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

Thermal Broadening of Phonon Spectral Function in Classical Lattice Models: Projective Truncation Approximation

Hu-Wei Jia School of Physics, Renmin University of China, Beijing 100872, China Key Laboratory of Quantum State Construction and Manipulation (Ministry of Education), Renmin University of China, Beijing 100872, China    Wen-Jun Liu School of Physics, Renmin University of China, Beijing 100872, China Key Laboratory of Quantum State Construction and Manipulation (Ministry of Education), Renmin University of China, Beijing 100872, China    Yue-Hong Wu School of Physics, Renmin University of China, Beijing 100872, China Key Laboratory of Quantum State Construction and Manipulation (Ministry of Education), Renmin University of China, Beijing 100872, China    Kou-Han Ma International Center for Quantum Materials and School of Physics, Peking University, Beijing 100871, China    Lei Wang School of Physics, Renmin University of China, Beijing 100872, China Key Laboratory of Quantum State Construction and Manipulation (Ministry of Education), Renmin University of China, Beijing 100872, China    Ning-Hua Tong [email protected] School of Physics, Renmin University of China, Beijing 100872, China Key Laboratory of Quantum State Construction and Manipulation (Ministry of Education), Renmin University of China, Beijing 100872, China
Abstract

Thermal broadening of the quasi-particle peak in the spectral function is an important physical feature in many statistical systems, but it is difficult to calculate. To tackle this problem, we propose the HH-expanded basis within the projective truncation approximation (PTA) of the Green’s function equation of motion. A zeros-removing technique is introduced to stabilize the iterative solution of the PTA equations. Benchmarking calculations on the classical one-variable anharmonic oscillator model and the one-dimensional ϕ4\phi^{4} lattice model show that the thermal broadened quasi-particle peak in the spectral function can be produced on a semi-quantitative level. Using this method, we discuss the low- and high- temperature power-law behaviors of the spectral width Γk(T)\Gamma_{k}(T) of the one-dimensional ϕ4\phi^{4} model, finding it in contradiction with the assumption of effective phonon theory. A short-chain limit of this model is also discovered. Issues of extending the HH-expanded basis to quantum systems and of the applicability of the Debye formula for thermal conductivity are discussed.

I Introduction

The Green’s function (GF) is an important tool for studying the physical features of many-body systems in condensed-matter physics. The idea of projection [1, 2, 3, 4, 5] has been widely used in the calculation of the GF, leading to theories such as the two-pole approximation [6, 7], the composite operator method [8, 9], the self-consistent projection operator approach [10], the operator projection method [11], continued fraction [12, 13], the irreducible Green’s function method [14], mode-coupling theory [15, 16], and the dynamical projective operatorial approach [17], etc.. However, in practical applications, complicated analytical derivation and additional uncontrolled approximations are often required in these theories, which hampers systematic improvement of the results.

Recently, we developed a GF equation of motion (EOM) projective truncation approximation (PTA) method [18, 19, 20, 21]. The PTA overcomes the arbitrariness of traditional Tyablicov type truncation [22] by guaranteeing the analytical structure of the GF, diagonal-positivity of the spectral function, and the equilibrium state conservation of energy [18]. Compared to traditional projection-based methods, the PTA avoids the complicated analytical derivations to a large extent, and minimizes the need for additional uncontrolled approximations. Within the PTA, the study of a given system is reduced to selecting a basis of operators and solving the generalized eigenvalue problem on that basis. The power of modern computers is thus combined with the idea of projection. By a clever selection of the basis and systematically enlarging its size, exact results can be approached efficiently [19].

In the present work, we address the issue of describing the damping of quasi particles in the PTA. The damping (decay, relaxation, etc.) of quasi particles is ubiquitous in the experimental observations of spectra and transport. Modern numerical methods such as quantum Monte Carlo (QMC), density matrix renormalization group (DMRG), and tensor-network renormalization group (TNRG) methods have their own limitations in dealing with the damping. For examples, QMC requires an additional procedure to produce the real frequency spectral function, while DMRG and TNRG are more suitable for low dimensional systems at zero temperature. Theories based on Feynman diagrams usually apply only to weak correlation regime. In the traditional EOM formalism [4, 5] or the generalized Langevin equation approach [1, 2, 3], one needs to evaluate the imaginary part of the self-energy or the memory function to obtain a microscopic description of the damping, which is usually a difficult task. Within the PTA, although the quasi-particle energy can be obtained straightforwardly, the damping is also a challenge. This is because the PTA produces an approximate spectral function composed of NN discrete delta peaks, with NN being the dimension of the operator basis. Naively increasing NN to infinity is shown to be very inefficient for describing the broadening the spectral function and damping of quasi particles. See below.

Here, we propose the PTA on an HH-expanded basis to solve a part of this problem, i.e., the damping of quasi particles due to the thermal fluctuations at finite temperature. In the spectral function, it corresponds to the thermal broadening of the quasi-particle peak. The same idea can be extended to tackle the quantum damping and this issue will be addressed in the discussion part. In the framework of the PTA, our idea is first to produce the approximate eigen excitation operator OO, with [O,H]ϵO[O,H]\approx\epsilon O, from a low-order PTA. The corresponding spectral function has a delta peak at ω=ϵ\omega=\epsilon. Then we construct the new basis of operators by expanding OO with (exact or approximate) conserved quantities such as the Hamiltonian HH of the studied system or the mode Hamiltonian HνkH_{\nu k} of mode (νk)(\nu k). That is, we choose the new basis operators in the form {Ofi(H)}\{Of_{i}(H)\}, with fif_{i} (i=1,2,,Ni=1,2,...,N) being linearly independent functions. On the level of low-order PTA, these basis operators are degenerate excitation operators corresponding to the excitation energy ϵ\epsilon. A new PTA calculation on the expanded basis will take into account the thermal fluctuation beyond the low-order PTA and hence split the degeneracy, producing refined spectral functions composed of NN closely distributed delta peaks around the low-order quasi-particle peak δ(ωϵ)\delta(\omega-\epsilon). It provides a means to describe the thermal broadening of the quasi particle.

To illustration the implementation and effect of the above idea, we study the phonon damping problem in the classical lattice models which have only thermal damping of quasi particles. We consider two models. One is the one-variable anharmonic oscillator (AHO) model, whose static averages and spectral function are exactly solvable [23]. The other is the one-dimensional (1D) ϕ4\phi^{4} lattice model which is frequently used in the study of thermal transport in low-dimensional materials. For this model, molecular dynamical (MD) simulation results are available for benchmark.

At finite temperature, the phonon peak in the spectral function of 1D ϕ4\phi^{4} lattice will be broadened by thermal fluctuations due to the existence of nonlinear interaction. For general models, various methods have been developed to study the phonon decay and its consequence on the thermal transport, including the effective phonon theory [24], the anharmonic phonons [25] and the resonance phonon approaches [26, 27] based on MD simulation, non-equilibrium GF [28], series expansion [29], mode-coupling theory [15, 16], and the Boltzmann equation approach [30]. But a systematic study of the phonon decay for the 1D classical ϕ4\phi^{4} model, in particular, its low- and high-temperature asymptotic behaviors, is still lacking.

Our results show that PTA on the HH-expanded basis can produce semi-quantitatively accurate Γk\Gamma_{k}, the peak width of the phonon spectral function. It works for all temperatures, including the low temperature limit where MD simulation is too time consuming. Using this method, we find the low- and high-temperature asymptotic power laws in the temperature dependence of Γk(T)\Gamma_{k}(T) of the 1D ϕ4\phi^{4} model.

Notably, in contrast to the assumption of the effective phonon theory ΓkϵΩk\Gamma_{k}\propto\epsilon\Omega_{k} [24], our results suggest a different relation Γk(T)c[Ωk(T)Ωk(0)]\Gamma_{k}(T)\approx c\left[\Omega_{k}(T)-\Omega_{k}(0)\right], with cc being weakly dependent on TT and momentum kk. When inserted into the Debye formula [31, 32] for thermal conductivity κ\kappa, it gives different temperature dependence of κ(T)\kappa(T) from that of MD. This raised question on the applicability of Debye formula for the 1D ϕ4\phi^{4} lattice model. We also find a short-chain limit of the ϕ4\phi^{4} model with distinct temperature dependence of Γk(T)\Gamma_{k}(T), which is amenable to experimental test in low dimensional materials. The method could be extended to quantum many-body systems where the damping of quasi particles may occur at zero temperature due to quantum fluctuations. We address this issue in the discussion section.

The rest of this paper is arranged as follows. For completeness, in Sec.II, we first review the formalism of GF EOM and PTA for classical systems. In Sec.III, we develop the formalism of the PTA on an HH-expanded basis for the AHO model. In Sec.IV, we apply this method to 1D ϕ4\phi^{4} lattice model, and we analyze the thermal broadening effect in phonon spectral function. Section V is devoted to a discussion and summary.

II Green’s Function EOM and PTA for classical systems

In this section, we review the formalism of the PTA for classical system, which was developed in Ref.[33]. This is to set the stage for developing the PTA on the HH-expanded basis.

For two dynamical variables at time tt and tt^{\prime}, A[q(t),p(t)]A[q(t),p(t)] and B[q(t),p(t)]B[q(t^{\prime}),p(t^{\prime})] (qq and pp are abbreviation of canonical coordinates qiq_{i}’s and momenta pip_{i}’s of the studied system), the retarded GF is defined as [34, 35, 36, 37]

Gr[A(t)|B(t)]θ(tt){A(t),B(t)}.G^{r}\left[A(t)|B(t^{\prime})\right]\equiv\theta(t-t^{\prime})\langle\{A(t),B(t^{\prime})\}\rangle. (1)

Here θ(x)\theta(x) is the Heaviside step function. O\langle O\rangle is the average of variable OO in the equilibrium state of Hamiltonian H(q,p)H(q,p) at temperature TT. {,}\{\,,\,\} is the Poisson bracket.

The Fourier transformation of the above GF is

Gr(A|B)ω=Gr[A(t)|B(t)]ei(tt)(ω+iη)d(tt).G^{r}(A|B)_{\omega}=\int_{-\infty}^{\infty}G^{r}\left[A(t)|B(t^{\prime})\right]e^{i(t-t^{\prime})(\omega+i\eta)}d(t-t^{\prime}). (2)

Here, η\eta is an infinitesimal positive number. By making derivative to time tt and tt^{\prime}, we obtain the EOM as [33]

ωG(A|B)ω\displaystyle\omega G(A|B)_{\omega} =\displaystyle= i{A,B}+iG({A,H}|B)ω\displaystyle i\langle\{A,B\}\rangle+iG(\{A,H\}|B)_{\omega} (3)
=\displaystyle= i{A,B}iG(A|{B,H})ω.\displaystyle i\langle\{A,B\}\rangle-iG(A|\{B,H\})_{\omega}.

Here, the Zubarev GF G(A|B)ωG(A|B)_{\omega} without superscript rr is related to the retarded GF Gr(A|B)ωG^{r}(A|B)_{\omega} by Gr(A|B)ω=G(A|B)ω+iηG^{r}(A|B)_{\omega}=G(A|B)_{\omega+i\eta}.

The average value BA\langle BA\rangle can be obtained from GF through the fluctuation-dissipation theorem [34, 37]

BA=1βρA,B(ω)ω𝑑ω+B0A0.\langle BA\rangle=\frac{1}{\beta}\int_{-\infty}^{\infty}\frac{\rho_{A,B}(\omega)}{\omega}d\omega+\langle B_{0}A_{0}\rangle. (4)

Here, A0A_{0} and B0B_{0} are the zero-frequency components of A(t)A(t) and B(t)B(t), respectively [33]. The spectral function ρA,B(ω)\rho_{A,B}(\omega) in the above equation is defined as

ρA,B(ω)=i2π[G(A|B)ω+iηG(A|B)ωiη].\rho_{A,B}(\omega)=\frac{i}{2\pi}\left[G(A|B)_{\omega+i\eta}-G(A|B)_{\omega-i\eta}\right]. (5)

To solve the EOM of GF by the PTA, we first choose a set of linearly-independent basis variables A1,A2,,ANA_{1},A_{2},...,A_{N}, which span an NN-dimensional subspace of variables. Due to the symmetry of the Hamiltonian of the type H=ipi2/2μi+V(q)H=\sum_{i}p_{i}^{2}/2\mu_{i}+V(q), we can select the basis in one of the subspaces below,

{Oe}={f(q)ipimi|imi=2k,k},\displaystyle\{O_{e}\}=\{f(q)\prod_{i}p_{i}^{m_{i}}\,\,\Big{|}\,\,\sum_{i}m_{i}=2k,\,\,k\in\mathbb{Z}\},
{Oo}={g(q)ipini|ini=2k+1,k}.\displaystyle\{O_{o}\}=\{g(q)\prod_{i}p_{i}^{n_{i}}\,\,\Big{|}\,\,\sum_{i}n_{i}=2k+1,\,\,k\in\mathbb{Z}\}.

Here, f(q)f(q) and g(q)g(q) are arbitrary functions of qq. One can prove that the GF matrix G(A|A)ωG(\vec{A}|\vec{A})_{\omega}, defined for the vector of variables A=(A1,A2,,AN)T\vec{A}=\left(A_{1},A_{2},...,A_{N}\right)^{T}, fulfils the EOM

ω2G(A|A)ω={{A,H},A}G({{A,H},H}|A)ω.\displaystyle\omega^{2}G(\vec{A}\big{|}\vec{A}^{{\dagger}})_{\omega}=-\langle\{\{\vec{A},H\},\vec{A}^{{\dagger}}\}\rangle-G(\{\{\vec{A},H\},H\}\big{|}\vec{A}^{{\dagger}})_{\omega}.

To truncate the EOM in Eq.(II), we introduce the inner product of two classical variables AA and BB as

(A|B){A,{B,H}}.(A|B)\equiv\langle\{A^{\ast},\{B,H\}\}\rangle. (8)

We then truncate the EOM as

{{A,H},H}𝐌T(AA0),\{\{\vec{A},H\},H\}\approx-{\bf M}^{T}\left(\vec{A}-\vec{A}_{0}\right), (9)

Projecting this equation to each basis operator AkA_{k}, one obtains the matrix equation 𝐌=𝐈1𝐋{\bf M}={\bf I}^{-1}{\bf L}. Here 𝐈{\bf I} is the inner product matrix, with element Iij(Ai|Aj)I_{ij}\equiv(A_{i}|A_{j}). 𝐋{\bf L} is the Liouville matrix, with element Lij=(Ai|{{Aj,H},H})L_{ij}=-(A_{i}|\{\{A_{j},H\},H\}). Both are Hermitian and positive semi-definite matrices. 𝐌{\bf M} is thus guaranteed to have real positive eigenvalues. The GF matrix is formally expressed as

G(A|A)ω(ω2𝐌T)1𝐈T.G(\vec{A}\big{|}\vec{A^{{\dagger}}})_{\omega}\approx\left(\omega^{2}-{\bf M}^{T}\right)^{-1}{\bf I}^{T}. (10)

The spectral function obtained from Eqs.(5) and (10) reads

ρAi,Aj(ω)\displaystyle\rho_{A_{i},A_{j}^{\ast}}(\omega)
k(𝐈𝐔)ik(𝐈𝐔)jk2ϵk[δ(ωϵk)δ(ω+ϵk)].\displaystyle\approx\sum_{k}\frac{\left({\bf IU}\right)^{\ast}_{ik}\left({\bf IU}\right)_{jk}}{2\sqrt{\epsilon_{k}}}\left[\delta\left(\omega-\sqrt{\epsilon_{k}}\right)-\delta\left(\omega+\sqrt{\epsilon_{k}}\right)\right].

In this equation, 𝐔{\bf U} and 𝚲=diag(ϵ1,ϵ2,,ϵN){\bf\Lambda}=\text{diag}(\epsilon_{1},\epsilon_{2},...,\epsilon_{N}) are eigenvector and eigenvalue matrices of 𝐌{\bf M}, respectively. That is, 𝐔1𝐌𝐔=𝚲{\bf U}^{-1}{\bf M}{\bf U}={\bf\Lambda}. The eigenvalues ϵk\epsilon_{k}’s are real and ϵk0\epsilon_{k}\geqslant 0 for all kk. They can be obtained by solving the generalized eigen-value problem,

𝐋𝐔=𝐈𝐔𝚲.{\bf LU}={\bf IU\Lambda}. (12)

𝐔{\bf U} can be made to satisfy the generalized unitary condition 𝐔𝐈𝐔=𝟏{\bf U}^{\dagger}{\bf I}{\bf U}={\bf 1}.

From the above expression for the spectral function, Eq.(4) gives the averages CijAiAjC_{ij}\equiv\langle A_{i}^{\ast}A_{j}\rangle as

𝐂1β𝐈𝐋1𝐈+𝐂0.{\bf C}\approx\frac{1}{\beta}{\bf I}{\bf L}^{-1}{\bf I}+{\bf C}_{0}. (13)

Here (C0)ij=Ai0Aj0(C_{0})_{ij}=\langle A_{i0}^{\ast}A_{j0}\rangle is the correlation of the static component of variables. This equation, once being used to calculate the averages appearing in 𝐈{\bf I} and 𝐋{\bf L}, closes the set of non-linear equations for the averages. The PTA solution to the averages can be obtained by self-consistently solving this set of equations. PTA solution to GF and spectral function are then obtained from Eqs.(10) and (II).

III One-Variable Anharmonic Oscillator Model

In this section, the idea of expanding the basis with HH to describe the thermal broadening of quasi-particle peak is applied to a simple classical statistical model, the one-variable AHO model. Its Hamiltonian reads

H=12μp2+V(x),\displaystyle H=\frac{1}{2\mu}p^{2}+V(x),
V(x)=12μω02x2+αx4.\displaystyle V(x)=\frac{1}{2}\mu\omega_{0}^{2}x^{2}+\alpha x^{4}. (14)

This model describes the motion of a single oscillator of mass μ\mu in a quadratic potential with the basic frequency ω0\omega_{0} and an additional quartic potential of strength α\alpha. The spectral function ρx,x(ω)\rho_{x,x}(\omega) of variable xx in the equilibrium state was solved exactly in the canonical ensemble [23]. This is an ideal model for demonstrating the applicability of the PTA on the HH-expanded basis. At finite temperature, the spectral function ρx,x(ω)\rho_{x,x}(\omega) has a peak around the effective oscillating frequency Ωp\Omega_{p} with width Γ\Gamma. The broadening of the spectral peak can be understood as follows. When the nonlinear potential is present, the oscillating frequency of an oscillator depends on its initial energy E0E_{0}. In the equilibrium state of a finite temperature TT, E0E_{0} has a thermal fluctuation of order TT in the canonical ensemble. This leads to an extended distribution of the oscillating frequencies, and subsequently to the broadening of the spectral peak.

