Calculating Eigenvalues of Many-Body Systems From Partition Functions
Calculating Eigenvalues of Many-Body Systems From Partition Functions
Calculating Eigenvalues of Many-Body Systems From Partition Functions
Abstract: A method for calculating the eigenvalue of a many-body system without solving
the eigenfunction is suggested. In many cases, we only need the knowledge of eigenvalues
rather than eigenfunctions, so we need a method solving only the eigenvalue, leaving alone
the eigenfunction. In this paper, the method is established based on statistical mechanics.
In statistical mechanics, calculating thermodynamic quantities needs only the knowledge of
eigenvalues and then the information of eigenvalues is embodied in thermodynamic quan-
tities. The method suggested in the present paper is indeed a method for extracting the
eigenvalue from thermodynamic quantities. As applications, we calculate the eigenvalues
for some many-body systems. Especially, the method is used to calculate the quantum
exchange energies in quantum many-body systems. Using the method, we also calculate
the influence of the topological effect on eigenvalues. Moreover, we improve the result of
the relation between the counting function and the heat kernel in literature.
∗
daiwusheng@tju.edu.cn.
Contents
1 Introduction 1
11 Conclusions 20
1 Introduction
Calculating the eigenvalue of a many-body system is often difficult. The most direct way
is to solve the eigenequation of the Hamiltonian H,
–1–
Once the eigenequation is solved, both the eigenvalue En and the eigenfunction φn are
solved simultaneously. Nevertheless, often one only needs the knowledge of eigenvalues
{En } rather than eigenfunctions. In such cases, solving eigenfunctions is redundant. This
inspires us to develop an approach which only focus on solving eigenvalues.
Statistical mechanics is essentially an averaging method. In statistical mechanics the
information of eigenfunctions is statistically averaged out; calculating thermodynamic quan-
tities needs only eigenvalues. For example, in canonical ensembles, all the thermodynamic
property of an N -particle system is embodied in the canonical partition function which is
determined only by eigenvalues {En }:
X
Z (β) = e−βEn , (1.2)
n
1
where β = kT with k the Boltzmann constant and T the temperature.
The knowledge of eigenvalues is embodied in the canonical partition function Z (β).
Our problem is how to extract the eigenvalue En from the partition function (1.2), or, more
generally, from thermodynamic quantities.
The canonical partition function (1.2), from another perspective, is the global heat
kernel of the Hamiltonian operator H. The global heat kernel
X
K (t) = e−tEn (1.3)
n
is the trace of the local heat kernel which is the Green function of the initial-value problem of
the heat-type equation [1]. Obviously, the global heat kernel is just the canonical partition
function with the replacement t → β.
Recently, a relation between the heat kernel and the spectral counting function is
revealed [2, 3]. The counting function N (E) counts the number of the eigenstates whose
eigenvalues are smaller than E. The relation between the heat kernel and the counting
function allows us to calculate the counting function N (E) from the heat kernel K (t), or,
the canonical partition function Z (β), directly.
The eigenvalue can be calculated from the counting function N (E) [2]. This implies
that one can calculate the eigenvalue from the canonical partition function Z (β) or other
thermodynamic quantities.
It is often difficult to calculate the eigenvalue of noninteracting quantum systems and
interacting classical and quantum systems. In noninteracting quantum systems, there exist
quantum exchange interactions; in classical interacting systems, there exist classical inter-
particle interactions, and in quantum interacting systems, there exist both classical inter-
particle interactions and quantum exchange interactions [4]. The method developed in the
present paper allows us to calculated eigenvalues from the thermodynamic quantity which
is obtained by the statistical mechanical method.
In quantum many-body systems, the most important factor is the quantum exchange
interaction. The effect of quantum exchange interaction in eigenvalue is the exchange en-
ergy. In statistical mechanics, the quantum exchange effect is taken into account by simply
employing Bose-Einstein statistics, Fermi-Dirac statistics, and various kinds of intermedi-
ate statistics through imposing various maximum occupation numbers: ∞ for Bose-Einstein
–2–
statistics, 1 for Fermi-Dirac statistics, and an integer n for Gentile statistics [5–12]. The
method suggested in the present paper allows us to calculate the exchange energy in eigen-
values from the partition function obtained in statistical mechanics. In other words, we
can calculate the exchange energy in virtue of the statistical mechanics. For example, in
quantum ideal and interacting gases, the contribution of the exchange energy to the eigen-
value is represented by the second virial coefficient which can be obtained by statistical
mechanical method.
In quantum mechanics, the exchange energy is reckoned in by symmetrizing or anti-
symmetrizing the wavefunctions, in quantum filed theory, the exchange energy is reckoned
in by imposing the quantization condition on the fields, while, in statistical mechanics, the
exchange energy is reckoned in by only simply setting the value of the maximum occupa-
tion number, since in statistical mechanics only the information of eigenvalues other than
the information of wavefunctions is needed. That is to say, in statistical mechanics, the
exchange energy can be calculated relatively simply. Therefore, calculating the exchange
energy through statistical mechanics is a more simple approach.
Moreover, using the method, we calculate the influence of the topological effect on
eigenvalues for two-dimensional non-interacting classical and quantum systems. In two-
dimensional systems, the topological property is described by connectivity which is de-
scribed by the Euler-Poincaré characteristic number [13].
In statistical mechanics, there are many solved models, in which the partition functions
are solved exactly or approximately. Using the method, we calculate the eigenvalues for such
models from the solved partition functions. In this paper, we consider classical and quantum
non-interacting systems, classical interacting systems with the Lennard-Jones interaction
and quantum interacting systems with the hard-sphere interaction, the one-dimensional
Ising models, and the one-dimensional Potts model.
