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

What can we learn from the experiment of electrostatic conveyor belt for excitons?

T. T. Zhao Key Laboratory for Microstructural Material Physics of Hebei Province, School of Science, Yanshan University, Qinhuangdao 066004, China    Rui Li Key Laboratory for Microstructural Material Physics of Hebei Province, School of Science, Yanshan University, Qinhuangdao 066004, China    C. S. Liu [email protected] Key Laboratory for Microstructural Material Physics of Hebei Province, School of Science, Yanshan University, Qinhuangdao 066004, China
Abstract

Motivated by the experiment of electrostatic conveyor belt for indirect excitons [A. G. Winbow, et al., Phys. Rev. Lett. 106, 196806 (2011)], we study the exciton patterns for understanding the exciton dynamics. By analyzing the exciton diffusion, we find that the patterns mainly come from the photoluminescence of two kinds of excitons. The patterns near the laser spot come from the hot excitons which can be regarded as the classical particles. However, the patterns far from the laser spot come from the cooled excitons or coherent excitons. Taking into account of the finite lifetime of Bosonic excitons and of the interactions between them, we build a time-dependent nonlinear Schrödinger equation including the non-Hermitian dissipation to describe the coherent exciton dynamics. The real-time and imaginary-time evolutions are used alternately to solve the Schrödinger equation in order to simulate the exciton diffusion accompanied with the exciton cooling in the moving lattices. By calculating the escape probability, we obtain the transport distances of the coherent excitons in the conveyor which are consistent with the experimental data. The cooling speed of excitons is found to be important in the coherent exciton transport. Moreover, the plateau in the average transport distance cannot be explained by the dynamical localization-delocalization transition induced by the disorders.

coupled quantum wells, indirect excitons, Bose-Einstein condensation, exciton transport, dynamical localization

I Introduction

A particle is usually called as a quantum walkers when its time evolution follows the rules of quantum mechanics. Standard quantum mechanics assumes Hermiticity of the Hamiltonian. However, non-Hermitiancity is a universal phenomenon in many fields of physics, such as the dissipative systems with gain and loss Ashida et al. (2021). As a result, the non-Hermitian quantum walker is of universal significance and has attracted attention recently Wang et al. (2021); Xue et al. (2022); Yuce and Ramezani (2023); Xiao et al. (2023). The direct observation of a quantum walker is very difficult whether it is in Hermitian or in non-Hermitian systems. Taking the particle in a one-dimensional periodic potential as an example, if the potential is in time, the number of the particle transport for the simplest quantum walker during the time cycle equals to the Chern number. The Chern number characterizes the topological structure of the model mapping from the parameter space of wave vector and time (k,tk,t) to the momentum space (kx,kyk_{x},k_{y}) Thouless (1983). The gedanken experiment considered by Thouless demands adiabatic approximation, i.e. very slowly changing potential. In particular, the particles must be the Fermions occupying in a filled Bloch bands at zero temperature.

The excitons, Bosonic quasiparticles, are the electron-hole bound pairs in semiconductor which are expected to realize Bose-Einstein condensation Keldysh and Kozlov (1968). Due to the finite lifetime, the exciton systems have the endogenously non-Hermitianity. Both the short lifetime and the low cooling rate hinder the realizing of the Bose-Einstein condensation. In order to overcome the above two disadvantages, the indirect excitons, i.e. the spatially separated electron-hole bound pairs, were generated in coupled quantum wells Butov et al. (2002a). As their neutral overall, they are harder to manipulate electrically. However, the energy of the indirect excitons can be controlled by voltage due to the dipole moment. By applying an alternating voltage to an electrode grid that covers the device, a wavelike potential was created for the excitons. The excitons are modulated by the wavelike potential which acts as the conveyor belt across the sample. Since the high density of excitons has the high luminous intensity, the exciton pattern allows us to track the location of the excitons. The exciton pattern in the moving lattices, or the conveyor belt, gives the opportunity to observe the non-Hermitian quantum walker.

The data of electrostatic conveyor belt for indirect excitons were reported in Ref. Winbow et al. (2011). [(a) and (b)], [(c) and (d)] and [(e) and (f)] in Fig. 1 show the xx-yy photoluminescence (PL) images, xx-energy PL images and xx- PL intensity respectively for conveyer off and on. The left column of the last row [Fig. 1 (g)] shows the average transport distance of indirect excitons via conveyer M1M_{1} as a function of the conveyer amplitude. Fig. 1 (h) shows the distance of indirect excitons via the conveyer M1M_{1} as a function of density. The first moment of the PL intensity M1=xI(x)𝑑x/I(x)𝑑xM_{1}=\int xI(x)dx/\int I(x)dx, which characterizes the average transport distance of the indirect excitons via conveyer. I(x)I(x) is the PL intensity profile obtained by the integration of the xx-energy images over the emission wavelength. Major features of the indirect exciton transport are summarized as follow. (i) There exist the dynamical localization-delocalization transitions. (ii) Crossing the transition point, the transport distance increases with the conveyer amplitude and tends to saturation. (iii) The exciton transport is less efficient for higher velocity. (iv) The efficient exciton transport via the conveyer only occurs at intermediate densities. (v) Several bright stripes are shown in the PL images.

