose support is on the smaller set B(n) in (4.1), where x1,,xn1B1, xn1+1,,xn1+n2B2, etc., and which are antisymmetric only with respect to permutations of the yi corresponding to xi in the same box. This isometry is simply

ψnN!j=1Mnj!1/2χB(n)ψn 4.9

Note that the normalization factor is chosen such that both sides have the same norm, and the left side can be obtained from the right by a suitable antisymmetrization over all variables yi. Moreover, both functions yield the same value when plugged into Eg. Let AqN,(n) denote the set {χB(n)ψ:ψAqN}, i.e., functions supported in B(n) that are antisymmetric in the variables corresponding to the same box. The bound (4.8) and the above observation imply that

Fg(β,N,L)-Tlnnsup{ψkAqN,(n)}ψi|ψjg=δijke-βEg(ψk) 4.10

Energy and norm bounds

Our goal in this next step is to derive a lower bound on Eg(ψ) for ψAqN,(n), i.e., functions supported in B(n), and to compare the norm of such a ψ with the standard, unweighted L2 norm. For this purpose, we shall need a certain version of the Hardy inequality, which will be derived in the next subsection.

Hardy inequalities

Recall the usual Hardy inequality

R3|f(x)|2dx14R3|f(x)|2|x|2dx 5.1

for functions fH˙1(R3). We shall need a local version of (5.1) on balls.

Lemma 5.1

Let BR3 denote the open centered ball with radius . For any fH1(B)

2B|f(x)|2dx+922B|f(x)|2dx14B|f(x)|2|x|2dx 5.2

Proof

We apply the Hardy inequality (5.1) to the function h(x)=f(x)[1-|x|/]+, where [·]+ denotes the positive part. For the right side of (5.1), we obtain

14B|h(x)|2|x|2dx=14B|f(x)|2|x|21-2|x|+|x|22dx1-ε4B|f(x)|2|x|2dx-1-ε4ε2B|f(x)|2dx 5.3

for any ε>0. For the left side of (5.1), a simple Schwarz inequality yields

B|h(x)|2dx(1+δ)B|f(x)|2dx+1+δδ2B|f(x)|2dx 5.4

for δ>0. In combination, we obtain the desired inequality (5.2) by choosing ε=1/6 and δ=2/3.

For later use, we need a version of Lemma 5.1 on cubes with arbitrary location relative to the singularity.

Lemma 5.2

Let C=[0,]3. For any yR3 and any fH1(C),

c0C|f(x)|2dx+c12C|f(x)|2dx14C|f(x)|2|x-y|2dx 5.5

with c016 and c1144.

The stated bounds on the constants c0 and c1 are presumably far from optimal, but suffice for our purpose.

Proof

If yC, we can replace it by the point in C closest to y. This can only increase the right side. Hence, we may assume that yC. Let B denote the ball of radius /2 around y. Then

14C\B|f(x)|2|x-y|2dx12C\B|f(x)|2dx 5.6

Define a function f~ by extending f to [-,2]3 as

f~(x1,x2,x3)=f(τ(x1),τ(x2),τ(x3)) 5.7

where

τ(x):=-xx[-,0]xx[0,]2-xx[,2] 5.8

Then, f~H1([-,2]3). Since B[-,2]3, we get with the aid of the Hardy inequality (5.2) on B (with /2 in place of )

14CB|f(x)|2|x-y|2dx14B|f~(x)|2|x-y|2dx2B|f~(x)|2dx+182B|f~(x)|2dx82CB|f(x)|2dx+182CB|f(x)|2dx 5.9

In the last step, we used that B intersects, besides C, at most 7 other translates of C, and that the intersection of B with these translates is, when reflected back to C, contained in CB (see Fig. 1). In combination, (5.6) and (5.9) imply (5.5).

Fig. 1.