III.1 basis B1{x2m1p2n2}B_{1}\equiv\{x^{2m-1}p^{2n-2}\}

Before we present the main subject of this paper, i.e., PTA calculation of the spectral function on the HH-expanded basis, let us first explore the convergence behavior of the spectral function from an ordinary basis, i.e., the basis composed of powers of basic variables xx and pp. The results will be contrasted with the PTA results from the HH-expanded basis.

If we start from the GF Gx,x(ω)G_{x,x}(\omega) and carry out the EOM an even number of times, we will get the higher order GFs Gx2m1p2n2,x(ω)G_{x^{2m-1}p^{2n-2},x}(\omega) (m,n=1,2,m,n=1,2,...). The specific form of the generated variable x2m1p2n2x^{2m-1}p^{2n-2} is due to the parity symmetry of the potential V(x)=V(x)V(-x)=V(x). We therefore define our basis variable as

Amn=x2m1p2n2,(m=1,2,,M;n=1,2,,N)A_{mn}=x^{2m-1}p^{2n-2},\,\,\,\,(m=1,2,...,M;\,n=1,2,...,N) (15)

They form the basis B1={Amn}B_{1}=\{A_{mn}\} (m=1,2,,M;n=1,2,,Nm=1,2,...,M;\,n=1,2,...,N), which span a D=M×ND=M\times N dimensional subspace. Due to the symmetry V(x)=V(x)V(-x)=V(x), the static components of these variables are zero, (Amn)0=0\left(A_{mn}\right)_{0}=0.

Following the standard formalism of PTA summarized in Sec.II, on this basis we have the inner product matrix

Irs,mn\displaystyle I_{rs,mn} \displaystyle\equiv (Ars|Amn)\displaystyle\left(A_{rs}|A_{mn}\right) (16)
=\displaystyle= c1x2r+2m4p2s+2n4\displaystyle c_{1}\langle x^{2r+2m-4}\rangle\langle p^{2s+2n-4}\rangle
+c2x2r+2m3V(x)p2s+2n6\displaystyle+c_{2}\langle x^{2r+2m-3}V^{\prime}(x)\rangle\langle p^{2s+2n-6}\rangle
+c3x2r+2m2V′′(x)p2s+2n6,\displaystyle+c_{3}\langle x^{2r+2m-2}V^{\prime\prime}(x)\rangle\langle p^{2s+2n-6}\rangle,

with c1=[(2m1)/μ][(2r1)(2n1)(2s2)(2m2)]c_{1}=[(2m-1)/\mu]\left[(2r-1)(2n-1)-(2s-2)(2m-2)\right], c2=(2n2)[(2r1)(2n3)(2s2)(2m1)]c_{2}=-(2n-2)\left[(2r-1)(2n-3)-(2s-2)(2m-1)\right], and c3=(2n2)(2s2)c_{3}=(2n-2)(2s-2). The Liouville matrix is obtained as

Lrs,mn\displaystyle L_{rs,mn} =\displaystyle= 1μ2(2m1)(2m2)Irs,m1n+1\displaystyle-\frac{1}{\mu^{2}}(2m-1)(2m-2)\,I_{rs,m-1\,n+1} (17)
+ω02[(2m1)(4n3)+2n2]Irs,mn\displaystyle+\omega_{0}^{2}\left[(2m-1)(4n-3)+2n-2\right]\,I_{rs,mn}
+4αμ[(2m1)(4n3)+3(2n2)]Irs,m+1n\displaystyle+\frac{4\alpha}{\mu}\left[(2m-1)(4n-3)+3(2n-2)\right]\,I_{rs,m+1\,n}
μ2ω04(2n2)(2n3)Irs,m+1n1\displaystyle-\mu^{2}\omega_{0}^{4}(2n-2)(2n-3)\,I_{rs,m+1\,n-1}
16α2(2n2)(2n3)Irs,m+3n1\displaystyle-16\alpha^{2}(2n-2)(2n-3)\,I_{rs,m+3\,n-1}
8αμω02(2n2)(2n3)Irs,m+2n1.\displaystyle-8\alpha\mu\omega_{0}^{2}(2n-2)(2n-3)\,I_{rs,m+2\,n-1}.

The calculation of matrices 𝐈{\bf I} and 𝐋{\bf L} requires evaluating the averages of the type x2k\langle x^{2k}\rangle and p2k\langle p^{2k}\rangle. The latter can be obtained exactly from the Gaussian integral, which gives p2k=(2k1)!!(μT)k\langle p^{2k}\rangle=(2k-1)!!(\mu T)^{k} (k=1,2,k=1,2,...). x2k\langle x^{2k}\rangle needs to be computed self-consistently from GFs. However, since our purpose is to check the convergence behavior of the spectral function on this basis, we use the numerically exact value of x2k\langle x^{2k}\rangle.

From 𝐈{\bf I} and 𝐋{\bf L}, we can calculate the spectral function ρx,x(ω)\rho_{x,x}(\omega) from the following formalism,

ρx,x(ω)=k(𝐈𝐔)𝟏𝐤𝟐2ϵk[δ(ωϵk)δ(ω+ϵk)].\rho_{x,x}(\omega)=\sum_{k}\frac{(\bf{IU})_{1k}^{2}}{2\sqrt{\epsilon_{k}}}\left[\delta(\omega-\sqrt{\epsilon_{k}})-\delta(\omega+\sqrt{\epsilon_{k}})\right]. (18)

In the above equation, the matrix 𝐔\bf{U} and the excitation energy ϵk\epsilon_{k}’s are determined by the generalized eigen-value problem

𝐋𝐔=𝐈𝐔𝚲,\bf{LU}=\bf{IU\Lambda}, (19)

where 𝐔\bf{U} is the generalized eigen vector matrix, and 𝚲=diag(ϵ1,ϵ2,,ϵD){\bf\Lambda}=\text{diag}(\epsilon_{1},\epsilon_{2},...,\epsilon_{D}) is the generalized eigen value matrix.

III.2 HH-expanded basis B2{xeλiH}B_{2}\equiv\{xe^{\lambda_{i}H}\} and B3{x,x3}{eλiH}B_{3}\equiv\{x,x^{3}\}\otimes\{e^{\lambda_{i}H}\}

In this subsection, we implement PTA on two successively larger HH-expanded bases, B2{xeλiH}B_{2}\equiv\{xe^{\lambda_{i}H}\} and B3{x,x3}{eλiH}B_{3}\equiv\{x,x^{3}\}\otimes\{e^{\lambda_{i}H}\} (i=1,2,,Ni=1,2,...,N). Here {λi}\{\lambda_{i}\} are NN different real numbers. We choose the exponential function eλiHe^{\lambda_{i}H} to expand the low order basis variables {x}\{x\} and {x,x3}\{x,x^{3}\}, because with this functional form, the calculation of 𝐈{\bf I} and 𝐋{\bf L} can be simplified significantly. For details, see Appendix A. The PTA results from bases B2B_{2} and B3B_{3} will be compared with the exact ones.

III.2.1 basis B2{xeλiH}B_{2}\equiv\{xe^{\lambda_{i}H}\}

We first derive PTA equations based on the HH-expanded basis B2B_{2} composed of variables

Aixfi,(i=1,2,,N).A_{i}\equiv xf_{i},\,\,\,\,\,(i=1,2,...,N). (20)

Here, fi=eλiHf_{i}=e^{\lambda_{i}H}. To put the physically interesting variable xx into the basis, we assign λ10\lambda_{1}\equiv 0, meaning A1=xA_{1}=x. {Ai}\{A_{i}\} form a NN-dimensional basis, with zero static components Ai0=0A_{i0}=0. The parametrization of {λi}\{\lambda_{i}\} will be specified in Sec.III.D.

The inner product matrix 𝐈\bf{I} and Liouville matrix 𝐋\bf{L} are derived as

Iij=βμ(βλiλj)fifj,\displaystyle I_{ij}=\frac{\beta}{\mu(\beta-\lambda_{i}-\lambda_{j})}\langle f_{i}f_{j}\rangle,
Lij=βμ2(βλiλj)[μω02fifj+12αx2fifj].\displaystyle L_{ij}=\frac{\beta}{\mu^{2}(\beta-\lambda_{i}-\lambda_{j})}\left[\mu\omega_{0}^{2}\langle f_{i}f_{j}\rangle+12\alpha\langle x^{2}f_{i}f_{j}\rangle\right].

Both 𝐈{\bf I} and 𝐋{\bf L} contain the correlation among the expanding factors {eλiH}\{e^{\lambda_{i}H}\} that takes into account the energy fluctuation effect. Details of the derivation are given in Appendix A.

To evaluate the averages appearing in IijI_{ij} and LijL_{ij}, we first note that x2fifj\langle x^{2}f_{i}f_{j}\rangle in LijL_{ij} can be expressed in terms of CijAiAjC_{ij}\equiv\langle A_{i}A_{j}\rangle and thus can be calculated self-consistently from the fluctuation-dissipation theorem Eq.(13) as

Cij\displaystyle C_{ij} =\displaystyle= x2fifj\displaystyle\langle x^{2}f_{i}f_{j}\rangle (22)
=\displaystyle= 1β(𝐈𝐋1𝐈)ij.\displaystyle\frac{1}{\beta}\left({\bf I}{\bf L}^{-1}{\bf I}\right)_{ij}.

Here, 𝐂0=0{\bf C}_{0}=0 has been used. In particular, we have x2=C11\langle x^{2}\rangle=C_{11}. The other unknown average fifj\langle f_{i}f_{j}\rangle cannot be written in the form AiAj\langle A_{i}A_{j}\rangle. For this specific model, it can be evaluated by numerical integration, or by a quadratic variational approximation [38]. We have compared the two methods and find little difference in the final results. Here we use the latter method,

fifj\displaystyle\langle f_{i}f_{j}\rangle \displaystyle\approx fifjvari\displaystyle\langle f_{i}f_{j}\rangle_{vari} (23)
=\displaystyle= 𝑑x𝑑pe(βλiλj)Hvari𝑑x𝑑peβHvari.\displaystyle\frac{\int\int_{-\infty}^{\infty}dxdp\,e^{-(\beta-\lambda_{i}-\lambda_{j})H_{vari}}}{\int\int_{-\infty}^{\infty}dxdp\,e^{-\beta H_{vari}}}.

The variational Hamiltonian HvariH_{vari} is obtained from a quadratic variational approximation as

Hvari=12μp2+θx2,H_{vari}=\frac{1}{2\mu}p^{2}+\theta x^{2}, (24)

with θ=14μω02+(14μω02)2+3αT\theta=\frac{1}{4}\mu\omega_{0}^{2}+\sqrt{\left(\frac{1}{4}\mu\omega_{0}^{2}\right)^{2}+3\alpha T}. Using this variational Hamiltonian, it is easy to obtain the expression for fifj\langle f_{i}f_{j}\rangle as

fifjββλiλj,\langle f_{i}f_{j}\rangle\approx\frac{\beta}{\beta-\lambda_{i}-\lambda_{j}}, (25)

being independent of the parameters in HH.

The spectral function ρx,x(ω)\rho_{x,x}(\omega) is then calculated from Eqs.(18) and (19).

III.2.2 basis B3{x,x3}{eλiH}B_{3}\equiv\{x,x^{3}\}\otimes\{e^{\lambda_{i}H}\}

Here, we derive the 𝐈{\bf I} and 𝐋{\bf L} matrices on basis B3B_{3}. We denote the basis operator as

Aμi=aμfi,(μ=1,2;i=1,2,,N).A_{\mu i}=a_{\mu}f_{i},\,\,\,\,(\mu=1,2;\,\,i=1,2,...,N). (26)

where a1=xa_{1}=x and a2=x3a_{2}=x^{3}, and fi=eλiHf_{i}=e^{\lambda_{i}H} (i=1,2,,Ni=1,2,...,N). These variables also have zero static components (Aμi)0=0\left(A_{\mu i}\right)_{0}=0. The basis B3{aμfi}B_{3}\equiv\{a_{\mu}f_{i}\} (μ=1,2;i=1,2,,N\mu=1,2;\,\,i=1,2,...,N) is a 2N2N dimensional subspace, being larger than B2B_{2}.

We denote the matrix elements of 𝐈{\bf I} and 𝐋{\bf L} as Iμνij=(Aμi|Aνj)I_{\mu\nu}^{ij}=\left(A_{\mu i}|A_{\nu j}\right), and Lμνij=(Aμi|{{Aνj,H},H})L_{\mu\nu}^{ij}=-\left(A_{\mu i}|\{\{A_{\nu j},H\},H\}\right), respectively. Using the same method for deriving 𝐈{\bf I} and 𝐋{\bf L} on the B2B_{2} basis and some additional simplifications, we obtain

𝐈ij=βμ(βλiλj)(fifj3x2fifj3x2fifj9x4fifj).\displaystyle{\bf I}^{ij}=\frac{\beta}{\mu\left(\beta-\lambda_{i}-\lambda_{j}\right)}\left(\begin{array}[]{cc}\langle f_{i}f_{j}\rangle&3\langle x^{2}f_{i}f_{j}\rangle\\ 3\langle x^{2}f_{i}f_{j}\rangle&9\langle x^{4}f_{i}f_{j}\rangle\\ \end{array}\right). (29)

The matrix elements of 𝐋{\bf L} read

L11ij\displaystyle L_{11}^{ij} =\displaystyle= βμ2(βλiλj)[μω02fifj+12αx2fifj],\displaystyle\frac{\beta}{\mu^{2}\left(\beta-\lambda_{i}-\lambda_{j}\right)}\left[\mu\omega_{0}^{2}\langle f_{i}f_{j}\rangle+12\alpha\langle x^{2}f_{i}f_{j}\rangle\right],
L12ij\displaystyle L_{12}^{ij} =\displaystyle= L21ij\displaystyle L_{21}^{ij}
=\displaystyle= 3βμ2(βλiλj)[μω02x2fifj+12αx4fifj],\displaystyle\frac{3\beta}{\mu^{2}\left(\beta-\lambda_{i}-\lambda_{j}\right)}\left[\mu\omega_{0}^{2}\langle x^{2}f_{i}f_{j}\rangle+12\alpha\langle x^{4}f_{i}f_{j}\rangle\right],
L22ij\displaystyle L_{22}^{ij} =\displaystyle= 54βμ2(βλiλj)2x2fifj\displaystyle-\frac{54\beta}{\mu^{2}\left(\beta-\lambda_{i}-\lambda_{j}\right)^{2}}\langle x^{2}f_{i}f_{j}\rangle
+ββλiλj[63ω02μx4fifj+324αμ2x6fifj].\displaystyle+\frac{\beta}{\beta-\lambda_{i}-\lambda_{j}}\left[\frac{63\omega_{0}^{2}}{\mu}\langle x^{4}f_{i}f_{j}\rangle+\frac{324\alpha}{\mu^{2}}\langle x^{6}f_{i}f_{j}\rangle\right].

Details of the derivation are given in Appendix A.

Cμνij=aμaνfifjC_{\mu\nu}^{ij}=\langle a_{\mu}a_{\nu}f_{i}f_{j}\rangle is still given by Eq.(13), with 𝐂0=0{\bf C}_{0}=0. Some unknown averages in 𝐈{\bf I} and 𝐋{\bf L} can be calculated self-consistently through

x2fifj=C11ij,\displaystyle\langle x^{2}f_{i}f_{j}\rangle=C_{11}^{ij},
x4fifj=C12ij,\displaystyle\langle x^{4}f_{i}f_{j}\rangle=C_{12}^{ij},
x6fifj=C22ij.\displaystyle\langle x^{6}f_{i}f_{j}\rangle=C_{22}^{ij}. (32)

The other quantities, including fifj\langle f_{i}f_{j}\rangle, GF, and the spectral function ρx,x(ω)\rho_{x,x}(\omega), can be calculated similarly as for basis B2B_{2}.

III.3 removing the redundant basis variables

In our numerical solution of the above PTA equations, we supplemented a technique that we called the zeros-removing method to stabilize the iteration process. It will be useful in general PTA calculation. Therefore, in this subsection, we introduce this method in the general framework of PTA.

In principle, the general PTA formalism works for arbitrary basis dimension NN. In practice, however, if the PTA calculation is implemented naively, it is limited to some dimension NN by numerical instability. This is because the orthogonality of the basis operators are difficult to control and they easily become linearly dependent when NN is large. As a result, the inner product matrix 𝐈\bf{I}, which should be positive-definite in theory, often becomes nearly singular for large NN. This makes the iterative process unstable. This problem, arising from the non-orthogonality of the basis variables, is quite common in PTA calculations. For example, in our study of Anderson impurity model with the PTA [18], the singularity of 𝐈\bf{I} also appears.

To partly overcome this difficulty, here we propose a method commonly used in the ab initio calculation with unorthogonal basis [39]. The idea of the method is to first figure out the redundant basis variables by diagonalizing 𝐈{\bf I} and find certain smallest eigen values and the associated eigen vectors. Then one removes these vectors from the basis and redoes the PTA calculation on the remaining basis. Our calculation shows that this method can significantly stabilize the iteration and makes it possible to do PTA calculation on a much larger basis. Below, we present the formalism of this method.