There are many studies devoted to the problem of eigenvalue spectra, such as the
eigenvalue spectrum of the Rabi model [14], the eigenvalue spectrum of the open spin-1/2
XXZ quantum chains with non-diagonal boundary terms [15], the eigenvalue spectrum of
the antiperiodic spin-1/2 XXZ quantum chains [16], the statistical property of the eigenvalue
spectrum [17, 18], the structure of the eigenvalue spectrum [19, 20], the ground-state energy
of the Heisenberg-Ising lattice [21], the ground state and the excited state of many-body
localized Hamiltonians [22], and the ground state energy of a system of N bosons [23].
In statistical mechanics, many methods are developed for the calculation of partition
functions. For example, the canonical partition function for quon statistics [24], general
formulas for the canonical partition function of parastatistical systems [25], the canonical
partition function of the freely jointed chain model [26], the partition function of the inter-
acting calorons ensemble [27], the algorithm for computing the exact partition function of
lattice polymer models [28], the exact partition function for the q-state Potts Model [29],
the partition function for the antiferromagnetic Ising model and the hard-core models [30],
and the canonical partition functions for different gaseous systems [31] are investigated.
The relation between the counting function and the heat kernel is the basics of the
method used in the present paper. However, the relation given by Ref. [2] neglects a
special case. In this paper, we improve the result in Ref. [2].
–3–
This paper is organized as follows. In section 2, we describe the method of calculating
the eigenvalue from the canonical partition function. In section 3, we illustrate the method
by examples. In sections 4 and 5, we calculate the eigenvalue, especially the exchange
energy, of identical particles in a box and in an external field. In section 6, we calculate
the influence of the topological effect on eigenvalues. In sections 7 and 8, we calculate the
eigenvalue of interacting particles with the Lennard-Jones potential and the hard-sphere
potential. In sections 9 and 10, we calculate the eigenvalues of the Ising system and the
Potts system. Conclusions are summarized in section 11. In the appendix, we provide a
complete result for the relation between the counting function and the heat kernel.
c+i∞
1 Z (β) βE
Z
N (E) = e dβ. (2.1)
2πi c−i∞ β
The n-th eigenvalue En can be obtained from the counting function by the equation
[32]
N (En ) = n. (2.2)
–4–
Consequently, the eigenvalue En can be solved by the following equation:
c+i∞
1 Z (β) βEn
Z
e dβ = n. (2.3)
2πi c−i∞ β
In the following, we solve eigenvalues for some many-body systems from the corre-
sponding canonical partition functions.
In should be emphasized that, in the relation between the counting function and the
heat kernel (canonical partition function), Eq. (2.1), there is a constant term − 12 when
E = En , (see Appendix A). We will not take the contribution into account, because its
influence is often small enough to be ignored, especially for highly-excited states, .
In this section, we illustrate the method by some models whose eigenvalues are already
known, including a particle in a box, a harmonic oscillator, and N bosonic harmonic oscil-
lators.
where z (β) is the single-particle partition function. The single-particle partition function
for free particles is [33]
V
z (β) = D , (3.2)
λ
q
β
where λ = h 2πm is the thermal wavelength with m the mass of the particle and h the
Planck constant, V is the volume of the container, and D is the spatial dimension. The
canonical partition function of an N -particle classical ideal gas, by Eqs. (3.2) and (3.1), is
then
V N
Z (β) = . (3.3)
λD
By the relation between the counting function N (λ) and the canonical partition func-
tion Z (β), Eq. (2.1), we can obtain the counting function,
2πm DN/2
cl N 1 DN/2
N En = V Encl , (3.4)
h2 Γ (1 + DN/2)
–5–
where Γ (x) is the Gamma function. The eigenvalue is determined by the equation obtained
by substituting Eq. (3.4) into Eq. (2.2):
2πm DN/2
N 1 DN/2
V Encl = n. (3.5)
h2 Γ (1 + DN/2)
Solving the equation gives the eigenvalue
h2
DN
Encl = Γ 2/(DN )
1+ n2/(DN ) . (3.6)
2πmV 2/D 2
Now let us see a familiar special case: the one-dimensional single-particle case. In this
case, D = 1, N = 1, and V = L. Eq. (3.6) then becomes
h2 n 2
Encl = . (3.7)
N =1 8m L
This is just the eigenvalue of a particle in a one-dimensional periodic box with a side length
L. In a one-dimensional periodic box with a side length L, the momentum of the particle
h
is pn = 2L n, so the eigenvalue (3.7) becomes
p2
Encl = n. (3.8)
N =1 2m
3.2 One harmonic oscillator
The harmonic oscillator in quantum mechanics corresponds to a classical ideal harmonic
oscillator gas in statistical mechanics.
In order to show the validity of the method and illustrate the method, we take the
harmonic oscillator as an example.
The eigenvalue of a harmonic oscillator is exactly known: En = ~ω n + 21 . The
Now we show how to obtain the eigenvalue from the partition function by the method.
By the relation between the counting function N (En ) and the canonical partition
function Z (β), Eq. (2.1), we can obtain the counting function.
