Diffusive dynamics of fractionalized particles and the enhanced conductivity at the border of the neutral-ionic transition

Yuta Sakai    Chisa Hotta The University of Tokyo, Komaba, Meguro-ku, Tokyo 153-8902, Japan
Abstract

We study the diffusive dynamics of the classical one-dimensional lattice model of mobile particles featuring the incoherent metallic state of the organic TTF-CA found in the vicinity of the neutral-ionic(NI)-transition. The particles are strongly correlated and feel the alternating site potentials, exhibiting uniform ionic (I) Mott insulating and neutral band insulating (N) phases when the Coulomb interaction and site potentials are large, respectively. We focus on the neutral-ionic domain walls (NIDW) activated by their competition, which is typically regarded as fractionalized particles. The finite temperature phase diagram reveals a thermodynamically stable NIDW phase not found in previous literature that emerges at the triple point, where the transition line splits into two first-order lines toward the critical endpoints. We analyze the long-time behavior of dynamics of the NIDWs and find that it captures the incoherent transport of the system. The conductivity derived from the diffusion constant and the number population of NIDWs shows a strong enhancement inside the NIDW phase, which can explain the large conductivity at the NI crossover recently found in TTF-CA.

I Introduction

Enhanced correlations near the boundaries between different material phases have provided a compelling and fruitful platform for condensed matter research. Intriguing aspects of such phases are often captured by the anomalous behavior of transport properties. The anomalous electrical resistivity linear-in-temperature resistivity beyond the Mott-Ioffe-Regel (MIR) limit known as “bad metals” Bruin et al. (2013) are the longstanding topic across high Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT cupratesKordyuk (2015); Keimer et al. (2015), vanadium dioxideAllen et al. (1993); Qazilbash et al. (2006), and alkali-doped fulleridesGunnarsson et al. (2003). Similar behavior is found at much lower temperatures in heavy fermionic systemsTrovarelli et al. (2000) referred to as strange metals due to their non-Fermi liquid natureSchofield (1999). Perovskite and Ruddlesden-Popper manganites show a large sensitivity in response to an applied magnetic fieldSalamon and Jaime (2001). Their electrical resistivity grows by orders of magnitudes as a consequence of the interplay of charge, spin and orbitals accompanying phase segregation or inhomogeneityDagotto et al. (2001). These phenomena are more or less governed by many integrants, which makes it overwhelming to establish a simple understanding of their underlying mechanisms beyond the phenomenologyHartnoll (2014).

To this end, organic materials can be an ideal platform, because they show as rich a phase diagram as the aforementioned materials while they are mostly based on the single molecular orbital picture and relatively strong correlation effectsSeo et al. (2004). Aside from the most famous κ𝜅\kappaitalic_κ-ET families that show Mott insulator, superconductivity, criticality, and spin liquid behaviors, TTF-CA has also captured a long-standing interest in featuring neutral-ionic (NI) transition, which embodies the competition of band insulator and Mott insulatorSunami et al. (2022). The present paper focuses on the anomalous enhancement of electrical conductivity observed in TTF-CA at the NI crossover region that develops on the higher temperature part of the phase boundaryTakehara et al. (2019).

Experimentally, the anomalous conduction of TTF-CA has been attributed to topological low-energy excitation called neutral-ionic domain wall (NIDW). The NIDWs carry fractionalized elementary charges, ±e/2plus-or-minus𝑒2\pm e/2± italic_e / 2, defined as the averaged charge density of neighboring two sites on both sides of the bond. In an applied DC electric field, thermally excited NIDW pairs carry fractional charges of opposite signs. In the most elementary picture, they may drift in opposite directions and contribute to the electrical conductivity. Such fractionalization is typical of one- or quasi-one-dimensional systems, e.g. XXZ spin-1/2 model with a finite spin gapMayr and Horsch (2006) and both the bosonic and fermionic models on the anisotropic triangular latticeKohno et al. (2007); Hotta and Pollmann (2008), which are sometimes called fractons, where they typically yield a continuum in their excitation spectrum. For TTF-CA, it was discussed a long time agoNagaosa and Takimoto (1986) using the XXZ spin-1/2 model as an effective model in the limit of strong on-site Coulomb interactions and potentials.

Indeed, the studies on TTF-CA go back to the 1980’s and the experimental phase diagram for temperature and pressure has been clarified, mostly based on transportTakehara et al. (2019), NQRTakehara et al. (2018), NMRSunami et al. (2018), and optical measurementsMasino et al. (2007); Okamoto et al. (1989). At ambient pressure upon cooling, the neural phase transforms to the ionic state at 81 K to gain the Madelung energy including the long-range Coulomb interactions. By applying pressure, the high-temperature neutral phase crossovers to the paraelectric ionic state where the charge transfer occurs from TTF to CA, and at lower temperatures, the lattice distortion associated with the spin degrees of freedom of the Mott state drives the system to a ferroelectric ionic phase by breaking the inversion symmetry.

In theories, the ionic extended Hubbard model is applied, which includes the on-site and nearest neighbor Coulomb interactions U𝑈Uitalic_U and V𝑉Vitalic_V, the uniform transfer integral t𝑡titalic_t, and the alternating difference in the energy level ΔΔ\Deltaroman_Δ between TTF and CA Nagaosa and Takimoto (1986). The basic feature of this model was studied in the initial series of works based on the quantum Monte Carlo simulation (QMC), where the degree of charge transfer denoted as ρ𝜌\rhoitalic_ρ shows a discontinuity at the phase transition at low temperatures. For the ionic Hubbard model without V𝑉Vitalic_V, a more extensive studies have followed that clarified the ground state phase diagram on the plane of U𝑈Uitalic_U and ΔΔ\Deltaroman_Δ, where the competition happens between the band insulator and the Mott insulator, namely, neutral and ionic phases, at around ΔUsimilar-toΔ𝑈\Delta\sim Uroman_Δ ∼ italic_U. Fabrizio, et.al. proposed a scenario based on field theory that there is a spontaneously dimerized phase in between the band insulator and the Mott insulatorFabrizio et al. (1999). The dimerization destroys the spin-density-wave (SDW) Mott state and transforms it into the bond charge-density-wave (BCDW) state before entering the CDW band insulator. Despite several controversial numerical results in early periodsTakada and Kido (2001); Lou et al. (2003); Anusooya-Pati et al. (2001); Wilkens and Martin (2001); Gidopoulos et al. (2000) later numerical and analytical works have confirmed the existence of very narrow dimerized phaseTorio et al. (2001); Zhang et al. (2003); Manmana et al. (2004); Tsuchiizu and Furusaki (2004). This means that without the introduction of lattice dimerization, there is an underlying tendency to form a dimer toward the magnetically gapped singlet state with ferroelectricityWilkens and Martin (2001). The latest QMC study at finite temperatureOtsuka et al. (2012) implementing the lattice dimerization showed the double free-energy-minima in the low-temperature dimerized ionic phase. There, the dimerization disappears at high temperatures after they may undergo a first-order transition with a jump in the degree of charge transfer and hysteresis 111The role of dimerization on the order of transition is not clear in the literature. The ionic-to-neutral transition can be of first order when V𝑉Vitalic_V is present, or even by the choice of the boundary conditions at finite size. . The strong coupling limit of the ionic Hubbard model was studied by the effective spin-1 model Legeza et al. (2006) in agreement with a narrow intermediate phaseTincani et al. (2009).