Fig. 1

Two-dimensional illustration of the reflection technique used in the proof of Lemma 5.2. The box C and two of its neighbor boxes are shown, as well as the ball B around yC. Using the extended function f~, we can mirror C\B back into CB. There are at most 8 reflected components in three dimensions, the worst case being if the ball B intersects with a corner of C

A lower bound on Eg

Let ψ be an L2((R3×{1,,q})N,g(x)2dy)-normalized function in AqN,(n), defined just above (4.10). Let djk denote the distance between boxes Bj and Bk. For xB(n), we can bound

g(x)1j<kMnjnkdjk+23+j=1Mnj(nj-1)23K-+V43 5.10

where

K-=1j<kMdjk>0njnkdjk+23andV=j=1Mnj(nj+mj-1) 5.11

Here, mk denotes the total number of particles in the 26 neighboring boxes of Bk. The bound (5.10) immediately leads to the lower bound

Eg(ψ)K-+V432E(ψ) 5.12

for ψAqN,(n), where E on the right side stands for the energy functional for noninteracting particles, corresponding to g1 in (4.6).

Bounds on norms

In the following, it will be necessary to compare the norm ·g=·|·g1/2 with the standard L2 norm · without weight. For ψAqN,(n), the bound (5.10) immediately implies the lower bound

ψgK-+V43ψ 5.13

To obtain a corresponding upper bound, we proceed as follows. For given i, corresponding to xiBk for some box Bk, let N[i] be the set of js with ji such that xj is either in the same box Bk or in one of the 26 neighboring boxes touching Bk. With mk as defined above, |N[i]|=nk+mk-1 for xiBk. Then

g(x)12i=1NjN[i]1|xi-xj|+K+withK+=1j<kMdjk>0njnkdjk 5.14

for xB(n). The Cauchy–Schwarz inequality implies

ψg2(1+ε)K+2ψ2+1+ε-1V4i=1NjN[i]|ψ(y)|2|xi-xj|2dy 5.15

for any ε>0, where V is defined in (5.11). In the last term, we use the Hardy inequality (5.5) for the integration over xi, and obtain

ψg2(1+ε)K+2+c121+ε-1V2ψ2+1+ε-1c0Vi=1NN[i]|iψ(y)|2dy 5.16

If we reinsert g(x)2 into the last integrand using (5.10), we thus obtain the following lemma.

Lemma 5.3

For ψAqN,(n), we have the bounds

K-+V432ψ2ψg2(1+ε)K+2+c1ε2V2ψ2+(1+ε-1)c0VK-+V432i=1NN[i]|iψ(y)|2g(x)2dy 5.17

for any ε>0, where K± and V are defined in (5.11) and (5.14), respectively.

A bound on the number of particles in a box

Let again ψ be a wavefunction in AqN,(n) and let us assume it is normalized, i.e., ψg=1. We have the following a priori bound.

Proposition 5.4

There exists a constant κ>0 such that for any normalized ψAqN,(n) and any >0 we have

Eg(ψ)κq2/3j=1Mnj-q+5/32 5.18

Here, [·]+=max{0,·} denotes the positive part. The bound (5.18) allows us to conclude that for all normalized ψAqN,(n) with Eg(ψ)<E we have njq for all j if we choose such that E2q2/3κ. Furthermore, for large E2 we get the bound maxjnjq2/5(E2)3/5.

Proof

We use Lemma 3 from [10] which states that for a subset A{1,,N} corresponding to particles xkBj for kA,

iABj|A|g(x)2|iψ(y)|2dyAκ~2|A|-q+Bj|A|g(x)2|ψ(y)|2dyA 5.19

for some κ~>0 independent of A, and ψ. Here, yA is short for {yi}iA. Integrating this over the {yj}jA and summing over j yield (5.18) with the exponent 5/3 replaced by 1, and κ=κ~q2/3.

