This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Universal Energy Functionals for Trapped Fermi Gases in Low Dimensions

Jiansen Zhang [email protected]    Shina Tan [email protected] International Center for Quantum Materials, Peking University, Beijing 100871, China
Abstract

We study the system of trapped two-component Fermi gases with zero-range interaction in two dimensions (2D) or one dimension (1D). We calculate the one-particle density matrices of these systems at small displacements, from which we show that the NN-body energies are linear functionals of the occupation probabilities of single-particle energy eigenstates. A universal energy functional was first derived in 2011 [1] for trapped zero-range interacting two-component Fermi gases in three dimensions (3D). We also calculate the asymptotic behaviors of the occupation probabilities of single-particle energy eigenstates at high energies.

preprint: APS/123-QED

I Introduction

The zero-range interacting systems are good models for many physical systems, such as the ultracold Bose gases [2, 3, 4, 5, 6], ultracold Fermi gases [7, 8, 9, 10], and few-nucleon systems [11, 12, 13]. If the mean inter-particle distance dd and the thermal de Brogile wavelength λ\lambda are both much larger than the range rer_{e} of the interaction between the particles, the system may be approximated as a zero-range interacting system, and it has universal properties that do not depend on the details of the interaction. These universal properties depend on the interaction potential through the ss-wave scattering length aa, which characterizes the low-energy scattering properties. This universality exists in the Bose systems [14, 15, 16, 17, 18, 19], the Fermi systems [20, 21, 22, 23], and the mixtures [24, 25, 26, 27].

For the 3D two-component Fermi systems with ss-wave contact interaction, it was found that there exists a universal parameter 3D\mathcal{I}_{3\mathrm{D}}, called contact, characterizing the tail of the momentum distribution at large 𝐤\mathbf{k}, where 𝐤\hbar\mathbf{k} is the single-particle momentum, and \hbar is Planck’s constant over 2π2\pi, and that this tail is related to many other physical properties of the system through some exact relations [28, 29, 30, 31]. The name contact comes from the fact that it is a measure of the number of pairs of fermions in two different internal states with small separations. These exact relations [28, 29, 30, 31] have been generalized to the 1D two-component Fermi system [32], the 2D two-component Fermi system [33, 34, 35, 36], the spin-orbit-coupled Fermi system [37, 38], the Bose system [39, 40], and the mixtures [34, 40].

As a zero-range interacting system, the 3D two-component Fermi gas trapped in a smooth potential has an elegant property: its energy can be expressed as a linear functional of the occupation probabilities of single-particle energy eigenstates, i.e. [1]

E=23D4πma+limϵM(ϵν<ϵMϵνnν3Dπ2ϵM2m),E=\frac{\hbar^{2}\mathcal{I}_{3\mathrm{D}}}{4\pi ma}+\lim_{\epsilon_{\text{M}}\rightarrow\infty}\left(\sum_{\epsilon_{\nu}<\epsilon_{\text{M}}}\epsilon_{\nu}n_{\nu}-\frac{\hbar\mathcal{I}_{3\mathrm{D}}}{\pi^{2}}\sqrt{\frac{\epsilon_{\text{M}}}{2m}}\right), (1)

where mm is the mass of each fermion, ϵν\epsilon_{\nu} is the single-particle energy of the ν\nuth single-particle level in the specified smooth external potential, nν=nν+nνn_{\nu}=n_{\nu\uparrow}+n_{\nu\downarrow}, and nνn_{\nu\uparrow} (nνn_{\nu\downarrow}) is the occupation probability of the spin up (down) state in the ν\nuth level. This general functional can be regarded as a generalization of the energy of trapped non-interacting Fermi gases,

E=νσϵνnνσ.E=\sum_{\nu\sigma}\epsilon_{\nu}n_{\nu\sigma}. (2)

Since the zero-range interaction model is valid for lower spatial dimensions, a straightforward idea is to generalize the energy functional Eq. (1) to lower dimensions. The 1D and 2D two-component Fermi gases have been studied for many years. Experimentally, one can realize them by confining the particles in some transverse directions and allowing the particles to move freely in the remaining dimensions [41, 42, 43].

In this paper, we follow the method used in Ref. [1]. We first study the one-particle density matrices of the 2D and 1D trapped two-component Fermi gases with contact interactions. We then generalize the linear energy functional Eq. (1) to 2D and 1D.

This paper is organized as follows. In Sec. II, we introduce the normalized NN-body energy eigenstate and the 2D Bethe-Peierls boundary condition. Using the boundary condition, we expand the one-particle density matrix at small displacements. In Sec. III, we combine the one-particle density matrix with the single-particle imaginary time propagator to find the universal energy functional in 2D:

E=limϵM(ϵν<ϵMϵνnν22D4πmln(e2γma2D2ϵM22)),\displaystyle E=\lim_{\epsilon_{\mathrm{M}}\rightarrow\infty}\left(\sum_{\epsilon_{\nu}<\epsilon_{\mathrm{M}}}\epsilon_{\nu}n_{\nu}-\dfrac{\hbar^{2}\mathcal{I}_{2\mathrm{D}}}{4\pi m}\ln{\dfrac{e^{2\gamma}ma_{2\mathrm{D}}^{2}\epsilon_{\mathrm{M}}}{2\hbar^{2}}}\right), (3)

where γ=0.5772\gamma=0.5772\cdots is Euler’s constant, e=2.718e=2.718\cdots is the base of natural logarithm, 2D\mathcal{I}_{2\mathrm{D}} is the 2D contact, a2Da_{2\mathrm{D}} is the 2D scattering length between two fermions in different spin states, nν=σnνσn_{\nu}=\sum_{\sigma}n_{\nu\sigma}, σ=,\sigma=\uparrow,\downarrow, and nνσn_{\nu\sigma} is the occupation probability of the spin-σ\sigma state of the ν\nuth single-particle level. If the external potential is zero, the single-particle levels reduce to plane-wave states and Eq. (3) reduces to the energy theorem in Refs. [33, 34, 35, 36]. One can extract the contact 2D\mathcal{I}_{2\mathrm{D}} from the asymptotic behavior of ρ(ϵ)\rho(\epsilon) at large ϵ\epsilon, where

ρ(ϵ)νσnνσδ(ϵϵν),\rho(\epsilon)\equiv\sum_{\nu\sigma}n_{\nu\sigma}\delta(\epsilon-\epsilon_{\nu}), (4)

and the coarse-grained version of ρ(ϵ)\rho(\epsilon) has the following asymptotic expansion at large ϵ\epsilon:

ρ(ϵ)|cg=22D4πmϵ2+O(ϵ3).\rho(\epsilon)|_{\text{cg}}=\frac{\hbar^{2}\mathcal{I}_{2\mathrm{D}}}{4\pi m}\epsilon^{-2}+O(\epsilon^{-3}). (5)

We also derived the occupation probabilities of high energy states; see Eq. (50). In Sec. IV and Sec. V, we do analogous calculations for the 1D two-component Fermi system and find that

E\displaystyle E =\displaystyle= 2a1D1D2m+σνϵνnνσ,\displaystyle-\dfrac{\hbar^{2}a_{1\mathrm{D}}\mathcal{I}_{1\mathrm{D}}}{2m}+\sum_{\sigma\nu}\epsilon_{\nu}n_{\nu\sigma}, (6)
ρ(ϵ)|cg\displaystyle\rho(\epsilon)|_{\text{cg}} =\displaystyle= 31D22πm3/2ϵ5/2+O(ϵ7/2),\displaystyle\frac{\hbar^{3}\mathcal{I}_{1\mathrm{D}}}{2\sqrt{2}\pi m^{3/2}}\epsilon^{-5/2}+O(\epsilon^{-7/2}), (7)

where a1Da_{1\mathrm{D}} is the 1D scattering length between two fermions in different spin states, and 1D\mathcal{I}_{1\mathrm{D}} is the 1D contact. If the external potential is zero, the single-particle levels reduce to plane-wave states and Eq. (6) reduces to the energy theorem in Ref. [32]. We also derived the occupation probabilities of high energy states in 1D; see Eq. (72). In Sec. VI, we summarize our results and discuss the utilities and generalizations of our energy functionals.