Recently, some dynamic aspects of the NI transition was clarified by the experimental works on TTF-CA at the two characteristic phases in the temperature range of T280300similar-to𝑇280300T\sim 280-300italic_T ∼ 280 - 300 K. Based on the infrared spectroscopy Okamoto et al. (1989), NQR Takehara et al. (2019) and NMR Sunami et al. (2018) experiments, it is argued that, in the dimerized paraelectric ionic phase at P1035similar-to𝑃1035P\sim 10-35italic_P ∼ 10 - 35 kbar, there arise dynamical ferroelectric domains carrying opposite polarizations with their neighboring domains separated by the defects in the dimerization periods, which they call spin solitons. At the crossover region between the neutral and paraelectric ionic state, P810similar-to𝑃810P\sim 8-10italic_P ∼ 8 - 10 kbar, a large enhancement of conductivity whose magnitude is comparable to the metallic one is observedTakehara et al. (2019), which is attributed to NIDW conductivity. In theories, there had been some works based on the phase Hamiltonian and bosonization techniques Fukuyama and Ogata (2016); Tsuchiizu et al. (2016) to explain what kind of topological excitations and the gaps are available. However, the dynamic aspect of the NI system has not yet been explored so far, particularly in relevance with the dynamical inhomogeneity due to NIDWs and their many-body effects.

In the present paper, we study the one-dimensional classical model in the strong coupling limit of the ionic extended Hubbard model. We apply a Monte Carlo (MC) based Glauber dynamics and evaluate the dissipative nature of the NIDWs, showing that the NIDWs thermally activated in pairs across the charge transfer gap can fluctuate and propagate for a substantial timescale until they meet another NIDW carrying an opposite charge and disappear in pairs. The basic treatment given here is to find numerically in the classical limit of the quantum microscopic model that the hydrodynamic equations to describe a coarse-grained thermodynamic version of the microscopic model actually hold. We extract the electrical conductivity using the Einstein relation, finding that it is indeed enhanced at the phase boundary due to the substantial increase of the population of NIDWs, despite that the diffusion constant is suppressed by the correlation effect. Our calculation proves that such NIDW conductivity can happen without a lattice dimerization or quantum fluctuation effects, and that there is a thermodynamically stable NIDW phase at a relatively high temperature at least in the classical limit.

Refer to caption
Figure 1: (a) Schematic illustration of NI transition. (b) Particle configurations on D and A sites, whose charge densities are measured from a neutral(N) state with two particles on D sites as being zero. (c) Schematic illustration of the distribution of particles with NIDWs, shown in vertical bars which carry ±1/2plus-or-minus12\pm 1/2± 1 / 2 (red/blue) charges. (d) Example of creating or annihilating the NIDWs in pairs.

II Model and methods

II.1 Model

We consider a model of classical particles carrying up and down spins in one-dimensional(1D) chain consisting of two sublattices denoted as A and D, whose Hamiltonian is given as

cl=Ui=1Nni,ni,+Vi=1Nnini+1+ΔjAnj,subscriptcl𝑈superscriptsubscript𝑖1𝑁subscript𝑛𝑖subscript𝑛𝑖𝑉superscriptsubscript𝑖1𝑁subscript𝑛𝑖subscript𝑛𝑖1Δsubscript𝑗Asubscript𝑛𝑗\displaystyle\mathcal{H}_{\rm cl}=U\sum_{i=1}^{N}n_{i,\uparrow}n_{i,\downarrow% }+V\sum_{i=1}^{N}n_{i}n_{i+1}+\Delta\sum_{j\in\text{A}}n_{j},caligraphic_H start_POSTSUBSCRIPT roman_cl end_POSTSUBSCRIPT = italic_U ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_i , ↑ end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_i , ↓ end_POSTSUBSCRIPT + italic_V ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT + roman_Δ ∑ start_POSTSUBSCRIPT italic_j ∈ A end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , (1)

where niσ=0subscript𝑛𝑖𝜎0n_{i\sigma}=0italic_n start_POSTSUBSCRIPT italic_i italic_σ end_POSTSUBSCRIPT = 0 or 1111 is the number of particles with σ=,𝜎\sigma=\uparrow,\downarrowitalic_σ = ↑ , ↓ spins on-site i𝑖iitalic_i interacting via on-site and nearest neighbor Coulomb interactions, U𝑈Uitalic_U and V𝑉Vitalic_V. We set the energy level of site A to be higher than that of site D by ΔΔ\Deltaroman_Δ, as shown schematically in Fig. 1(a). We set U=7.5𝑈7.5U=7.5italic_U = 7.5 and V=3.5𝑉3.5V=3.5italic_V = 3.5 throughout this paper following the previous works on TTF-CANagaosa and Takimoto (1986).

Previous studies on the NI transition have mainly focused on the ionic extended Hubbard model given as IEHM=cl+kinsubscriptIEHMsubscriptclsubscriptkin\mathcal{H}_{\mathrm{IEHM}}=\mathcal{H}_{\rm cl}+\mathcal{H}_{\rm kin}caligraphic_H start_POSTSUBSCRIPT roman_IEHM end_POSTSUBSCRIPT = caligraphic_H start_POSTSUBSCRIPT roman_cl end_POSTSUBSCRIPT + caligraphic_H start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT which, in addition to Eq.(1), includes the quantum kinetic hopping term given as kin=ti=1Nσ=,(ci,σci+1,σ+h.c.)subscriptkin𝑡superscriptsubscript𝑖1𝑁subscript𝜎superscriptsubscript𝑐𝑖𝜎subscript𝑐𝑖1𝜎h.c.{\mathcal{H}}_{\rm kin}=-t\sum_{i=1}^{N}\sum_{\sigma=\uparrow,\downarrow}(c_{i% ,\sigma}^{\dagger}c_{i+1,\sigma}+\text{h.c.})caligraphic_H start_POSTSUBSCRIPT roman_kin end_POSTSUBSCRIPT = - italic_t ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_σ = ↑ , ↓ end_POSTSUBSCRIPT ( italic_c start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_i + 1 , italic_σ end_POSTSUBSCRIPT + h.c. ) with the electron creation/annihilation operators, ci,σ/ci,σsuperscriptsubscript𝑐𝑖𝜎subscript𝑐𝑖𝜎c_{i,\sigma}^{\dagger}/c_{i,\sigma}italic_c start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT / italic_c start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT, where the aforementioned values of U𝑈Uitalic_U and V𝑉Vitalic_V are taken as a unit of t=1𝑡1t=1italic_t = 1 whose values are assumed as 0.1-0.2eVIshibashi and Terakura (2010). Our model serves as its strong coupling limit t/U0𝑡𝑈0t/U\rightarrow 0italic_t / italic_U → 0, and accordingly, we assume that the particles follow Pauli’s principle. The NI transition is characterized by the degree of charge transfer from site D to site A given as,

ρ=2jAnj/N,𝜌2subscript𝑗Asubscript𝑛𝑗𝑁\rho=2\sum_{j\in\text{A}}n_{j}/N,italic_ρ = 2 ∑ start_POSTSUBSCRIPT italic_j ∈ A end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT / italic_N , (2)

which is called ionicity. Although ρ𝜌\rhoitalic_ρ can range between 0 and 2 in principle, ρ>1𝜌1\rho>1italic_ρ > 1 is unlikely, because the two limiting cases, the perfect N and I phases, have ρ=0𝜌0\rho=0italic_ρ = 0 and 1, respectively. We consider clsubscriptcl\mathcal{H}_{\rm cl}caligraphic_H start_POSTSUBSCRIPT roman_cl end_POSTSUBSCRIPT at half-filling, namely the total number of particles is Ne=Nsubscript𝑁𝑒𝑁N_{e}=Nitalic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = italic_N consisting of the same number of particles with up and down spins, and apply a periodic boundary condition. The NI transition is understood as a consequence of the competition between U𝑈Uitalic_U and ΔΔ\Deltaroman_Δ.