Refer to caption
Refer to caption
Refer to caption
Figure 1: The figures are taken from Ref. Winbow et al. (2011). (a-d) xyx-y and xenergyx-energy PL images, and (e, f) PL intensity profiles I(x)I(x) for conveyer off and on here Pex=20P_{\mathrm{ex}}=20 μ\muW, Aconv=7.5A_{\mathrm{conv}}=7.5 meV, vconv=0.7v_{\mathrm{conv}}=0.7 μ\mum/ns. The blue dashed lines in (c) and (e) are used to approximately separate the hot excitons near the center and the cooled excitons far from the center when the conveyer is off. The green dashed line across (b), (d) and (f) is used to indicate the hot (incoherent and coherent) excitons near the center and the coherent excitons far from the center when the conveyer is on. The green solid line across (b), (d) and (f) indicates the cooled excitons (degenerated excitons) traveling about 40 μ\mum. (g) The average transport distance of indirect excitons via conveyer M1M_{1} as a function of the conveyer amplitude, AconvA_{\mathrm{conv}}. Lines are a guide to the eye. The intersection point of the line is defined as AturnonA_{\mathrm{turn-on}} where the dynamical localization-delocalization transitions occur as is shown in (g1). Pex=20P_{\mathrm{ex}}=20 μ\mu W. (h) The measured average transport distance of indirect excitons via conveyer M1M_{1} as a function of density. Aconv=4.9A_{\mathrm{conv}}=4.9 meV, vconv=0.7v_{\mathrm{conv}}=0.7 μ\mum/ns.

The above experimental facts raise several interesting questions which need to be clarified. For example, whether the efficiency of the exciton transport for lower conveyer velocity is also related to the Chern number, or not Thouless (1983)? The other important issue is whether the several bright stripes come from the exciton coherence or not. The excitons in the coherent state or the degenerate state mean the off-diagonal long-range order exists in the system and the excitons are in a condensed state. Previously, a nonlinear Schrödinger equation including the attractive two-body interaction and the repulsive three-body interaction was proposed to explain the complex exciton patterns where the excitons are assumed to be in a coherent state Liu et al. (2006); Xu et al. (2012); Liu et al. (2014, 2009). One may ask naturally if the nonlinear Schrödinger equation can explain the exciton transport. In Ref. Winbow et al. (2011), the indirect excitons were considered as classical particles. The conveyer velocity and amplitude, as well as the exciton density, dependencies of the transport distances are all explained by a classical diffusion equation. However, the exciton degeneracy hidden in the PL patterns were not studied. As will be seen, the spatial and spectral patterns give the opportunity to coherence of the indirect excitons.

Since any realistic material inevitably contain a certain degree of impurities and defects, and the interactions between particles are almost always present, the combined effect of disorder and interactions can lead to novel quantum states such as the Bose glass phase in disordered Bosonic systems Scalettar et al. (1991); Lugan et al. (2007). Recently, the phase transition between the Bose glass and the superfluid was directly observed in the two-dimensional Bose glass in an optical quasicrystal Yu et al. (2023). The observed dynamical localization-delocalization transitions in conveyer for excitons gives another platform to answer more fundamental questions about the delocalization-localization transition High et al. (2007).

By analysing the PL patterns, we find the exciton transport can be simplified as the diffusion of Bosons with the finite lifetime in a moving lattice. The main findings of this work are summarized as follow. (i) The pattern can be divided approximately into the incoherent part and coherent part which can be described by the diffusion differential equation used in Ref. Winbow et al. (2011) and the nonlinear Schrödinger equation proposed by us respectively. (ii) The bright stripes are the interplay between the coherent excitons and the moving lattices. (iii) The cooling rate is the key factor in the exciton transport. (iv) The lifetime and cooling rate of the coherent excitons are estimated. (v) The sample is very clean and free of impurity. We discuss the data and build a time-dependent nonlinear Schrödinger equation in Sec. II to describe the coherent excitons. We also show how to obtain the PL intensity profiles I(x)I(x) by calculating the escape rate. In Sec. III, numerical calculations and detailed discussions on the exciton distribution in moving lattices are given. Sec. IV is devoted to a brief summary.