Suppose we have a set of basis variables {Ai}\{A_{i}\} (i=1,2,,Ni=1,2,...,N). When 𝐈\bf{I} and 𝐋\bf{L} matrices are obtained, we first diagonalize 𝐈\bf{I} using the unitary transformation,

(𝐔I)1𝐈𝐔I=𝚲I.({\bf U}_{I})^{-1}{\bf I}{\bf U}_{I}={\bf\Lambda}_{I}. (33)

Here 𝐔I{\bf U}_{I} is an unitary matrix and 𝚲I{\bf\Lambda}_{I} is diagonal. Among the NN eigen values of 𝐈{\bf I}, suppose the smallest N1N_{1} of them are close to zero, with their absolute value below a criterion, say, 101210^{-12}. The other N2=NN1N_{2}=N-N_{1} eigenvalues {i1,i2,,iN2}\{i_{1},i_{2},...,i_{N_{2}}\} are larger than this criterion. Then we can approximately write the matrix 𝚲I{\bf\Lambda}_{I} in the following block form,

𝚲I(𝟎𝟎𝟎𝚲I2),{\bf\Lambda}_{I}\approx\left(\begin{array}[]{cc}{\bf 0}&{\bf 0}\\ {\bf 0}&{\bf\Lambda}_{I2}\\ \end{array}\right), (34)

with 𝚲I2=diag(i1,i2,,iN2){\bf\Lambda}_{I2}={\text{diag}}(i_{1},i_{2},...,i_{N_{2}}) and ik>0i_{k}>0 (k=1,2,,N2k=1,2,...,N_{2}). We can also write 𝐔I{\bf U}_{I} as

𝐔I=(𝐔I1𝐔I2),{\bf U}_{I}=\left(\begin{array}[]{cc}{\bf U}_{I1}&{\bf U}_{I2}\end{array}\right), (35)

where (𝐔I1)N×N1\left({\bf U}_{I1}\right)_{N\times N_{1}} contains the N1N_{1} column vectors of the zero-length variables, while (𝐔I2)N×N2\left({\bf U}_{I2}\right)_{N\times N_{2}} contains the N2N_{2} column vectors of finite-length variables. From the orthonormal and complete conditions 𝐔I𝐔I=𝐔I𝐔I=𝟏{\bf U}_{I}^{\dagger}{\bf U}_{I}={\bf U}_{I}{\bf U}_{I}^{\dagger}={\bf 1}, we have the following properties for 𝐔I1{\bf U}_{I1} and 𝐔I2{\bf U}_{I2},

𝐔I1𝐔I1+𝐔I2𝐔I2=𝟏N×N,\displaystyle{\bf U}_{I1}{\bf U}_{I1}^{\dagger}+{\bf U}_{I2}{\bf U}_{I2}^{\dagger}={\bf 1}_{N\times N},
𝐔I1𝐔I1=𝟏N1×N1,𝐔I2𝐔I2=𝟏N2×N2,\displaystyle{\bf U}_{I1}^{\dagger}{\bf U}_{I1}={\bf 1}_{N_{1}\times N_{1}},\,\,\,\,\,\,\,\,{\bf U}_{I2}^{\dagger}{\bf U}_{I2}={\bf 1}_{N_{2}\times N_{2}},
𝐔I1𝐔I2=𝟎N1×N2,𝐔I2𝐔I1=𝟎N2×N1.\displaystyle{\bf U}_{I1}^{\dagger}{\bf U}_{I2}={\bf 0}_{N_{1}\times N_{2}},\,\,\,\,\,\,\,\,{\bf U}_{I2}^{\dagger}{\bf U}_{I1}={\bf 0}_{N_{2}\times N_{1}}. (36)

We normalize the finite-length variables by defining matrix 𝐔~2=𝐔I21𝚲I2\tilde{\bf U}_{2}={\bf U}_{I2}\frac{1}{\sqrt{{\bf\Lambda}_{I2}}}. Now we can write down the orthogonalized variables. The zero-length variables are denoted as O1kO_{1k} and the normalized finite-length variables as O2kO_{2k}. They read

O1k=i=1NAi(UI1)ik,\displaystyle O_{1k}=\sum_{i=1}^{N}A_{i}\left(U_{I1}\right)_{ik},
O2k=i=1NAi(U~2)ik.\displaystyle O_{2k}=\sum_{i=1}^{N}A_{i}\left(\tilde{U}_{2}\right)_{ik}. (37)

Since O1kO_{1k} has zero length, one can prove that the GF involving O1kO_{1k} are zero, i.e., G(O1k|O1p)ω=G(O1k|O2p)ω=0G(O_{1k}|O_{1p})_{\omega}=G(O_{1k}|O_{2p})_{\omega}=0. Its static correlation function is also zero, i.e., O1kX=0\langle O_{1k}X\rangle=0 for arbitrary variable XX. We can therefore focus only on the remaining N2N_{2}-dimensional subspace. The original basis variable AiA_{i} is now expressed in terms of O1kO_{1k} and O2kO_{2k} as

Ai=k=1N1O1k(𝐔I1)ki+k=1N2O2k(𝚲I2𝐔~2)ki.A_{i}=\sum_{k=1}^{N_{1}}O_{1k}\left({\bf U}_{I1}^{\dagger}\right)_{ki}+\sum_{k=1}^{N_{2}}O_{2k}\left({\bf\Lambda}_{I2}\tilde{\bf U}_{2}^{\dagger}\right)_{ki}. (38)

The Liouville matrix on the normalized basis O2kO_{2k} reads

𝐋~22=𝐔~2𝐋𝐔~2.\tilde{\bf L}_{22}=\tilde{\bf U}_{2}^{\dagger}{\bf L}\tilde{\bf U}_{2}. (39)

Here 𝐋~22\tilde{\bf L}_{22} is a N2×N2N_{2}\times N_{2} Hermitian matrix. We diagonalize it by a unitary transformation

[𝐕22]1𝐋~22𝐕22=𝚲2,[{\bf V}_{22}]^{-1}\tilde{\bf L}_{22}{\bf V}_{22}={\bf\Lambda}_{2}, (40)

with 𝚲2=diag(ϵ1,ϵ2,,ϵN2){\bf\Lambda}_{2}=\text{diag}(\epsilon_{1},\epsilon_{2},...,\epsilon_{N_{2}}) being the square of the N2N_{2} eigen excitation energies.

In the N2N_{2}-dimensional space of the normalized basis variables, we employ the standard PTA to write down the GF matrix as

𝐆~22(ω)\displaystyle\tilde{\bf G}_{22}(\omega) =\displaystyle= (ω2𝟏𝐋~22T)1\displaystyle\left(\omega^{2}{\bf 1}-\tilde{\bf L}_{22}^{T}\right)^{-1} (41)
=\displaystyle= (𝐕221)T(ω2𝟏𝚲2)1𝐕22T.\displaystyle\left({\bf V}_{22}^{-1}\right)^{T}\left(\omega^{2}{\bf 1}-{\bf\Lambda}_{2}\right)^{-1}{\bf V}_{22}^{T}.

In the above equation, (G~22)kp(ω)=G(O2k|O2p)ω(\tilde{G}_{22})_{kp}(\omega)=G(O_{2k}|O_{2p}^{\ast})_{\omega}. Using Eq.(38), we obtain the GF of original basis variables AiA_{i} as

G(Ai|Aj)ω=k,p=1N2ikip(UI2)ik(UI2)jp(G~22)kp(ω).G(A_{i}|A_{j}^{\ast})_{\omega}=\sum_{k,p=1}^{N_{2}}\sqrt{i_{k}i_{p}}\left(U_{I2}\right)^{\ast}_{ik}\left(U_{I2}\right)_{jp}(\tilde{G}_{22})_{kp}(\omega). (42)

The spectral function is obtained similarly as

ρ~O2k,O2p(ω)\displaystyle\tilde{\rho}_{O_{2k},O_{2p}^{\ast}}(\omega)
=m=1N2(V22)km(V22)pm2ϵm[δ(ωϵm)δ(ω+ϵm)],\displaystyle=\sum_{m=1}^{N_{2}}\frac{\left(V_{22}\right)^{\ast}_{km}\left(V_{22}\right)_{pm}}{2\sqrt{\epsilon_{m}}}\left[\delta(\omega-\sqrt{\epsilon_{m}})-\delta(\omega+\sqrt{\epsilon_{m}})\right],

and

ρAi,Aj(ω)=k,p=1N2ikip(UI2)ik(UI2)jpρ~O2k,O2p(ω).\rho_{A_{i},A_{j}^{\ast}}(\omega)=\sum_{k,p=1}^{N_{2}}\sqrt{i_{k}i_{p}}\left(U_{I2}\right)^{\ast}_{ik}\left(U_{I2}\right)_{jp}\tilde{\rho}_{O_{2k},O_{2p}^{\ast}}(\omega). (44)

From the spectral theorem, the averages (C~22)kp=O2kO2p(\tilde{C}_{22})_{kp}=\langle O_{2k}^{\ast}O_{2p}\rangle are calculated as

𝐂~22=1β(𝐋~22)1.\tilde{\bf C}_{22}=\frac{1}{\beta}\left(\tilde{\bf L}_{22}\right)^{-1}. (45)

The correlation function between the original basis variables reads

AjAi=k,p=1N2ikip(UI2)ik(UI2)jp[C~22]pk.\langle A_{j}^{\ast}A_{i}\rangle=\sum_{k,p=1}^{N_{2}}\sqrt{i_{k}i_{p}}\left(U_{I2}\right)^{\ast}_{ik}\left(U_{I2}\right)_{jp}\left[\tilde{C}_{22}\right]_{pk}. (46)

The closed equations for AjAi\langle A_{j}^{\ast}A_{i}\rangle are obtained by expressing 𝐈{\bf I} and 𝐋{\bf L} in terms ofAjAi\langle A_{j}^{\ast}A_{i}\rangle. They can be solved iteratively. Once AjAi\langle A_{j}^{\ast}A_{i}\rangle are obtained, the GF and spectral functions can be calculated from Eqs.(42) and (44).

III.4 results for one-variable AHO model

In this subsection, we present the results for the one-variable AHO model, obtained from the basis B1{x2m1p2n2}B_{1}\equiv\{x^{2m-1}p^{2n-2}\} (1mM1\leq m\leq M, 1nN1\leq n\leq N), B2{xeλiH}B_{2}\equiv\{xe^{\lambda_{i}H}\} (1iN1\leq i\leq N), and B3{x,x3}{eλiH}B_{3}\equiv\{x,x^{3}\}\otimes\{e^{\lambda_{i}H}\} (1iN1\leq i\leq N). We compare these results with those from exact numerical integration (for averages) and exact formula [23] (for the spectral function). We want to show that, although the B1B_{1} basis becomes complete in the limit M=N=M=N=\infty, the shape of the primary peak in the spectral function converges very slowly. The B2B_{2} basis quite faithfully produces the shape of the main peak with moderate dimension NN. The B3B_{3} basis produces both the main peak and the secondary peak. These results demonstrate that the HH-expanded basis can efficiently describe the thermal broadening effect in the spectral function. For the data shown below, we typically use the smallness criterion 101410910^{-14}\sim 10^{-9} for truncating the spectrum of 𝐈{\bf I} and the convergence criterion 10910710^{-9}\sim 10^{-7} for iterative solution of AiAj\langle A_{i}^{\ast}A_{j}\rangle. All our results in this paper are in the unit kB=1k_{B}=1.

The HH-expanded bases B2B_{2} and B3B_{3} are characterized by NN real parameters {λi}\{\lambda_{i}\} (i=1,2,,Ni=1,2,...,N). We find that the value of these parameters will influence the convergence speed of PTA calculation and the quality of the spectral function. Since in fifjTr[e(βλiλj)]\langle f_{i}f_{j}\rangle\sim\text{Tr}[e^{-(\beta-\lambda_{i}-\lambda_{j})}], λi+λj<β\lambda_{i}+\lambda_{j}<\beta is required, we let λi(,β/2)\lambda_{i}\in(-\infty,\beta/2). The lower and higher λi\lambda_{i}’s are responsible for describing low and high energy thermal fluctuations, respectively. We use the following scheme for assigning them,

λi={12(1T1Ti)(2iN),0(i=1).\lambda_{i}=\left\{\begin{array}[]{lll}\frac{1}{2}\left(\frac{1}{T}-\frac{1}{T_{i}}\right)&\,\,\,(2\leq i\leq N),\\ \\ 0&\,\,\,(i=1).\end{array}\right. (47)

Here, we set λ1=0\lambda_{1}=0 to include xx in the basis. We let lnTi\ln{T_{i}} distribute equal-distantly in the interval [lnTΔ,lnT+Δ][\ln{T}-\Delta,\ln{T}+\Delta]. That is, for N>1N>1, we use

lnTi=lnTΔ+(i1)2ΔN1,(2iN).\ln{T_{i}}=\ln{T}-\Delta+(i-1)\frac{2\Delta}{N-1},\,\,\,\,\,\,\,\,(2\leq i\leq N). (48)

Typically Δ=4.09.0\Delta=4.0\sim 9.0 is used in our calculation.

Our test shows that the above scheme works quite well for the AHO model. The peaks in the spectral function ρx,x(ω)\rho_{x,x}(\omega) shift slightly with Δ\Delta in the range Δ=1.010.0\Delta=1.0\sim 10.0. The key features extracted from ρx,x(ω)\rho_{x,x}(\omega), such as the peak position Ωp\Omega_{p}, peak width Γ\Gamma, and the moments of the spectral function, change very little. As Δ\Delta decreases, Γ\Gamma stays qualitatively the same until Δ106\Delta\sim 10^{-6}, at which point Γ\Gamma jumps to zero, recovering the expected result for Δ=0\Delta=0. For tiny but finite Δ\Delta, ρx,x(ω)\rho_{x,x}(\omega) has two delta peaks, since the basis is effectively reduced to a two-dimensional basis {x,xH}\{x,xH\} in the small Δ\Delta limit. The same holds for the 1D ϕ4\phi^{4} lattice model. When we need to average the spectral function ρx,x(ω)\rho_{x,x}(\omega) over NrN_{r} random assignments of {λi}\{\lambda_{i}\}, we let lnTi\ln{T_{i}} take random numbers from the uniform distribution within [lnTΔ,lnT+Δ][\ln{T}-\Delta,\ln{T}+\Delta].

In this work, when presenting the spectral function, we always broaden the delta peaks in the spectral function by a Gaussian function with standard variance σ\sigma. Depending on the occasion, the value of σ\sigma is either fixed to be 0.20.050.2\sim 0.05, or it is taken to be proportional to the width of the main peak Γ\Gamma, σ=rΓ\sigma=r\Gamma with r=0.20.3r=0.2\sim 0.3.

Considering that the spectral function in the present study always has a single primary peak in the ω>0\omega>0 regime, and it is almost symmetric in shape, we extract the peak position Ωp\Omega_{p} of the primary peak through

Ωp=m1,\Omega_{p}=m_{1}, (49)

and the width Γ\Gamma of the primary peak through

Γ=m2m12.\Gamma=\sqrt{m_{2}-m_{1}^{2}}. (50)

Here, mkm_{k} is the kk-th order moment of the spectral function on the positive frequency axis. This method was used in, e.g., defining the spin wave line width [40]. mkm_{k} is evaluated from the weight wiw_{i} and pole pip_{i} of the raw data of spectral function ρx,x(ω)=iwiδ(ωpi)\rho_{x,x}(\omega)=\sum_{i}w_{i}\delta(\omega-p_{i}) as mk=(i,pi>0wipik)/(i,pi>0wi)m_{k}=(\sum_{i,p_{i}>0}w_{i}p_{i}^{k})/(\sum_{i,p_{i}>0}w_{i}). In this way, both Ωp\Omega_{p} and Γ\Gamma are independent of the broadening procedure used for plotting the curve of ρx,x(ω)\rho_{x,x}(\omega). Note that in the literature, there are various ways for extracting the width of phonon peak from the MD simulation data [25, 26, 41, 42].

Refer to caption
Figure 1: (color online) x2\langle x^{2}\rangle as functions of TT for AHO model. Solid lines are obtained from B1B_{1}-PTA with M=1M=1, 22, 33, and 66 from bottom to top (N=MN=M is always used). The dashed line is for the exact solution. Inset: The errors with respect to the exact results as functions of MM (N=MN=M) in x2\langle x^{2}\rangle (red squares), x4\langle x^{4}\rangle (green dots), and x6\langle x^{6}\rangle (blue triangles) at T=0.3T=0.3 . The dashed line is M2M^{-2} for guiding eyes. Model parameters are μ=1.0\mu=1.0, ω0=0.3\omega_{0}=0.3, and α=0.25\alpha=0.25.
Refer to caption
Figure 2: (color online) Comparison of spectral function ρx,x(ω)\rho_{x,x}(\omega) obtained from B1B_{1}-PTA (red solid line) with the exact curve (black dashed line). Model parameters are μ=1.0\mu=1.0, ω0=0.3\omega_{0}=0.3, α=0.25\alpha=0.25, and T=0.3T=0.3. The delta peaks in the PTA curves are broadened using Gaussian function with standard variance σ=0.02\sigma=0.02.

Below, we first examine the convergence behavior of the basis B1B_{1}. In the limit M=N=M=N=\infty, B1B_{1} becomes complete and the results from PTA are expected to become exact. In Fig.1, we show the average value x2\langle x^{2}\rangle as a functions of TT for various MM (fixing N=MN=M). It clearly shows that as M=NM=N increases, the curve uniformly approaches the exact dashed line. The inset of Fig.1 shows that the errors of x2\langle x^{2}\rangle, x4\langle x^{4}\rangle, and x6\langle x^{6}\rangle at T=0.3T=0.3 decrease with increasing MM as M2M^{-2} (fixing N=MN=M). It is also observed that at about M=N=10M=N=10, i.e., at the dimension of basis D=M×N=100D=M\times N=100, the error from PTA abruptly increases. We have observed that the largest (smallest) eigenvalue of 𝐈{\bf I} increases (decreases) exponentially with increasing M=NM=N. At about M=N=10M=N=10, the condition number of 𝐈{\bf I} deteriorates to such an extent that the iterative solution of PTA equations can no longer be carried out accurately. In our previous application of PTA [18], we also observed similar phenomena, hinting that this problem may be universal for PTA. Note that this problem can not be solved by the use of zeros-removing method discussed in Sec.III.C.