II One-Particle Density Matrix in 2D

We consider a trapped two-component Fermi system in 2D, with NN_{\uparrow} spin-up fermions and NN_{\downarrow} spin-down fermions. The total number is N=N+NN=N_{\downarrow}+N_{\uparrow}. Here the trapping potential V(𝐫)V(\mathbf{r}) is assumed to be smooth. First we calculate the one-particle density matrix. Consider a normalized NN-body energy eigenstate

|Φ\displaystyle\ket{\Phi} =\displaystyle= (N!N!)1/2D1D1Φ(𝐫1𝐫N𝐬1𝐬N)\displaystyle(N_{\uparrow}!N_{\downarrow}!)^{-1/2}\int D_{1}^{\uparrow}D_{1}^{\downarrow}\Phi(\mathbf{r}_{1}\dots\mathbf{r}_{N_{\uparrow}}\mathbf{s}_{1}\dots\mathbf{s}_{N_{\downarrow}}) (8)
×ψ(𝐫1)ψ(𝐫N)ψ(𝐬1)ψ(𝐬N)|0,\displaystyle\times\psi_{\uparrow}^{\dagger}(\mathbf{r}_{1})\dots\psi_{\uparrow}^{\dagger}(\mathbf{r}_{N_{\uparrow}})\psi_{\downarrow}^{\dagger}(\mathbf{s}_{1})\dots\psi_{\downarrow}^{\dagger}(\mathbf{s}_{N_{\downarrow}})\ket{0},

where 𝐫1,,𝐫N\mathbf{r}_{1},\dots,\mathbf{r}_{N_{\uparrow}} are the position vectors of the spin-up fermions, 𝐬1,,𝐬N\mathbf{s}_{1},\dots,\mathbf{s}_{N_{\downarrow}} are the position vectors of the spin-down fermions, ψ(𝐫)\psi_{\uparrow}^{\dagger}(\mathbf{r}) is the creation operator for a spin-up fermion at position 𝐫\mathbf{r}, ψ(𝐬)\psi_{\downarrow}^{\dagger}(\mathbf{s}) is the creation operator for a spin-down fermion at position 𝐬\mathbf{s}, Diμ=iNd2rμD_{i}^{\uparrow}\equiv\prod_{\mu=i}^{N_{\uparrow}}\mathrm{d}^{2}r_{\mu}, Diμ=iNd2sμD_{i}^{\downarrow}\equiv\prod_{\mu=i}^{N_{\downarrow}}\mathrm{d}^{2}s_{\mu}, and Φ(𝐫1𝐫N𝐬1𝐬N)\Phi(\mathbf{r}_{1}\dots\mathbf{r}_{N_{\uparrow}}\mathbf{s}_{1}\dots\mathbf{s}_{N_{\downarrow}}) is the NN-body wave function which is antisymmetric under the interchange of the positions of any two spin-up (spin-down) fermions. When 𝐫1\mathbf{r}_{1} and 𝐬1\mathbf{s}_{1} are close, the wave function satisfies the 2D Bethe-Peierls boundary condition

Φ\displaystyle\Phi =\displaystyle= A(𝐫1+𝐬12;𝐫2𝐫N𝐬2𝐬N)\displaystyle A\left(\frac{\mathbf{r}_{1}+\mathbf{s}_{1}}{2};\mathbf{r}_{2}\dots\mathbf{r}_{N_{\uparrow}}\mathbf{s}_{2}\dots\mathbf{s}_{N_{\downarrow}}\right) (9)
×12πln(a2D|𝐫1𝐬1|+O(|𝐫1𝐬1|)),\displaystyle\times\frac{1}{2\pi}\ln{\frac{a_{2\mathrm{D}}}{|\mathbf{r}_{1}-\mathbf{s}_{1}|}+O(|\mathbf{r}_{1}-\mathbf{s}_{1}|)},

where a2Da_{2\mathrm{D}} is the two-dimensional ss-wave scattering length, and AA is a function of (N1)(N-1) position vectors. The one-particle density matrix for the spin-σ\sigma fermions is defined as

pσ(𝐫,𝐫+𝐛)=Φ|ψσ(𝐫)ψσ(𝐫+𝐛)|Φ.p_{\sigma}(\mathbf{r},\mathbf{r}+\mathbf{b})=\bra{\Phi}\psi_{\sigma}^{\dagger}(\mathbf{r})\psi_{\sigma}(\mathbf{r}+\mathbf{b})\ket{\Phi}. (10)

In particular, by substituting Eq. (8) into the above definition, we find that

p(𝐫,𝐫+𝐛)\displaystyle p_{\uparrow}(\mathbf{r},\mathbf{r}+\mathbf{b}) =\displaystyle= ND2D1Φ(𝐫,𝐫2𝐫N𝐬1𝐬N)\displaystyle N_{\uparrow}\int D_{2}^{\uparrow}D_{1}^{\downarrow}\Phi^{*}(\mathbf{r},\mathbf{r}_{2}\dots\mathbf{r}_{N_{\uparrow}}\mathbf{s}_{1}\dots\mathbf{s}_{N_{\downarrow}}) (11)
×Φ(𝐫+𝐛,𝐫2𝐫N𝐬1𝐬N).\displaystyle\times\Phi(\mathbf{r+b},\mathbf{r}_{2}\dots\mathbf{r}_{N_{\uparrow}}\mathbf{s}_{1}\dots\mathbf{s}_{N_{\downarrow}}).

We will expand p(𝐫,𝐫+𝐛)p_{\uparrow}(\mathbf{r},\mathbf{r}+\mathbf{b}) through order O(b3){O(b^{3})} at small distance bb. Since Φ\Phi is singular when two fermions in different spin states are close, we divide the 2(N1)2(N-1)-dimensional integration domain into two regions: 𝒞η\mathcal{C}_{\eta} and 𝒟η\mathcal{D}_{\eta}. 𝒟η\mathcal{D}_{\eta} is the region in which every spin-down fermion lies outside of the circle of radius η\eta centered at 𝐫\mathbf{r}, that is, |𝐬μ𝐫|>η|\mathbf{s}_{\mu}-\mathbf{r}|>\eta for μ=1,,N\mu=1,\dots,N_{\downarrow}, which is shown in Fig. 1(a). 𝒞η\mathcal{C}_{\eta} is the complement of 𝒟η\mathcal{D}_{\eta}. We set η\eta small but η>b\eta>b. In 𝒞η\mathcal{C}_{\eta} the cases that two or more spin-down fermions come inside the small circle of radius η\eta centered at 𝐫\mathbf{r} are possible, but the contributions from these cases are suppressed by Fermi statistics and are of higher order than b4b^{4}, see Fig. 1(c). Next, we calculate the integrals in these two regions and add them up, then the dependencies on η\eta will be canceled.

Refer to caption
(a) 𝒟η\mathcal{D}_{\eta}
Refer to caption
(b) The 1st subregion of 𝒞η\mathcal{C}_{\eta}
Refer to caption
(c) 2 or more spin-down fermions inside the circle of radius η\eta centered at 𝐫\mathbf{r}
Figure 1: The configuration of NN fermions. 1(a) shows the region 𝒟η\mathcal{D}_{\eta}, where no spin-down fermion comes inside the circle centered at 𝐫\mathbf{r} with radius η\eta. 1(b) shows that a spin-down fermion at position 𝐬\mathbf{s} is within the circle of radius η\eta centered at 𝐫\mathbf{r}. 1(c) shows that two or more spin-down fermions come inside the small circle, and the probability amplitudes of these situations are suppressed by Fermi statistics.