First expand partition function (3.9) in power series of e−β~ω ,
∞
X 1
Z (β) = e− 2 β~ω e−β~ωk . (3.10)
k=1
Substituting Eq. (3.10) into Eq. (2.1) gives the counting function:
∞
X 1
N (E) = θ E− + k ~ω , (3.11)
2
k=1
where θ (x) is the Heaviside theta function. The eigenvalue is determined by N (En ) = n,
which directly gives the eigenvalue of a harmonic oscillator:
1
En = ~ω n + . (3.12)
2
–6–
3.3 N bosonic harmonic oscillators
An N bosonic harmonic oscillator system in quantum mechanics corresponds to a bosonic
harmonic oscillator gas in statistical mechanics.
In this example, we show the validity of the method.
where Z (β) is the single particle partition function given by Eq. (3.9).
As examples, consider two cases: N = 2 and N = 3.
Exact eigenvalues For N = 2, the partition function given by Eq. (3.13) reads
e2~ωβ
Z (β, 2) = 2 . (3.14)
(e~ωβ − 1) (e~ωβ + 1)
The counting function can be obtained by substituting Eq. (3.15) into Eq. (A.4):
X1h i
N (E) = 2k + 1 − (−1)k θ (E − k~ω) , (3.16)
4
k=1
The eigenvalue can be obtained by solving Eq. (2.2). It can be directly shown that the
exact solution of Eq. (2.2) with the counting function (3.16) is
E N =2 = (n + 1) ~ω (3.17)
e9~ωβ/2
Z (β, 3) = 3 . (3.19)
(e~ωβ − 1) (1 + e~ωβ ) (1 + e~ωβ + e2~ωβ )
–7–
Expanding Z (β, 3) given by Eq. (3.19) as a power series of e−β~ω gives
X 1
3
9
Z (β, 3) = 47 + 6 k − k+ − 9 (−1)k−1/2
72 2 2
n= 23 , 52 ,...
2π k
+16 cos k−π e−β~ω . (3.20)
3
The counting function can be obtained by substituting Eq. (3.20) into Eq. (A.4):
X 1
3
9
N (E) = 47 + 6 k − k+ − 9 (−1)k−1/2
72 2 2
n= 23 , 52 ,...
2π
+16 cos k − π θ (E − ~ωk) . (3.21)
3
The eigenvalue can be obtained by solving Eq. (2.2). It can be directly shown that the
exact solution of Eq. (2.2) with the counting function (3.21) is
N =3 3
E = n+ ~ω (3.22)
2
1 X − (1 + k) (~ω)k B2+k 1
Z c+i∞
N (E) = β k−1 eβE dβ
2 (2 + k)! 2πi c−i∞
k≥−2
k
1 X 2k (~ω) B1+k 21 1
Z c+i∞
+ β k−1 eβE dβ, (3.24)
2 (1 + k)! 2πi c−i∞
k≥−1
we arrive at
1 E 1 E2
N (E) = − + + 2 2, (3.26)
24 4 ~ω 4~ ω
–8–
En ℏω
20
●●●
●●●●
■■■■■■■■■■
●●●●●●
●
■■■■■■■■■
●●●
●●●
●●●
■■■■■■■■■
●●●●●
●●●
●●●
■■■■■■■■
●●●
15 ●●●
■■■■■■■■
●●
●
●●●
●●●
■■■■■■■
●●●●●
■■■■■■■
●●●
●●●
●●
■■■■■■
●●● ● The smoothed eigenvalue
●●
■■■■■■
●●
●●
10 ●●●●
■■■■■
■ The exact eigenvalue
●●
■■■■■
●●
●●●
■■■■
●●
●●
■■■■
●●●
■■■
●
●
5 ●
■■■
●
●
●
■■
●
n
20 40 60 80
Figure 1. Comparison of the smoothed eigenvalue and the exact eigenvalue of a two bosonic
oscillator system.
where B0 = 1, B1 = − 21 , B2 = 61 , B0 12 = 1, and B1 1
2 = 0 are used. Solving Eq. (2.2)
with Eq. (3.26) gives the smoothed eigenvalue,
r !
N =2 1 48n + 5
En, = − 1 ~ω. (3.27)
2 3
−5 13E 1 E2 E3
N (E) = + + 2 2+ . (3.28)
96 144 ~ω 8~ ω 36~3 ω 3
Solving Eq. (2.2) with Eq. (3.28) gives the smoothed eigenvalue,
7
EnN =3 = h p i1/3
61/3 648n + 6 (69984n2 − 343)
h p i1/3
2
648n + 6 (69984n − 343)
3
+ − ~ω. (3.29)
62/3 2
The exact eigenvalues, Eqs. (3.17), (3.18), (3.22), and (3.23), and the smoothed eigen-
value, Eq. (3.28) and (3.29) are compared in Figures (1) and (2).
–9–
En /ℏω
14 ●●●●
●●●●●
■■■■■■■■■■■■■■■■■
●●●●●●●●
●
●●●●
●●●●
■■■■■■■■■■■■■■■■
●●●●
12 ●●●●●●●
●●●
■■■■■■■■■■■■■■
●●●●
●●●
●●●●●
●
■■■■■■■■■■■■
●●
10 ●●●
●●●
●●●●●
■■■■■■■■■■
●●
●●●
●●
■■■■■■■■
●●
8 ●●
●● ● The smoothed eigenvalue
●●
■■■■■■■
●●
●●
■■■■■
● ■ The exact eigenvalue
●●
6 ●●
●
■■■■
●
●
●
■■■
●
4 ●
n
20 40 60 80
Figure 2. Comparison of the smoothed eigenvalue and the exact eigenvalue of a three bosonic
oscillator system.