To raise the exponent from 1 to 5/3, we partition Bj into μ3 disjoint cubes {Ck}k of side length /μ for some integer μ1. We use the identity

1=QAsQχCk(xs)tQcχCkc(xt) 5.20

for xABj|A|, where Qc denotes A\Q and Ckc=Bj\Ck. By plugging (5.20) into (5.19), we obtain

iABj|A|g(x)2|iψ(y)|2dyA=iAk=1μ3Bj|A|χCk(xi)g(x)2|iψ(y)|2dyA=iAk=1μ3QABj|A|sQχCk(xs)tQcχCkc(xt)χCk(xi)g(x)2|iψ(x)|2dxA=k=1μ3QAiQBj|A|sQχCk(xs)tQcχCkc(xt)g(x)2|iψ(x)|2dxA 5.21

For the integration over {ys}sQ we can again use (5.19), with suitably rescaled variables to replace the integration over Bj with the one over Ck. (Note that g is homogeneous of order -1 and satisfies the simple scaling property g(λx)=λ-1g(x) for λ>0.) This yields the bound

(5.21)k=1μ3QAμ2κ~2|Q|-qCk|Q|dyQCkc(|A|-|Q|)dyQcg(x)2|ψ(y)|2=μ2κ~2(|A|-μ3q)Bj|A|g(x)2|ψ(y)|2dyA 5.22

In the last step, we used again the identity (5.20) as well as

|A|=k=1μ3QA|Q|sQχCk(xs)tQcχCkc(xt) 5.23

Since the left side of (5.22) is obviously non-negative, we can replace |A|-μ3q by its positive part on the right side.

It remains to choose μ. If we ignore the restriction that μ1 is an integer, we would choose μ=(2/5)(|A|/q)1/3 to obtain the desired coefficient |A|5/3/q2/3. It is easy to see that

supμNμ2|A|-μ3q+cq2/3|A|-q+5/3 5.24

for some universal constant c>0. This proves the desired bound, with κ=κ~c.

A bound on the entropy

In this section, we shall use the estimates above to give a rough bound on

Ng(E)=trχHg<E, 6.1

that is, the maximal number of orthonormal functions in AqN with Eg(ψ)<E, for some (large) E. Its logarithm is, by definition, the entropy. Using the localization technique described in Sect. 4, the min–max principle implies that

Ng(E)nNgn(E) 6.2

where Ngn(E) is the maximal number of orthonormal functions in AqN,(n) with Eg(ψ)<E. Given E, we shall choose small enough such E2q2/3κ, with κ the constant in Proposition 5.4. As remarked there, this implies that njq for all 1jM.

We will actually show that if E2 is small enough, then the spectral gap for an excitation is larger than E, and hence Ngn(E) is simply equal to the dimension of the space of ground states.

Lemma 6.1

There exists a universal constant c>0 such that if we choose E2c, then

Ngn(E)=j=1Mqnj 6.3

Proof

With the aid of (5.12), we have

Eg(ψ)K-+V432E(ψ) 6.4

for ψAqN,(n). The ground states of the operator corresponding to the quadratic form E are all constant, i.e., they are simply products of anti-symmetric functions of the spin variables corresponding to each box, and have zero energy. The spectral gap above the ground state energy is given by (π/)2. With P0 denoting the projection in L2(B(n),dy) onto the ground state space, we thus have

E(ψ)π22(1-P0)ψ2 6.5

To bound the norm on the right side from below in terms of the weighted ·g norm, we shall use Lemma 5.3. In (5.17), we can simply bound

i=1NN[i]|iψ(y)|2g(x)2dy<Eψg2i=1NN[i]=EVψg2 6.6

to obtain

ψg2(1+ε)K+2+c121+ε-1V2ψ2+48c01+ε-1E2ψg2 6.7

for any ε>0 and any ψAqN,(n) with Eg(ψ)<Eψg2. If E2 is small, we can take ε=1 to conclude that

ψg2cK+2+V2-2ψ2 6.8

Moreover, note that K+(1+23)K-, since djk>0 actually implies djk. We thus also have that

ψg2cK-+V432ψ2 6.9

Applying this to (1-P0)ψ in (6.5) and inserting the resulting bound in (6.4), we obtain

Eg(ψ)c-2(1-P0)ψg2 6.10

Finally, note that the ground states of Eg and E actually agree, up to a multiplicative normalization constant. Hence, if ψ is orthogonal to a ground state with respect to the inner product ·|·g, then

(1-P0)ψg2=ψg2+P0ψg2ψg2 6.11

This concludes the proof.