In 𝒟η\mathcal{D}_{\eta}, we expand Φ\Phi in powers of 𝐛\mathbf{b} as

Φ(𝐫+𝐛,𝐑)\displaystyle\Phi(\mathbf{r+b},\mathbf{R}) =\displaystyle= Φ(𝐫,𝐑)+𝐫Φ(𝐫,𝐑)𝐛\displaystyle\Phi(\mathbf{r},\mathbf{R})+\nabla_{\mathbf{r}}\Phi(\mathbf{r},\mathbf{R})\cdot\mathbf{b} (12)
+12i,j=122rirjΦ(𝐫,𝐑)bibj\displaystyle+\dfrac{1}{2}\sum_{i,j=1}^{2}\dfrac{\partial^{2}}{\partial r_{i}\partial r_{j}}\Phi(\mathbf{r},\mathbf{R})b_{i}b_{j}
+T𝐛+O(b4),\displaystyle+T_{\mathbf{b}}+O(b^{4}),

where 𝐑(𝐫2𝐫N𝐬1𝐬N)\mathbf{R}\equiv(\mathbf{r}_{2}\dots\mathbf{r}_{N_{\uparrow}}\mathbf{s}_{1}\dots\mathbf{s}_{N_{\downarrow}}) and T𝐛13!i,j,k=123rirjrkΦ(𝐫,𝐑)bibjbkT_{\mathbf{b}}\equiv\dfrac{1}{3!}\sum_{i,j,k=1}^{2}\frac{\partial^{3}}{\partial r_{i}\partial r_{j}\partial r_{k}}\Phi(\mathbf{r},\mathbf{R})b_{i}b_{j}b_{k}. Let I𝒟I_{\mathcal{D}} be the integral evaluated in 𝒟η\mathcal{D}_{\eta}, and I𝒞I_{\mathcal{C}} be the integral evaluated in 𝒞η\mathcal{C}_{\eta}. We find

I𝒟\displaystyle I_{\mathcal{D}} =\displaystyle= n𝒟(𝐫)+𝐮(𝐫)𝐛+12i,j=12v,ij(𝐫)bibj\displaystyle n_{\uparrow}^{\mathcal{D}}(\mathbf{r})+\mathbf{u}_{\uparrow}(\mathbf{r})\cdot\mathbf{b}+\dfrac{1}{2}\sum_{i,j=1}^{2}v_{\uparrow,ij}(\mathbf{r})b_{i}b_{j} (13)
+T𝐛+O(b4),\displaystyle+T^{\prime}_{\mathbf{b}}+O(b^{4}),

where

n𝒟(𝐫)\displaystyle n_{\uparrow}^{\mathcal{D}}(\mathbf{r}) =Nlimη0𝒟ηD2D1|Φ(𝐫,𝐑)|2,\displaystyle=N_{\uparrow}\lim_{\eta\to 0}\int_{\mathcal{D}_{\eta}}D_{2}^{\uparrow}D_{1}^{\downarrow}|\Phi(\mathbf{r},\mathbf{R})|^{2}, (14)
𝐮(𝐫)\displaystyle\mathbf{u}_{\uparrow}(\mathbf{r}) =Nlimη0𝒟ηD2D1Φ(𝐫,𝐑)𝐫Φ(𝐫,𝐑),\displaystyle=N_{\uparrow}\lim_{\eta\to 0}\int_{\mathcal{D}_{\eta}}D_{2}^{\uparrow}D_{1}^{\downarrow}\Phi^{*}(\mathbf{r},\mathbf{R})\nabla_{\mathbf{r}}\Phi(\mathbf{r},\mathbf{R}), (15)
v,ij(𝐫)\displaystyle v_{\uparrow,ij}(\mathbf{r}) =Nlimη0𝒟ηD2D1Φ(𝐫,𝐑)2rirjΦ(𝐫,𝐑),\displaystyle=N_{\uparrow}\lim_{\eta\to 0}\int_{\mathcal{D}_{\eta}}D_{2}^{\uparrow}D_{1}^{\downarrow}\Phi^{*}(\mathbf{r},\mathbf{R})\dfrac{\partial^{2}}{\partial r_{i}\partial r_{j}}\Phi(\mathbf{r},\mathbf{R}), (16)
T𝐛\displaystyle T^{\prime}_{\mathbf{b}} =Nlimη0𝒟ηD2D1Φ(𝐫,𝐑)T𝐛.\displaystyle=N_{\uparrow}\lim_{\eta\to 0}\int_{\mathcal{D}_{\eta}}D_{2}^{\uparrow}D_{1}^{\downarrow}\Phi^{*}(\mathbf{r},\mathbf{R})T_{\mathbf{b}}. (17)

To calculate the contributions from the region 𝒞η\mathcal{C}_{\eta}, we use the Bethe-Peierls boundary condition (9). The region 𝒞η\mathcal{C}_{\eta} can be approximately partitioned into NN_{\downarrow} subregions, and in the μ\muth subregion (μ=1,,N\mu=1,\dots,N_{\downarrow}) 𝐬μ\mathbf{s}_{\mu} is within the circle of radius η\eta centered at 𝐫\mathbf{r}. The contributions to I𝒞I_{\mathcal{C}} from these subregions are equal due to Fermi statistics. In the first subregion (shown in Fig. 1(b)) we have

Φ(𝐫,𝐑)\displaystyle\Phi(\mathbf{r},\mathbf{R}^{\prime}) =\displaystyle= A(𝐫+𝐬2;𝐑)×12πln(a2D|𝐫𝐬|)\displaystyle A\left(\frac{\mathbf{r}+\mathbf{s}}{2};\mathbf{R}^{\prime}\right)\times\dfrac{1}{2\pi}\ln{\dfrac{a_{2\mathrm{D}}}{|\mathbf{r}-\mathbf{s}|}} (18)
+O(|𝐫𝐬|),\displaystyle+O(|\mathbf{r}-\mathbf{s}|),
Φ(𝐫+𝐛,𝐑)\displaystyle\Phi(\mathbf{r}+\mathbf{b},\mathbf{R}^{\prime}) =\displaystyle= A(𝐫+𝐬+𝐛2;𝐑)×12πln(a2D|𝐫+𝐛𝐬|)\displaystyle A\left(\frac{\mathbf{r}+\mathbf{s}+\mathbf{b}}{2};\mathbf{R}^{\prime}\right)\times\dfrac{1}{2\pi}\ln{\dfrac{a_{2\mathrm{D}}}{|\mathbf{r}+\mathbf{b}-\mathbf{s}|}} (19)
+O(|𝐫+𝐛𝐬|),\displaystyle+O(|\mathbf{r}+\mathbf{b}-\mathbf{s}|),

where 𝐬𝐬1\mathbf{s}\equiv\mathbf{s}_{1} and 𝐑(𝐫2𝐫N𝐬2𝐬N)\mathbf{R}^{\prime}\equiv(\mathbf{r}_{2}\dots\mathbf{r}_{N_{\uparrow}}\mathbf{s}_{2}\dots\mathbf{s}_{N_{\downarrow}}). We then do the following expansions:

A(𝐫+𝐬2;𝐑)=A(𝐫;𝐑)𝐪2𝐫A+O(q2),A\left(\frac{\mathbf{r}+\mathbf{s}}{2};\mathbf{R}^{\prime}\right)=A(\mathbf{r};\mathbf{R}^{\prime})-\dfrac{\mathbf{q}}{2}\cdot\nabla_{\mathbf{r}}A+O(q^{2}), (20)
A(𝐫+𝐬+𝐛2;𝐑)=\displaystyle A\left(\frac{\mathbf{r}+\mathbf{s}+\mathbf{b}}{2};\mathbf{R}^{\prime}\right)=\, A(𝐫;𝐑)+(𝐛2𝐪2)𝐫A\displaystyle A(\mathbf{r};\mathbf{R}^{\prime})+\left(\dfrac{\mathbf{b}}{2}-\dfrac{\mathbf{q}}{2}\right)\cdot\nabla_{\mathbf{r}}A
+O(|𝐛𝐪|2),\displaystyle+O(|\mathbf{b}-\mathbf{q}|^{2}), (21)

where 𝐪=𝐫𝐬\mathbf{q}=\mathbf{r}-\mathbf{s}. So we have

I𝒞=NND2D2q<ηd2qF𝐛+O(b4),I_{\mathcal{C}}=N_{\uparrow}N_{\downarrow}\int D_{2}^{\uparrow}D_{2}^{\downarrow}\int_{q<\eta}\mathrm{d}^{2}qF_{\mathbf{b}}+O(b^{4}), (22)

where

F𝐛\displaystyle F_{\mathbf{b}} =\displaystyle= 14π2ln(a2Dq)ln(a2D|𝐪+𝐛|)(A𝐫A𝐪2)\displaystyle\dfrac{1}{4\pi^{2}}\ln{\dfrac{a_{2\mathrm{D}}}{q}}\ln{\dfrac{a_{2\mathrm{D}}}{|\mathbf{q}+\mathbf{b}|}}\left(A^{*}-\nabla_{\mathbf{r}}A^{*}\cdot\dfrac{\mathbf{q}}{2}\right) (23)
×[A+𝐫A(𝐛2𝐪2)].\displaystyle\times\left[A+\nabla_{\mathbf{r}}A\cdot\left(\dfrac{\mathbf{b}}{2}-\dfrac{\mathbf{q}}{2}\right)\right].