II.2 Classical Monte Carlo dynamics

We are concerned with the dynamical properties of clsubscriptcl\mathcal{H}_{\rm cl}caligraphic_H start_POSTSUBSCRIPT roman_cl end_POSTSUBSCRIPT, which can be captured within the framework of Glauber dynamics of the classical Monte Carlo(MC) calculationGlauber (1963). Glauber dynamics was first introduced to describe the time evolution of the Ising model by regarding the local spin flip as “time” step, which constitutes a stochastic continuous Markov chain. This type of dynamic interpretation of MC sampling is generally known to be valid for the system which has no system-encoded intrinsic rule dominating the time evolution of the concerned degrees of freedom, instead, which can be modeled to evolve with stochastic dynamics induced by a weak coupling to a heat bathBinder et al. (1992); Binder (1997, 2022). Here, we call it (classical) MC dynamics. The MC dynamics has been applied to various classical systems so far, such as a Brownian motion of polymersBaumgärtner et al. (1983), relaxation phenomena in quadrupolar glassesCarmesin and Binder (1987), and a diffusion process of interstitial atoms in an alloyKehr et al. (1981, 1989) (for more information, see Refs.[Binder et al., 1992] and [Binder, 1997]). The aforementioned works using dynamic MC simulations aim to calculate dynamic properties as a function of Monte Carlo timestep (MCS) such as autocorrelation functions or a mean square displacement (MSD) of particles, and by analyzing those properties, a characteristic timescale and a diffusion constant are obtained. Their results show that MC dynamics can safely capture the dynamical properties of the systems that do not afford molecular dynamics (MD) simulations built on the equation of motion. Still, for some cases, the acceptance ratio at each MCS can be extremely low, which hinders the calculation, for which a similar framework called kinetic MC was developedBortz et al. (1975). We tested and found that our model has a high enough acceptance ratio to sustain the simpler Glauber dynamics.

We perform an MC calculation by the following steps;
(1) Randomly choose one particle with spin σ𝜎\sigmaitalic_σ and the direction to hop,
(2) If there already exists a particle with the same spin in the destination, discard the trial. Otherwise, transfer the particle with the transition probability of the heat-bath method,

P(ΔE)=11+eΔE/kBT,𝑃Δ𝐸11superscript𝑒Δ𝐸subscript𝑘𝐵𝑇P(\Delta E)=\frac{1}{1+e^{\Delta E/k_{B}T}},italic_P ( roman_Δ italic_E ) = divide start_ARG 1 end_ARG start_ARG 1 + italic_e start_POSTSUPERSCRIPT roman_Δ italic_E / italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG , (3)

where ΔEΔ𝐸\Delta Eroman_Δ italic_E is the energy difference between the states before and after the hop.
We repeat (1) and (2) over Ne/2subscript𝑁𝑒2N_{e}/2italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT / 2 times per σ=,𝜎\sigma=\uparrow,\downarrowitalic_σ = ↑ , ↓, which constitute a single MCS. We prepare an I-state as an initial state and perform the calculation over typically NMCS109similar-tosubscript𝑁MCSsuperscript109N_{\rm MCS}\sim 10^{9}italic_N start_POSTSUBSCRIPT roman_MCS end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT steps, while discarding the first 108superscript10810^{8}10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT MCSs in calculating the phase diagram. For the dynamical calculations, we set (NMCS,N)=(109,512)subscript𝑁MCS𝑁superscript109512(N_{\rm MCS},N)=(10^{9},512)( italic_N start_POSTSUBSCRIPT roman_MCS end_POSTSUBSCRIPT , italic_N ) = ( 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT , 512 ) and (5×108,1024)5superscript1081024(5\times 10^{8},1024)( 5 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT , 1024 ) for the calculation of ndw,Dsubscript𝑛dw𝐷n_{\rm dw},Ditalic_n start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT , italic_D and σdwsubscript𝜎dw\sigma_{\rm dw}italic_σ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT, and NMCS=1×109subscript𝑁MCS1superscript109N_{\rm MCS}=1\times 10^{9}italic_N start_POSTSUBSCRIPT roman_MCS end_POSTSUBSCRIPT = 1 × 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT for both N=512𝑁512N=512italic_N = 512 and N=1024𝑁1024N=1024italic_N = 1024 in obtaining σesubscript𝜎𝑒\sigma_{e}italic_σ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, both of which the 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT MCSs are discarded as a relaxation. We confirm that the results do not depend on the initial configurations.

II.2.1 Charge dynamics in an electric field

We apply two different ways of analysis to evaluate the thermally activated electrical conductivity at finite temperatures. The first method is to introduce the effect of the electric field as a constant energy potential bias {\cal E}caligraphic_E. This corresponds to adding +e𝑒+e{\cal E}+ italic_e caligraphic_E to ΔEΔ𝐸\Delta Eroman_Δ italic_E in Eq.(3) when calculating the transition probability of hopping the particle to the right-hand side, and adding e𝑒-e{\cal E}- italic_e caligraphic_E for the left. The site potential for charge e𝑒-e- italic_e is effectively higher on the right and serves as an electric field in the right direction that drives the particles toward the left. Similar simulations on classical transport using the MC dynamics with field gradient are implemented in the earlier works including Kawasaki dynamicsKawasaki (1966). In this framework, the current {\cal I}caligraphic_I can be defined as the number of particles passing through the periodic boundary divided by the total MCSs of the duration. We define the conductivity as

σe=/subscript𝜎𝑒\sigma_{e}={\cal I}/{\cal E}italic_σ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = caligraphic_I / caligraphic_E (4)

for a small enough electric field {\cal E}caligraphic_E that allows /{\cal I}/{\cal E}caligraphic_I / caligraphic_E to be nearly independent of {\cal E}caligraphic_E.

II.2.2 Diffusion of NIDW

The second method is to estimate the conductivity of NIDW, denoted as σdwsubscript𝜎dw\sigma_{\rm dw}italic_σ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT, assuming that each NIDW carries a charge of ±e/2plus-or-minus𝑒2\pm e/2± italic_e / 2. For this purpose, we calculate the population density of NIDW denoted as ndwsubscript𝑛dwn_{\rm dw}italic_n start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT and the diffusion constant D𝐷Ditalic_D. Here, the electric field is not applied. The following analysis is guaranteed when the NIDW shows diffusive dynamics, for which the Einstein relation holds. Our results indeed apply to this case.

To measure the charge density on each site in units of elementary charge e𝑒eitalic_e, we set the N-state as a charge-neutral vacuum state, i.e. with net zero charge for both D and A sites which are doubly occupied by particles and vacant, respectively. Figure 1(b) summarizes the realization of states and their charge densities. In the I state, each of the D and A sites is occupied by a single particle, which correspond to the charge density of +11+1+ 1 and 11-1- 1, respectively, where the minus sign denotes the “hole” measured from a vacuum. During the simulation, the D site can sometimes be vacant and the A site be doubly occupied which are denoted as +22+2+ 2 and 22-2- 2, respectively. The ones with ±2plus-or-minus2\pm 2± 2 are called ”highly ionic” (I2) and because their energies are much higher than other configurations observed rarely but with a finite probability (see Fig. 1(c)). The charge density on each bond is given by the average of charges of both sides, which can take the values of 00, ±1/2plus-or-minus12\pm 1/2± 1 / 2 and ±1plus-or-minus1\pm 1± 1. The amount of charges carried by the NIDWs is ±1/2plus-or-minus12\pm 1/2± 1 / 2. Accordingly, the bond that carries ±1plus-or-minus1\pm 1± 1 charge has two NIDWs. The NIDWs are created in pairs at some bond and are assigned ±1/2plus-or-minus12\pm 1/2± 1 / 2 charges as shown in Fig. 1(d). Once created, the NIDW randomly moves around until it meets another NIDW charged with the opposite sign and they disappear in pairs. By analyzing the trajectory of all the NIDWs created during the whole MC process, we can obtain ndwsubscript𝑛dwn_{\rm dw}italic_n start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT and D𝐷Ditalic_D as we explain in the following.