In combination with (6.2), Lemma 6.1 yields the bound

Ng(E)nj=1Mqnj=qMNqMeNN 6.12

for E2c. We recall that the number of boxes is M=(L/)3=N/(ρ3), which is large for E21 and EL-2. Hence, we get the upper bound

Ng(E)cqE3/2ρN 6.13

for a suitable constant c>0. This bound readily implies the following proposition.

Proposition 6.2

Let {Ej}j denote the eigenvalues of the Hamiltonian Hg associated with the quadratic form Eg in (2.2) on AqN. For given η=βρ2/3, there exists a cη>0 such that if E¯cηβ-1NlnN then

EjE¯e-βEj2e-12βE¯ 6.14

Proof

We have

EjE¯e-βEjk0Ng((k+2)E¯)e-(k+1)βE¯ 6.15

and, thus, the result follows if

Ng((k+2)E¯)e-(k+12)βE¯12k 6.16

for all k0. Using the bound (6.13), one easily checks that this is the case under the stated condition on E¯ for suitable cη.

For evaluating the free energy, we can thus limit our attention to eigenvalues Ej satisfying βEjcηNlnN for suitable cη>0. We shall show in the next section that in this low energy sector the eigenvalues are well approximated by the corresponding ones for non-interacting particles.

Comparison with non-interacting particles in the low-energy sector

We shall now investigate the bounds derived in Sect. 5 more closely and apply them to the low energy sector, where Eg(ψ)Eψg2 for some ENlnN. We again localize the particles into boxes, this time with much larger , however. We start with the estimate on the ratio of the norm ψg to the standard, non-weighted L2 norm ψ.

Proposition 7.1

Let ψAqN,(n) satisfy Eg(ψ)Eψg2 for some E with E21 for large N. Then

1K-+V432ψ2ψg21-δ 7.1

with

δcq1/5(E2)3/10N-1/3(ρ3)-1/6+q2/5(E2)11/10N-7/6(ρ3)-1/3 7.2

with K- and V defined in (5.11).

We note that δ is small if

E2min{N10/9(ρ3)5/9,N35/33(ρ3)10/33} 7.3

which gives us freedom to choose large while ENlnN. We will choose Nν for rather small ν below, in which case the first term in (7.2) will be dominating.

Proof

The first bound in (7.1) follows immediately (5.17). For the lower bound, we use

i=1NN[i]|iψ(y)|2g(x)2dy27n¯Eψg2 7.4

in (5.17), where we denote n¯=maxjnj. We can also bound V27n¯N and

K-+V43N(N-1)23L 7.5

The second bound in (5.17), thus, becomes

1-1+ε-1c012L2(27n¯)2N(N-1)2Eψg2(1+ε)K+2+c121+ε-1V2ψ2 7.6

for arbitrary ε>0. By assumption E2 is not small, hence we have n¯cq2/5(E2)3/5, as remarked after Proposition 5.4.

It remains to estimate the ratio K-/K+. We distinguish the contribution to the sum coming from djk<r3 and djkr3, respectively, for some large integer r to be chosen below. We have

K+-K-=1j<kMdjk>0njnkdjk23djk+23n¯1j<kM0<djk<r3njdjk23djk+23+1+r2-11j<kMdjkr3njnkdjkcn¯rN+1+r2-1K+ 7.7

By optimizing over r as well as ε and using that n¯cq2/5(E2)3/5, we arrive at the desired result.