Carrying out the integral I𝒞I_{\mathcal{C}} and adding it to I𝒟I_{\mathcal{D}}, we get

p(𝐫,𝐫+𝐛)=\displaystyle p_{\uparrow}(\mathbf{r},\mathbf{r}+\mathbf{b})= I𝒞+I𝒟\displaystyle\,I_{\mathcal{C}}+I_{\mathcal{D}}
=\displaystyle= n(𝐫)+𝐮(𝐫)𝐛+18πC2D(𝐫)b2ln(ba2De)\displaystyle\,n_{\uparrow}(\mathbf{r})+\mathbf{u}_{\uparrow}(\mathbf{r})\cdot\mathbf{b}+\frac{1}{8\pi}C_{2\mathrm{D}}(\mathbf{r})b^{2}\ln{\frac{b}{a_{2\mathrm{D}}e}}
+12i,j=12v,ij(𝐫)bibj\displaystyle+\dfrac{1}{2}\sum_{i,j=1}^{2}v_{\uparrow,ij}(\mathbf{r})b_{i}b_{j}
+116πb2(12ln(ba2D)38)𝐰𝐛\displaystyle+\dfrac{1}{16\pi}b^{2}\left(\dfrac{1}{2}\ln{\dfrac{b}{a_{2\mathrm{D}}}}-\dfrac{3}{8}\right)\mathbf{w}^{*}\cdot\mathbf{b}
+116πb2(32ln(ba2D)118)𝐰𝐛\displaystyle+\dfrac{1}{16\pi}b^{2}\left(\dfrac{3}{2}\ln{\dfrac{b}{a_{2\mathrm{D}}}}-\dfrac{11}{8}\right)\mathbf{w}\cdot\mathbf{b}
+T𝐛+O(b4),\displaystyle+T^{\prime}_{\mathbf{b}}+O(b^{4}), (24)

where

n(𝐫)\displaystyle n_{\uparrow}(\mathbf{r}) =\displaystyle= ND2D1|Φ(𝐫,𝐑)|2,\displaystyle N_{\uparrow}\int D_{2}^{\uparrow}D_{1}^{\downarrow}|\Phi(\mathbf{r},\mathbf{R})|^{2}, (25)
C2D(𝐫)\displaystyle C_{2\mathrm{D}}(\mathbf{r}) \displaystyle\equiv NND2D2|A(𝐫;𝐑)|2,\displaystyle N_{\uparrow}N_{\downarrow}\int D_{2}^{\uparrow}D_{2}^{\downarrow}\ |A(\mathbf{r};\mathbf{R}^{\prime})|^{2}, (26)
𝐰(𝐫)\displaystyle\mathbf{w}(\mathbf{r}) \displaystyle\equiv NND2D2A(𝐫;𝐑)𝐫A(𝐫;𝐑).\displaystyle N_{\uparrow}N_{\downarrow}\int D_{2}^{\uparrow}D_{2}^{\downarrow}\ A^{*}(\mathbf{r};\mathbf{R}^{\prime})\nabla_{\mathbf{r}}A(\mathbf{r};\mathbf{R}^{\prime}). (27)

n(𝐫)n_{\uparrow}(\mathbf{r}) is the spatial density of spin-up fermions at 𝐫\mathbf{r}, C2D(𝐫)C_{2\mathrm{D}}(\mathbf{r}) is the 2D contact density, and 𝐰(𝐫)\mathbf{w}(\mathbf{r}) is related to the center-of-mass motion of small-distance pairs of fermions in different spin states. We can also find a similar expansion for p(𝐫,𝐫+𝐛)p_{\downarrow}(\mathbf{r},\mathbf{r}+\mathbf{b}).

III Universal Energy Functional in 2D

We define an absolutely convergent series

Jσ(β)νnνσeβϵν=νΦ|cνσcνσ|Φeβϵν,J_{\sigma}(\beta)\equiv\sum_{\nu}n_{\nu\sigma}e^{-\beta\epsilon_{\nu}}=\sum_{\nu}\expectationvalue{c^{\dagger}_{\nu\sigma}c_{\nu\sigma}}{\Phi}e^{-\beta\epsilon_{\nu}}, (28)

where β\beta satisfies Reβ0\beta\geq 0, |Φ\ket{\Phi} is an NN-body energy eigenstate,

nνσ=Φ|cνσcνσ|Φn_{\nu\sigma}=\bra{\Phi}c_{\nu\sigma}^{\dagger}c_{\nu\sigma}\ket{\Phi} (29)

is the occupation probability of the spin-σ\sigma state of the ν\nuth single-particle level,

cνσ=d2rϕν(𝐫)ψσ(𝐫)c_{\nu\sigma}=\int\mathrm{d}^{2}r\phi^{*}_{\nu}(\mathbf{r})\psi_{\sigma}(\mathbf{r}) (30)

is the fermion annihilation operator of such a single-particle state, and ϕν(𝐫)\phi_{\nu}(\mathbf{r}) is the wave function of the ν\nuth single-particle level in the trapping potential V(𝐫)V(\mathbf{r}) and satisfies the single-particle Schrödinger equation

[22m2+V(𝐫)]ϕν(𝐫)=ϵνϕν(𝐫)\left[-\frac{\hbar^{2}}{2m}\nabla^{2}+V(\mathbf{r})\right]\phi_{\nu}(\mathbf{r})=\epsilon_{\nu}\phi_{\nu}(\mathbf{r}) (31)

and the normalization condition

|ϕν(𝐫)|2d2r=1.\int|\phi_{\nu}(\mathbf{r})|^{2}\mathrm{d}^{2}r=1. (32)

We rewrite Jσ(β)J_{\sigma}(\beta) as

Jσ(β)=d2rd2rUβ(𝐫,𝐫)pσ(𝐫,𝐫),J_{\sigma}(\beta)=\int\mathrm{d}^{2}r\mathrm{d}^{2}r^{\prime}U_{\beta}(\mathbf{r},\mathbf{r}^{\prime})p_{\sigma}(\mathbf{r},\mathbf{r}^{\prime}), (33)

where Uβ(𝐫,𝐫)=νeβϵνϕν(𝐫)ϕν(𝐫)U_{\beta}(\mathbf{r},\mathbf{r}^{\prime})=\sum_{\nu}e^{-\beta\epsilon_{\nu}}\phi_{\nu}(\mathbf{r})\phi^{*}_{\nu}(\mathbf{r}^{\prime}) is the propagator of a single particle moving in the potential V(𝐫)V(\mathbf{r}) whthin a time iβ-i\hbar\beta. For a small positive β\beta, at |𝐫𝐫|β/m|\mathbf{r}-\mathbf{r}^{\prime}|\gg\hbar\sqrt{\beta/m} the propagator is exponentially suppressed, while at |𝐫𝐫|β/m|\mathbf{r}-\mathbf{r}^{\prime}|\sim\hbar\sqrt{\beta/m} we have a short imaginary-time expansion

Uβ(𝐫,𝐫)\displaystyle U_{\beta}(\mathbf{r},\mathbf{r}^{\prime}) =\displaystyle= m2π2β[1V(𝐫)+V(𝐫)2β]\displaystyle\dfrac{m}{2\pi\hbar^{2}\beta}\left[1-\frac{V(\mathbf{r})+V(\mathbf{r}^{\prime})}{2}\beta\right] (34)
×exp[m(𝐫𝐫)222β]+O(β).\displaystyle\times\exp\left[-\dfrac{m(\mathbf{r}-\mathbf{r}^{\prime})^{2}}{2\hbar^{2}\beta}\right]+O(\beta).