In Fig.2, the spectral function ρx,x(ω)\rho_{x,x}(\omega) of AHO model at T=0.3T=0.3 is shown. In each panel, the red curve (from PTA on basis B1B_{1}) is compared to the black dashed line (from exact solution [23]). The exact ρx,x(ω)\rho_{x,x}(\omega) has a series of broad peaks at successively higher frequencies, whose weights decay exponentially with frequency. These multiple peaks are the overtone phenomenon due to the non-linear confinement, broadened by thermal fluctuations. At very low temperature, the peak positions tend to frequencies ω0\omega_{0}, 3ω03\omega_{0}, 5ω05\omega_{0}, etc.. The exact spectral function vanishes abruptly for ω<ω0\omega<\omega_{0}, in agreement with the fact that ω0\omega_{0} is the lowest oscillation frequency for all initial conditions.

It is seen that as M=NM=N increases up to M=N=8M=N=8, the PTA-produced spectral function contains more and more peaks (we broaden the delta peaks a little for presentation purpose). The peak positions and weights are distributed in such a way that the envelope curve uniformly approaches the exact one, including the primary peak and the higher order ones. Fig.1 and Fig.2 show that both the thermodynamical and the dynamical quantities indeed converge towards the exact value as the dimension of basis increases.

Refer to caption
Figure 3: (color online) Primary peak in the spectral function of AHO model, obtained from basis B1={x2m1p2n2}B_{1}=\{x^{2m-1}p^{2n-2}\} (m=1,2,,M,n=1,2,,Nm=1,2,...,M,\,\,n=1,2,...,N) with M=N=4M=N=4 (green solid line) and M=N=8M=N=8 (red dash-dotted line). The black dashed line is the exact curve. Inset: the peak position (black squares) and peak width (read circles) obtained using different MM (N=MN=M). The dashed lines are exact value. Model parameters are μ=1.0\mu=1.0, ω0=0.3\omega_{0}=0.3, α=0.25\alpha=0.25, and T=0.3T=0.3. For the main figure, we used a constant broadening parameter σ=0.02\sigma=0.02.

However, if we look at the primary peak alone in Fig.3 on the linear scale, we will find that with increasing M=NM=N, the primary peak of ρx,x(ω)\rho_{x,x}(\omega) from PTA receives little improvement. Up to M=N=8M=N=8, the primary peak has got only 44 discernible peaks on the linear scale. Clearly, on the naive basis B1B_{1}, the poles of GF distribute uniformly in the whole frequency axis. Among the M×NM\times N delta peaks, most of them have high frequencies and tiny weights, such that they contribute little to the broad primary peak. This leads to a very slow convergence in the shape of the primary peak with increasing basis dimension, although the extracted peak position Ωp\Omega_{p} and width Γ\Gamma converge fast with increasing M=NM=N, as shown in the inset of Fig.3.

Refer to caption
Figure 4: (color online) Comparison of spectral function ρx,x(ω)\rho_{x,x}(\omega) of AHO model, obtained from the HH-expanded basis B2={xeλiH}B_{2}=\{xe^{\lambda_{i}H}\} (i=1,2,,Ni=1,2,...,N) (red solid line) and from the exact solution (black dashed line). Model parameters are T=0.3T=0.3, μ=1.0\mu=1.0, ω0=0.3\omega_{0}=0.3, and α=0.25\alpha=0.25. We use Δ=6.0\Delta=6.0 and broaden parameter σ=0.02\sigma=0.02.
Refer to caption
Figure 5: (color online) Spectral function ρx,x(ω)\rho_{x,x}(\omega) of AHO model obtained by averaging the results of 500500 random realization of {λi}\{\lambda_{i}\} on basis B2={xeλiH}B_{2}=\{xe^{\lambda_{i}H}\} (i=1,2,,Ni=1,2,...,N) (solid lines). The broadening scheme σ=rΓ\sigma=r\Gamma is used with r=0.3r=0.3. The exact curve is dashed line. Inset: relative peak position Ωpω0\Omega_{p}-\omega_{0} (squares) and the peak width Γ\Gamma (circles) as functions of TT from PTA. The solid lines are extracted from exact spectral function. Dashed lines mark the power law T1/4T^{1/4} behavior. Basis dimension N=20N=20. Model parameters are μ=1.0\mu=1.0, ω0=0.3\omega_{0}=0.3, and α=0.25\alpha=0.25. Δ=6.0\Delta=6.0 is used.

Fig.4 shows the spectral function ρx,x(ω)\rho_{x,x}(\omega) obtained from PTA on the HH-expanded basis B2B_{2} (red solid lines), together with the exact curve (black dashed lines). They are shown in log-linear axes. In contrast to the case of basis B1B_{1} (shown in Fig.3), here as the dimension of basis increases, in the frequency regime of the primary peak, one gets significantly more delta peaks. The envelope tends to the exact broad primary peak much faster than that of B1B_{1} basis. A basis with dimension N=20N=20 already produces about 1010 visible peaks densely located within the frequency regime of primary peak, leading to qualitatively accurate description of the primary peak, if each delta peak of PTA is properly broadened with a Gaussian function. However, B2B_{2} basis also meets its limitation at about N=30N=30, beyond which the spectral function can hardly receive any improvement. This is due to the deterioration of the condition number of 𝐈{\bf I} matrix as the basis dimension increases.

In Fig.5, we compare the spectral function ρx,x(ω)\rho_{x,x}(\omega) averaged over 500500 random λi\lambda_{i} realizations (solid lines) with the exact curve (dashed lines) for different temperatures. Only the primary peak is shown in this plot with a linear-linear axis. The agreement is quite good in a wide temperature range. The exact ρx,x(ω)\rho_{x,x}(\omega) curve has a natural frequency cut off at ω=ω0\omega=\omega_{0}, below which there is no spectral weight. This feature is nicely captured by PTA on basis B2B_{2} with averaging on 500500 random {λi}\{\lambda_{i}\}’s (not shown in Fig.5). The inset shows that the relative peak position Ωpω0\Omega_{p}-\omega_{0} and the peak width Γ\Gamma from Eqs.(49) and (50) are in qualitative agreement with the exact one from low- to high-temperature limits. The quantitative deviation in Γ\Gamma becomes apparent only in the high TT regime, with PTA value lower by a factor 0.770.77. Ωpω0\Omega_{p}-\omega_{0} and Γ\Gamma have the same power law of TT, being T1T^{1} and T1/4T^{1/4} in the low and high temperature limits, respectively.

These power laws can be understood as follows. The peak position Ωp\Omega_{p} is qualitatively described by the N=1N=1 PTA expression

ωp=ω0+(12α/μ)x2.\omega_{p}=\sqrt{\omega_{0}+(12\alpha/\mu)\langle x^{2}\rangle}. (51)

In the low temperature limit, the system tends to be harmonic, x2Tω0\langle x^{2}\rangle\sim T\ll\omega_{0} due to equipartition theorem. In the high temperature limit, the particle’s movement is dominated by the x4x^{4} potential. Exact analysis gives x2T1/2ω0\langle x^{2}\rangle\sim T^{1/2}\gg\omega_{0} [33]. Equation (51) then gives the asymptotic behavior of Ωp(T)\Omega_{p}(T) observed in the inset of Fig.5,

Ωp(T){ω0+cT1,(TTcr)cT14,(TTcr).\Omega_{p}(T)\sim\left\{\begin{array}[]{lll}\omega_{0}+c\,T^{1},&\,\,\,(T\ll T_{cr})\\ \\ c^{\prime}\,T^{\frac{1}{4}},&\,\,\,(T\gg T_{cr}).\end{array}\right. (52)

Here, cc and cc^{\prime} are TT-independent coefficients. The crossover temperature Tcr1.0T_{cr}\approx 1.0.

The spectral width Γ(T)\Gamma(T) is generated by the thermal fluctuation of oscillating frequency which, due to the nonlinear potential, comes from the thermal fluctuation of the initial energy E0E_{0} of the system. Guided by Eq.(51), we assume that the particle oscillates with a frequency ωω02+12αx02\omega\sim\sqrt{\omega_{0}^{2}+12\alpha x_{0}^{2}}, with x0x_{0} being the initial coordinate. x0x_{0} depends on E0E_{0} through E012μω0x02E_{0}\sim\frac{1}{2}\mu\omega_{0}x_{0}^{2} at low TT, and E0αx04E_{0}\sim\alpha x_{0}^{4} at high TT. Inserting them into the expression for ω\omega and considering E0δE0TE_{0}\sim\delta E_{0}\sim T due to thermal fluctuation, we obtain the fluctuation of frequency δωδE0T\delta\omega\sim\delta E_{0}\sim T at low TT, and δωE03/4δE0T1/4\delta\omega\sim E_{0}^{-3/4}\delta E_{0}\sim T^{1/4}. This explains the observed asymptotic behavior of Γ(T)\Gamma(T) in the inset of Fig.5,

Γ(T){dT1,(TTcr)dT14,(TTcr).\Gamma(T)\sim\left\{\begin{array}[]{lll}d\,T^{1},&\,\,\,(T\ll T_{cr})\\ \\ d^{\prime}\,T^{\frac{1}{4}},&\,\,\,(T\gg T_{cr}).\end{array}\right. (53)
Refer to caption
Figure 6: (color online) Spectral function ρx,x(ω)\rho_{x,x}(\omega) of AHO model from basis B3={x,x3}{eλiH}B_{3}=\{x,x^{3}\}\otimes\{e^{\lambda_{i}H}\} (i=1,2,,Ni=1,2,...,N) (red solid line), broadened with σ=0.02\sigma=0.02. The exact curve is the black dashed line. Model parameters are μ=1.0\mu=1.0, ω0=0.3\omega_{0}=0.3, α=0.25\alpha=0.25, and T=0.3T=0.3. Δ=6.0\Delta=6.0 is used.

If we further enlarge the basis B2B_{2} by adding {x3eλiH}\{x^{3}e^{\lambda_{i}H}\}, we obtain basis B3B_{3}. Fig.6 shows the spectral function ρx,x(ω)\rho_{x,x}(\omega) from PTA on basis B3B_{3}. The exact curve is shown as the black dashed line. The basis dimension of B3B_{3} is D=2ND=2N. At N=1N=1, the two delta peaks are located at the central positions of the primary and the secondary peaks of the exact curve. With increasing NN, PTA produces more and more delta peaks in the frequency regime of the primary and the secondary peaks, forming two broadened curves. Combining the results for B2B_{2} and B3B_{3} bases, we conclude that the HH-expanding of basis can lift the degeneracy of the quasi-particle excitation energies obtained from a lower order PTA, and split the corresponding delta peak into a bunch of densely located peaks, which effectively describe a thermally broadened peak.

IV k\mathcal{H}_{k}-expanded basis for the 1D nonlinear ϕ4\phi^{4} lattice model

In this section, we apply the HH-expanded basis formalism to the classical 1D nonlinear ϕ4\phi^{4} lattice model, and we calculate the phonon spectral function ρQk,Qk\rho_{Q_{k},Q_{k}^{\ast}}, with a focus on the thermal broadening effect of the phonon peak at T>0T>0. Here, Qk=1/LjeijkxjQ_{k}=1/\sqrt{L}\sum_{j}e^{-ijk}x_{j} is the Fourier component of coordinates of particles. The Hamiltonian of the ϕ4\phi^{4} model reads

H=i=1L[pi22m+K2(xixi1)2+γ4xi4],H=\sum_{i=1}^{L}\Big{[}\frac{p_{i}^{2}}{2m}+\frac{K}{2}(x_{i}-x_{i-1})^{2}+\frac{\gamma}{4}x_{i}^{4}\Big{]}, (54)

Here, LL is the total number of particles, xix_{i} represents the deviation of the ii-th particle from its equilibrium position, KK is the nearest-neighbor coupling strength, and γ\gamma is the coefficient of the on-site potential. We use periodic boundary condition x0=xLx_{0}=x_{L}. This model is obtained by discretizing the classical ϕ4\phi^{4} field [43]. It has been widely used in the study of 1D heat transport [44, 45, 46, 47, 48, 26, 27, 49]. Similar to the FPU-β\beta model [50], the ϕ4\phi^{4} model has a scaling property which makes it sufficient to study only the parameter point K=γ=1K=\gamma=1 [33]. Below, we will focus on this parameter point.

For this model, the previous PTA study employs the basis QkQ_{k} [33]. To describe the thermal fluctuation effect, it is tempting to expand it into the basis AikQkeλiHA^{k}_{i}\equiv Q_{k}e^{\lambda_{i}H} (i=1,2,,Ni=1,2,...,N), similar to the case of one-variable AHO model. However, we find that this scheme has two problems. First, the peak width Γk\Gamma_{k} in the phonon dispersion function ρQk,Qk(ω)\rho_{Q_{k},Q_{k}^{\ast}}(\omega) obtained from this basis is proportional to 1/L1/\sqrt{L}, which vanishes in the large size limit LL\rightarrow\infty. This is unphysical. The reason is that, unlike in the one-variable AHO model, the thermal broadening of the phonon peak of ρQk,Qk(ω)\rho_{Q_{k},Q_{k}^{\ast}}(\omega) for a long chain of atoms is no longer contributed by the fluctuation of total energy HH, but mainly by the thermal fluctuation of energy k\mathcal{H}_{k} of mode kk. Indeed, in the microscopic canonical ensemble where the total energy HH is fixed, the HH-expanded basis becomes identical to the single variable QkQ_{k} and hence cannot produce a broadened phonon peak, while the MD simulation in the microscopic canonical ensemble can correctly produce the phonon broadened effect.

In fact, in the canonical ensemble in which our formalism is written, the thermal fluctuation δH\delta H of the total energy HH, when distributed equally to each mode, gives a magnitude δH/L\delta H/L to the fluctuations of each mode HkH_{k}. Due to the nonlinear coupling, it further gives an uncertainty of the frequency δωkδH/L\delta\omega_{k}\sim\sqrt{\delta H/L}. This explains the 1/L1/\sqrt{L} dependence of the incorrect results obtained from the basis QkeλiHQ_{k}e^{\lambda_{i}H}.

The second problem of the basis {QkeλiH}\{Q_{k}e^{\lambda_{i}H}\} is that the complex variable QkQ_{k} becomes real at momenta k=0k=0 and k=πk=\pi, which have distinct dynamical and statistical properties from the complex variables at k0,πk\neq 0,\pi. As a result, special care must be taken in PTA formalism to avoid the possible discontinuity in physical results at k=0k=0 and π\pi. This will complicate the formalism.

Due to the considerations above, in the following, we will work on the real basis ReQk\text{Re}Q_{k} and ImQk\text{Im}Q_{k}, and we expand the basis using the effective local energy k\mathcal{H}_{k} of the mode kk, instead of the total Hamiltonian HH. Although k\mathcal{H}_{k} is no longer a strictly conserving quantity like HH, it is approximately conserving on the level of QkQ_{k}-PTA.

IV.1 {Qk}\{Q_{k}\}-PTA reformulated on a real basis

Since we will construct the k\mathcal{H}_{k}-expanded real basis, in this subsection, we first derive the PTA formalism on the real basis {ReQk,ImQk}\{\text{Re}{Q_{k}},\text{Im}{Q_{k}}\}. The obtained equations are equivalent to those of PTA on the complex basis {Qk}\{Q_{k}\} discussed in Ref.[33]. Then we expand it with eλike^{\lambda_{i}\mathcal{H}_{k}}, where k\mathcal{H}_{k} is the effective energy of the real mode kk (see below). In this part, we reformulate this PTA in terms of the real basis, because we will need the results of this lower order PTA to identify the approximate eigen-exciatation variable in order to introduce the proper mode energy k\mathcal{H}_{k} and to expand the former with the latter.

We first define the following dynamical variables,

Qk1Ljxjeijk,\displaystyle Q_{k}\equiv\frac{1}{\sqrt{L}}\sum_{j}x_{j}e^{-ijk},
Pk1Ljpjeijk,\displaystyle P_{k}\equiv\frac{1}{\sqrt{L}}\sum_{j}p_{j}e^{ijk},
Rk1Ljxj3eijk.\displaystyle R_{k}\equiv\frac{1}{\sqrt{L}}\sum_{j}x^{3}_{j}e^{-ijk}. (55)

The lattice momentum kk takes LL different values k=(2π/L)nkk=(2\pi/L)n_{k} (nk=0,1,,L1n_{k}=0,1,...,L-1). Here and below, we assume that LL is an even number so that k=πk=\pi is possible. We take the real and imaginary part of QkQ_{k} and PkP_{k} as our dynamical variables. Considering the independence of variables, we will confine the momentum kk to [0,π][0,\pi], and we take the following definition for the dynamical variables

Q1kαkReQk,k[0,π],\displaystyle Q_{1k}\equiv\alpha_{k}\text{Re}{Q_{k}},\,\,\,\,\,\,\,k\in[0,\pi],
Q2kαkImQk,k(0,π),\displaystyle Q_{2k}\equiv-\alpha_{k}\text{Im}{Q_{k}},\,\,\,\,\,\,\,k\in(0,\pi),
P1kαkRePk,k[0,π],\displaystyle P_{1k}\equiv\alpha_{k}\text{Re}{P_{k}},\,\,\,\,\,\,\,k\in[0,\pi],
P2kαkImPk,k(0,π).\displaystyle P_{2k}\equiv\alpha_{k}\text{Im}{P_{k}},\,\,\,\,\,\,\,k\in(0,\pi). (56)

In the above equations, the factor αk\alpha_{k} is for normalizing the variables. αk=1\alpha_{k}=1 for k=0k=0 and π\pi, and αk=2\alpha_{k}=\sqrt{2} for k0k\neq 0 or π\pi. Note that since Qk=0Q_{k=0} and Qk=πQ_{k=\pi} are real variables, for Q1kQ_{1k} and P1kP_{1k}, kk takes L/2+1L/2+1 values, k=(2π/L)nkk=(2\pi/L)n_{k} (nk=0,1,,L/2n_{k}=0,1,...,L/2). For Q2kQ_{2k} and P2kP_{2k}, kk takes L/21L/2-1 values, k=(2π/L)nkk=(2\pi/L)n_{k} (nk=1,,L/21n_{k}=1,...,L/2-1). Altogether we have LL independent real variables.