In combination with (5.12), Proposition 7.1 yields the lower bound

Eg(ψ)ψg2E(ψ)ψ21-δ 7.8

for ψAqN,(n) in the low energy sector Eg(ψ)<E. This allows us to compare our model directly with non-interacting particles. Note that the eigenfunctions of the operator corresponding to the quadratic form on the right side are tensor products over different boxes and, in particular, the eigenvalues are simply sums over the corresponding eigenvalues of free fermions in each box. The bound (7.8) does not directly give us lower bounds on the eigenvalues of Hg, except for the lowest one, however. To complete the proof, we have to estimate the difference between the inner product ·|·g and the standard inner product on L2, denoted by ·|· in the following.

We define the multiplication operator

G=K-+V43-1g(x) 7.9

which is larger or equal to 1 by (5.10). The bound (5.12) thus reads

Eg(ψ)ψg2E(ψ)Gψ2=ϕ|G-1HG-1|ϕϕ2 7.10

where we introduced ϕ=Gψ and denoted by H the Hamiltonian for non-interacting particles, i.e., the Laplacian on B(n) with Neumann boundary conditions. Note that the orthogonality condition ψj|ψkg=0 is equivalent to ϕj|ϕk=0. Given some E0>0, we define the cutoff Hamiltonian

Hc=Hθ(E0-H), 7.11

with θ denoting the Heaviside step function. This is clearly a bounded operator with HcE0. Obviously

ϕ|G-1HG-1|ϕHc1/2G-1ϕ2 7.12

which we further bound as:

Hc1/2G-1ϕ2Hc1/2ϕ-Hc1/2(1-G-1)ϕ2Hc1/2ϕ2-2Hc1/2ϕHc1/2(1-G-1)ϕHc1/2ϕ2-2E0(1-G-1)ϕϕ 7.13

Now

(1-G-1)ϕ(1-G-2)1/2ϕδ1/2ϕ 7.14

where we used G1 in the first and Proposition 7.1 in the second step. We conclude that

Eg(ψ)ψg2ϕ|Hc-2E0δ1/2|ϕϕ2 7.15

under the conditions stated in Proposition 7.1.

Convergence of the free energy

We now have all the necessary tools to complete the proof of Theorem 2.1. Proposition 6.2 implies that if we choose E¯=cηβ-1NlnN for a suitable constant cη>0, then

Fg(β,N,L)-Tln2e-12βE¯+sup{ψkAqN}ψi|ψjg=δijk=1Ng(E¯)e-βEg(ψk) 8.1

Here, Ng(E¯) denotes the number of states with energy below E¯, which was estimated in (6.13). We can write, alternatively,

sup{ψk}ψi|ψjg=δijk=1Ng(E¯)e-βEg(ψk)=sup{ψk},Eg(ψk)<E¯ψi|ψjg=δijke-βEg(ψk) 8.2

By localizing into small boxes of side length with Neumann boundary conditions, as detailed in Sect. 4, we further have by the min–max principle

(8.2)nsup{ψAqN,(n)},Eg(ψ)E¯ψi|ψjg=δijke-βEg(ψk) 8.3

If we choose E¯21, we can apply the bound (7.15) from the previous subsection. It implies

(8.3)e2βE0δ1/2nsup{ϕGAqN,(n)},ϕk|Hc|ϕkE¯+2E0δ1/2ϕi|ϕj=δijke-βϕk|Hc|ϕk 8.4

with δ defined in Proposition 7.1. If we choose E0 such that E¯+2E0δ1/2E0, which is possible for δ<1/4, we can drop the cutoff in Hc and replace Hc by H, the Laplacian on (jBj)N with Neumann boundary conditions. To obtain an upper bound on (8.4), we can then further neglect the bound on ϕk|H|ϕk, and sum over all eigenvalues. We obtain