Recall that when |𝐫𝐫||\mathbf{r}-\mathbf{r}^{\prime}| is small, we also have an expansion of pσ(𝐫,𝐫)p_{\sigma}(\mathbf{r},\mathbf{r}^{\prime}). Defining Gσ=Uβ(𝐫,𝐫)pσ(𝐫,𝐫)G_{\sigma}=U_{\beta}(\mathbf{r},\mathbf{r}^{\prime})p_{\sigma}(\mathbf{r},\mathbf{r}^{\prime}), we have

Gσ\displaystyle G_{\sigma} =\displaystyle= m2π2β[1V(𝐫)+V(𝐫)2β]exp[mb222β]\displaystyle\dfrac{m}{2\pi\hbar^{2}\beta}\left[1-\frac{V(\mathbf{r})+V(\mathbf{r}^{\prime})}{2}\beta\right]\exp\left[-\dfrac{mb^{2}}{2\hbar^{2}\beta}\right] (35)
×[n(𝐫)+𝐮(𝐫)𝐛+i,j=12v,ij(𝐫)bibj2\displaystyle\times\Bigg{[}n_{\uparrow}(\mathbf{r})+\mathbf{u}_{\uparrow}(\mathbf{r})\cdot\mathbf{b}+\sum_{i,j=1}^{2}v_{\uparrow,ij}(\mathbf{r})\dfrac{b_{i}b_{j}}{2}
+b2C2D(𝐫)8πln(ba2De)+b2𝐰𝐛16π(32ln(ba2D)118)\displaystyle+\dfrac{b^{2}C_{2\mathrm{D}}(\mathbf{r})}{8\pi}\ln{\dfrac{b}{a_{2\mathrm{D}}e}}+\dfrac{b^{2}\mathbf{w}\cdot\mathbf{b}}{16\pi}\left(\dfrac{3}{2}\ln{\dfrac{b}{a_{2\mathrm{D}}}}-\dfrac{11}{8}\right)
+b2𝐰𝐛16π(12ln(ba2D)38)]+O(β),\displaystyle+\dfrac{b^{2}\mathbf{w}^{*}\cdot\mathbf{b}}{16\pi}\left(\dfrac{1}{2}\ln{\dfrac{b}{a_{2\mathrm{D}}}}-\dfrac{3}{8}\right)\Bigg{]}+O(\beta),

where 𝐛=𝐫𝐫\mathbf{b}=\mathbf{r}^{\prime}-\mathbf{r}. Substituting the above result into Eq. (33), we find

Jσ(β)\displaystyle J_{\sigma}(\beta) =\displaystyle= Nσβd2rV(𝐫)nσ(r)\displaystyle N_{\sigma}-\beta\int\mathrm{d}^{2}rV(\mathbf{r})n_{\sigma}(\mathrm{r}) (36)
22Dβ8πm(1+γ+ln(ma2D222β))\displaystyle-\dfrac{\hbar^{2}\mathcal{I}_{2\mathrm{D}}\beta}{8\pi m}\left(1+\gamma+\ln{\dfrac{ma_{2\mathrm{D}}^{2}}{2\hbar^{2}\beta}}\right)
+2β2md2ri=12vσ,ii(𝐫)+O(β2),\displaystyle+\dfrac{\hbar^{2}\beta}{2m}\int\mathrm{d}^{2}r\sum_{i=1}^{2}v_{\sigma,ii}(\mathbf{r})+O(\beta^{2}),

where

Nσ\displaystyle N_{\sigma} =\displaystyle= d2rnσ(𝐫),\displaystyle\int\mathrm{d}^{2}rn_{\sigma}(\mathbf{r}), (37)
2D\displaystyle\mathcal{I}_{2\mathrm{D}} =\displaystyle= d2rC2D(𝐫).\displaystyle\int\mathrm{d}^{2}rC_{2\mathrm{D}}(\mathbf{r}). (38)

Outside of the tiny range of two-body interactions, the NN-body Schrödinger equation is simplified as

EΦ\displaystyle E\Phi =\displaystyle= μ=1N[22m𝐫μ2+V(𝐫μ)]Φ\displaystyle\sum_{\mu=1}^{N_{\uparrow}}\left[-\dfrac{\hbar^{2}}{2m}\nabla^{2}_{\mathbf{r}_{\mu}}+V(\mathbf{r}_{\mu})\right]\Phi (39)
+μ=1N[22m𝐬μ2+V(𝐬μ)]Φ,\displaystyle+\sum_{\mu^{\prime}=1}^{N_{\downarrow}}\left[-\dfrac{\hbar^{2}}{2m}\nabla^{2}_{\mathbf{s}_{\mu^{\prime}}}+V(\mathbf{s}_{\mu^{\prime}})\right]\Phi,

where 𝐫μ𝐬μ\mathbf{r}_{\mu}\neq\mathbf{s}_{\mu^{\prime}} for all μ,μ\mu,\mu^{\prime}. Multiplying both sides of Eq. (39) by Φ\Phi^{*}, integrating them over 𝐫1,,𝐫N,𝐬1,,𝐬N\mathbf{r}_{1},\dots,\mathbf{r}_{N_{\uparrow}},\mathbf{s}_{1},\dots,\mathbf{s}_{N_{\downarrow}} for 𝐫μ𝐬μ\mathbf{r}_{\mu}\neq\mathbf{s}_{\mu^{\prime}} for all μ,μ\mu,\mu^{\prime}, we get

σd2r[V(𝐫)nσ(𝐫)22mi=12vσ,ii(𝐫)]=E.\sum_{\sigma}\int\mathrm{d}^{2}r\left[V(\mathbf{r})n_{\sigma}(\mathbf{r})-\dfrac{\hbar^{2}}{2m}\sum_{i=1}^{2}v_{\sigma,ii}(\mathbf{r})\right]=E. (40)

Summing Eq. (36) over σ\sigma, we find

σJσ(β)=\displaystyle\sum_{\sigma}J_{\sigma}(\beta)= νσnνσeβϵν\displaystyle\,\sum_{\nu\sigma}n_{\nu\sigma}e^{-\beta\epsilon_{\nu}}
=\displaystyle= N22Dβ8πm(1+γ+ln(ma2D222β))\displaystyle\,N-\dfrac{\hbar^{2}\mathcal{I}_{2\mathrm{D}}\beta}{8\pi m}\left(1+\gamma+\ln{\dfrac{ma_{2\mathrm{D}}^{2}}{2\hbar^{2}\beta}}\right)
βσd2r[V(𝐫)nσ22mi=12vσ,ii]\displaystyle-\beta\sum_{\sigma}\int\mathrm{d}^{2}r\left[V(\mathbf{r})n_{\sigma}-\dfrac{\hbar^{2}}{2m}\sum_{i=1}^{2}v_{\sigma,ii}\right]
+O(β2)\displaystyle+O(\beta^{2})
=\displaystyle= NEβ22Dβ4πm(1+γ+ln(ma2D222β))\displaystyle\,N-E\beta-\dfrac{\hbar^{2}\mathcal{I}_{2\mathrm{D}}\beta}{4\pi m}\left(1+\gamma+\ln{\dfrac{ma_{2\mathrm{D}}^{2}}{2\hbar^{2}\beta}}\right)
+O(β2).\displaystyle+O(\beta^{2}). (41)

Let

ρ(ϵ)σρσ(ϵ)=νσnνσδ(ϵϵν).\rho(\epsilon)\equiv\sum_{\sigma}\rho_{\sigma}(\epsilon)=\sum_{\nu\sigma}n_{\nu\sigma}\delta(\epsilon-\epsilon_{\nu}). (42)