One can prove that the above-defined variables QνkQ_{\nu k} and PνkP_{\nu k} (ν=1,2\nu=1,2) fulfill the properties of canonical variables. That is, {Qνk,Qμp}={Pνk,Pμp}=0\{Q_{\nu k},Q_{\mu p}\}=\{P_{\nu k},P_{\mu p}\}=0 and {Qνk,Pμp}=δνμδkp\{Q_{\nu k},P_{\mu p}\}=\delta_{\nu\mu}\delta_{kp}. They are a set of complete and independent real dynamical variables in place of {xi,pi}\{x_{i},p_{i}\} or {Qk,Pk}\{Q_{k},P_{k}\}.

The PTA formalism on the basis QνkQ_{\nu k} (k[0,π]k\in[0,\pi] for ν=1\nu=1, and k(0,π)k\in(0,\pi) for ν=2\nu=2) can be obtained from the previous PTA equations on basis QkQ_{k} [33]. We obtain the inner product matrix Iνk,μpI_{\nu k,\mu p} and the Liouville matrix Lνk,μpL_{\nu k,\mu p} as

Iνk,μp=1mδνμδkp,\displaystyle I_{\nu k,\mu p}=\frac{1}{m}\delta_{\nu\mu}\delta_{kp},
Lνk,μp=1mωk2δνμδkp.\displaystyle L_{\nu k,\mu p}=\frac{1}{m}\omega_{k}^{2}\delta_{\nu\mu}\delta_{kp}. (57)

The phonon excitation energy ωk\omega_{k} is given by

ωk=1m[2K(1cosk)+3γx2].\omega_{k}=\sqrt{\frac{1}{m}\left[2K\left(1-\cos{k}\right)+3\gamma\langle x^{2}\rangle\right]}. (58)

From these expressions, one obtains the spectral functions ρQνk,Qμp(ω)\rho_{Q_{\nu k},Q_{\mu p}}(\omega) as

ρQνk,Qμp(ω)=δνμδkp12mωk[δ(ωωk)δ(ω+ωk)].\rho_{Q_{\nu k},Q_{\mu p}}(\omega)=\delta_{\nu\mu}\delta_{kp}\frac{1}{2m\omega_{k}}\left[\delta(\omega-\omega_{k})-\delta(\omega+\omega_{k})\right]. (59)

The PTA self-consistent equation for x2\langle x^{2}\rangle is obtained from the spectral theorem Eq.(4) as

x2\displaystyle\langle x^{2}\rangle =\displaystyle= 1Lν=12kνQνkν2\displaystyle\frac{1}{L}\sum_{\nu=1}^{2}\sum_{k_{\nu}}\langle Q_{\nu k_{\nu}}^{2}\rangle (60)
=\displaystyle= 1Lk[0,2π)1βmωk2.\displaystyle\frac{1}{L}\sum_{k\in[0,2\pi)}\frac{1}{\beta m\omega_{k}^{2}}.

In the last line above, we have used the fact that Q1k2=Q2k2=Q1 2πk2\langle Q_{1k}^{2}\rangle=\langle Q_{2k}^{2}\rangle=\langle Q_{1\,2\pi-k}^{2}\rangle for k(0,π)k\in(0,\pi). The final expression is same as that of the {Qk}\{Q_{k}\} basis PTA.

IV.2 effective Hamiltonian νk\mathcal{H}_{\nu k} of the mode (νk)(\nu k)

In order to expand the basis QνkQ_{\nu k} using the mode energy, we first have to construct an effective energy νk\mathcal{H}_{\nu k} for mode (νk)(\nu k). We will do this construction approximately based on the real basis {Q1k,Q2k}\{Q_{1k},Q_{2k}\} PTA equations discussed above. The effective energy νk\mathcal{H}_{\nu k} is constructed by the following principles. First, it is a real variable with the unit of energy; second, it generates the expected dynamics, i.e., {Qνk,νk}{Qνk,H}\{Q_{\nu k},\mathcal{H}_{\nu k}\}\approx\{Q_{\nu k},H\} and {Pνk,νk}{Pνk,H}\{P_{\nu k},\mathcal{H}_{\nu k}\}\approx\{P_{\nu k},H\}; and third, it is a zero-th order conserving quantity, i.e., {νk,H}0\{\mathcal{H}_{\nu k},H\}\approx 0. The third principle is required because the broadening of the phonon peak will be regarded in our theory as splitting the degeneracy of the excitation energy ωk\omega_{k} by high order fluctuations beyond the zero-th order approximation, as discussed in the case of the one-variable AHO model. In this work, the two last principles are fulfilled at the level of PTA on the basis {Q1k,Q2k}\{Q_{1k},Q_{2k}\}.

From Eq.(IV.1) of {Q1k,Q2k}\{Q_{1k},Q_{2k}\}-PTA and the general principle of PTA, we have the approximate EOM as

{Qνk,H}=1mPνk,\displaystyle\{Q_{\nu k},H\}=\frac{1}{m}P_{\nu k},
{Pνk,H}mωk2Qνk.(ν=1,2)\displaystyle\{P_{\nu k},H\}\approx-m\omega_{k}^{2}\,Q_{\nu k}.\,\,\,\,\,\,\,(\nu=1,2) (61)

It is then easy to find the two eigen excitation variables associated to mode (νk)(\nu k) by diagonalizing the EOM matrix. We obtain

aνk(1)=ck(mωk2QνkiωkPνk),\displaystyle a_{\nu k}^{(1)}=c_{k}\left(m\omega_{k}^{2}\,Q_{\nu k}-i\omega_{k}P_{\nu k}\right),
aνk(2)=c~k(mωk2Qνk+iωkPνk).\displaystyle a_{\nu k}^{(2)}=\tilde{c}_{k}\left(m\omega_{k}^{2}\,Q_{\nu k}+i\omega_{k}P_{\nu k}\right). (62)

They obey {aνk(1),H}iωkaνk(1)\{a_{\nu k}^{(1)},H\}\approx i\omega_{k}\,a_{\nu k}^{(1)}, and {aνk(2),H}iωkaνk(1)\{a_{\nu k}^{(2)},H\}\approx-i\omega_{k}\,a_{\nu k}^{(1)}, respectively.

One can construct the approximately conserving quantity νk\mathcal{H}_{\nu k} as νk=fkaνk(1)aνk(1)\mathcal{H}_{\nu k}=f_{k}a_{\nu k}^{(1)\ast}a_{\nu k}^{(1)}. It satisfies the approximate conservation {νk,H}0\{\mathcal{H}_{\nu k},H\}\approx 0. Fixing the coefficients ckc_{k} and c~k\tilde{c}_{k} from the second principle, i.e., {Qνk,νk}{Qνk,H}\{Q_{\nu k},\mathcal{H}_{\nu k}\}\approx\{Q_{\nu k},H\}, we obtain the effective Hamiltonian of mode (νk)(\nu k) as

νk=12mωk2Qνk2+12mPνk2,\mathcal{H}_{\nu k}=\frac{1}{2}m\omega_{k}^{2}\,Q_{\nu k}^{2}+\frac{1}{2m}P_{\nu k}^{2}, (63)

which has the desired properties

{νk,H}0,\displaystyle\{\mathcal{H}_{\nu k},H\}\approx 0,
{Qνk,νk}{Qνk,H}\displaystyle\{Q_{\nu k},\mathcal{H}_{\nu k}\}\approx\{Q_{\nu k},H\}
{Pνk,νk}{Pνk,H}.\displaystyle\{P_{\nu k},\mathcal{H}_{\nu k}\}\approx\{P_{\nu k},H\}. (64)

IV.3 PTA on νk\mathcal{H}_{\nu k}-expanded basis

Following the same practice of the PTA on the HH-expanded basis for the one-variable AHO model, we construct the expanded basis for the 1D ϕ4\phi^{4} lattice model in the subspace (νk)(\nu k) as

Aiνk=Qνkeλiνk.(λireal,i=1,2,,N)A^{\nu k}_{i}=Q_{\nu k}e^{\lambda_{i}\mathcal{H}_{\nu k}}.\,\,\,\,\,(\lambda_{i}\,\,\,\text{real},\,\,\,\,i=1,2,...,N) (65)

Here, (νk)=(1,k1)(\nu k)=(1,k_{1}) and (2,k2)(2,k_{2}), with k1[0,π]k_{1}\in[0,\pi] and k2(0,π)k_{2}\in(0,\pi), respectively. We assign λ1=0\lambda_{1}=0 to make sure A1νk=QνkA^{\nu k}_{1}=Q_{\nu k}.

The inner product matrix 𝐈νk{\bf I}^{\nu k} and the Liouville matrix 𝐋νk{\bf L}^{\nu k} are derived from straightforward calculation, along the line for the AHO model. In the derivation, we have used the properties Eq.(IV.2) of νk\mathcal{H}_{\nu k} . We obtain

Iijνk\displaystyle I^{\nu k}_{ij} =\displaystyle= 1mββλiλje(λi+λj)νk,\displaystyle\frac{1}{m}\frac{\beta}{\beta-\lambda_{i}-\lambda_{j}}\langle e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\nu k}}\rangle,
Lijνk\displaystyle L^{\nu k}_{ij} =\displaystyle= 1m2K(1cosω)Iijνk\displaystyle\frac{1}{m}2K(1-\cos{\omega})I^{\nu k}_{ij} (66)
+1m2ββλiλj3γx2e(λi+λj)νk.\displaystyle+\frac{1}{m^{2}}\frac{\beta}{\beta-\lambda_{i}-\lambda_{j}}3\gamma\langle x^{2}e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\nu k}}\rangle.

In the above equations, like in the study of AHO model, we evaluate e(λi+λj)νk\langle e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\nu k}}\rangle using the quadratic variational approximation

e(λi+λj)νk\displaystyle\langle e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\nu k}}\rangle \displaystyle\approx e(λi+λj)νkvari\displaystyle\langle e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\nu k}}\rangle_{vari} (67)
=\displaystyle= ββλiλj,\displaystyle\frac{\beta}{\beta-\lambda_{i}-\lambda_{j}},

with the variational Hamiltonian

Hvari=kHk,\displaystyle H_{vari}=\sum_{k}H_{k},
Hk=12mPkPk+12mωk2QkQk.\displaystyle H_{k}=\frac{1}{2m}P_{k}^{\ast}P_{k}+\frac{1}{2}m\omega_{k}^{2}Q_{k}^{\ast}Q_{k}. (68)

To calculate x2e(λi+λj)νk\langle x^{2}e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\nu k}}\rangle self-consistently, we write it as

x2e(λi+λj)νk\displaystyle\langle x^{2}e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\nu k}}\rangle (69)
=\displaystyle= 1LμpQμp2e(λi+λj)νk\displaystyle\frac{1}{L}\sum_{\mu p}\langle Q_{\mu p}^{2}e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\nu k}}\rangle
=\displaystyle= 1LμpQμp2e(λi+λj)μpe(λi+λj)νke(λi+λj)μp\displaystyle\frac{1}{L}\sum_{\mu p}\langle Q_{\mu p}^{2}e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\mu p}}\frac{e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\nu k}}}{e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\mu p}}}\rangle
\displaystyle\approx 1Lμp[Qμp2e(λi+λj)μpe(λi+λj)νke(λi+λj)μp]\displaystyle\frac{1}{L}\sum_{\mu p}\left[\langle Q_{\mu p}^{2}e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\mu p}}\rangle\frac{\langle e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\nu k}}\rangle}{\langle e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\mu p}}\rangle}\right]
\displaystyle\approx 1LμpAiμpAjμp.\displaystyle\frac{1}{L}\sum_{\mu p}\langle A^{\mu p}_{i}A^{\mu p}_{j}\rangle.

In the above approximation, the momentum dependence of x2e(λi+λj)νk\langle x^{2}e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\nu k}}\rangle is neglected. This makes the interaction part of the 𝐋\bf{L} matrix kk-independent, reminiscent of the local self-energy in the dynamical mean-field theory [51, 52]. Note that the additional approximation used in Eq.(69) exaggerates the coupling between different modes and tends to overestimate the broadening. See Fig.11. The final expression can be evaluated from 𝐈νk{\bf I}^{\nu k} and 𝐋νk{\bf L}^{\nu k} via

AiνkAjνk1β[𝐈νk(𝐋νk)1𝐈νk]ij.\langle A^{\nu k}_{i}A^{\nu k}_{j}\rangle\approx\frac{1}{\beta}\left[{\bf I}^{\nu k}\left({\bf L}^{\nu k}\right)^{-1}{\bf I}^{\nu k}\right]_{ij}. (70)

From the self-consistent solution of the above equations, one can finally obtain the physical quantities of interest, ı.e., the spectral function ρQk,Qk(ω)\rho_{Q_{k},Q_{k}^{\ast}}(\omega). It is expressed by the spectral functions of real variables ρQ1k,Q1k(ω)\rho_{Q_{1k},Q_{1k}}(\omega) as

ρQk,Qk(ω)={ρQ1k,Q1k(ω),k[0,π]ρQ1 2πk,Q1 2πk(ω),k(π,2π).\rho_{Q_{k},Q_{k}^{\ast}}(\omega)=\left\{\begin{array}[]{lll}\rho_{Q_{1k},Q_{1k}}(\omega),&\,\,\,k\in[0,\pi]\\ \\ \rho_{Q_{1\,2\pi-k},Q_{1\,2\pi-k}}(\omega),&\,\,\,k\in(\pi,2\pi).\end{array}\right. (71)

The spectral function ρQ1k,Q1k(ω)\rho_{Q_{1k},Q_{1k}}(\omega) is then calculated from the standard PTA formalism Eqs.(II) and (12). Note that in the case of N=1N=1, the νk\mathcal{H}_{\nu k}-expanded basis recovers the basis {Qνk}\{Q_{\nu k}\}, and the excitation energy is given by the ωk\omega_{k} in Eq.(58).

In the practical calculation, the above equations are combined with the zeros-removing method described in section III.C.

IV.4 spectral function ρQk,Qk(ω)\rho_{Q_{k},Q_{k}^{\ast}}(\omega) for 1D ϕ4\phi^{4} lattice model

Refer to caption
Figure 7: (color online) (a) Phonon dispersion Ωk\Omega_{k} curves at T=5.0T=5.0 from PTA on the νk\mathcal{H}_{\nu k}-expanded basis, for various dimension NN’s (solid lines). The dashed line (denoted as LH) is the lower bound harmonic variational result [53]. (b) Relative error of x2\langle x^{2}\rangle from PTA with respect to that from cluster variational method (CVM) [56] versus basis dimension NN, for different temperatures. (c) Spectral width Γk\Gamma_{k} as functions of temperature. Filled and empty symbols with guiding lines are data from PTA and MD, respectively. The dashed lines mark the power law. Model parameters are m=1.0m=1.0, K=1.0K=1.0, γ=1.0\gamma=1.0, and L=1000L=1000. L=4000L=4000 is used for T102T\leq 10^{-2} in panel (c).
Refer to caption
Figure 8: (color online) Spectral function ρQk,Qk(ω)\rho_{Q_{k},Q_{k}^{\ast}}(\omega) at various kk for 1D ϕ4\phi^{4} lattice model at T=1.0T=1.0. In both panels, from left to right, k=0k=0, 0.1π0.1\pi, 0.2π0.2\pi, 0.35π0.35\pi, 0.5π0.5\pi, 0.65π0.65\pi, 0.8π0.8\pi, and π\pi. (a) linear scale, with broadening parameter r=0.3r=0.3; (b) log scale, with r=0.01r=0.01, obtained by averaging results of 200200 random {λi}\{\lambda_{i}\}’s. Inset of panel (b) shows the cut-off frequency ωc(k)\omega_{c}(k) extracted from PTA (squares) and Eq.(76) (line). Model parameters are m=1.0m=1.0, K=1.0K=1.0, γ=1.0\gamma=1.0, and L=1000L=1000. Δ=7.0\Delta=7.0 for k0k\neq 0 and Δ=12.0\Delta=12.0 for k=0k=0.

Here, we present the numerical results obtained from the above formula for 1D ϕ4\phi^{4} lattice model. We use the scheme Eqs.(47) and (48) for assigning {λi}\{\lambda_{i}\} values in our calculation.

First, in Fig.7, we benchmark PTA on the νk\mathcal{H}_{\nu k}-expanded basis by checking the convergence of its results with increasing basis dimension NN, taking other results as references. In Fig.7(a), we plot the dispersion Ωk\Omega_{k} curve obtained from successively larger bases. The curve from the lower bound harmonic variational approximation (LH) [53] (dashed line) is shown for comparison. It is seen that as NN increases, the long-wavelength part of Ωk\Omega_{k} shifts towards the LH curve which is expected to be more accurate, but the short-wavelength part shifts only slightly. The Ωk\Omega_{k} curve converges at about N=10N=10. Since our basis is not complete in the limit N=N=\infty, we do not expect the dispersion to converge to the exact curve. We see that the PTA with N=1N=1 already produces quite accurate Ωk\Omega_{k} and for this quantity only limited improvement can be gained by enlarging NN.

Fig.7(b) shows the relative error in x2\langle x^{2}\rangle calculated from PTA, taking the value from a two-site cluster variational method (CVM) [54, 55] as a reference. The CVM was originally designed for Ising-like lattice models with discrete degrees of freedom and was applied to the Klein-Gordan lattice model with continuous degrees of freedom [56]. It is known that the two-site CVM produces exact static averages for 1D models with nearest-neighbour interaction [57]. So here we take the CVM results as a reference. As NN increases from N=1N=1, the error of PTA first decreases and then saturates at about N=10N=10. The saturated error increases with increasing temperature. This shows both the effect and the limitation of the νk\mathcal{H}_{\nu k}-expanded basis.