Each NIDW has their own lifetime, and suppose that τavesubscript𝜏ave\tau_{\rm ave}italic_τ start_POSTSUBSCRIPT roman_ave end_POSTSUBSCRIPT is the value averaged over all NIDWs that appear during the entire MCSs. The total number of NIDW generated is measured as Navetotsuperscriptsubscript𝑁avetot{N}_{\rm ave}^{\rm tot}italic_N start_POSTSUBSCRIPT roman_ave end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_tot end_POSTSUPERSCRIPT, regardless of the lifetime of NIDWs. Using these values, the NIDW density is defined as the average number density of existing NIDWs at a certain point in time as

ndw=NavetotτaveNMCS1N.subscript𝑛dwsuperscriptsubscript𝑁avetotsubscript𝜏avesubscript𝑁MCS1𝑁n_{\rm dw}={N}_{\rm ave}^{\rm tot}\frac{\tau_{\rm ave}}{N_{\rm MCS}}\frac{1}{N}.italic_n start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT roman_ave end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_tot end_POSTSUPERSCRIPT divide start_ARG italic_τ start_POSTSUBSCRIPT roman_ave end_POSTSUBSCRIPT end_ARG start_ARG italic_N start_POSTSUBSCRIPT roman_MCS end_POSTSUBSCRIPT end_ARG divide start_ARG 1 end_ARG start_ARG italic_N end_ARG . (5)

It is natural to expect that the dynamics of NIDW follow the standard diffusion equation, in which case the following relationship applies;

D=limtxt22t.𝐷subscript𝑡delimited-⟨⟩superscriptsubscript𝑥𝑡22𝑡D=\lim_{t\rightarrow\infty}\frac{\langle x_{t}^{2}\rangle}{2t}.italic_D = roman_lim start_POSTSUBSCRIPT italic_t → ∞ end_POSTSUBSCRIPT divide start_ARG ⟨ italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ end_ARG start_ARG 2 italic_t end_ARG . (6)

Here, xtsubscript𝑥𝑡x_{t}italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is the displacement of NIDW from the generated position in a unit of lattice constant after the elapsed time t𝑡titalic_t, and xt2delimited-⟨⟩superscriptsubscript𝑥𝑡2\langle x_{t}^{2}\rangle⟨ italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ is the mean square displacement (MSD). The sample average delimited-⟨⟩\langle\cdot\rangle⟨ ⋅ ⟩ is taken over all the NIDWs trajectories with the lifetime τ>t𝜏𝑡\tau>titalic_τ > italic_t. Once we obtain the diffusion constant, the mobility μ𝜇\muitalic_μ of NIDW is obtained by applying Einstein relation as

μ=eD2kBT.𝜇𝑒𝐷2subscript𝑘𝐵𝑇\mu=\frac{eD}{2k_{B}T}.italic_μ = divide start_ARG italic_e italic_D end_ARG start_ARG 2 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG . (7)

Here, only for these equations we introduce explicitly the elementary charge e=1𝑒1e=1italic_e = 1 for clarification, where we apply e/2𝑒2e/2italic_e / 2 as the absolute effective charge of NIDW. Accordingly, the electrical conductivity σdwsubscript𝜎dw\sigma_{\rm dw}italic_σ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT is calculated as

σdw=ndwe2μ=e2ndwD4kBT.subscript𝜎dwsubscript𝑛dw𝑒2𝜇superscript𝑒2subscript𝑛dw𝐷4subscript𝑘𝐵𝑇\sigma_{\rm dw}=n_{\rm dw}\frac{e}{2}\mu=\frac{e^{2}n_{\rm dw}D}{4k_{B}T}.italic_σ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT divide start_ARG italic_e end_ARG start_ARG 2 end_ARG italic_μ = divide start_ARG italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT italic_D end_ARG start_ARG 4 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG . (8)
Refer to caption
Figure 2: (a) Finite temperature phase diagram of Eq.(1) at U=7.5𝑈7.5U=7.5italic_U = 7.5 and V=3.5𝑉3.5V=3.5italic_V = 3.5. (b) Free energy landscape F(ρ;T,Δ)𝐹𝜌𝑇ΔF(\rho;T,\Delta)italic_F ( italic_ρ ; italic_T , roman_Δ ) at N=128𝑁128N=128italic_N = 128 as a function of ρ𝜌\rhoitalic_ρ for three different phases that have minimum point at ρ0=0,1subscript𝜌001\rho_{0}=0,1italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0 , 1 and 0.6250.6250.6250.625. (c) The location of the free-energy minimum, ρ0subscript𝜌0\rho_{0}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, for T=0.65,0.70,0.80𝑇0.650.700.80T=0.65,0.70,0.80italic_T = 0.65 , 0.70 , 0.80, and 0.950.950.950.95 at N=128𝑁128N=128italic_N = 128.
Refer to caption
Figure 3: (a) Mdwsubscript𝑀dwM_{\rm dw}italic_M start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT in Eq.(10) numerically obtained and averaged as a function of t𝑡titalic_t (time step after NIDW was created) for Δ=1.17Δ1.17\Delta=1.17roman_Δ = 1.17, T=0.5𝑇0.5T=0.5italic_T = 0.5 and N=1024𝑁1024N=1024italic_N = 1024. The shaded region is the contribution from the short-lifetime NIDWs showing pair-recombination, and the broken line is Eq.(12) with τdw=2.19×103subscript𝜏dw2.19superscript103\tau_{\rm dw}=2.19\times 10^{3}italic_τ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT = 2.19 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT and Ndwtot=3.39×106superscriptsubscript𝑁dwtot3.39superscript106N_{\rm dw}^{\rm tot}=3.39\times 10^{6}italic_N start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_tot end_POSTSUPERSCRIPT = 3.39 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT. Inset shows the corresponding values of Ndwsubscript𝑁dwN_{\rm dw}italic_N start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT, which is the lifetime distribution of NIDWs. The broken line in the inset is the fitted curve using Eq.(11). (b) xt2delimited-⟨⟩superscriptsubscript𝑥𝑡2\langle x_{t}^{2}\rangle⟨ italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ as a function of t𝑡titalic_t for Δ=1.17Δ1.17\Delta=1.17roman_Δ = 1.17, T=0.5𝑇0.5T=0.5italic_T = 0.5 and N=1024𝑁1024N=1024italic_N = 1024. The broken line is the result of the data fitting with Eq.(6) in the range of t1.5×104𝑡1.5superscript104t\geq 1.5\times 10^{4}italic_t ≥ 1.5 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT, where we find D=3.5×102𝐷3.5superscript102D=3.5\times 10^{-2}italic_D = 3.5 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT

.

III Results

III.1 Finite temperature phase diagram

We first present the phase diagram on the plane of T𝑇Titalic_T and ΔΔ\Deltaroman_Δ in Fig. 2(a). At zero temperature, the NI transition takes place at Δ=U2VΔ𝑈2𝑉\Delta=U-2Vroman_Δ = italic_U - 2 italic_V, which is determined by the difference in the ground state energies of the N and I phases. At finite temperature, the I phase is expected to be stabilized by the additional entropic term originating from the spin degeneracy of CN2NsubscriptsubscriptC𝑁2𝑁{}_{N}\mathrm{C}_{\frac{N}{2}}start_FLOATSUBSCRIPT italic_N end_FLOATSUBSCRIPT roman_C start_POSTSUBSCRIPT divide start_ARG italic_N end_ARG start_ARG 2 end_ARG end_POSTSUBSCRIPT. Assuming that both I and N phases are homogeneous, free of any domains or defects, the phase boundary in the thermodynamic limit is obtained as

Δc=(U2V)+(2ln(2))Tc,subscriptΔ𝑐𝑈2𝑉22subscript𝑇𝑐\Delta_{c}=\left(U-2V\right)+\left(2\ln{2}\right)T_{c},roman_Δ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = ( italic_U - 2 italic_V ) + ( 2 roman_ln ( start_ARG 2 end_ARG ) ) italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT , (9)

which is shown in the broken line in the phase diagram.