Equation (41) can be rewritten as

+ρ(ϵ)eβϵdϵ=\displaystyle\int_{-\infty}^{+\infty}\rho(\epsilon)e^{-\beta\epsilon}\mathrm{d}\epsilon= N22Dβ4πm(1+γ+ln(ma2D222β))\displaystyle\,N-\dfrac{\hbar^{2}\mathcal{I}_{2\mathrm{D}}\beta}{4\pi m}\left(1+\gamma+\ln{\dfrac{ma_{2\mathrm{D}}^{2}}{2\hbar^{2}\beta}}\right)
Eβ+O(β2).\displaystyle-E\beta+O(\beta^{2}). (43)

Setting β=η+is\beta=\eta+is where η\eta is a positive infinitesimal and ss is real, we see that the above equation shows the Fourier transform of ρ(ϵ)\rho(\epsilon) at small ss, and this Fourier transform has a singular term proportional to slnss\ln s. This singular term is caused by a power law tail of the coarse-grained version of ρ(ϵ)\rho(\epsilon) at ϵ\epsilon\to\infty. Taking the inverse Fourier transform of this singular term, we find the power law tail shown in Eq. (5).

Applying ddβ\frac{\mathrm{d}}{\mathrm{d}\beta} to both sides of Eq. (III), we find

E=ρ(ϵ)ϵeβϵdϵ22D4πm(γ+ln(a2D2m22β))+O(β).E=\int_{-\infty}^{\infty}\rho(\epsilon)\epsilon e^{-\beta\epsilon}\mathrm{d}\epsilon-\dfrac{\hbar^{2}\mathcal{I}_{2\mathrm{D}}}{4\pi m}\left(\gamma+\ln{\dfrac{a_{2\mathrm{D}}^{2}m}{2\hbar^{2}\beta}}\right)+O(\beta). (44)

We divide the domain of integration over ϵ\epsilon into two regions: one is (,ϵM)(-\infty,\epsilon_{\mathrm{M}}) and the other is (ϵM,)(\epsilon_{\mathrm{M}},\infty), where ϵM\epsilon_{\mathrm{M}} is an energy scale such that ϵM\epsilon_{\mathrm{M}} is very large but ϵMβ1\epsilon_{\mathrm{M}}\beta\ll 1. In (,ϵM)(-\infty,\epsilon_{\mathrm{M}}) we have

ϵMρ(ϵ)ϵeβϵdϵϵMρ(ϵ)ϵdϵ,\int_{-\infty}^{\epsilon_{\mathrm{M}}}\rho(\epsilon)\epsilon e^{-\beta\epsilon}\mathrm{d}\epsilon\approx\int_{-\infty}^{\epsilon_{\mathrm{M}}}\rho(\epsilon)\epsilon\mathrm{d}\epsilon, (45)

while in (ϵM,)(\epsilon_{\mathrm{M}},\infty) we use Eq. (5) to do the integral:

ϵMρ(ϵ)ϵeβϵdϵ\displaystyle\int_{\epsilon_{\mathrm{M}}}^{\infty}\rho(\epsilon)\epsilon e^{-\beta\epsilon}\mathrm{d}\epsilon =22D4πmϵMϵ1eβϵdϵ\displaystyle=\dfrac{\hbar^{2}\mathcal{I}_{2\mathrm{D}}}{4\pi m}\int_{\epsilon_{\mathrm{M}}}^{\infty}\epsilon^{-1}e^{-\beta\epsilon}\mathrm{d}\epsilon
=22D4πmΓ(0,ϵMβ)\displaystyle=\dfrac{\hbar^{2}\mathcal{I}_{2\mathrm{D}}}{4\pi m}\mathrm{\Gamma}(0,\epsilon_{\mathrm{M}}\beta)
=22D4πm[γln((ϵMβ))]+O(ϵMβ).\displaystyle=\dfrac{\hbar^{2}\mathcal{I}_{2\mathrm{D}}}{4\pi m}[-\gamma-\ln{(\epsilon_{\mathrm{M}}\beta)}]+O(\epsilon_{\mathrm{M}}\beta). (46)

Thus, taking β0\beta\rightarrow 0, we get

E\displaystyle E =limϵM[ϵMρ(ϵ)ϵdϵ22D4πmln(e2γma2D2ϵM22)]\displaystyle=\lim_{\epsilon_{\mathrm{M}}\rightarrow\infty}\left[\int_{-\infty}^{\epsilon_{\mathrm{M}}}\rho(\epsilon)\epsilon\mathrm{d}\epsilon-\dfrac{\hbar^{2}\mathcal{I}_{2\mathrm{D}}}{4\pi m}\ln{\dfrac{e^{2\gamma}ma_{2\mathrm{D}}^{2}\epsilon_{\mathrm{M}}}{2\hbar^{2}}}\right]
=limϵM(ϵν<ϵMϵνnν22D4πmln(e2γma2D2ϵM22)),\displaystyle=\lim_{\epsilon_{\mathrm{M}}\rightarrow\infty}\left(\sum_{\epsilon_{\nu}<\epsilon_{\mathrm{M}}}\epsilon_{\nu}n_{\nu}-\dfrac{\hbar^{2}\mathcal{I}_{2\mathrm{D}}}{4\pi m}\ln{\dfrac{e^{2\gamma}ma_{2\mathrm{D}}^{2}\epsilon_{\mathrm{M}}}{2\hbar^{2}}}\right), (47)

which is Eq. (3).

According to Eqs. (10), (29), and (30), we have

nνσ=d2rd2bϕν(𝐫)ϕν(𝐫+𝐛)pσ(𝐫,𝐫+𝐛).n_{\nu\sigma}=\int\text{d}^{2}r\int\text{d}^{2}b\,\phi_{\nu}(\mathbf{r})\phi_{\nu}^{*}(\mathbf{r}+\mathbf{b})p_{\sigma}(\mathbf{r},\mathbf{r}+\mathbf{b}). (48)

When ϵν\epsilon_{\nu} is large, the integrand as a function of 𝐛\mathbf{b} oscillates rapidly, which implies that the only important contribution is from the singular term in the expansion of pσ(𝐫,𝐫+𝐛)p_{\sigma}(\mathbf{r},\mathbf{r}+\mathbf{b}) at 𝐛0\mathbf{b}\rightarrow 0 [1], and this singular term is 18πC2D(𝐫)b2ln(ba2De)\frac{1}{8\pi}C_{2\mathrm{D}}(\mathbf{r})b^{2}\ln{\frac{b}{a_{2\mathrm{D}}e}}. Since ϕν(𝐫)\phi_{\nu}(\mathbf{r}) satisfies the single-particle Schrödinger equation, Eq. (31), we have

ϕν(𝐫+𝐛)(2/2mϵν)2𝐛4ϕν(𝐫+𝐛)\phi^{*}_{\nu}(\mathbf{r}+\mathbf{b})\approx(\hbar^{2}/2m\epsilon_{\nu})^{2}\nabla^{4}_{\mathbf{b}}\phi^{*}_{\nu}(\mathbf{r}+\mathbf{b}) (49)

with relative error O(ϵν1)\sim O(\epsilon_{\nu}^{-1}) at b2/2mϵνb\sim\sqrt{\hbar^{2}/2m\epsilon_{\nu}}. Substituting Eq. (49) into Eq. (48) and carrying out the integral over 𝐛\mathbf{b}, we find

nνσ=1kν4d2rC2D(𝐫)|ϕν(𝐫)|2+O(ϵν5/2),n_{\nu\sigma}=\frac{1}{k_{\nu}^{4}}\int\text{d}^{2}rC_{2\mathrm{D}}(\mathbf{r})|\phi_{\nu}(\mathbf{r})|^{2}+O(\epsilon_{\nu}^{-5/2}), (50)

where kν=2mϵν/k_{\nu}=\sqrt{2m\epsilon_{\nu}}/\hbar.