(8.4)e2βE0δ1/2e-βF(β,N,L,) 8.5

where F(β,N,L,) denotes the free energy of non-interacting fermions in jBj (with Neumann boundary conditions on the boundaries of the Bj). In particular, in combination (8.1)–(8.5) imply

Fg(β,N,L)F(β,N,L,)-2E0δ1/2-Tln1+2e-12βE¯e-2βE0δ1/2eβF(β,N,L,) 8.6

We will choose 1, in which case F(β,N,L,)N and hence the last term in (8.6) is, in fact, exponentially small in N, since E¯NlnN. To complete the proof, it suffices to observe that

F(β,N,L,)F(β,N,L)-cηNρ1/3 8.7

which is an easy exercise. To minimize the total error, we shall choose

ρ-1/3N1/63lnN-23/21 8.8

to obtain

Fg(β,N,L)F(β,N,L)-cηρ2/3N62/63lnN23/21 8.9

This completes the proof of Theorem 2.1.

Acknowledgements

Open access funding provided by Institute of Science and Technology (IST Austria). Financial support by the Austrian Science Fund (FWF), Project Nr. P 27533-N27, is gratefully acknowledged.

Contributor Information

Thomas Moser, Email: thomas.moser@ist.ac.at.

Robert Seiringer, Email: robert.seiringer@ist.ac.at.