In Fig.7(c), the spectral widths Γk=0(T)\Gamma_{k=0}(T) and Γk=0.8π(T)\Gamma_{k=0.8\pi}(T) extracted from ρQk,Qk(ω)\rho_{Q_{k},Q_{k}^{\ast}}(\omega) of PTA are compared with those from MD simulations. In MD simulations, Γk(T)\Gamma_{k}(T) is extracted from the power spectrum using the half-area method [41]. For this purpose, we first define the complex normal variables

ak=iP~k+Q~k,\displaystyle a_{k}=i\tilde{P}_{k}+\tilde{Q}_{k}, (72)

where

P~k\displaystyle\tilde{P}_{k} =2Lj=1Lpjsin(jk),\displaystyle=\sqrt{\frac{2}{L}}\,\sum_{j=1}^{L}p_{j}\sin(jk), (73)
Q~k\displaystyle\tilde{Q}_{k} =2Lj=1Lxjsin(jk).\displaystyle=\sqrt{\frac{2}{L}}\,\sum_{j=1}^{L}x_{j}\sin(jk). (74)

are momentum and coordinate variables. We calculate |a^k(ω)|2|\hat{a}_{k}(\omega)|^{2}, the Fourier transform of the complex normal variable ak(t)a_{k}(t) in Eq. (72). Γk\Gamma_{k} is then defined as the width of the frequency window around the center ω~k\tilde{\omega}_{k} of the power spectrum |a^k(ω)|2|\hat{a}_{k}(\omega)|^{2}, in which the power is equal to half of the total power. That is,

ω~kΓk/2ω~k+Γk/2|a^k(ω)|2𝑑ω0|a^k(ω)|2𝑑ω=12.\displaystyle\frac{\displaystyle\int_{\tilde{\omega}_{k}-\Gamma_{k}/2}^{\tilde{\omega}_{k}+\Gamma_{k}/2}|\hat{a}_{k}(\omega)|^{2}\,d\omega}{\displaystyle\int_{0}^{\infty}|\hat{a}_{k}(\omega)|^{2}\,d\omega}=\frac{1}{2}. (75)

Here, ω~k\tilde{\omega}_{k} is the center frequency of the peak in |a^k(ω)|2|\hat{a}_{k}(\omega)|^{2}. In the case of large fluctuations, this definition gives a much more robust result in the numerical simulation than the standard method does [41].

In Fig.7(c), for both k=0k=0 and k=0.8πk=0.8\pi, we see a qualitative agreement between PTA and MD results. Both curves for k=0k=0 (squares) and k=0.8πk=0.8\pi (up triangles) show qualitative agreement in all temperature regime. In particular, the asymptotic power law behaviors, i.e., Ωk(T)T1/4\Omega_{k}(T\rightarrow\infty)\sim T^{1/4}, Ωk=0(T0)T1/3\Omega_{k=0}(T\rightarrow 0)\sim T^{1/3}, and Ωk0(T0)T2/3\Omega_{k\neq 0}(T\rightarrow 0)\sim T^{2/3} are observed in both results. The source of these asymptotic power laws will be analysed in Fig.12. Part of the quantitative deviation between PTA and MD results may be traced back to the truncation error and additional errors introduced in Eq.(69). We find that the present PTA on the HνkH_{\nu k}-expanded basis slightly underestimates Γk\Gamma_{k} in the high-temperature regime, and overestimate it by a factor of 5.05.0 in the low-TT regime. This is similar to the case of the AHO model, as shown in the inset of Fig.5. Another source of the discrepancy may lie in the different definition and extraction method of Γk\Gamma_{k}. Other spectral width data from MD simulation in the literatures [25, 27, 42] are also compared with our results. We find qualitative agreement in both the kk and TT dependences. See Fig.11.

In Fig.8, we show the evolution of spectral function ρQk,Qk(ω)\rho_{Q_{k},Q_{k}^{\ast}}(\omega) of 1D ϕ4\phi^{4} lattice model with momentum kk at T=1.0T=1.0, obtained from νk\mathcal{H}_{\nu k}-expanded basis PTA and averaging over 200200 random realization of {λi}\{\lambda_{i}\}’s. Panels (a) and (b) show the curve on the linear and log y axis, respectively. As kk increases, the peak position increases and the width of the peak shrinks. The heights of the peaks do not change much, since the total weight of ρQk,Qk(ω)\rho_{Q_{k},Q_{k}^{\ast}}(\omega) in ω>0\omega>0 is not conserved. In (b), the spectral function shows a feature similar to that of AHO model: a low frequency cut-off at frequency ωc(k)\omega_{c}(k). In AHO model, ρx,x(ω)\rho_{x,x}(\omega) has a low frequency cut-off at ωc=ω0\omega_{c}=\omega_{0}, which is well described by the HH-expanded PTA with random averaging. The similarity between PTA equations of the AHO model, Eqs.(III.2.1), (22), and (25), and those of the 1D ϕ4\phi^{4} lattice model, Eqs.(IV.3), (67), (69), and (70) can explain the low frequency cut-off in the ϕ4\phi^{4} lattice model and gives the cut-off frequency

ωc(k)=(2K/m)(1cosk).\omega_{c}(k)=\sqrt{(2K/m)(1-\cos{k})}. (76)

The inset of Fig.8(b) compares the observed cut off frequency (symbols) with this equation (line) and perfect agreement is obtained.

Refer to caption
Figure 9: (color online) Spectral function ρQk,Qk(ω)\rho_{Q_{k},Q_{k}^{\ast}}(\omega) at k=0k=0 for 1D ϕ4\phi^{4} lattice model at different temperatures. Curves are obtained from averaging the results of 400400 random {λi}\{\lambda_{i}\}’s with Δ=7.0\Delta=7.0. Broadening parameter r=0.3r=0.3. Model parameters are m=1.0m=1.0, K=1.0K=1.0, γ=1.0\gamma=1.0, and L=1000L=1000.

Fig.9 shows the evolution of the phonon spectral function ρQk=0,Qk=0(ω)\rho_{Q_{k=0},Q_{k=0}^{\ast}}(\omega) with temperature. We used basis dimension N=20N=20. For a given TT, the spectral function curve has a single, almost symmetric peak in the positive frequency regime, located around ω=Ωk=0\omega=\Omega_{k=0}, with width Γk=0\Gamma_{k=0}. Both Ωk=0\Omega_{k=0} and Γk=0\Gamma_{k=0} increases with increasing temperature. We will see from Fig.12 that both increase as T1/3T^{1/3} in the low-temperature regime, while they increase as T1/4T^{1/4} in the high-temperature limit. Note that the frequency with the largest ρQk,Qk(ω)\rho_{Q_{k},Q_{k}^{\ast}}(\omega) is numerically very close to the Ωk\Omega_{k} calculated from Eq.(49) from one-shot raw ρQk,Qk(ω)\rho_{Q_{k}^{\ast},Q_{k}}(\omega) data. Therefore, in this paper, we always use the latter definition for the dispersion.

Comparing Ωk\Omega_{k} to the dispersion ωk\omega_{k} Eq.(58) obtained from a single QkQ_{k}-basis calculation (equivalent to the νk\mathcal{H}_{\nu k}-expanded basis at N=1N=1 as well as to the quadratic variation [33]), we find the same qualitative behavior and small quantitative deviation. We conclude that the phonon dispersion from N=1N=1 PTA is already quite accurate and quantitative improvement of Ωk\Omega_{k} with increasing NN is limited.

Refer to caption
Figure 10: (color online) (a) Dispersion Ωk\Omega_{k} and (b) the spectral width Γk\Gamma_{k} for 1D ϕ4\phi^{4} lattice model. They are extracted from the phonon spectral function ρQ1k,Q1k(ω)\rho_{Q_{1k},Q_{1k}}(\omega) from PTA on HνkH_{\nu k}-expanded basis with N=20N=20. Temperatures are marked out in the figure. Model parameters are m=1.0m=1.0, K=1.0K=1.0, γ=1.0\gamma=1.0. We use L=1000L=1000 for T=0.01T=0.01 and L=100L=100 for other temperatures.
Refer to caption
Figure 11: (color online) Comparison of (a) τk\tau_{k}, and (b) the mean free path lkl_{k} at T=1.0T=1.0 obtained from PTA and various MD simulations in Refs. [27, 25, 42]. Model parameters are m=1.0m=1.0, K=1.0K=1.0, γ=1.0\gamma=1.0, and L=1000L=1000. PTA results are multiplied by a factor 5.05.0 in the figure.

Fig.10 shows the dispersion Ωk\Omega_{k} and width Γk\Gamma_{k}, extracted from ρQk,Qk(ω)\rho_{Q_{k},Q_{k}^{\ast}}(\omega) using Eqs.(49) and (50), respectively. We present them as functions of kk for various temperatures. The dispersion Ωk\Omega_{k} shown in Fig.10(a) is quantitatively close to Eq.(58), as checked in Fig.7(a). The gap at k=0k=0 increases with temperature in a two-section power-law fashion, as summarized in Fig.12(a). The spectral width shown in Fig.10(b) is a decreasing function of kk for all TT, in agreement with the MD result [42]. The long wave length phonons that are important to the thermal transport are strongly damped by thermal fluctuations. This leads to a finite thermal conductivity in the thermodynamic limit at T>0T>0, in contrast to the lattice models with continuous spatial translational invariance. Note that the data in Fig.10 converge with respect to increasing chain length LL. To obtain reliable data for the thermodynamic limit, we used a chain length up to L=4000L=4000, which is larger than the correlation length of the lowest studied temperature.

In Fig.11, we make quantitative comparisons between the results from the PTA and various MD simulations. Fig.11(a) shows the phonon life time τk1/(2Γk)\tau_{k}\equiv 1/(2\Gamma_{k}) and Fig.11(b) shows the mean free path lkτkvkl_{k}\equiv\tau_{k}v_{k} at T=1.0T=1.0. Note that the extraction of τk\tau_{k} and lkl_{k} from MD simulation is a nontrivial task and there are various ways of doing it [25, 27, 41, 42]. This leads to quantitative deviations among different MD results. The PTA results for τk\tau_{k} and lkl_{k} are smaller than those of MD by a factor of 55 or so. When multiplied by a factor 55, PTA result has qualitatively the same kk-dependence with the MD curves. We believe that, apart from the difference in the definition of Γk\Gamma_{k}, the approximation in Eq.(69) is a major source for the underestimation of τk\tau_{k} and lkl_{k} in PTA. In Eq.(69), the correlation between Qμp2Q_{\mu p}^{2} and e(λi+λj)νke^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\nu k}} for modes (μp)(νk)(\mu p)\neq(\nu k) is replaced with those of the same mode. This tends to exaggerate x2e(λi+λj)νk\langle x^{2}e^{(\lambda_{i}+\lambda_{j})\mathcal{H}_{\nu k}}\rangle and hence overestimates the broadening Γk\Gamma_{k} at low temperatures (see Fig.7(c)). Further improvement of our theory in this aspect is in progress.

From the benchmark comparison with MD in Fig.7 and Fig.11, we pinpoint certain aspects where the present method of PTA can compete with and has some advantages over MD. For the low temperature limit where the equilibrium state needs to be built in a much longer time, MD is time-consuming. Both the low and high frequency limits of the dynamical correlation function are resource-demanding for MD. The methods of extracting relevant physical quantities are yet to be unified in MD, leading to quantitative deviations among different MD studies, as shown in Fig.11. In contrast, PTA produces results for well-defined quantities within the same order of time for arbitrary temperature and frequency with well-controlled precision.

Refer to caption
Figure 12: (color online) (a) Dispersions Ωk\Omega_{k} and (b) the spectral widths Γk\Gamma_{k} as functions of TT for various momenta. They are extracted from ρQk,Qk(ω)\rho_{Q_{k},Q_{k}^{*}}(\omega) of 1D ϕ4\phi^{4} lattice model, obtained from PTA on HνkH_{\nu k}-expanded basis with N=20N=20. The low and high temperature asymptotic power laws are marked in the figure. The stars are results for AHO model. Inset: ratio of Γk(T)\Gamma_{k}(T) and δΩk(T)=Ωk(T)(2K/m)(1cosk)\delta\Omega_{k}(T)=\Omega_{k}(T)-\sqrt{(2K/m)(1-\cos{k})} as functions of TT for the same kk values as the main figure. Model parameters are m=1.0m=1.0, K=1.0K=1.0, and γ=1.0\gamma=1.0. We used L=100L=100 (for high TT) \sim 40004000 (for low TT) which are much larger than the correlation length at the corresponding temperature.

Having presented the benchmarking results, below we focus on new physical results obtained by PTA for the 1D ϕ4\phi^{4} lattice model in Figs.12, 13, and 14. Fig.12 shows the temperature dependence of Ωk\Omega_{k} and Γk\Gamma_{k}, for a series kk from 0 to π\pi. The dispersion Ωk(T)\Omega_{k}(T) shown in Fig.12(a) is a three-section power law curve of temperature except at k=0k=0. In the high temperature regime TTcr1.0T\gg T_{\text{cr}}\sim 1.0, Ωk(T)T1/4\Omega_{k}(T)\sim T^{1/4} and the curves for all momentum merge into the curve Ω(T)\Omega(T) of the AHO model (stars in the figure). This shows that TTcrT\gg T_{\text{cr}} is a local-oscillator regime where oscillations are dominated by local x4x^{4} potential and are dispersionless. As temperature decreases below TcrT_{cr}, Ωk(T)T1/3\Omega_{k}(T)\sim T^{1/3}, before it finally enter a plateau at very low temperature TTkT\ll T_{k}. The crossover temperature TkT_{k} decreases when kk decreases. In particular, Tk=0=0T_{k=0}=0 and the intermediate temperature power law Ωk=0(T)T1/3\Omega_{k=0}(T)\sim T^{1/3} is thus maintained to T=0T=0.

Fig.12(b) shows the spectral width Γk(T)\Gamma_{k}(T). It is also an increasing function with three-section power-law behavior, except at k=0k=0. Being different from Ωk\Omega_{k}, Γk\Gamma_{k} decreases with increasing kk. It tends to zero at T=0T=0 for all momentum. At high temperatures TTcr1.0T\gg T_{cr}\sim 1.0, Γk(T)T1/4\Gamma_{k}(T)\sim T^{1/4} for all momenta and they merge into the Γ(T)\Gamma(T) curve of AHO model (stars). As temperature lowers, Γk(T)T1/3\Gamma_{k}(T)\sim T^{1/3} in a temperature window TkTTcrT_{k}\ll T\ll T_{cr}. The TkT_{k} here is the same kk-dependent crossover temperature in the Ωk(T)\Omega_{k}(T) curve. Finally, when TTkT\ll T_{k}, we have the third section with power law Γk(T)T2/3\Gamma_{k}(T)\sim T^{2/3}. For k=0k=0, Tk=0=0T_{k=0}=0. So Γk=0(T)T1/3\Gamma_{k=0}(T)\sim T^{1/3} is maintained to the low temperature limit. As shown in Fig.7(c), these power laws are confirmed by the MD results.

The inset of Fig.12(b) shows the ratio Γk(T)/δΩk(T)\Gamma_{k}(T)/\delta\Omega_{k}(T) as functions of TT for the same kk’s as in the main figure. Here δΩk(T)Ωk(T)Ωk(T=0)\delta\Omega_{k}(T)\equiv\Omega_{k}(T)-\Omega_{k}(T=0), with Ωk(T=0)=(2K/m)(1cosk)\Omega_{k}(T=0)=\sqrt{(2K/m)(1-\cos{k})} being the dispersion at zero temperature. It is seen that the power law TT-dependence in the numerator and denominator cancels, giving a ratio that is weakly dependent on TT and kk. This is consistent with the asymptotic expressions for Ωk(T)\Omega_{k}(T) and Γk(T)\Gamma_{k}(T) in Eqs.(77) and (78).

The peculiar temperature dependence of Ωk(T)\Omega_{k}(T) and Γk(T)\Gamma_{k}(T) can be understood in terms of the quadratic variation result for ωk\omega_{k} in Eq.(58), by a similar analysis as for the AHO model. It is well known [43, 33, 56] that for ϕ4\phi^{4} model, x2T2/3\langle x^{2}\rangle\sim T^{2/3} for TTcrT\ll T_{cr} and x2T1/2\langle x^{2}\rangle\sim T^{1/2} for TTcrT\gg T_{cr}. TcrT_{cr} depends on the ratio between KK and γ\gamma, and Tcr1.0T_{cr}\sim 1.0 in our calculation. Putting this into Eq.(58), one obtains the asymptotic TT dependence of dispersion as

Ωk(T){Ωk(T=0)+cT23,(TTk)cT13,(TkTTcr)c′′T14,(TTcr).\Omega_{k}(T)\sim\left\{\begin{array}[]{lllll}\Omega_{k}(T=0)+c\,T^{\frac{2}{3}},&\,\,\,(T\ll T_{k})\\ \\ c^{\prime}\,T^{\frac{1}{3}},&\,\,\,(T_{k}\ll T\ll T_{cr})\\ \\ c^{\prime\prime}\,T^{\frac{1}{4}},&\,\,\,(T\gg T_{cr}).\end{array}\right. (77)

Here, TkT_{k} is obtained by equating 2K(1cosk)2K(1-\cos{k}) with 3γx23\gamma\langle x^{2}\rangle in Eq.(58), which gives Tk[(K/γ)(1cosk)]3/2T_{k}\sim\left[(K/\gamma)(1-\cos{k})\right]^{3/2}. It tends to zero in the limit k=0k=0. Therefore, we call the three regimes of temperature the local oscillator regime (TTcrT\gg T_{cr}), the pure harmonic regime (TTkT\ll T_{k}), where anharmonic interaction does not appear in the dispersion Ωkωk(γ=0)\Omega_{k}\approx\omega_{k}(\gamma=0), and the renormalized harmonic regime TkTTcrT_{k}\ll T\ll T_{cr}, where the oscillators still form collective waves but with renormalized dispersion since 3γx2ωk3\gamma\langle x^{2}\rangle\gg\omega_{k} in Eq.(58).