IV One-Particle Density Matrix in 1D

The calculation procedure in 1D is similar to the one in 2D. We define a normalized NN-body energy eigenstate |Φ\ket{\Phi} in 1D,

|Φ=\displaystyle\ket{\Phi}= (N!N!)1/2D~1D~1Φ(x1xNy1yN)\displaystyle(N_{\uparrow}!N_{\downarrow}!)^{-1/2}\int\tilde{D}_{1}^{\uparrow}\tilde{D}_{1}^{\downarrow}\Phi(x_{1}\dots x_{N_{\uparrow}}y_{1}\dots y_{N_{\downarrow}})
×ψ(x1)ψ(xN)ψ(y1)ψ(yN)|0.\displaystyle\times\psi_{\uparrow}^{\dagger}(x_{1})\dots\psi_{\uparrow}^{\dagger}(x_{N_{\uparrow}})\psi_{\downarrow}^{\dagger}(y_{1})\dots\psi_{\downarrow}^{\dagger}(y_{N_{\downarrow}})\ket{0}. (51)

where N=N+NN=N_{\uparrow}+N_{\downarrow}, NN_{\uparrow} is the number of spin-up fermions, NN_{\downarrow} is the number of spin-down fermions, x1,,xNx_{1},\dots,x_{N_{\uparrow}} are the coordinates of the spin-up fermions, y1,,yNy_{1},\dots,y_{N_{\downarrow}} are the coordinates of the spin-down fermions, ψ(x)\psi_{\uparrow}^{\dagger}(x) is the creation operator for a spin-up fermion at position xx, ψ(y)\psi_{\downarrow}^{\dagger}(y) is the creation operator for a spin-down fermion at position yy, D~iμ=iNdxμ\tilde{D}_{i}^{\uparrow}\equiv\prod_{\mu=i}^{N_{\uparrow}}\mathrm{d}x_{\mu}, D~iμ=iNdyμ\tilde{D}_{i}^{\downarrow}\equiv\prod_{\mu=i}^{N_{\downarrow}}\mathrm{d}y_{\mu}, and Φ(x1xNy1yN)\Phi(x_{1}\dots x_{N_{\uparrow}}y_{1}\dots y_{N_{\downarrow}}) is the NN-body wave function which is antisymmetric under the interchange of any two spin-up (spin-down) fermions. The 1D Bethe-Peierls boundary condition is

|Φ\displaystyle\ket{\Phi} =\displaystyle= A(x1+y12;x2xNy2yN)\displaystyle A\left(\frac{x_{1}+y_{1}}{2};x_{2}\dots x_{N_{\uparrow}}y_{2}\dots y_{N_{\downarrow}}\right) (52)
×(1|x1y1|a1D)+O(|x1y1|2),\displaystyle\times\left(1-\frac{|x_{1}-y_{1}|}{a_{1\mathrm{D}}}\right)+O(|x_{1}-y_{1}|^{2}),

which is satisfied by the wave function when x1x_{1} and y1y_{1} are close. The one-particle density matrix for spin-σ\sigma fermions in 1D is defined as

pσ(x,x+b)=Φ|ψσ(x)ψσ(x+b)|Φ.p_{\sigma}(x,x+b)=\bra{\Phi}\psi_{\sigma}^{\dagger}(x)\psi_{\sigma}(x+b)\ket{\Phi}. (53)

For spin-up fermions, we substitute Eq. (IV) into the above definition and find

p(x,x+b)=\displaystyle p_{\uparrow}(x,x+b)= ND~2D~1Φ(x,x2xNy1yN)\displaystyle\,N_{\uparrow}\int\tilde{D}_{2}^{\uparrow}\tilde{D}_{1}^{\downarrow}\Phi^{*}(x,x_{2}\dots x_{N_{\uparrow}}y_{1}\dots y_{N_{\downarrow}})
×Φ(x+b,x2xNy1yN).\displaystyle\times\Phi(x+b,x_{2}\dots x_{N_{\uparrow}}y_{1}\dots y_{N_{\downarrow}}). (54)

After finishing calculations analogous to those for the 2D one-particle density matrix, we find

p(x,x+b)=\displaystyle p_{\uparrow}(x,x+b)= n(x)+u(x)b+12v(x)b2\displaystyle n_{\uparrow}(x)+u_{\uparrow}(x)b+\frac{1}{2}v_{\uparrow}(x)b^{2}
+C1D(x)(b2a1D4+|b|312)\displaystyle+C_{1\mathrm{D}}(x)\left(-\frac{b^{2}a_{1\mathrm{D}}}{4}+\frac{|b|^{3}}{12}\right)
+w(x)2b33a1D+w(x)b36a1D+Tb+O(b4),\displaystyle+w(x)\frac{2b^{3}}{3a_{1\mathrm{D}}}+w^{*}(x)\frac{b^{3}}{6a_{1\mathrm{D}}}+T^{\prime}_{b}+O(b^{4}), (55)

where

n(x)\displaystyle n_{\uparrow}(x) =ND~2D~1|Φ(x,𝐗)|2,\displaystyle=N_{\uparrow}\int\tilde{D}_{2}^{\uparrow}\tilde{D}_{1}^{\downarrow}|\Phi(x,\mathbf{X})|^{2}, (56)
u(x)\displaystyle u_{\uparrow}(x) =Nlimη0𝒟ηD~2D~1Φ(x,𝐗)xΦ(x,𝐗),\displaystyle=N_{\uparrow}\lim_{\eta\to 0}\int_{\mathcal{D}_{\eta}}\tilde{D}_{2}^{\uparrow}\tilde{D}_{1}^{\downarrow}\Phi^{*}(x,\mathbf{X})\frac{\partial}{\partial x}\Phi(x,\mathbf{X}), (57)
v(x)\displaystyle v_{\uparrow}(x) =Nlimη0𝒟ηD~2D~1Φ(x,𝐗)2x2Φ(x,𝐗),\displaystyle=N_{\uparrow}\lim_{\eta\to 0}\int_{\mathcal{D}_{\eta}}\tilde{D}_{2}^{\uparrow}\tilde{D}_{1}^{\downarrow}\Phi^{*}(x,\mathbf{X})\dfrac{\partial^{2}}{\partial x^{2}}\Phi(x,\mathbf{X}), (58)
Tb\displaystyle T^{\prime}_{b} =Nb33!limη0𝒟ηD~2D~1Φ(x,𝐗)3x3Φ(x,𝐗),\displaystyle=\frac{N_{\uparrow}b^{3}}{3!}\lim_{\eta\to 0}\int_{\mathcal{D}_{\eta}}\tilde{D}_{2}^{\uparrow}\tilde{D}_{1}^{\downarrow}\Phi^{*}(x,\mathbf{X})\dfrac{\partial^{3}}{\partial x^{3}}\Phi(x,\mathbf{X}), (59)
C1D(x)\displaystyle C_{1\mathrm{D}}(x) 4NNa1D2D~2D~2|A(x;𝐗)|2,\displaystyle\equiv\frac{4N_{\uparrow}N_{\downarrow}}{a_{1\mathrm{D}}^{2}}\int\tilde{D}_{2}^{\uparrow}\tilde{D}_{2}^{\downarrow}\ |A(x;\mathbf{X}^{\prime})|^{2}, (60)
w(x)\displaystyle w(x) NND~2D~2A(x;𝐗)A(x;𝐗)x,\displaystyle\equiv N_{\uparrow}N_{\downarrow}\int\tilde{D}_{2}^{\uparrow}\tilde{D}_{2}^{\downarrow}\ A^{*}(x;\mathbf{X}^{\prime})\frac{\partial A(x;\mathbf{X}^{\prime})}{\partial x}, (61)

and 𝐗(x2xNy1yN)\mathbf{X}\equiv(x_{2}\dots x_{N_{\uparrow}}y_{1}\dots y_{N_{\downarrow}}), 𝐗(x2xNy2yN)\mathbf{X}^{\prime}\equiv(x_{2}\dots x_{N_{\uparrow}}y_{2}\dots y_{N_{\downarrow}}). n(x)n_{\uparrow}(x) is the spatial density of spin-up fermions at position xx, C1D(x)C_{1\mathrm{D}}(x) is the 1D contact density, and w(x)w(x) is related to the center-of-mass motion of small-distance pairs of fermions in different spin states. We can also find a similar expansion for p(x,x+b)p_{\downarrow}(x,x+b).