1 1
Z (β) ≃ z (β)N ± z (β)N −2 z (2β) + . . . , (4.1)
N! 2 (N − 2)!
where "+" stands for Bose gases, "−" stands for Fermi gases, and z (β) is the single-
particle partition function given by Eq. (3.2). Substituting Eq. (3.2) into Eq. (4.1) gives
the canonical partition function of a quantum ideal gas,
N N −1
1 V 1 V
Z (β) ≃ ± 1+D/2 + ··· . (4.2)
N! λD 2 (N − 2)! λD
– 10 –
The counting function is given by substituting Eq. (4.2) into Eq. (2.1):
V N DN/2
1
N (E) ≃ En
N !Γ (1 + DN/2) ΛD
V N −1 DN/2−D/2
1
± 1+D/2 En + ··· , (4.3)
2 (N − 2)!Γ (1 + D (N − 1) /2) ΛD
where Λ = √h . Then solving Eq. (2.2) with the counting function (4.3) gives the
2πm
eigenvalue,
– 11 –
5 Noninteracting identical particles in external fields
In this section, we calculate the eigenvalue of identical particles in an external field with
the help of statistical mechanics.
For an ideal gas in an external field, the single-particle eigenvalue is determined by
~2
the Hamiltonian H = − 2m ∇2 + U (x). The partition function for a classical gas can be
expressed as [1, 2, 35]
V X
z (β) = D 1 + Bk β k , (5.1)
λ
k=1/2,1,...
where Bk is the heat kernel coefficient, e.g., B1/2 = V1 dxU (x). Eq. (5.1) is known as the
R
1 V 2
2
V p 1 V
Z (β) ≃ D
+ D
βB1/2 ± 1+D/2 D + . . . , (5.3)
2 λ λ 2 λ
where the second term is the contribution of the external field, the third term is the con-
tribution of the quantum exchange interaction.
The counting function can be obtained by substituting Eq. (5.3) into Eq. (2.1):
V 2 D V 2
1 1
N (E) ≃ E + B1/2 E D−1/2
2Γ (1 + D) ΛD Γ (1/2 + D) ΛD
1 V D/2
± 1+D/2 E + .... (5.4)
2 Γ (1 + D/2) ΛD
Solving Eq (2.2) with the counting function Eq. (5.4) gives
" #
±2c1 n1/(2D)−1/2 + 2 B1/2 ΛV −1/D
2/(N D)
En |N =2 ≃ Encl (N !) 1− ,
c2 n1/(2D) ± Dc1 n1/(2D)−1/2 + (2D − 1) B1/2 ΛV −1/D
(5.5)
where c1 = 2−3/2+1/(2D)−D/2 (D!)−1/2+1/(2D) Γ(1/2+D)
Γ(1+D/2) and c2 = 21/(2D) D (D!)−1+1/(2D) Γ (1/2 + D).
From Eq. (5.5), one can see the influence of the external field, B1/2 ΛV and the −1/D ,
influence of the quantum exchange interaction, partly reflected in the terms with the sign
"±", on the eigenvalue of a quantum ideal gas.
For one-dimensional cases, the eigenvalue
" √
−1/D ± 2 2
#
h2 8 B ΛV
1/2
En |D=1
N =2 ≃ 1− √ √ n, (5.6)
πmL2 2 2πn1/2 + 4 B1/2 ΛV −1/D ± 2
– 12 –
for two-dimensional cases, the eigenvalue
" √ #
32 B1/2 ΛV −1/D n1/4 ± 3 2π
D=2 h2
En |N =2 ≃ 1− √ √ n1/2 , (5.7)
πmΩ
12 2πn1/2 + 48 B1/2 ΛV −1/D n1/4 ± 3 2π
and for three-dimensional cases, the eigenvalue
1/3
D=3 3 h2
En |N =2 ≃ [1
2 πmV 2/3
#
192 B1/2 ΛV −1/D n1/3 ± 10 × 25/6 × 32/3
− 1/3 1/6
√ 1/2 −1/D
1/3 5/6 2/3
n1/3 .
90 × 2 × 3 πn + 480 B1/2 ΛV n ± 15 × 2 × 3
(5.8)
In this section, we discuss the topology effect on the eigenvalue of classical and quantum
particles in nontrivial topological containers. Classical and quantum particles in quantum
mechanics corresponds to classical and quantum gases in statistical mechanics. In statistical
mechanics, the geometric effect and the topology effect are systematically studied [36–41].
In the following, we calculate the eigenvalue from the result given by statistical mechanics.
– 13 –
From the expression of the eigenvalue, Eq. (6.3), we can see the geometric effect, reflected
in the terms with the factor √LΩ , and the topological effect, reflected in the terms with the
factor χ, explicitly.
The counting function is given by substituting Eq. (6.4) into Eq. (2.1):
Ω 2 2
1 1 L Ω 3/2 1 Ω
N (E) ≃ 2
E − √ E ± E
4 Λ 3 π Λ Λ2 2 Λ2
1 L 2
χ Ω
+ E+ E + .... (6.5)
32 Λ 6 Λ2
Solving Eq. (2.2) with the counting function Eq. (6.5) gives the eigenvalue
√ √
1 L2
2 2 √L 3/4 1 1/2
2
h − 3 Ω
n + π ±1 + 3 χ + 16 Ω n
En |N =2 ≃ 1− √ √ L n1/2 . (6.6)
πmΩ 3/4
√ 1 1 L 2
1/2
2 πn − 2 √ n + π ±1 + χ + Ω
n3 16 Ω
From Eq. (6.6), one can see that the eigenvalue of a two-dimensional ideal quantum
gas in a nontrivial topological box is modified by the geometric effect described by √LΩ ,
and by the topological effect described by χ. Moreover, for such quantum cases, there exist
exchange energies partly reflected in the terms with the sign "±".