For the thermal broadening Γk(T)\Gamma_{k}(T), we invoke the expression ωk(E0)1m[2K(1cosk)+3γx02]\omega_{k}(E_{0})\sim\sqrt{\frac{1}{m}\left[2K(1-\cos{k})+3\gamma x_{0}^{2}\right]}, according to Eq.(58). Here E0E_{0} is the initial energy of the mode kk, which has a fluctuation magnitude of TT. In the low-temperature limit, x02x2T2/3E02/3x_{0}^{2}\sim\langle x^{2}\rangle\sim T^{2/3}\sim E_{0}^{2/3}. In the high-temperature limit, the energy is dominated by the x4x^{4} local potential and we have x04E0x_{0}^{4}\sim E_{0}. Inputting these relations into ωk(E0)\omega_{k}(E_{0}) and taking E0δE0TE_{0}\sim\delta E_{0}\sim T, we obtain the expression

Γk(T){dT23,(TTk)dT13,(TkTTcr)d′′T14,(TTcr).\Gamma_{k}(T)\sim\left\{\begin{array}[]{lllll}d\,T^{\frac{2}{3}},&\,\,\,(T\ll T_{k})\\ \\ d^{\prime}\,T^{\frac{1}{3}},&\,\,\,(T_{k}\ll T\ll T_{cr})\\ \\ d^{\prime\prime}\,T^{\frac{1}{4}},&\,\,\,(T\gg T_{cr}).\end{array}\right. (78)

Here, we have assumed TkTcrT_{k}\ll T_{cr}. Eqs.(77) and (78) summarize the observed results in Fig.12.

Eqs.(77) and (78) suggests a relation between the width Γk(T)\Gamma_{k}(T) and dispersion Ωk(T)\Omega_{k}(T),

Γk(T)c[Ωk(T)Ωk(0)].\Gamma_{k}(T)\approx c\left[\Omega_{k}(T)-\Omega_{k}(0)\right]. (79)

Here, cc is weakly dependent on TT and kk. See the inset of Fig.12(b) for its behavior. This relation is in contrast to the assumption of the effective phonon theory [24] that Γk(T)ϵ(T)Ωk(T)\Gamma_{k}(T)\approx\epsilon(T)\Omega_{k}(T). Here, ϵ(T)=En/En+El\epsilon(T)=\langle E_{n}\rangle/\langle E_{n}+E_{l}\rangle, with EnE_{n} and ElE_{l} being the nonlinear and linear potential energies, respectively. The results Γk(T)\Gamma_{k}(T) and Ωk(T)\Omega_{k}(T) shown in Fig.12 do not agree with this assumption both in the momentum and in the temperature dependences.

Refer to caption
Figure 13: (color online) x2\langle x^{2}\rangle as functions of temperature for various chain length LL (symbols with guiding lines). The exact curve for L=2L=2 is shown as stars. The low temperature asymptotic powers are marked in the figure. The arrows mark out the temperature at which the correlation length ξ(T)\xi(T) equals to LL. The dashed line shows T2/3T^{2/3} power law. The inset shows the correlation length ξ\xi as a function of TT obtained by cluster variation method [56]. We use basis dimension N=20N=20 and parameter Δ=7.0\Delta=7.0. Model parameters are m=1.0m=1.0, K=1.0K=1.0, and γ=1.0\gamma=1.0.
Refer to caption
Figure 14: (color online) Dispersion Ωk(T)\Omega_{k}(T) and broadening Γk(T)\Gamma_{k}(T) as functions of temperature for various chain length LL. (a) Ωk=0\Omega_{k=0}, (b) Γk=0\Gamma_{k=0}, (c) Ωk=π(T)\Omega_{k=\pi}(T) (empty symbols) and δΩk=π(T)=Ωk=π(T)Ωk=π(0)\delta\Omega_{k=\pi}(T)=\Omega_{k=\pi}(T)-\Omega_{k=\pi}(0) (solid symbols), and (d) Γk=π(T)\Gamma_{k=\pi}(T). In each panel, from top to bottom, L=2L=2, 2020, 200200, and 20002000. Lines are for guiding eyes. We use basis dimension N=20N=20 and parameter Δ=7.0\Delta=7.0. Model parameters are m=1.0m=1.0, K=1.0K=1.0, and γ=1.0\gamma=1.0.

IV.5 short-chain limit of 1D ϕ4\phi^{4} lattice model

So far, we have focused solely on the thermodynamical limit where the chain length LL is much larger than the correlation length ξ(T)\xi(T). Our results presented above are obtained with sufficiently large LL to avoid any finite-size effect. Here, we report that the ϕ4\phi^{4} model has also a short-chain limit characterized by Lξ(T)L\ll\xi(T), where physical quantities have distinct power laws from those in the thermodynamical limit. This regime is pertinent to the thermal transport experiments with small size low-dimensional systems, but to our knowledge, has not been discussed in the literature.

In Fig.13, x2(T)\langle x^{2}\rangle(T) is plotted for chain size ranging from L=2L=2 to L=2000L=2000. For each LL, in the high temperature limit TTcr1.0T\gg T_{cr}\approx 1.0, we have x2(T)T1/2\langle x^{2}\rangle(T)\sim T^{1/2}, the power law of the atomic limit discussed in previous subsections. As TT decreases below TcrT_{cr}, x2T2/3\langle x^{2}\rangle\sim T^{2/3}. Here TcrT_{cr} is the LL-independent local-nonlocal crossover temperature. Since the correlation length ξ\xi increases with decreasing TT as ξT1/3\xi\sim T^{-1/3} in the low temperature limit [56] (see the inset of Fig.13), when TT decreases further and is below a short-chain crossover temperature TshT_{sh} where ξ=L\xi=L (marked by the arrows in the figure), ξ>L\xi>L occurs and x2(T)\langle x^{2}\rangle(T) begins to deviate from the T2/3T^{2/3} law. In the low TT limit where Lξ(T)L\ll\xi(T), x2T1/2\langle x^{2}\rangle\sim T^{1/2}, being different from the T2/3T^{2/3} behavior of the thermodynamical limit. We have checked that the dependence of TshT_{sh} on LL is determined by ξ(Tsh)=L\xi(T_{sh})=L, giving Tsh(L)L3T_{sh}(L)\sim L^{-3}, with a coefficient about 1.01.0. In the thermodynamical limit, Tsh(L=)=0T_{sh}(L=\infty)=0 and the T2/3T^{2/3} behavior extends to zero temperature. Note that Tsh(L)Tcr1.0T_{sh}(L)\leq T_{cr}\approx 1.0 holds for its largest value L=2L=2. This behavior is summarized as

x2(T){T12,(TTsh(L))T23,(Tsh(L)TTcr)T12,(TTcr).\langle x^{2}\rangle(T)\sim\left\{\begin{array}[]{lllll}T^{\frac{1}{2}},&\,\,\,(T\ll T_{sh}(L))\\ \\ T^{\frac{2}{3}},&\,\,\,(T_{sh}(L)\ll T\ll T_{cr})\\ \\ T^{\frac{1}{2}},&\,\,\,(T\gg T_{cr}).\end{array}\right. (80)

This behavior can be understood from the N=1N=1 solution, where x2\langle x^{2}\rangle is determined by the self-consistent equations Eqs.(58) and (60). In the limit of small LL and low TT, the summation of kk in Eq.(60) is dominated by the k0k\sim 0 terms. The result is therefore similar to that of the atomic limit at K=0K=0 or TTcrT\gg T_{cr}.

A similar short-chain limit exists in the dynamical quantities Ωk(T)\Omega_{k}(T) and Γk(T)\Gamma_{k}(T). In Fig.14(a) and (b), we show the curves of Ωk=0(T)\Omega_{k=0}(T) and Γk=0(T)\Gamma_{k=0}(T) for a series LL. For momentum k=0k=0, both quantities have a common three-section power-law behavior. As TT decreases, they change from the high temperature atomic-limit behavior T1/4T^{1/4}, to the non-local long-chain behavior T1/3T^{1/3} in Tsh(L)TTcrT_{sh}(L)\ll T\ll T_{cr}, and to the short-chain behavior T1/4T^{1/4} in the low temperature limit TTsh(L)TcrT\ll T_{sh}(L)\leq T_{cr}. The behavior is summarized as

Ωk=0(T),Γk=0(T){T14,(TTsh(L))T13,(Tsh(L)TTcr)T14,(TTcr).\Omega_{k=0}(T),\,\,\,\Gamma_{k=0}(T)\sim\left\{\begin{array}[]{lllll}T^{\frac{1}{4}},&\,\,\,(T\ll T_{sh}(L))\\ \\ T^{\frac{1}{3}},&\,\,\,(T_{sh}(L)\ll T\ll T_{cr})\\ \\ T^{\frac{1}{4}},&\,\,\,(T\gg T_{cr}).\end{array}\right. (81)

In Fig.14(c) and (d), we show the quantities for k=πk=\pi. In the low TT limit, Ωk=π(T)\Omega_{k=\pi}(T) is a finite constant, not showing the short-chain behaviors. However, the quantities δΩk=π(T)Ωk=π(T)Ωk=π(0)\delta\Omega_{k=\pi}(T)\equiv\Omega_{k=\pi}(T)-\Omega_{k=\pi}(0) and Γk=π(T)\Gamma_{k=\pi}(T) do have short-chain limit behavior. For other k>0k>0, similar behaviors are found and they are summarized as

δΩk>0(T),Γk>0(T){T12,(TTsh(L))T23,(Tsh(L)TTcr)T14,(TTcr).\delta\Omega_{k>0}(T),\,\,\,\Gamma_{k>0}(T)\sim\left\{\begin{array}[]{lllll}T^{\frac{1}{2}},&\,\,\,(T\ll T_{sh}(L))\\ \\ T^{\frac{2}{3}},&\,\,\,(T_{sh}(L)\ll T\ll T_{cr})\\ \\ T^{\frac{1}{4}},&\,\,\,(T\gg T_{cr}).\end{array}\right. (82)

These results are in principle amenable to experimental tests on low dimensional materials with varying length and temperature. They also show that one must be careful when interpreting the temperature behavior of the experimental or MD data related to thermal transport in low dimensional systems, since at least two energy scales are present to separate the scaling regimes. One is related to the length of chain, and the other is related to the competition between local and non-local interactions. In the present work, we fix parameters m=K=γ=1m=K=\gamma=1 and hence have a definite Tcr1.0T_{cr}\approx 1.0 and Tsh(L)1.0×L3T_{sh}(L)\approx 1.0\times L^{-3}. For a system with arbitrary parameters, the concrete values of the two crossover temperatures will shift according to the exact or approximate scaling properties of the model. Therefore before one can carry out reliable power-law fitting, one must collect data in a sufficiently wide range of temperature to discern the possible crossovers.

V Discussions and Summary

We first discuss the implication of our results for the temperature dependence of thermal conductivity κ(T)\kappa(T). In the Debye formula κ=kCkvklk\kappa=\sum_{k}C_{k}v_{k}l_{k} [31, 32], κ\kappa is expressed in terms of the specific heat capacity Ck=(1/L)Hk/TC_{k}=(1/L)\partial\langle H_{k}\rangle/\partial T, the phonon group velocity vk=Ωk/kv_{k}=\partial\Omega_{k}/\partial k, and the mean free path lk=vk/(2Γk)l_{k}=v_{k}/(2\Gamma_{k}).

PTA results for Ωk\Omega_{k} and Γk\Gamma_{k} can be used to calculate κ\kappa through the Debye formula. It is known that Ck>0C_{k>0} ranges between 1.01.0 (low TT) and 0.750.75 (high TT) and is weakly kk dependent, and Ck=0=0.75C_{k=0}=0.75 [33]. Approximating CkC_{k} as a constant, and inputting Eqs.(77) and (78) into the Debye formula, we obtain the asymptotic TT dependence of κ\kappa as

κ(T){T23,(TTcr)T34,(TTcr),\kappa(T)\sim\left\{\begin{array}[]{lll}T^{-\frac{2}{3}},&\,\,\,(T\ll T_{cr})\\ \\ T^{-\frac{3}{4}},&\,\,\,(T\gg T_{cr}),\end{array}\right. (83)

with Tcr1.0T_{cr}\approx 1.0. Direct numerical evaluation of the Debye formula using PTA input confirms this result. However, this behavior does not agree with the high temperature result κ(T)T1.35\kappa(T)\sim T^{-1.35} from MD simulation in Ref.[44] and the expression κ(T)T4/3\kappa(T)\sim T^{-4/3} from the effective phonon theory [48]. This poses question on the applicability of the Debye formula for describing the thermal conductivity of the ϕ4\phi^{4} lattice model. A detailed answer to this question is beyond the scope of this paper and we leave it for a future study.

In this work, we developed the idea of expanding a low level eigen-operator QνkQ_{\nu k} by multiplying it with functions of approximate mode energy νk\mathcal{H}_{\nu k}, to describe the thermal-broadened quasi-particle peak in the spectral function. Here, we have used the basis variable QνkeλiνkQ_{\nu k}e^{\lambda_{i}\mathcal{H}_{\nu k}}, but it is interesting to know the importance of the inter-mode coupling for the effectiveness of basis. For example, PTA on the more complete basis QνkeiμpλμpiμpQ_{\nu k}e^{\sum_{i\mu p}\lambda^{i}_{\mu p}\mathcal{H}_{\mu p}} needs to be investigated to answer this question. We expect that it can give quantitative improvement of Γk\Gamma_{k}. It is also noted that the present HH-expanded basis does not cover all types of basis operators or scattering processes. For example, the variable Qν1k1Qν2k2Qν3k3Q_{\nu_{1}k_{1}}Q_{\nu_{2}k_{2}}Q_{\nu_{3}k_{3}} (k1+k2+k3=kmod2π)k_{1}+k_{2}+k_{3}=k\mod{2\pi}) with different k1k3k_{1}\sim k_{3} is not covered. Scrutinization is therefore needed for applying the present HH-expansion of the basis to general situations.

For quantum systems, quantum fluctuation is an important source of quasi-particle broadening. In particular, in systems such as frustrated quantum antiferromagnetism or system close to quantum criticality, quantum damping plays a dominant role. The present scheme can be generalized to study the quantum damping of quasi particles or quantum broadening of the quasi-particle peak in the spectral function. A direct application of the present formalism needs additional care, because for quantum systems the conservation identity Eqs.(84) and (85), used to simplify the matrices 𝐈{\bf I} and 𝐋{\bf L}, must be modified. The GF and the whole formalism of the PTA needs to be replaced with the quantum version. Other than this, there seem to be no other constraints. In case the scattering processes other than those contained in the HH-expanded basis play a vital role, one must consider the expanding schemes with other (approximately) conserved quantities other than energy, such as the particle number, momentum, spin, etc., to describe the decay of quasi-particles due to various fluctuations. What we have in mind is the spin-wave damping in frustrated antiferromagnets [40]. Further study of this interesting subject is in progress.

In summary, we developed the PTA on the HH-expanded basis to describe the thermal broadening of quasi particles. A zeros-removing technique is used to stabilize the iterative solution of PTA equations. Benchmarking calculations on the classical AHO model and the 1D ϕ4\phi^{4} model show that this method can produce semi-quantitative thermal broadening of a quasi-particle peak in the spectral function. Using this method, we disclose the low- and high-temperature power laws of the phonon spectral width Γk\Gamma_{k} for the 1D ϕ4\phi^{4} lattice model, contradicting the assumption in effective phonon theory and raising questions on the applicability of the Debye formula for this model. A short-chain limit of this model is also discovered. Future development and possible extension of this method are discussed.

VI Acknowledgments

We acknowledge helpful discussions with Y.J. Wu, C.L. Ji, X.G. Ren, L. Wan, Y. Wan, and T. Li. We are grateful to Professor Ren for showing us the method to treat the redundant basis variables. This work is supported by NSFC (Grant No.11974420 and No.12075316).

Appendix A Derivation of 𝐈{\bf I} and 𝐋{\bf L} matrices for AHO model on bases B2B_{2} and B3B_{3}

In this Appendix, we present the derivation of 𝐈{\bf I} and 𝐋{\bf L} matrices for AHO on the bases B2B_{2} and B3B_{3}.

Before presenting the details, we present the identity to be used for simplifying the expressions [33]. For arbitrary variables AA and BB, we have

{A,B}=β{A,H}B,\langle\{A,B\}\rangle=\beta\langle\{A,H\}B\rangle, (84)

and for arbitrary number θ\theta,

{A,H}BeθH=1βθ{A,B}eθH.\langle\{A,H\}Be^{\theta H}\rangle=\frac{1}{\beta-\theta}\langle\{A,B\}e^{\theta H}\rangle. (85)

Eq.(84) is proved in Appendix B. Eq.(85) is obtained by letting K=1K=1, X1=HX_{1}=H, and g(X1)=eθHg(X_{1})=e^{\theta H} in the identity Eq.(99).

A.1 Basis B2B_{2}

For basis B2={Ai=xfi}B_{2}=\{A_{i}=xf_{i}\} (fi=eλiHf_{i}=e^{\lambda_{i}H}), the matrix elements Iij=(xfi|xfj)I_{ij}=(xf_{i}|xf_{j}) is calculated through Eq.(84) as

Iij\displaystyle I_{ij} =\displaystyle= {xfi,{xfj,H}}\displaystyle\langle\{xf_{i},\{xf_{j},H\}\}\rangle (86)
=\displaystyle= β{x,H}{x,H}fifj.\displaystyle\beta\langle\{x,H\}\{x,H\}f_{i}f_{j}\rangle.

Using Eq.(85) (letting A=xA=x, B={x,H}B=\{x,H\}, and θ=λi+λj\theta=\lambda_{i}+\lambda_{j}), and {x,{x,H}}=1/μ\{x,\{x,H\}\}=1/\mu, we obtain

Iij\displaystyle I_{ij} =\displaystyle= {x,{x,H}}fifj+(λi+λj){x,H}{x,H}fifj\displaystyle\langle\{x,\{x,H\}\}f_{i}f_{j}\rangle+(\lambda_{i}+\lambda_{j})\langle\{x,H\}\{x,H\}f_{i}f_{j}\rangle (87)
=\displaystyle= βμ(βλiλj)fifj.\displaystyle\frac{\beta}{\mu\left(\beta-\lambda_{i}-\lambda_{j}\right)}\langle f_{i}f_{j}\rangle.

The matrix element Lij=(xfi|{{xfj,H},H})L_{ij}=-(xf_{i}|\{\{xf_{j},H\},H\}) is calculated similarly as

Lij\displaystyle L_{ij} =\displaystyle= {xfi,{{{x,H},H},H}fj}\displaystyle-\langle\{xf_{i},\{\{\{x,H\},H\},H\}f_{j}\}\rangle (88)
=\displaystyle= β{x,H}{{{x,H},H},H}fifj.\displaystyle-\beta\langle\{x,H\}\{\{\{x,H\},H\},H\}f_{i}f_{j}\rangle.