To unbiasedly elucidate the phase diagram, we perform an MC calculation and take the histogram of distribution of states against ρ𝜌\rhoitalic_ρ for given T𝑇Titalic_T and ΔΔ\Deltaroman_Δ, denoted as H(ρ;T,Δ)𝐻𝜌𝑇ΔH(\rho;T,\Delta)italic_H ( italic_ρ ; italic_T , roman_Δ ). Because the distribution scales with the partition function, we can safely obtain the free energy profile as, F(ρ;T,Δ)=TlnH(ρ;T,Δ)+C𝐹𝜌𝑇Δ𝑇𝐻𝜌𝑇Δ𝐶F(\rho;T,\Delta)=-T\ln H(\rho;T,\Delta)+Citalic_F ( italic_ρ ; italic_T , roman_Δ ) = - italic_T roman_ln italic_H ( italic_ρ ; italic_T , roman_Δ ) + italic_C, and then we can determine the value of the unknown constant C𝐶Citalic_C using the value of free energy density of N phase, f(ρ=0,T,Δ)=U/2𝑓𝜌0𝑇Δ𝑈2f(\rho=0,T,\Delta)=U/2italic_f ( italic_ρ = 0 , italic_T , roman_Δ ) = italic_U / 2. The free energy density for each value of ρ𝜌\rhoitalic_ρ is shown in Fig. 2(b).

At low temperatures, a two-minima structure of the free energy is observed. However, at relatively high temperatures near the NI phase boundary, three local minima appear, which means that another metastable phase appears and compete with the N and I phases. By analyzing these free energy landscapes, we extract the value ρ=ρ0𝜌subscript𝜌0\rho=\rho_{0}italic_ρ = italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT that has the lowest value of F(ρ;T,Δ)𝐹𝜌𝑇ΔF(\rho;T,\Delta)italic_F ( italic_ρ ; italic_T , roman_Δ ) as the thermal equilibrium state and obtain the phase diagram, where we denote the three phases as N phase (ρ0=0subscript𝜌00\rho_{0}=0italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0), I phase (ρ0=1subscript𝜌01\rho_{0}=1italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1) and NIDW phase (0<ρ0<10subscript𝜌010<\rho_{0}<10 < italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT < 1). The numerically obtained NI phase boundary at low temperatures almost coincides with the broken line, Eq.(9), within the error bars and are not shown for simplicity.

The phase boundaries between the NIDW-phase and the other two phases obtained by the MC calculations are shown for N=32,64,128𝑁3264128N=32,64,128italic_N = 32 , 64 , 128, and 256256256256. We extrapolate these results by the polynomial fitting and obtain the boundaries for N=512,1024𝑁5121024N=512,1024italic_N = 512 , 1024, and the bulk limit N𝑁N\rightarrow\inftyitalic_N → ∞ as a reference for the dynamical simulation we see shortly. We find in the free energy landscape that these transition lines are of first order, have two endpoints, and meet at the triple point reminiscent of the water. Figure 2(b) shows the example of three cases, ρ0=0,1subscript𝜌001\rho_{0}=0,1italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0 , 1 and 0.6250.6250.6250.625, which correspond to the N, I, and NIDW phases, respectively. At high temperatures above the two endpoints, these phases show crossover behavior.

By extracting ρ0subscript𝜌0\rho_{0}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT which gives the minimum of f(ρ;T,Δ)𝑓𝜌𝑇Δf(\rho;T,\Delta)italic_f ( italic_ρ ; italic_T , roman_Δ ), we obtain Fig. 2(c) as functions of ΔΔ\Deltaroman_Δ for several choices of T𝑇Titalic_T. When crossing the first order transition line, ρ0subscript𝜌0\rho_{0}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT shows a clear discontinuity, which does not vanish with increasing system size, as we confirm by the extrapolation curves in the phase diagram. The values of jump decrease as the state approaches the endpoints and we find continuous changes with ΔΔ\Deltaroman_Δ to the I phase at T=0.804𝑇0.804T=0.804italic_T = 0.804 and to I and N phases at T=0.95𝑇0.95T=0.95italic_T = 0.95, which are the temperatures of the two endpoints for N=128𝑁128N=128italic_N = 128.

III.2 Diffusion of NIDWs

III.2.1 Two types of NIDWs

In evaluating ndwsubscript𝑛dwn_{\rm dw}italic_n start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT, it is useful to know how observables related to each NIDWs behave before evaluating Eq.(5). For this purpose, we first consider the lifetime distribution of NIDWs denoted as Ndw(t)subscript𝑁dw𝑡N_{\rm dw}(t)italic_N start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT ( italic_t ), and the cumulative lifetime distribution is defined as

Mdw(t)=t=t+1Ndw(t).subscript𝑀dw𝑡superscriptsubscriptsuperscript𝑡𝑡1subscript𝑁dwsuperscript𝑡M_{\rm dw}(t)=\sum_{t^{\prime}=t+1}^{\infty}{N}_{\rm dw}(t^{\prime}).italic_M start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_t + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) . (10)

Here t𝑡titalic_t is the time step measured for each NIDW from its creation, where we take into account all of the NIDWs that appeared during the measurements over the entire MCSs, i.e. over NMCS109similar-tosubscript𝑁MCSsuperscript109N_{\rm MCS}\sim 10^{9}italic_N start_POSTSUBSCRIPT roman_MCS end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT steps. Figures 3(a) and 3(b) show Mdwsubscript𝑀dw{M}_{\rm dw}italic_M start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT, Ndwsubscript𝑁dw{N}_{\rm dw}italic_N start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT and x2delimited-⟨⟩superscript𝑥2\langle x^{2}\rangle⟨ italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ as functions of t𝑡titalic_t. Some of the NIDWs recombine with the counterpart at its creation after a short timescale and disappear, and since they are localized, they do not contribute to the conductivity. We call it “pair-recombination”, which is related to the case observed in the 1D random walk problem; the number of walkers that take timescale τ𝜏\tauitalic_τ to return for the first time to its origin is known as τ32proportional-toabsentsuperscript𝜏32\propto\tau^{-\frac{3}{2}}∝ italic_τ start_POSTSUPERSCRIPT - divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT. Therefore, the distribution of localized NIDWs and those of longer lifetimes differ. At the same time, without considering this pair recombination, we can naturally expect that there is a characteristic lifetime of NIDWs τdwsubscript𝜏dw\tau_{\rm dw}italic_τ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT that may depend on temperature or model parameters, and the number of NIDWs will follow et/τdwproportional-toabsentsuperscript𝑒𝑡subscript𝜏dw\propto e^{-t/\tau_{\rm dw}}∝ italic_e start_POSTSUPERSCRIPT - italic_t / italic_τ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT end_POSTSUPERSCRIPT. By combining these two factors, we expect

Ndw(t)(const.+t32)etτdw.N_{\rm dw}(t)\propto({\rm const.}+t^{-\frac{3}{2}})e^{-\frac{t}{\tau_{\rm dw}}}.italic_N start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT ( italic_t ) ∝ ( roman_const . + italic_t start_POSTSUPERSCRIPT - divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT ) italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_t end_ARG start_ARG italic_τ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT end_ARG end_POSTSUPERSCRIPT . (11)

The inset of Fig. 3(a) shows that Ndwsubscript𝑁dwN_{\rm dw}italic_N start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT is fitted very well with Eq.(11). Because the second term from the pair-recombination will be suppressed at a long enough timescale in integrating Mdw(t)t(const.+t32)etτdwdtM_{\rm dw}(t)\propto\int_{t}^{\infty}\left({\rm const.}+t^{-\frac{3}{2}}\right% )e^{-\frac{t}{\tau_{\rm dw}}}dtitalic_M start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT ( italic_t ) ∝ ∫ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( roman_const . + italic_t start_POSTSUPERSCRIPT - divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT ) italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_t end_ARG start_ARG italic_τ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT end_ARG end_POSTSUPERSCRIPT italic_d italic_t, we expect

Mdw(t)Ndwtotet/τdw,(tτdw).similar-to-or-equalssubscript𝑀dw𝑡superscriptsubscript𝑁dwtotsuperscript𝑒𝑡subscript𝜏dwgreater-than-or-equivalent-to𝑡subscript𝜏dwM_{\rm dw}(t)\simeq N_{\rm dw}^{\rm tot}e^{-t/\tau_{\rm dw}},\quad(t\gtrsim% \tau_{\rm dw}).italic_M start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT ( italic_t ) ≃ italic_N start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_tot end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_t / italic_τ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , ( italic_t ≳ italic_τ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT ) . (12)