A system consisting of particles interacted with each other through the Lennard-Jones
potential in quantum mechanics corresponds to an interacting gas with the Lennard-Jones
interaction in statistical mechanics.
The Lennard-Jones inter-particle potential reads
σ 12 σ 6
U = 4ε − , (7.1)
r r
where ε is the depth of the potential well and σ is the finite distance at which the inter-
particle potential is zero [42]. The canonical partition function of a classical interacting gas
with N particles is given by [31]
N N −1
1 V 1 V
Z (β) ≃ + b2 + . . . , (7.2)
N! λ3 (N − 2)! λ3
– 14 –
where b2 = 2π
R 3
d r e−βU − 1 .
λ3
The coefficient b2 for the Lennard-Jones interaction is [42]
2π r03
b2 ≃ (u0 β − 1) , (7.3)
3 λ3
where r0 = 21/6 σ. Then the canonical partition reads
N N 3 N 3
1 V 2π V r0 2π V r0
Z (β) ≃ 3
− 3
+ 3
u0 β + . . . . (7.4)
N! λ 3 (N − 2)! λ V 3 (N − 2)! λ V
The counting function can be obtained by substituting Eq. (7.4) into Eq. (2.1),
N
r03 V N 3N/2
1 V 3N/2 2π
N (E) ≃ E − E
N !Γ (1 + 3N/2) Λ3 3 (N − 2)!Γ (1 + 3N/2) V Λ3
N
2π r03 V
+ u0 E 3N/2−1 + . . . . (7.5)
3 (N − 2)!Γ (3N/2) V Λ3
The eigenvalue can be obtained by solving Eq. (2.2) with the counting function (7.5):
En ≃ Encl (N !)2/(3N ) [1
4 (N !)2/(3N ) Γ2/(3N ) (1 + 3N/2) r03 n2/(3N ) − 6N r03 V 2/3 u0 Λ−2
+ h i .
3 (N !)2/(3N ) Γ2/(3N ) (1 + 3N/2) (N −1)π3
V − 2N r03 n2/(3N ) + 3N (3N − 2) r03 V 2/3 u0 Λ−2
(7.6)
Especially, for N = 1, there is of course no inter-particle interactions, so the eigenvalue
(7.6) recovers the eigenvalue of a free particle
32/3 h2
En |N =1 ≃ n2/3 . (7.7)
27/3 mπ 2/3 V 2/3
For N = 2, the eigenvalue (7.6) becomes
1/3 " #
3 h2 121/3 × 43 πr03 n1/3 − 4πr03 V 2/3 u0 Λ−2
En |N =2 ≃ 1+ n1/3 .
πmV 2/3 3 × 121/3 V − 34 πr03 n1/3 + 8πr03 V 2/3 u0 Λ−2
2
(7.8)
A system consisting of particles interacted with each other through the hard-sphere potential
in quantum mechanics corresponds to an interacting gas with the hard-sphere interaction
in statistical mechanics.
In a quantum hard-sphere gas, there exist both the quantum exchange interaction
and the classical hard-sphere interaction. The canonical partition function of a quantum
interacting gas is [31]
N N −1
1 V 1 V
Z (β) ≃ + b2 + . . . , (8.1)
N ! λ3 (N − 2)! λ3
– 15 –
where
1
Z
b2 = d6 qU2 (q1 , q2 ) (8.2)
2V λ3
~2
with U2 (q1 , q2 ) = λ6 hq1 , q2 |e−βH2 |q1 , q2 i − hq1 |e−βH1 |q1 ihq2 |e−βH1 |q2 i , H1 = − 2m ∇2 ,
2 2
H2 = − 2m~
∇21 − 2m
~
∇22 + 4πa~2 /mδ (q1 − q2 ), and a is the radius of the particle.
2 a5
For Bose gases, Eq. (8.2) becomes b2 = 2−5/2 − 2 λa − 10π 3 λ5 [42]; for Fermi gases, Eq.
3 5
(8.2) becomes b2 = −2−5/2 − 6π λa3 + 18π 2 λa5 [42].
Then the canonical partition function of a Bose gas is
N N −1
a 10π 2 a5
1 V 1 V −5/2
ZB (β) ≃ + 2 −2 − + ..., (8.3)
N ! λ3 (N − 2)! λ3 λ 3 λ5
and the canonical partition function of a Fermi gas is
N N −1
a3 5
1 V 1 V −5/2 2a
ZF (β) ≃ + −2 − 6π 3 + 18π 5 + . . . , (8.4)
N ! λ3 (N − 2)! λ3 λ λ
respectively.
Then the counting function of a Bose gas by Eq. (2.1) is
NB (E)
(2π)3N/2 m3N/2 V N 3N/2 23N/2−4 π 3(N −1)/2 m3(N −1)/2 V N −1 3N/2−3/2
≃ E + E
N !Γ (1 + 3N/2) h3N (N − 2)!Γ (3N/2 − 1/2) h3(N −1)
(8.5)
23N/2 π (3N −2)/2 a m(3N −2)/2 V N −1 3N/2−1 10a5 (2π)3N/2+3 m(3N +2)/2 V N −1 3N/2+1
− 3N −2
E − E + ...,
(N − 2)!Γ (3N/2) h 3 (N − 2)!Γ (3N/2 + 2) h3N +2
(8.6)
and the counting function of a Fermi gas by Eq. (2.1) is
NF (E)
3 N−1
(2π)3N/2 m3N/2 V N 3N/2 23N/2−4 π 3N/2−3/2 m3N/2− 2 V
≃ E − E 3N/2−3/2
N !Γ (1 + 3N/2) h3N (N − 2)!Γ (3N/2 − 1/2) h3N −3
(8.7)
3× 23N/2+1 π 3N/2+1 a3 m3N/2 V N −1 9× 23N/2+2 π 3N/2+3 a5 m3N/2+1 V N −1
− E 3N/2 + E 3N/2+1 + . . . .