V Universal Energy Functional in 1D

We define Jσ(β)J_{\sigma}(\beta) in 1D:

Jσ(β)νnνσeβϵν=νΦ|cνσcνσ|Φeβϵν,J_{\sigma}(\beta)\equiv\sum_{\nu}n_{\nu\sigma}e^{-\beta\epsilon_{\nu}}=\sum_{\nu}\expectationvalue{c^{\dagger}_{\nu\sigma}c_{\nu\sigma}}{\Phi}e^{-\beta\epsilon_{\nu}}, (62)

where

cνσ=ϕν(x)ψσ(x)dx,c_{\nu\sigma}=\int_{-\infty}^{\infty}\phi^{*}_{\nu}(x)\psi_{\sigma}(x)\mathrm{d}x, (63)

and ϕν(x)\phi_{\nu}(x) is the wave function of the ν\nuth single-particle level in the trapping potential V(x)V(x) and satisfies the single-particle Schrödinger equation

[22md2dx2+V(x)]ϕν(x)=ϵνϕν(x)\left[-\frac{\hbar^{2}}{2m}\frac{\mathrm{d}^{2}}{\mathrm{d}x^{2}}+V(x)\right]\phi_{\nu}(x)=\epsilon_{\nu}\phi_{\nu}(x) (64)

and the normalization condition

|ϕν(x)|2dx=1.\int_{-\infty}^{\infty}|\phi_{\nu}(x)|^{2}\mathrm{d}x=1. (65)

We can rewrite Jσ(β)J_{\sigma}(\beta) as

Jσ(β)=dxdxUβ(x,x)pσ(x,x),J_{\sigma}(\beta)=\int_{-\infty}^{\infty}\mathrm{d}x\int_{-\infty}^{\infty}\mathrm{d}x^{\prime}U_{\beta}(x,x^{\prime})p_{\sigma}(x,x^{\prime}), (66)

where Uβ(x,x)=νeβϵνϕν(x)ϕν(x)U_{\beta}(x,x^{\prime})=\sum_{\nu}e^{-\beta\epsilon_{\nu}}\phi_{\nu}(x)\phi^{*}_{\nu}(x^{\prime}) is the propagator of a single particle moving in the potential V(x)V(x) within a time iβ-i\hbar\beta. We find

Jσ(β)=\displaystyle J_{\sigma}(\beta)= NσβdxV(x)nσ(x)2a1D1Dβ4m\displaystyle N_{\sigma}-\beta\int\mathrm{d}xV(x)n_{\sigma}(x)-\dfrac{\hbar^{2}a_{1\mathrm{D}}\mathcal{I}_{1\mathrm{D}}\beta}{4m}
+2β2mdxvσ(x)+31Dβ3/232πm3+O(β2),\displaystyle+\dfrac{\hbar^{2}\beta}{2m}\int\mathrm{d}xv_{\sigma}(x)+\frac{\hbar^{3}\mathcal{I}_{1\mathrm{D}}\beta^{3/2}}{3\sqrt{2\pi m^{3}}}+O(\beta^{2}), (67)

where

Nσ\displaystyle N_{\sigma} =\displaystyle= dxnσ(x),\displaystyle\int_{-\infty}^{\infty}\mathrm{d}x\,n_{\sigma}(x), (68)
1D\displaystyle\mathcal{I}_{1\mathrm{D}} =\displaystyle= dxC1D(x).\displaystyle\int_{-\infty}^{\infty}\mathrm{d}x\,C_{1\mathrm{D}}(x). (69)

With the help of the NN-body Schrödinger equation, we find

σJσ(β)\displaystyle\sum_{\sigma}J_{\sigma}(\beta) =\displaystyle= νσnνσeβϵν\displaystyle\sum_{\nu\sigma}n_{\nu\sigma}e^{-\beta\epsilon_{\nu}} (70)
=\displaystyle= Nβσdx[V(x)nσ(x)22mvσ(x)]\displaystyle N-\beta\sum_{\sigma}\int\mathrm{d}x\left[V(x)n_{\sigma}(x)-\dfrac{\hbar^{2}}{2m}v_{\sigma}(x)\right]
2a1D1Dβ2m+23β3/21D32πm3\displaystyle-\dfrac{\hbar^{2}a_{1\mathrm{D}}\mathcal{I}_{1\mathrm{D}}\beta}{2m}+\frac{2\hbar^{3}\beta^{3/2}\mathcal{I}_{1\mathrm{D}}}{3\sqrt{2\pi m^{3}}}
=\displaystyle= NEβ2a1D1Dβ2m\displaystyle N-E\beta-\dfrac{\hbar^{2}a_{1\mathrm{D}}\mathcal{I}_{1\mathrm{D}}\beta}{2m}
+231Dβ3/232πm3+O(β2).\displaystyle+\frac{2\hbar^{3}\mathcal{I}_{1\mathrm{D}}\beta^{3/2}}{3\sqrt{2\pi m^{3}}}+O(\beta^{2}).

Applying ddβ\frac{\mathrm{d}}{\mathrm{d}\beta} to the above expansion and taking β0\beta\rightarrow 0, we get the energy functional shown in Eq. (6). Clearly, the energy functional only gains an extra finite shift, 2a1D1D2m-\frac{\hbar^{2}a_{1\mathrm{D}}\mathcal{I}_{1\mathrm{D}}}{2m}, due to the interaction. In 1D, the energy theorem is [32]

E=σdk2π2k22mnσ(k)2a1D1D2m+V.E=\sum_{\sigma}\int\frac{\mathrm{d}k}{2\pi}\frac{\hbar^{2}k^{2}}{2m}n_{\sigma}(k)-\frac{\hbar^{2}a_{1\mathrm{D}}\mathcal{I}_{1\mathrm{D}}}{2m}+\expectationvalue{V}. (71)

If there is no external potential, namely if V0V\equiv 0, the energy functional in Eq. (6) reduces to this energy theorem.

We also calculate the asymptotics of ρ(ϵ)\rho(\epsilon) and nνσn_{\nu\sigma} in 1D, and the results are Eq. (7) and

nνσ=1kν4dxC1D(x)|ϕν(x)|2+O(ϵν5/2),n_{\nu\sigma}=\frac{1}{k_{\nu}^{4}}\int\text{d}x\,C_{1\mathrm{D}}(x)|\phi_{\nu}(x)|^{2}+O(\epsilon_{\nu}^{-5/2}), (72)

where kν=2mϵν/k_{\nu}=\sqrt{2m\epsilon_{\nu}}/\hbar.

VI Summary and discussion

In this work, we have generalized the universal energy functional for trapped two-component Fermi gases from 3D to lower spatial dimensions. We have shown that in lower dimensions the total energy of two-component fermions with zero-range interaction trapped in any smooth potential can be expressed as linear functionals of the occupation probabilities of one-particle energy eigenstates, just like in 3D. We first calculated the one-particle density matrix of two-component fermions by using the Bethe-Peierls boundary conditions. We have also calculated the asymptotic formulas of the occupation probabilities of single-particle levels at high energy.

The energy functional [Eq. (3) in 2D, or Eq. (6) in 1D] is a universal functional, and it holds for all finite-energy states, i.e. both few-body and many-body states, both pure and mixed states, both zero-temperature and finite-temperature states. It will be important to understand the nontrivial constraints on the occupation probabilities of the single-particle levels, because such understanding will enable one to determine the many-body ground state energies by minimizing the energy functional in the presence of such constraints. One might be able to generalize the energy functional to multi-component fermions, to fermions with unequal masses, and to bosons. Future experiments might be able to measure both the occupation probabilities of single-particle levels and the many-body energies of the systems we have studied. Such experiments should verify the energy functionals that we have derived.

Acknowledgements.
This work was supported by the National Key R&\&D Program of China (Grants No. 2019YFA0308403 and No. 2021YFA1400902).

References