Using Eq.(85) (letting A=xA=x, B={{{x,H},H},H}B=\{\{\{x,H\},H\},H\}, and θ=λi+λj\theta=\lambda_{i}+\lambda_{j}) and {x,{{{x,H},H},H}}=(1/μ2)(μω02+12αx2)\{x,\{\{\{x,H\},H\},H\}\}=-(1/\mu^{2})\left(\mu\omega_{0}^{2}+12\alpha x^{2}\right), we obtain

Lij\displaystyle L_{ij} =\displaystyle= βμ2(βλiλj)[μω02fifj+12αx2fifj].\displaystyle\frac{\beta}{\mu^{2}\left(\beta-\lambda_{i}-\lambda_{j}\right)}\left[\mu\omega_{0}^{2}\langle f_{i}f_{j}\rangle+12\alpha\langle x^{2}f_{i}f_{j}\rangle\right].

Eqs.(87) and (A.1) are Eq.(III.2.1) in the main text.

A.2 Basis B3B_{3}

For basis B3B_{3}, the basis variables are denoted in Eq.(26) as Aμi=aμfiA_{\mu i}=a_{\mu}f_{i}, with a1=xa_{1}=x, a2=x3a_{2}=x^{3}, and fi=eλiHf_{i}=e^{\lambda_{i}H}. The matrix elements of 𝐈{\bf I} and 𝐋{\bf L} read

Iμνij={Aμi,{Aνj,H}},I_{\mu\nu}^{ij}=\langle\{A_{\mu i},\{A_{\nu j},H\}\}\rangle, (90)

and

Lμνij={Aμi,{{{Aνj,H},H},H}},L_{\mu\nu}^{ij}=-\langle\{A_{\mu i},\{\{\{A_{\nu j},H\},H\},H\}\}\rangle, (91)

respectively. Using the same method used for basis B2B_{2} above, we obtain

Iμνij\displaystyle I_{\mu\nu}^{ij} =\displaystyle= ββλiλjIμν(0)fi(H)fj(H),\displaystyle\frac{\beta}{\beta-\lambda_{i}-\lambda_{j}}\langle I_{\mu\nu}^{(0)}f_{i}(H)f_{j}(H)\rangle,
Lμνij\displaystyle L_{\mu\nu}^{ij} =\displaystyle= ββλiλjLμν(0)fi(H)fj(H).\displaystyle\frac{\beta}{\beta-\lambda_{i}-\lambda_{j}}\langle L_{\mu\nu}^{(0)}f_{i}(H)f_{j}(H)\rangle. (92)

Here, the variables Iμν(0)={aμ,{aν,H}}I^{(0)}_{\mu\nu}=\{a_{\mu},\{a_{\nu},H\}\} and Lμν(0)={aμ,{{{aν,H},H},H}}L^{(0)}_{\mu\nu}=-\{a_{\mu},\{\{\{a_{\nu},H\},H\},H\}\}. Calculation gives

I11(0)=1μ,\displaystyle I_{11}^{(0)}=\frac{1}{\mu},
I12(0)=I21(0)=3μx2,\displaystyle I_{12}^{(0)}=I^{(0)}_{21}=\frac{3}{\mu}x^{2},
I22(0)=9μx4,\displaystyle I_{22}^{(0)}=\frac{9}{\mu}x^{4}, (93)

and

L11(0)=1μ2V′′(x),\displaystyle L_{11}^{(0)}=\frac{1}{\mu^{2}}V^{{\prime}{\prime}}(x),
L12(0)=18μ3p2+18μ2xV(x)+3μ2x2V′′(x),\displaystyle L_{12}^{(0)}=-\frac{18}{\mu^{3}}p^{2}+\frac{18}{\mu^{2}}xV^{\prime}(x)+\frac{3}{\mu^{2}}x^{2}V^{\prime\prime}(x),
L21(0)=3μ2x2V′′(x),\displaystyle L_{21}^{(0)}=\frac{3}{\mu^{2}}x^{2}V^{\prime\prime}(x),
L22(0)=54μ3x2p2+54μ2x3V(x)+9μ2x4V′′(x).\displaystyle L_{22}^{(0)}=-\frac{54}{\mu^{3}}x^{2}p^{2}+\frac{54}{\mu^{2}}x^{3}V^{\prime}(x)+\frac{9}{\mu^{2}}x^{4}V^{\prime\prime}(x).

Here V(x)V(x) is the potential function Eq.(III). Inserting these equations into Eq.(A.2), we get the expressions for IμνijI_{\mu\nu}^{ij} and LμνijL_{\mu\nu}^{ij}. They contain the averages fifj\langle f_{i}f_{j}\rangle, x2fifj\langle x^{2}f_{i}f_{j}\rangle, x4fifj\langle x^{4}f_{i}f_{j}\rangle, p2fifj\langle p^{2}f_{i}f_{j}\rangle, and x2p2fifj\langle x^{2}p^{2}f_{i}f_{j}\rangle. The first three can be calculated self-consistently. The last two can be simplified by the conservation identity Eq.(85).

In Eq.(85), we let A=xA=x, B=pB=p, and θ=λi+λj\theta=\lambda_{i}+\lambda_{j}, we obtain

p2fifj=μβλiλjfifj.\langle p^{2}f_{i}f_{j}\rangle=\frac{\mu}{\beta-\lambda_{i}-\lambda_{j}}\langle f_{i}f_{j}\rangle. (95)

Similarly, letting A=x2A=x^{2} and B=xpB=xp, we obtain

x2p2fifj=μβλiλjx2fifj.\langle x^{2}p^{2}f_{i}f_{j}\rangle=\frac{\mu}{\beta-\lambda_{i}-\lambda_{j}}\langle x^{2}f_{i}f_{j}\rangle. (96)

Using Eqs.(95) and (96), we finally obtain Eqs.(29) and (III.2.2). Note that the conservation relation Eq.(85) is also used to reduce L12ijL^{ij}_{12} to L21ijL^{ij}_{21}. The final result for 𝐋{\bf L} is Hermitian.

Appendix B Proof of a conservation identity

In this Appendix, we derive an identity to be used in the derivation of expressions for 𝐈{\bf I} and 𝐋{\bf L} in Appendix A.

First, we cite the conservation identity Eq.(15) of Ref.[33], i.e., for arbitrary variables AA and BB, we have

{A(t),B(t)}\displaystyle\langle\{A(t),B(t^{\prime})\}\rangle =\displaystyle= β{A,H}(t)B(t)\displaystyle\beta\langle\{A,H\}(t)B(t^{\prime})\rangle (97)
=\displaystyle= βA(t){B,H}(t).\displaystyle-\beta\langle A(t)\{B,H\}(t^{\prime})\rangle.

In the above equation, the average is defined as the canonical ensemble average in an equilibrium state at temperature TT. β=1/kBT\beta=1/k_{B}T is the inverse temperature. In the equal time situation t=tt=t^{\prime}, we get

{A,B}\displaystyle\langle\{A,B\}\rangle =\displaystyle= β{A,H}B\displaystyle\beta\langle\{A,H\}B\rangle (98)
=\displaystyle= βA{B,H}.\displaystyle-\beta\langle A\{B,H\}\rangle.

In the first equation of Eq.(98), we replace BB with Bg(X1,X2,,XK)Bg(X_{1},X_{2},...,X_{K}). Here XiX_{i} is a conserved quantity satisfying {Xi,H}=0\{X_{i},H\}=0 (i=1,2,,Ki=1,2,...,K) and gg is an arbitrary KK-variable function. We get the following conservation identity,

{A,B}g(X1,X2,,XK)\displaystyle\langle\{A,B\}g(X_{1},X_{2},...,X_{K})\rangle (99)
=\displaystyle= β{A,H}Bg(X1,X2,,XK)\displaystyle\beta\langle\{A,H\}Bg(X_{1},X_{2},...,X_{K})\rangle
i=1K{A,Xi}BXig(X1,X2,,XK).\displaystyle-\sum_{i=1}^{K}\langle\{A,X_{i}\}B\frac{\partial}{\partial X_{i}}g(X_{1},X_{2},...,X_{K})\rangle.

Eq.(85) is a special case of Eq.(99), with K=1K=1, X1=HX_{1}=H, and g(X1)=eθHg(X_{1})=e^{\theta H}.

References

  • [1] H. Mori, Transport, collective motion, and Brownian motion, Prog. Theor. Phys. 33, 423 (1965).
  • [2] H. Mori, A continued-fraction representation of the time-correlation functions, Prog. Theor. Phys. 34, 399 (1965).
  • [3] R. Zwanzig, Nonequilibrium Statistical Mechanics (Oxford University Press, New York, 2001).
  • [4] Yu. A. Tserkovnikov, A method of solving infinite systems of equations for two-time thermal Green’s functions, Theor. Math. Phys. 49, 993 (1981).
  • [5] Yu. A. Tserkovnikov, Two-time temperature Green’s functions in kinetic theory and molecular hydrodynamics: I. The chain of equations for the irreducible functions, Theor. Math. Phys. 118, 85 (1999).
  • [6] L. M. Roth, New method for linearizing many-body equations of motion in statistical mechanics, Phys. Rev. Lett. 20, 1431 (1968).
  • [7] L. M. Roth, Electron correlation in narrow energy bands. I. The two-pole approximation in a narrow s band, Phys. Rev. 184, 451 (1969).
  • [8] For a review, see A. Avella, Composite operator method analysis of the underdoped cuprates puzzle, Adv. Condens. Matter Phys. 2014, 515698 (2014).
  • [9] L. Haurie et al., Bands renormalization and superconductivity in the strongly correlated Hubbard model using composite operators method, J. Phys.: Condens. Matter 36, 255601 (2024).
  • [10] Y. Kakehashi and P. Fulde, Self-consistent projection operator approach to nonlocal excitation spectra, Phys. Rev. B 70, 195102 (2004).
  • [11] S. Onoda and M. Imada, Operator projection method applied to the single-particle Green’s function in the Hubbard model, J. Phys. Soc. Jpn 70, 632 (2001).
  • [12] M. H. Lee, Orthogonalization process by recurrence relations, Phys. Rev. Lett. 49, 1072 (1982).
  • [13] M. H. Lee, Solutions of the generalized Langevin equation by a method of recurrence relations, Phys. Rev. B 26, 2547 (1982).
  • [14] A. L. Kuzemsky, Irreducible Green functions method and many-particle interacting systems on a lattice, Riv. Nuovo Cimento 25, 1 (2002).
  • [15] S. Lepri, Relaxation of classical many-body Hamiltonians in one dimension, Phys. Rev. E 58, 7165 (1998).
  • [16] A. Castellano, J. P. A. Batista, and M. J. Verstraete, Mode-coupling theory of lattice dynamics for classical and quantum crystals, J. Chem. Phys. 159, 234501 (2023).
  • [17] A. Eskandari-asl and A. Avella, Time-resolved ARPES signal in pumped semiconductors within the dynamical projective operatorial approach, Phys. Rev. B 110, 094309 (2024).
  • [18] P. Fan, K. Yang, K. H. Ma, and N. H. Tong, Projective truncation approximation for equations of motion of two-time Green’s functions, Phys. Rev. B 97, 165140 (2018).
  • [19] P. Fan and N. H. Tong, Controllable precision of the projective truncation approximation for Green’s functions, Chin. Phys. B 28, 047102 (2019).
  • [20] K. H. Ma and N. H. Tong, Interacting spinless fermions on the square lattice: Charge order, phase separation, and superconductivity, Phys. Rev. B 104, 155116 (2021).
  • [21] P. Fan, N. H. Tong, and Z. G. Zhu, Spatial spin-spin correlations of the single-impurity Anderson model with a ferromagnetic bath, Phys. Rev. B 106, 155130 (2022).
  • [22] For a review, see Y. G. Rudoy, The Bogoliubov–Tyablikov Green’s function method in the quantum theory of magnetism, Theor. Math. Phys. 168, 1318 (2011).
  • [23] Y. Onodera, Dynamic susceptibility of classical anharmonic oscillator - A unified oscillator model for order-disorder and displacive ferroelectrics, Prog. Theor. Phys. 44, 1477 (1970).
  • [24] N. Li, P. Tong, and B. Li, Effective phonons in anharmonic lattices: Anomalous vs. normal heat conduction, Europhys. Lett. 75, 49 (2006).
  • [25] S. Liu, J. Liu, P. Ha¨\ddot{a}nggi, C. Wu, and B. Li, Triggering waves in nonlinear lattices: Quest for anharmonic phonons and corresponding mean-free paths, Phys. Rev. B 90, 174304 (2014).
  • [26] L. Xu and L. Wang, Dispersion and absorption in one-dimensional nonlinear lattices: A resonance phonon approach, Phys. Rev. E 94, 030101(R) (2016).
  • [27] L. Xu and L. Wang, Resonance phonon approach to phonon relaxation time and mean free path in one-dimensional nonlinear lattices, Phys. Rev. E 95, 042138 (2017).
  • [28] J. S. Wang, J. Wang, and N. Zeng, Nonequilibrium Green’s function approach to mesoscopic thermal transport, Phys. Rev. B 74, 033408 (2006).
  • [29] R. G. Della Valle and P. Procacci, Computer-aided series expansion for phonon self-energy, J. Comp. Phys. 165, 428 (2000).
  • [30] B. Nickel, The solution to the 4-phonon Boltzmann equation for a 1D chain in a thermal gradient, J. Phys. A 40, 1219 (2007).
  • [31] J. M. Ziman, Electrons and Phonons (Oxford at the Clarendon Press, 1960), Chaps. 7 and 8.
  • [32] P. B. Allen and J. L. Feldman, Thermal conductivity of disordered harmonic solids, Phys. Rev. B 48, 12581 (1993).
  • [33] K. H. Ma, Y. J. Guo, L. Wang, and N. H. Tong, Projective-truncation-approximation study of the one-dimensional ϕ4\phi^{4} lattice model, Phys. Rev. E 106, 014110 (2022).
  • [34] J. C. Herzel, Green’s functions and double-time distribution functions in classical statistical mechanics, J. Math. Phys. 8, 1650 (1967).
  • [35] M. J. Smith, Classical Green’s functions for the ideal gas, Phys. Rev. Lett. 24, 1398 (1970).
  • [36] A. Cavallo, F. Cosenza, and L. De. Cesare, Classical Heisenberg ferromagnetic chain with long-range interactions: A spectral density approach, Phys. Rev. B 66, 174439 (2002).
  • [37] For a review, see A. Cavallo, F. Cosenza, and L. De. Cesare, in New Developements in Ferromagnetism Research, edited by V. N. Murray (Nova Science 2005), Chap 6.
  • [38] T. Dauxois, M. Peyrard, and A. R. Bishop, Dynamics and thermodynamics of a nonlinear model for DNA denaturation, Phys. Rev. E 47, 684 (1993).
  • [39] X. G. Ren (private discussion).
  • [40] S. Winterfeldt and D. Ihle, Spin dynamics and antiferromagnetic short-range order in the two-dimensional Heisenberg model, Phys. Rev. B 59, 6010 (1999).
  • [41] Y. J. Guo, Y. C. Sun, and L. Wang, Thermalization process in a one-dimensional lattice with two-dimensional motions: The role of angular momentum conservation, Phys. Rev. E 108, 014127 (2023).
  • [42] Y. Liu and D. He, Approach to phonon relaxation time and mean free path in nonlinear lattices, Chin. Phys. Lett. 38, 044401 (2021).
  • [43] D. Boyanovsky, C. Destri, and H. J. de Vega, Approach to thermalization in the classical ϕ4\phi^{4} theory in 1+11+1 dimensions: Energy cascades and universal scaling, Phys. Rev. D 69, 045003 (2004).
  • [44] K. Aoki and D. Kusnezov, Bulk properties of anharmonic chains in strong thermal gradients: non-equilibrium ϕ4\phi^{4} theory, Phys. Lett. A 265, 250 (2000).
  • [45] B. Hu, B. Li, and H. Zhao, Heat conduction in one-dimensional nonintegrable systems, Phys. Rev. E 61, 3828 (2000).
  • [46] K. Aoki, J. Lukkarinen, and H. Spohn, Energy transport in weakly anharmonic chains, J. Stat. Phys. 124, 1105 (2006).
  • [47] A. Dhar, Heat transport in low-dimensional systems, Adv Phys. 57, 457 (2008).
  • [48] N. Li and B. Li, Scaling of temperature-dependent thermal conductivities for one-dimensional nonlinear lattices, Phys. Rev. E 87, 042125 (2013).
  • [49] N. Li, J. Liu, C. Wu, and B. Li, Temperature and frequency dependentmean free paths of renormalized phonons in nonlinear lattices, New J. Phys. 20, 023006 (2018).
  • [50] K. Aoki and D. Kusnezov, Fermi-Pasta-Ulam β\beta Model: Boundary jumps, Fourier’s law, and scaling, Phys. Rev. Lett. 86, 4029 (2001).
  • [51] W. Metzner and D. Vollhardt, Correlated lattice fermions in d=d=\infty dimensions, Phys. Rev. Lett. 62, 324 (1989).
  • [52] A. Georges, G. Kotliar, W. Krauth, and M. J. Rozenberg, Dynamical mean-field theory of strongly correlated fermion systems and the limit of infinite dimensions, Rev. Mod. Phys. 68, 13 (1996).
  • [53] J. Liu, S. Liu, N. Li, B. Li, and C. Wu, Renormalized phonons in nonlinear lattices: A variational approach, Phys. Rev. E 91, 042910 (2015).
  • [54] R. Kikuchi, A theory of cooperative phenomena, Phys. Rev. 81, 988 (1951).
  • [55] For a review, see Theory and Applications of the Cluster Variation and Path Probability Methods, edited by J. L. Morán-López and J. M. Sanchez (Plenum, New York 1996).
  • [56] H. W. Jia and N. H. Tong, Thermodynamics of classical one-dimensional Klein-Gordon lattice model, arXiv:2410.19249 (unpublished).
  • [57] A. Pelizzolla, Cluster variation method in statistical physics and probabilistic graphical models, J. Phys. A 38, R309 (2005).