(N − 2)!Γ (3N/2 + 1) h3N (N − 2)!Γ (3N/2 + 2) h3N +2
(8.8)
Solving Eq. (2.2) gives the eigenvalue of a hard-sphere Bose gas,
a5 a
" #
h n 5/(3N ) + 2h n 1/(3N ) − 2h
2/3N 1 5/3 2 1/3 3
En ≃ Encl (N !) 1+ 5
V V
(8.9)
h1 Va5/3 n5/(3N ) + h4 n1/N + 5h2 V 1/3 a
n1/(3N ) + 3 (N − 1) h3
10π 2 (N !)5/(3N) Γ5/(3N) (1+3N/2) 1/(3N) 1/(3N)
with h1 = 3Γ(3N/2+2) , h2 = 2(N !) Γ
Γ(3N/2)
(1+3N/2) 1
, h3 = 25/2 Γ(3N/2−1/2) ,
1/N 1/N−1
3(N !) Γ (1+3N/2)
and h4 = N −1 . The eigenvalue of a hard-sphere Fermi gas,
5 3
−2t1 Va5/3 n5/(3N ) + 2t2 aV n1/N + 2t3
" #
En ≃ Encl (N !)2/(3N ) 1 + 5 3
(2 + 3N ) t1 Va5/3 n5/(3N ) + t4 n1/N − 3N t2 aV n1/N − 3 (N − 1) t3
(8.10)
– 16 –
18π 2 (N !)5/(3N) Γ5/(3N) (1+3N/2) 6π(N !)1/(3N) Γ1/(3N) (1+3N/2) 1
with t1 = Γ(3N/2+2) , t2 = Γ(3N/2+1) , t3 = 25/2 Γ(3N/2−1/2)
,
1/N 1/N−1
and t4 = 3 (N !) Γ N −1 (1+3N/2) .
In order to illustrate the influence of the quantum exchange interaction, we consider a
system consisting of two bosons or two fermions. For the Bose case, the eigenvalue (4.5)
becomes
D=3 h2
En |Bose
N =2 ∼ 12
1/3
n1/3 [1
2πmV 2/3
5 √ a 1/6
10 × 21/3 × 35/6 π 5/2 Va5/3 n5/6 + 36 × 31/6 π V 1/3
#
n − 6 × 21/6
+ 5 √ a 1/6 √ √ ;
−40 × 21/3 × 35/6 π 5/2 Va5/3 n5/6 − 72 × 31/6 π V 1/3 n + 9 × 21/6 6π n + 9 × 21/6
(8.11)
D=3 h2
En |FNermi
=2 ∼ 121/3 n1/3 [1 (8.12)
2πmV 2/3
5 √ 3√ √
−18 × 22/3 × 35/6 π 5/2 Va5/3 n5/6 + 24 3π 3/2 aV n + 2 2
#
+ 5 √ 3√ √ √ √ . (8.13)
72 × 22/3 × 35/6 π 5/2 Va5/3 n5/6 − 72 3π 3/2 × aV n + 6 3π n − 3 2
The eigenvalue of the one-dimensional Ising model without interactions can be calculated
from the corresponding canonical partition function directly.
For a one-dimensional N -particle Ising model without the interaction between spins, the
canonical partition function reads [33]
where B is the magnetic field strength and µ is the spin magnetic moment.
The counting function can be obtained by substituting Eq. (9.1) into Eq. (2.1),
N +1
X N
N (E) = θ (E − Bµ [2 (s − 1) − N ]) , (9.2)
s=1
s−1
where θ (x) is the step function. Substituting Eq. (9.2) into Eq. (2.2) and solving the
equation give the eigenvalue
E = Bµ [2 (n − 1) − N ] (9.3)
– 17 –
Table 1. The eigenvalue E and the degree of degeneracy ω : N = 3
E −3Bµ − 3ǫ −Bµ + ǫ Bµ + ǫ 3Bµ − 3ǫ
ω 1 3 3 1
Z (β, N ) = λN N
+ + λ− , (9.5)
where q
βǫ 2
λ± = e cosh (βµB) ± cosh (βµB) − 2e−2βǫ sinh (2βǫ) . (9.6)
The counting function can be obtained by substituting Eq. (9.5) into Eq. (2.1). In the
following, we list the eigenvalues for N = 3, 4, 5, . . . , 9.
For example, for N = 3, the counting function is
The Potts model in statistical mechanics is a generalization of the Ising model, which is a
model of interacting spins on a crystalline lattice. The Hamiltonian of the one-dimensional
– 18 –
Table 5. The eigenvalue E and the degree of degeneracy ω : N = 7
E −7Bµ − 7ǫ −5Bµ − 3ǫ −3Bµ + ǫ −3Bµ − 3ǫ −Bµ + 5ǫ −Bµ + ǫ
ω 1 7 14 7 7 21
E −Bµ − 3ǫ Bµ + 5ǫ Bµ + ǫ Bµ − 3ǫ 3Bµ + ǫ 3Bµ − 3ǫ
ω 7 7 21 7 14 7
E 5Bµ − 3ǫ 7Bµ − 7ǫ
ω 7 1
P
Potts model is H = −J hi,ji δ (σi , σj ), where σi = 1, 2, 3, . . . , q, δ (σi , σj ) = 1 if i = j,
and (σi , σj ) = 0 other wise. The canonical partition function of the one-dimensional Potts
model of N particles is [44]
N −1
Z (β, N ) = q q − 1 + eβJ . (10.1)
The counting function can be obtained by substituting Eq. (10.1) into Eq. (2.1),
N −1
N −1
(q − 1)N −1−l θ (Jl + E) .