II Model Hamiltonian

The clues of understanding the PL patterns in the conveyor belt come from the various exciton patterns found by the previous experiments, in particular, the two puzzling exciton rings, inner and external, and periodic bright spots in the external ring Butov et al. (2002a, b); Snoke et al. (2002). A charge-separated transport mechanism was proposed and gave a satisfactory explanation both of the formation of the two exciton rings and of the dark region between the inner and the external ring Butov et al. (2004); Rapaport et al. (2004). This mechanism was further confirmed by PL images of single quantum well Rapaport et al. (2004).

As pointed out in Refs. Butov et al. (2004); Rapaport et al. (2004), when electrons and holes are first excited by high-power laser, they are actually charge-separated and have a small recombination rate. No true excitons are formed at this stage. They can travel a long distance from the laser spot before recombination. After a long-distance traveling, the hot electrons and holes collide with the semiconductor lattices and are cooled down. The cooled electrons and holes forms the excitons in the annular region near the the laser spot. The recombination of this part of excitons generates a PL ring. As the drift speed of electrons (with smaller effective mass) is larger than that of holes (with larger effective mass), there are always some electrons and holes escaping from this recombination. Also due to the neutrality of the coupled quantum wells, the negative charges will slow down and begin to accumulate at the sites far away from the PL ring. The cold electrons and holes then eventually meet to form the excitons. This part of the excitons recombine and also form the exciton ring at the boundary of the opposite charges. The second ring is far away the laser spot comparing with the former. The two rings were called as inner ring and external ring respectively.

As the excitons forming in the external ring are from the cooled electrons and holes, they have a low kinetic energy and temperature. While whether these excitons are condensed is in debate Butov et al. (2004); Balatsky et al. (2004); Vörös et al. (2006); High et al. (2012); Semkat et al. (2012); Levitov et al. (2005); Chernyuk and Sugakov (2006); Sugakov (2007), we reasonably assumed that the excitions were in the highly degenerated states, and proposed a self-trapped interaction model involving an attractive two-body interaction and a repulsive three-body interaction Liu et al. (2006); Xu et al. (2012); Liu et al. (2009). The mechanism gave a good account to the periodic bright spots in the external ring. In addition, it also explained well the abnormal exciton distribution in an impurity potential, in which the PL pattern became much more compact, and exhibited an annular shape with a darker central region Lai et al. (2004). Moreover, the model also captured some experimental details. For instance, the dip can turn into a tip at the center of the annular cloud when the sample was excited by a higher power lasers.

Inspired by both the above experimental data and theories, we are ready to investigate the exciton PL patterns in the moving lattices. As we know, the PL intensity is approximately proportional to the exciton number and the PL energy is approximately proportional to the exciton energy which includes the kinetic , the bound and the interaction energies. However, the bound energy remains unchanged due to the special structure of the spatially separated electrons and holes. The interaction energy between excitons is also negligible in the low-density case. As a result, the change of PL energy is only related to the kinetic energy. The PL energies shown in Fig. 1 (c) (1.534 eV marked by blue dashed line) and PL intensities in Fig. 1 (e) far from the center are lower than those near the center. Therefore, it indicates that the excitons near the center have larger kinetic energy and higher particle density. We argue that the spatial distribution of excitons here is also consistent with that of the exciton rings reported in Ref. Butov et al. (2002a, b). The hot excitons are formed at the center [dark region -10 μ\mum << x << 10 μ\mum in Fig. 1 (e)] and the cooled excitons are located far from the center. So there exist two kinds of excitons, one is the hot excitons near the center and can be regarded as classical particles, the other is the cooled excitons far away the center cannot be regarded as classical particles.

When the conveyer is turned on [see Fig. 1 (b)], the moving potential modifies both the exciton potential energy and the PL energy [see Fig. 1 (d)]. While the moving lattices drag the hot excitons, they are cooled down further by inelastic collision to the semiconductor phonons. In comparison with the high temperature case near the center, the excitons far from the center have a long de Broglie wavelength in the low temperature region. As we know, a quantum phase transition occurs when the Bosons condense from the non-degenerate state to degenerate state. Whether the finite life-time excitons can experience a Bose-Einstein condensation transition is an interesting issue. Neglecting the above complexity, we roughly divide the exciton distribution into two parts by the green dashed line shown in Fig. 1 (b), (d) and (f). On the left side of the green dashed line, part of the excitons are in coherent state and part of excitons are in incoherent state. On the right side of the green dashed line, however, we consider all the excitons are in coherent states or degenerate states. According to the above analysis, the coherent excitons may come from two parts. One part directly comes from the hot exciton near the center by the cooling. The other comes directly from the pairing of the cooled electrons and holes.