References

  • 1.Albeverio S, Gesztesy F, Høegh-Krohn R, Holden H. Solvable Models in Quantum Mechanics. 2. Providence: American Mathematical Society; 2004. [Google Scholar]
  • 2.Albeverio S, Høegh-Krohn R, Streit L. Energy forms, Hamiltonians, and distorted Brownian paths. J. Math. Phys. 1977;18:907–917. doi: 10.1063/1.523359. [DOI] [Google Scholar]
  • 3.Braaten E, Hammer HW. Universality in few-body systems with large scattering length. Phys. Rep. 2006;428:259–390. doi: 10.1016/j.physrep.2006.03.001. [DOI] [Google Scholar]
  • 4.Burovski E, Prokof’ev N, Svistunov B, Troyer M. Critical temperature and thermodynamics of attractive fermions at unitarity. Phys. Rev. Lett. 2006;96:160402. doi: 10.1103/PhysRevLett.96.160402. [DOI] [PubMed] [Google Scholar]
  • 5.Correggi M, Dell’Antonio G, Finco D, Michelangeli A, Teta A. Stability for a system of N fermions plus a different particle with zero-range interactions. Rev. Math. Phys. 2012;24:1250017. doi: 10.1142/S0129055X12500171. [DOI] [Google Scholar]
  • 6.Correggi, M., Dell’Antonio, G., Finco, D., Michelangeli, A., Teta, A.: A class of Hamiltonians for a three-particle fermionic system at unitarity. Math. Phys. Anal. Geom. 18, Art. 32 (2015)
  • 7.Dell’Antonio G, Figari R, Teta A. Hamiltonians for systems of N particles interacting through point interactions. Ann. Inst. Henri Poincare. 1994;60:253–290. [Google Scholar]
  • 8.Efimov VN. Weakly-bound states of three resonantly-interacting particles. Sov. J. Nucl. Phys. 1971;12:589–595. [Google Scholar]
  • 9.Finco D, Teta A. Quadratic forms for the fermionic unitary gas model. Rep. Math. Phys. 2012;69:131–159. doi: 10.1016/S0034-4877(12)60022-6. [DOI] [Google Scholar]
  • 10.Frank RL, Seiringer R. Lieb–Thirring inequality for a model of particles with point interactions. J. Math. Phys. 2012;53:095201. doi: 10.1063/1.3697416. [DOI] [Google Scholar]
  • 11.Goulko O, Wingate M. Thermodynamics of balanced and slightly spin-imbalanced Fermi gases at unitarity. Phys. Rev. A. 2010;82:053621. doi: 10.1103/PhysRevA.82.053621. [DOI] [Google Scholar]
  • 12.Jonsell S, Heiselberg H, Pethick CJ. Universal behavior of the energy of trapped few-boson systems with large scattering length. Phys. Rev. Lett. 2002;89:250401. doi: 10.1103/PhysRevLett.89.250401. [DOI] [PubMed] [Google Scholar]
  • 13.Lieb EH, Seiringer R. The Stability of Matter in Quantum Mechanics. Cambridge: Cambridge Univ. Press; 2010. [Google Scholar]
  • 14.Lieb, E.H., Seiringer, R., Solovej, J.P., Yngvason, J.: The Mathematics of the Bose Gas and its Condensation, Oberwolfach Seminars, vol. 34. Birkhäuser, Basel (2005)
  • 15.Lieb, E.H., Thirring, W.: Bound for the kinetic energy of fermions which proves the stability of matter. Phys. Rev. Lett. 35, 687–689 (1975). [Errata ibid., 1116 (1975)]
  • 16.Lieb EH, Thirring W. Inequalities for the moments of the eigenvalues of the Schrödinger Hamiltonian and their relation to Sobolev inequalities. In: Lieb EH, Simon B, Wightman A, editors. Studies in Mathematical Physics. Princeton: Princeton University Press; 1976. pp. 269–303. [Google Scholar]
  • 17.Lundholm D, Nam PT, Portmann F. Fractional Hardy–Lieb–Thirring and related inequalities for interacting systems. Arch. Ration. Mech. Anal. 2016;219:1343–1382. doi: 10.1007/s00205-015-0923-5. [DOI] [Google Scholar]
  • 18.Lundholm D, Portmann F, Solovej JP. Lieb–Thirring bounds for interacting Bose gases. Commun. Math. Phys. 2015;335:1019–1056. doi: 10.1007/s00220-014-2278-4. [DOI] [Google Scholar]
  • 19.Lundholm D, Solovej JP. Local exclusion and Lieb–Thirring inequalities for intermediate and fractional statistics. Ann. Henri Poincaré. 2014;15:1061–1107. doi: 10.1007/s00023-013-0273-5. [DOI] [Google Scholar]
  • 20.Lundholm D, Solovej JP. Hardy and Lieb–Thirring inequalities for anyons. Commun. Math. Phys. 2013;322:883–908. doi: 10.1007/s00220-013-1748-4. [DOI] [Google Scholar]
  • 21.Moser, T., Seiringer, R.: Stability of a fermionic N+1 particle system with point interactions. Preprint, arXiv:1609.08342
  • 22.Tamura H. The Efimov effect of three-body Schrödinger operators. J. Funct. Anal. 1991;95:433–459. doi: 10.1016/0022-1236(91)90038-7. [DOI] [Google Scholar]
  • 23.Thirring W. Quantum Mathematical Physics. 2. Berlin: Springer; 2002. [Google Scholar]
  • 24.Thomas LH. The interaction between a neutron and a proton and the structure of H3. Phys. Rev. 1935;12:903–909. doi: 10.1103/PhysRev.47.903. [DOI] [Google Scholar]
  • 25.Werner F, Castin Y. Unitary gas in an isotropic harmonic trap: symmetry properties and applications. Phys. Rev. A. 2006;74:053604. doi: 10.1103/PhysRevA.74.053604. [DOI] [Google Scholar]
  • 26.Yafaev, D.R.: On the theory of discrete spectrum of the three-particle Schrödinger operator. Mat. Sb. (N.S.) 94(136), 567–593 (1974)
  • 27.Zwerger, W. (ed.).: The BCS–BEC Crossover and the Unitary Fermi Gas. Springer Lecture Notes in Physics, vol. 836. Springer, New York (2012)

Articles from Letters in Mathematical Physics are provided here courtesy of Springer

RESOURCES