X
N (E) = q (10.2)
l
l=0
Substituting Eq. (10.2) into Eq. (2.2) and solving the equation give the eigenvalue
E = nJ (10.3)
– 19 –
11 Conclusions
In this paper, we suggest a method for calculating the eigenvalue of a many-body system
from the corresponding canonical partition function. The advantage of the method is that
it allows us to merely calculate the eigenvalue without solving the eigenfunction simultane-
ously. Recalling that in many approximate methods, although only needing the eigenvalue,
one has to solve the eigenfunction in the meantime. Solving eigenfunctions, however, is
always a difficult task.
In statistical mechanics, the calculation of thermodynamic quantities only needs the
knowledge of eigenvalues. Only the information of eigenvalue is embodied in thermody-
namic quantities. The method suggested in the present paper is an approach for extracting
the eigenvalue from the thermodynamic quantity which obtained by statistical mechanical
method.
In the present paper, we calculate the eigenvalue from the canonical partition function.
In future works, one can generalizes the method to calculate the eigenvalue from the other
thermodynamic quantities.
Moreover, we improve the result of the relation between the counting function and the
heat kernel given in [2].
In Ref. [2], we only consider the counting function which counts the number of eigenstates
with eigenvalue smaller than a given number λ, but a special case is ignored: the given
number is just a eigenvalue, i.e., λ = λn with λn the n-th eigenvalue. In the following, we
provide a complete version of the relation between N (λ) and K (t).
Theorem 1
Z c+i∞
eλt
1
K (t) dt, when λn 6= λ (A.3)
2πi
c−i∞ t
N (λ) = Z c+i∞
1 eλt 1
K (t) dt − , when λn = λ (A.4)
2πi c−i∞ t 2
ln n
with c > limn→∞ λn .
– 20 –
is a generalization of the Dirichlet series. The function
∞
X an
f (s + ω) = (A.6)
µs+ω
n=1 n
1
R c+iT x ω dω
The integral in the right-hand side of Eq. (A.7), 2πi c−iT µn ω , should be considered
in different situations, i.e., µn < x, µn = x, and µn > x. In the limitation T → ∞, the
integral reads [45]
Z c+iT ω 1, µn < x,
1 x dω 1
lim = , µn = x, (A.8)
T →∞ 2πi c−iT µn ω
2
0, µ > x.
n
Comparing the definition of the counting function and the global heat kernel, Eqs. (A.1)
and (A.2), with Eq. (A.10) proves Eqs. (A.3) and (A.4).
Acknowledgments
We are very indebted to Dr G. Zeitrauman for his encouragement. This work is supported
in part by NSF of China under Grant No. 11575125 and No. 11675119.
References
[1] D. V. Vassilevich, Heat kernel expansion: user’s manual, Physics reports 388 (2003), no. 5
279–360.
– 21 –
[2] W.-S. Dai and M. Xie, The number of eigenstates: counting function and heat kernel,
Journal of High Energy Physics 2009 (2009), no. 02 033.
[3] W.-S. Dai and M. Xie, An approach for the calculation of one-loop effective actions, vacuum
energies, and spectral counting functions, Journal of High Energy Physics 2010 (2010), no. 6
1–29.
[4] W. Dai and M. Xie, Hard-sphere gases as ideal gases with multi-core boundaries: An
approach to two-and three-dimensional interacting gases, EPL (Europhysics Letters) 72
(2005), no. 6 887.
[5] G. Gentile j, itosservazioni sopra le statistiche intermedie, Il Nuovo Cimento (1924-1942) 17
(1940), no. 10 493–497.
[6] A. Khare, Fractional statistics and quantum theory. World Scientific, 2005.
[7] W.-S. Dai and M. Xie, Calculating statistical distributions from operator relations: The
statistical distributions of various intermediate statistics, Annals of Physics 332 (2012)
166–179.
[8] A. Algin and A. Olkun, Bose–einstein condensation in low dimensional systems with
deformed bosons, Annals of Physics 383 (2017) 239–256.
[9] W.-S. Dai and M. Xie, Intermediate-statistics spin waves, Journal of Statistical Mechanics:
Theory and Experiment 2009 (2009), no. 04 P04021.
[10] V. P. Maslov, The relationship between the fermi–dirac distribution and statistical
distributions in languages, Mathematical Notes 101 (2017), no. 3-4 645–659.
[11] W.-S. Dai and M. Xie, Gentile statistics with a large maximum occupation number, Annals of
Physics 309 (2004), no. 2 295–305.
[12] W.-S. Dai and M. Xie, An exactly solvable phase transition model: generalized statistics and
generalized bose–einstein condensation, Journal of Statistical Mechanics: Theory and
Experiment 2009 (2009), no. 07 P07034.
[13] M. Kac, Can one hear the shape of a drum?, The american mathematical monthly 73 (1966),
no. 4 1–23.
[14] A. J. Maciejewski, M. Przybylska, and T. Stachowiak, Full spectrum of the rabi model,
Physics Letters A 378 (2014), no. 1 16–20.