In Fig. 3(a), we plot Mdw(t)subscript𝑀dw𝑡M_{\rm dw}(t)italic_M start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT ( italic_t ) obtained by our MC calculation for N=1024𝑁1024N=1024italic_N = 1024, Δ=1.17Δ1.17\Delta=1.17roman_Δ = 1.17 and T=0.5𝑇0.5T=0.5italic_T = 0.5, near the boundary of the NI transition. We find that Eq.(12) gives a good approximation, which gives τdw=2.19×103subscript𝜏dw2.19superscript103\tau_{\rm dw}=2.19\times 10^{3}italic_τ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT = 2.19 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT and Ndwtot=3.39×106superscriptsubscript𝑁dwtot3.39superscript106N_{\rm dw}^{\rm tot}=3.39\times 10^{6}italic_N start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_tot end_POSTSUPERSCRIPT = 3.39 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT. The physical implication of these parameters is the typical lifetime and the total number of NIDWs created during the entire MCSs, which safely avoided the pair-recombination. We notice that within our calculation, the maximum displacement of NIDWs is about less than 100 in a unit of lattice constant. We have prepared the system size of N=512,1024𝑁5121024N=512,1024italic_N = 512 , 1024, which is much larger than these displacements.

Previously, we have introduced ndwsubscript𝑛dwn_{\rm dw}italic_n start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT in Eq.(5) by using the values τavesubscript𝜏𝑎𝑣𝑒\tau_{ave}italic_τ start_POSTSUBSCRIPT italic_a italic_v italic_e end_POSTSUBSCRIPT and Navetotsuperscriptsubscript𝑁𝑎𝑣𝑒totN_{ave}^{\rm tot}italic_N start_POSTSUBSCRIPT italic_a italic_v italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_tot end_POSTSUPERSCRIPT averaged over all NIDWs. However, we want to extract the NIDWs that contribute to the conduction, and accordingly, we adopt the following formula which excludes the effect of pair recombination as

ndw=NdwtotτdwNMCS1N.subscript𝑛dwsuperscriptsubscript𝑁dwtotsubscript𝜏dwsubscript𝑁MCS1𝑁n_{\rm dw}=N_{\rm dw}^{\rm tot}\frac{\tau_{\rm dw}}{N_{\rm MCS}}\frac{1}{N}.italic_n start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_tot end_POSTSUPERSCRIPT divide start_ARG italic_τ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT end_ARG start_ARG italic_N start_POSTSUBSCRIPT roman_MCS end_POSTSUBSCRIPT end_ARG divide start_ARG 1 end_ARG start_ARG italic_N end_ARG . (13)

Figure 3(b) shows xt2delimited-⟨⟩superscriptsubscript𝑥𝑡2\langle x_{t}^{2}\rangle⟨ italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ as a function of t𝑡titalic_t. As mentioned earlier, as long as NIDW follows the normal diffusive mechanism, xt2delimited-⟨⟩superscriptsubscript𝑥𝑡2\langle x_{t}^{2}\rangle⟨ italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ becomes proportional to t𝑡titalic_t. Indeed, when t104greater-than-or-equivalent-to𝑡superscript104t\gtrsim 10^{4}italic_t ≳ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT, it extrapolates to a linear line of t𝑡titalic_t as xt20.07tsimilar-todelimited-⟨⟩superscriptsubscript𝑥𝑡20.07𝑡\langle x_{t}^{2}\rangle\sim 0.07t⟨ italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ ∼ 0.07 italic_t and the diffusion constant D𝐷Ditalic_D can be extracted as D=3.5×102𝐷3.5superscript102D=3.5\times 10^{-2}italic_D = 3.5 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT.

Refer to caption
Figure 4: Number density of NIDWs ndwsubscript𝑛dwn_{\rm dw}italic_n start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT and diffusion constant D𝐷Ditalic_D obtained by a set of analysis given in Fig. 3 at T=0.4,0.5𝑇0.40.5T=0.4,0.5italic_T = 0.4 , 0.5 and N=512,1024𝑁5121024N=512,1024italic_N = 512 , 1024 as functions of ΔΔ\Deltaroman_Δ. The dotted line shows the NI transition point ΔcsubscriptΔ𝑐\Delta_{c}roman_Δ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for each temperature calculated by Eq.(9), derived by assuming the homogeneous states of I- and N- phases.

III.2.2 Diffusion constant and number density

Figure 4(a) shows the ΔΔ\Deltaroman_Δ dependence of ndwsubscript𝑛dwn_{\rm dw}italic_n start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT obtained from Eq.(13) for T=0.4,0.5𝑇0.40.5T=0.4,0.5italic_T = 0.4 , 0.5 and N=512,1024𝑁5121024N=512,1024italic_N = 512 , 1024. At T=0.4𝑇0.4T=0.4italic_T = 0.4, ndwsubscript𝑛dwn_{\rm dw}italic_n start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT shows a sharp peak at around ΔcsubscriptΔ𝑐\Delta_{c}roman_Δ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, indicating that the the number of NIDWs increases significantly toward the transition from both sides, signaling the first-order transition between the N and I phases. At T=0.5𝑇0.5T=0.5italic_T = 0.5, the emergent NIDW phase mentioned in Sec.III.1 suppresses these peak structures and the number of NIDWs increases, taking a maximum at Δm1.22similar-to-or-equalssubscriptΔm1.22\Delta_{\text{m}}\simeq 1.22roman_Δ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT ≃ 1.22. As we see from the phase diagram in Fig. 2(a), ΔmsubscriptΔm\Delta_{\text{m}}roman_Δ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT falls on the NIDW-to-N phase boundary, and is off the ideal NI phase boundary line, Δc=1.19subscriptΔ𝑐1.19\Delta_{c}=1.19roman_Δ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 1.19, shown in broken line.

Figure 4(b) shows the corresponding values of D𝐷Ditalic_D obtained from xt2delimited-⟨⟩superscriptsubscript𝑥𝑡2\langle x_{t}^{2}\rangle⟨ italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ using Eq.(6). We find a suppression of D𝐷Ditalic_D at around ΔcsubscriptΔ𝑐\Delta_{c}roman_Δ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. The value of D𝐷Ditalic_D is larger for the I phase than the N phase and takes the minimum at Δmsimilar-toabsentsubscriptΔm\sim\Delta_{\text{m}}∼ roman_Δ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT, i.e. the same point as the maximum of ndwsubscript𝑛dwn_{\rm dw}italic_n start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT.

Refer to caption
Figure 5: Conductivity σesubscript𝜎𝑒\sigma_{e}italic_σ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT and σdwsubscript𝜎dw\sigma_{\rm dw}italic_σ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT derived from two different frameworks at T=0.4,0.5𝑇0.40.5T=0.4,0.5italic_T = 0.4 , 0.5 and N=512,1024𝑁5121024N=512,1024italic_N = 512 , 1024 as functions of ΔΔ\Deltaroman_Δ. The dotted line shows the NI transition point ΔcsubscriptΔ𝑐\Delta_{c}roman_Δ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for each temperature calculated by Eq.(9), derived by assuming the homogeneous states of I- and N- phases.

III.3 Electrical conductivity

Finally, we derive the conductivity calculated by the two MC schemes in Sec.II.2. Figure 5 shows σesubscript𝜎𝑒\sigma_{e}italic_σ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT and σdwsubscript𝜎dw\sigma_{\rm dw}italic_σ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT for T=0.4,0.5𝑇0.40.5T=0.4,0.5italic_T = 0.4 , 0.5 for the two system sizes. Here, σesubscript𝜎𝑒\sigma_{e}italic_σ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT obtained from Eq.(4) has roughly twice as large amplitudes as σdwsubscript𝜎dw\sigma_{\rm dw}italic_σ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT evaluated via Eq.(8) based on Einstein relation, while except for this factor-2, they agree with each other. All these results show a clear enhancement of σe/dwsubscript𝜎𝑒dw\sigma_{e/\text{dw}}italic_σ start_POSTSUBSCRIPT italic_e / dw end_POSTSUBSCRIPT at around the phase boundary, successfully reproducing the experimentally predicted enhancement of electrical conductivity. The maximum of σe/dwsubscript𝜎𝑒dw\sigma_{e/\text{dw}}italic_σ start_POSTSUBSCRIPT italic_e / dw end_POSTSUBSCRIPT takes place at ΔesubscriptΔ𝑒\Delta_{e}roman_Δ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT which is very close to the maximum/minimum position ΔmsubscriptΔm\Delta_{\text{m}}roman_Δ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT of ndw/Dsubscript𝑛dw𝐷n_{\text{dw}}/Ditalic_n start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT / italic_D.

Refer to caption
Figure 6: Energy gaps of NIDW for our two calculation schemes are plotted to modified ionic potential, ΔΔmΔsubscriptΔm\Delta-\Delta_{\text{m}}roman_Δ - roman_Δ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT (ΔmsubscriptΔm\Delta_{\text{m}}roman_Δ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT is the location where σ𝜎\sigmaitalic_σ takes the maximum and ΔeΔmsimilar-tosubscriptΔesubscriptΔm\Delta_{\text{e}}\sim\Delta_{\text{m}}roman_Δ start_POSTSUBSCRIPT e end_POSTSUBSCRIPT ∼ roman_Δ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT), calculated by using Arrhenius plots of the electric conductivities with referring to Takehara et al. (2019)

III.4 Energy gap of NIDW

We finally evaluate the activation gap of the conductivity from the obtained conductivities. We adopt the Arrhenius plot of σesubscript𝜎𝑒\sigma_{e}italic_σ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT and σdwsubscript𝜎dw\sigma_{\text{dw}}italic_σ start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT for T=0.400.48𝑇0.400.48T=0.40-0.48italic_T = 0.40 - 0.48 at N=256𝑁256N=256italic_N = 256, as shown in Fig. 6(a). The activation gap obtained as in Fig. 6(b) shows very rapidly suppressed toward ΔesubscriptΔ𝑒\Delta_{e}roman_Δ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT from both sides. The minimum value of the gap is Eg=5.1±0.2subscript𝐸𝑔plus-or-minus5.10.2E_{g}=5.1\pm 0.2italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = 5.1 ± 0.2 for σesubscript𝜎𝑒\sigma_{e}italic_σ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT and Eg=3.9±0.3subscript𝐸𝑔plus-or-minus3.90.3E_{g}=3.9\pm 0.3italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = 3.9 ± 0.3 for σdwsubscript𝜎dw\sigma_{\rm dw}italic_σ start_POSTSUBSCRIPT roman_dw end_POSTSUBSCRIPT.

IV Summary and discussion

We analyzed the dynamics of the classical one-dimensional model featuring neutral-ionic(NI) transition based on the Monte Carlo(MC) simulation to clarify the origin of the enhancement of electrical conductivity observed in the experiments of TTF-CA. Our model corresponds to the strong coupling limit (t/U0𝑡𝑈0t/U\rightarrow 0italic_t / italic_U → 0) of the ionic extended Hubbard model. We focus on the degree of freedom called neutral ionic domain wall (NIDW), which is the bonds that have average ±e/2plus-or-minus𝑒2\pm e/2± italic_e / 2 charges on two adjacent sites measured from the neutral configuration. Although naively the decrease of Δ/(U2V)Δ𝑈2𝑉\Delta/(U-2V)roman_Δ / ( italic_U - 2 italic_V ) will drive the N-to-I transition as understood from the previous studies, the finite temperature phase diagram shows another thermodynamically stable phase, the NIDW phase, the mixture of local N and I states separated by the domain walls, in between the N and I phases when the temperature is higher than the triple point. At this point, the NI phase boundary branches to the two N-to-NIDW and NIDW-to-I phase boundaries which are of first order up to the endpoint where the jump in the degree of charge transfer ρ𝜌\rhoitalic_ρ vanishes; this ρ𝜌\rhoitalic_ρ was determined as the global minima of the free energy obtained by the histogram MC calculation.

Based on this phase diagram, we performed the MC calculation in the vicinity of the phase boundaries to analyze the dynamics of NIDWs by applying two schemes, the Kawasaki-like dynamics that apply the electric field as gradients of the site potentials, tracking the motion of particles, and the Glauber dynamics in a zero field tracking the NIDWs from their pair creation to the annihilation. The ΔΔ\Deltaroman_Δ dependence of the two electric conductivities, σesubscript𝜎𝑒\sigma_{e}italic_σ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT and σdwsubscript𝜎dw\sigma_{\text{dw}}italic_σ start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT, obtained from the two schemes agree well. In particular, the one at a temperature slightly higher than the triple point shows an enhancement at ΔmsubscriptΔm\Delta_{\text{m}}roman_Δ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT that approximately falls on the NIDW-to-N phase boundary. Their values are about five times larger than the ones at the NI phase boundary at low temperatures, which cannot be simply explained by the σdwT1proportional-tosubscript𝜎dwsuperscript𝑇1\sigma_{\text{dw}}\propto T^{-1}italic_σ start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT ∝ italic_T start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT dependence expected for metals.

Let us expand the discussion on the nature of activated NIDW conductivity in more detail. In our results, the conductivity σesubscript𝜎𝑒\sigma_{e}italic_σ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT of charge e𝑒eitalic_e of the particles driven by the electric field gives the same model-parameter dependence as the conductivity σdwsubscript𝜎dw\sigma_{\text{dw}}italic_σ start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT of the NIDWs carrying ±e/2plus-or-minus𝑒2\pm e/2± italic_e / 2 charges. Therefore, it is natural to understand the origin of this conduction as the diffusive motion of NIDWs. Here, σdwsubscript𝜎dw\sigma_{\text{dw}}italic_σ start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT is the product of the number of active domain walls ndwsubscript𝑛dwn_{\text{dw}}italic_n start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT and the diffusion constant D𝐷Ditalic_D, where we find that ndw/Dsubscript𝑛dw𝐷n_{\text{dw}}/Ditalic_n start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT / italic_D increases/decreases at ΔeΔmsimilar-tosubscriptΔ𝑒subscriptΔm\Delta_{e}\sim\Delta_{\text{m}}roman_Δ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ∼ roman_Δ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT; the increase of ndwsubscript𝑛dwn_{\text{dw}}italic_n start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT is dominant and enhances σdwsubscript𝜎dw\sigma_{\text{dw}}italic_σ start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT. At the same time, the activation gap evaluated from the conductivity shows a significant decrease at ΔmsubscriptΔm\Delta_{\text{m}}roman_Δ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT. The physical implication of these results is that the number of NIDWs is activated by the suppression of the excitation gap, while the mobility of NIDW is suppressed. The mean distance between NIDWs is too large to interact with each other even at ΔΔmsimilar-toΔsubscriptΔm\Delta\sim\Delta_{\text{m}}roman_Δ ∼ roman_Δ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT, it is unlikely that the suppression of mobility is caused by the interactions between NIDWs. However, the true object that moves around is the particle, which suffers a competition between N and I configuration particularly near the NIDW, which makes the NIDWs less mobile.

Such an effect seems more distinct when the system is closer to the N phase than the I phase inside the NIDW phase, because ΔmsubscriptΔm\Delta_{\text{m}}roman_Δ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT is at the NIDW-to-N boundary. One possible cause could be the increasing number of immobile structures interrupting the movements of NIDWs, which is the “parallel-spin I-domain”; the two adjacent D and A sites are occupied by a single particle with the same spin orientation, which cannot hop to each other due to the Pauli blocking effect. We indeed observe the duration of dynamics and find that the parallel-spin I-domains interrupt the movements of NIDWs. Because the N phase is nonmagnetic, it is much easier to create such a structure stochastically at the repetitive creation and annihilation of NIDWs.