[15] S. Faldella, N. Kitanine, and G. Niccoli, The complete spectrum and scalar products for the
open spin-1/2 xxz quantum chains with non-diagonal boundary terms, Journal of Statistical
Mechanics: Theory and Experiment 2014 (2014), no. 1 P01011.
[16] G. Niccoli, Antiperiodic spin-1/2 xxz quantum chains by separation of variables: complete
spectrum and form factors, Nuclear Physics B 870 (2013), no. 2 397–420.
[17] M. Mierzejewski, T. Prosen, D. Crivelli, and P. Prelovšek, Eigenvalue statistics of reduced
density matrix during driving and relaxation, Physical review letters 110 (2013), no. 20
200602.
[18] Y. B. Lev, G. Cohen, and D. R. Reichman, Absence of diffusion in an interacting system of
spinless fermions on a one-dimensional disordered lattice, Physical review letters 114 (2015),
no. 10 100601.
[19] C. J. Callias, Spectra of fermions in monopole fields—exactly soluble models, Physical Review
D 16 (1977), no. 10 3068.
– 22 –
[20] M. Christandl, B. Doran, S. Kousidis, and M. Walter, Eigenvalue distributions of reduced
density matrices, Communications in mathematical physics 332 (2014), no. 1 1–52.
[21] C. Yang and C. Yang, Ground-state energy of a heisenberg-ising lattice, Physical Review 147
(1966), no. 1 303.
[22] X. Yu, D. Pekker, and B. K. Clark, Finding matrix product state representations of highly
excited eigenstates of many-body localized hamiltonians, Physical review letters 118 (2017),
no. 1 017201.
[23] M. Lewin, P. T. Nam, S. Serfaty, and J. P. Solovej, Bogoliubov spectrum of interacting bose
gases, Communications on Pure and Applied Mathematics 68 (2015), no. 3 413–471.
[24] J. Goodison and D. J. Toms, The canonical partition function for quons, Physics Letters A
195 (1994), no. 1 38–42.
[25] S. Chaturvedi, Canonical partition functions for parastatistical systems of any order, Physical
Review E 54 (1996), no. 2 1378.
[26] M. Mazars, Canonical partition functions of freely jointed chains, Journal of Physics A:
Mathematical and General 31 (1998), no. 8 1949.
[27] S. Deldar and M. Kiamari, Partition function of interacting calorons ensemble, in AIP
Conference Proceedings, vol. 1701, p. 100003, AIP Publishing, 2016.
[28] Y.-H. Hsieh, C.-N. Chen, and C.-K. Hu, Efficient algorithm for computing exact partition
functions of lattice polymer models, Computer Physics Communications 209 (2016) 27–33.
[29] S.-C. Chang and R. Shrock, Exact partition functions for the q-state potts model with a
generalized magnetic field on lattice strip graphs, Journal of Statistical Physics 161 (2015),
no. 4 915–932.
[30] A. Galanis, D. Štefankovič, and E. Vigoda, Inapproximability of the partition function for the
antiferromagnetic ising and hard-core models, Combinatorics, Probability and Computing 25
(2016), no. 4 500–559.
[31] C.-C. Zhou and W.-S. Dai, Canonical partition functions: ideal quantum gases, interacting
classical gases, and interacting quantum gases, Journal of Statistical Mechanics: Theory and
Experiment 2018 (2018), no. 2 023105.
[32] R. Courant and D. Hilbert, Methods of Mathematical Physics, Volume 2: Differential
Equations. John Wiley & Sons, 2008.
[33] L. Reichl, A Modern Course in Statistical Physics.
[34] C.-C. Zhou and W.-S. Dai, A statistical mechanical approach to restricted integer partition
functions, Journal of Statistical Mechanics: Theory and Experiment 2018 (2018) 053111.
[35] M. Bordag, E. Elizalde, and K. Kirsten, Heat kernel coefficients of the laplace operator on the
d-dimensional ball, Journal of Mathematical Physics 37 (1996), no. 2 895–916.
[36] W. S. Dai and M. Xie, Quantum statistics of ideal gases in confined space, Physics Letters A
311 (2003), no. 4–5 340–346.
[37] A. Aydin and A. Sisman, Discrete density of states, Physics Letters A 380 (2016), no. 13
1236–1240.
[38] C. Firat and A. Sisman, Quantum forces of a gas confined in nano structures, Physica
Scripta 87 (2013), no. 4 045008.
– 23 –
[39] W.-S. Dai and M. Xie, Interacting quantum gases in confined space: Two-and
three-dimensional equations of state, Journal of Mathematical Physics 48 (2007), no. 12
123302.
[40] A. Aydin and A. Sisman, Discrete nature of thermodynamics in confined ideal fermi gases,
Physics Letters A 378 (2014), no. 30 2001–2007.
[41] A. Aydin and A. Sisman, Dimensional transitions in thermodynamic properties of ideal
maxwell–boltzmann gases, Physica Scripta 90 (2015), no. 4 045208.
[42] R. Pathria, Statistical Mechanics. Elsevier Science, 2011.
[43] R. Feynman, Statistical Mechanics: A Set Of Lectures. Advanced Books Classics. Avalon
Publishing, 1998.
[44] G. Mussardo, Statistical field theory: an introduction to exactly solved models in statistical
physics. Oxford University Press, 2010.
[45] G. Tenenbaum, Introduction to analytic and probabilistic number theory, vol. 163. American
Mathematical Soc., 2015.
– 24 –