We now discuss the relevance to the experimental findings. The temperature-pressure phase diagram of the TTF-CA consists of three phases, N phase, Iparapara{}_{\text{para}}start_FLOATSUBSCRIPT para end_FLOATSUBSCRIPT phase and Iferroferro{}_{\text{ferro}}start_FLOATSUBSCRIPT ferro end_FLOATSUBSCRIPT phase. At ambient pressure, the N phase at high temperature undergoes a direct first-order phase transition to the Iferroferro{}_{\text{ferro}}start_FLOATSUBSCRIPT ferro end_FLOATSUBSCRIPT phase observed by the kink in the resistivity. The NQR shows the shift and split of peaks signalling the charge transfer as well as the dimerization, and there, the ferroelectricity can happen when the inversion symmetry is broken by the lattice distortion which makes the charge transfer more rigid. When increasing pressure, the high-temperature N phase crossovers to the Iparapara{}_{\text{para}}start_FLOATSUBSCRIPT para end_FLOATSUBSCRIPT phase, because the value of ΔΔ\Deltaroman_Δ is considered to be effectively suppressed. The Iparapara{}_{\text{para}}start_FLOATSUBSCRIPT para end_FLOATSUBSCRIPT phase is basically a Mott insulator at high temperature which is paramagnetic, and the ferroelectricity is absent. The lowering of temperature drives the spin-Peierls transition and the system transform to the dimerized Iferroferro{}_{\text{ferro}}start_FLOATSUBSCRIPT ferro end_FLOATSUBSCRIPT phase. The recent experiment reports a strong enhancement of conductivity by one order of magnitude from the one in the Iparapara{}_{\text{para}}start_FLOATSUBSCRIPT para end_FLOATSUBSCRIPT phase at around the N-Iparapara{}_{\text{para}}start_FLOATSUBSCRIPT para end_FLOATSUBSCRIPT crossover regionTakehara et al. (2019). In this region, the infrared spectroscopy shows the active agsubscript𝑎𝑔a_{g}italic_a start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT mode related to the dimerizationOkamoto et al. (1989); Masino et al. (2007) which was interpreted as dimer liquid, a fluctuating dimerization. The detailed analysis of the temperature dependence of conductivity along the NI crossover line shows an activating behaviorSunami et al. (2022), whose activation energy is 0.055similar-toabsent0.055\sim 0.055∼ 0.055 eV, one order of magnitude smaller than the charge-transfer excitation energy 0.60.70.60.70.6-0.70.6 - 0.7 eV. The average ndwsubscript𝑛dwn_{\text{dw}}italic_n start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT distribution is derived as one NIDW per 10 sites.

The present results may apply to the NI crossover region of the TTF-CA phase diagram, T280300similar-to𝑇280300T\sim 280-300italic_T ∼ 280 - 300 K and P810similar-to𝑃810P\sim 8-10italic_P ∼ 8 - 10kbar. The energy scale of this temperature is 0.03similar-toabsent0.03\sim 0.03∼ 0.03 eV, which is smaller than t0.1similar-to𝑡0.1t\sim 0.1italic_t ∼ 0.1eV Ishibashi and Terakura (2010) and the band gap 0.15similar-toabsent0.15\sim 0.15∼ 0.15eV by few factors. This means that the insulating property is well-preserved and our classical approximation can at least capture the hydrodynamic nature of the activating types of conductivity driven by NIDWs. Experimentally, it has been discussed that the lattice dimerization that allows for another domain wall of spins, called spin soliton, are intrinsic to the conductivity as they emerge at the relatively high density of one soliton per 20-50 sites. This was explained as such that the NIDWs created in pairs from the I phase are recombined by the electric current rather than separate furtherSunami et al. (2022). However, our calculation shows that the stochastic process of creating the NIDWs are not as simple, and with the aid of their thermal fluctuation and correlation, and apart from the NIDWs that disappear by a pair-recombination at a relatively short lifetime, there are substantial portion of ndwsubscript𝑛dwn_{\text{dw}}italic_n start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT that contribute the conductivity. It is also notable that the σesubscript𝜎𝑒\sigma_{e}italic_σ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is basically larger than σdwsubscript𝜎dw\sigma_{\text{dw}}italic_σ start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT showing that the application of an electric field does not particularly support the pair-recombination, contrary to the previous intuitive claims. The present results indicate that the activated conductivity can solely be explained without the dimerization effect. Still, the effect of dimerization on the polarization and conductivity is an issue of wide interest not only in NI systems Egami et al. (1993); Resta and Sorella (1995). Our model can be extended further to include the effect of lattice distortion which would be our future perspective.

We finally refer to the implication of the diffusive transport observed in our model. Our conductivity relies on the local fluctuation of particles following Fick’s law, while neglecting the contributions from the heat transport and the couplings of particles with heat carriers. It is naturally considered that the electronic properties at high enough temperature with conductivity above the MIR bound are no longer controlled by quasiparticle physics, and the rapid momentum relaxation does not account for their transportHartnoll (2014). Such incoherent metal is dominated by the charge(classical particle) diffusion constants, to which the present model targeting the room temperature region of TTF-CA shall apply. The long-timescale behavior of NIDWs, xt2tproportional-todelimited-⟨⟩superscriptsubscript𝑥𝑡2𝑡\langle x_{t}^{2}\rangle\propto t⟨ italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ ∝ italic_t, supports this picture.

Let us remind that the relevant ionic Hubbard model with t,U,𝑡𝑈t,U,italic_t , italic_U , and ΔΔ\Deltaroman_Δ is reduced in the strong coupling limit to the spin-1 model with assisted spin-exchange (e.g. SjzSj+Sj+1subscriptsuperscript𝑆𝑧𝑗subscriptsuperscript𝑆𝑗subscriptsuperscript𝑆𝑗1S^{z}_{j}S^{+}_{j}S^{-}_{j+1}italic_S start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_S start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_S start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT terms) and the magnetic anisotropy (Sjz)2superscriptsubscriptsuperscript𝑆𝑧𝑗2(S^{z}_{j})^{2}( italic_S start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT termsLegeza et al. (2006). Our model corresponds to adding the V𝑉Vitalic_V term as SjzSj+1zsubscriptsuperscript𝑆𝑧𝑗subscriptsuperscript𝑆𝑧𝑗1S^{z}_{j}S^{z}_{j+1}italic_S start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_S start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT types of interactions and deleting the quantum fluctuation to this model (t=0𝑡0t=0italic_t = 0) to make them an extended three-state Potts model. Recently, the nature of dynamics of 1D quantum spin models at infinite-temperature is found to host three classes, depending on their long-time decay in the correlation function t1/zproportional-toabsentsuperscript𝑡1𝑧\propto t^{-1/z}∝ italic_t start_POSTSUPERSCRIPT - 1 / italic_z end_POSTSUPERSCRIPTDupont and Moore (2020); the diffusive case with dynamical exponent z=2𝑧2z=2italic_z = 2, ballistic one with z=1𝑧1z=1italic_z = 1, and the superdiffusive z=3/2𝑧32z=3/2italic_z = 3 / 2 ones. It was shown that most of the S=1𝑆1S=1italic_S = 1 isotropic and anisotropic quantum spin models behave diffusiveDupont and Moore (2020), in contrast to the S=1/2𝑆12S=1/2italic_S = 1 / 2 models with robust anomalous superdiffusive natureDe Nardis et al. (2021); Roy et al. (2023). Although to which class the system belongs depends much on the types of Hamiltonian, our S=1𝑆1S=1italic_S = 1 related model can fit this scenario, because its quantum version seriously loses the integrability and hosts high anisotropy. The class of transport may further change depending on what kind of terms are added to such effective Hamiltonian, e.g. the lattice dimerization that naturally arise in TTF-CA, which will be one of the interesting perspectives.

Acknowledgements

This work was supported by JST SPRING, Grant Number JPMJSP2108.

References