It is useful to estimate the exciton parameters according to the above analysis. As is indicated by the green dashed line in Fig. 1, after a 10 μ\mum traveling, the hot excitons can become highly degenerated excitons. As is also marked by a blue solid line, after another 30 μ\mum traveling, the coherent excitons recombine. If the excitons follow the conveyor belt with a low velocity vconv=0.7v_{\text{conv}}=0.7 μm/ns\mu\text{m/ns}, it takes about 14 ns for the hot excitons to the highly degenerated excitons. So the cooling time of the excitons from classical particles to coherent particles is about 14 ns. It takes another 40 ns before the degenerated excitons recombine. So the lifetime of the coherent excitons is estimated to be about 40 ns. This is the first time to obtain the lifetime of indirect exciton experimentally.

Based on the above analysis, it is reasonable to model the highly degenerate exciton gas by a time-dependent nonlinear Schrödinger equation Liu et al. (2006, 2009)

id|Ψ(t)dt=H|Ψ(t),i\frac{d\left|\Psi\left(t\right)\right\rangle}{dt}=H\left|\Psi\left(t\right)\right\rangle, (1)

where |Ψ=|ψ1,ψ2,,ψL\left|\Psi\right\rangle=\left|\psi_{1},\psi_{2},\cdots,\psi_{L}\right\rangle and LL is the lattice length. The tight-binding Hamiltonian reads

H\displaystyle H =\displaystyle= j{t~(|ψjψj+1|+h.c.)\displaystyle\sum_{j}\left\{\tilde{t}\left(\left|\psi_{j}\right\rangle\left\langle\psi_{j+1}\right|+\text{h.c.}\right)\right.
+[Vconv,j(t)+F(nj)iγ]|ψjψj|},\displaystyle+\left.\left[V_{\text{conv},{j}}\left(t\right)+F\left(n_{j}\right)-i\gamma\right]\left|\psi_{j}\right\rangle\left\langle\psi_{j}\right|\right\},

where

Vconv,j(t)\displaystyle V_{\text{conv},j}\left(t\right) =\displaystyle= Aconvcos[2π(jvconvt)/λ]\displaystyle A_{\text{conv}}\cos\left[2\pi\left(j-v_{\text{conv}}t\right)\right/\lambda] (3)
=\displaystyle= Aconvcos(2πκjωt)\displaystyle A_{\text{conv}}\cos\left(2\pi\kappa j-\omega t\right)

denotes the moving external potential created by a set of ac gate voltages. When vconv=0v_{\text{conv}}=0 and ω=0\omega=0, the moving lattice reduces to the static lattice. The parameter γ=1/Γ>0\gamma=1/\Gamma>0 is the loss rate and Γ\Gamma is the exciton average lifetime. t~\tilde{t} is the hopping amplitude and is set to be 1 in the following discussions. nj=|ψj|2n_{j}=\left|\psi_{j}\right|^{2} is the local probability density. F(nj)F\left(n_{j}\right) is the effective interaction between the indirect excitons.

In general, the excitons are considered as the weak repulsive interactions F(nj)=gnjF\left(n_{j}\right)=gn_{j}. Note that F(nj)=g1nj+g2nj2F\left(n_{j}\right)=-g_{1}n_{j}+g_{2}n_{j}^{2} was used to explain various exciton patterns Liu et al. (2006); Xu et al. (2012); Liu et al. (2014, 2009). The above phenomenological interaction may come from the dipolar interaction and the exchange interaction. Since the effective parameter g1Ng_{1}\propto N and g2N2g_{2}\propto N^{2} (NN is the particle number), we have the relationship g2g12g_{2}\propto g_{1}^{2}. We take g2=ϵg12g_{2}=\epsilon g_{1}^{2} here ϵ\epsilon is the parameter describing the complex interactions. As will be seen in Sec. III B, the effects of these two kinds of interactions are identical except for the cases of the very low and very high particle density. Eq. (1) is the modified Gross-Pitaevskii equation when γ=0\gamma=0. Unlike the Gross-Pitaevskii equation which is usually used to determine the ground states of a low-temperature Bosonic gas with a short- or zero-range two-body interaction, Eq. (1) contains the effect of the finite particle lifetime.

As the norm of the wavefunction decreases with time

ddtI(t)\displaystyle\frac{d}{dt}I(t) =\displaystyle= ddtΨ(t)|Ψ(t)=iΨ(t)|(HH)|Ψ(t)\displaystyle\frac{d}{dt}\langle\Psi(t)|\Psi(t)\rangle=i\left\langle\Psi(t)\right|(H^{{\dagger}}-H)\left|\Psi(t)\right\rangle
=\displaystyle= 2γΨ(t)|Ψ(t)=2γI(t),\displaystyle-2\gamma\langle\Psi(t)|\Psi(t)\rangle=-2\gamma I(t),

the time-dependent wavefunction norm can be written as

I(t)=Ψ(t)|Ψ(t)=j|ψj(t)|2=exp(2γt),\displaystyle I(t)=\langle\Psi(t)|\Psi(t)\rangle=\sum_{j}|\psi_{j}(t)|^{2}=\exp\left(-2\gamma t\right), (4)

when the initial-state is normalized Ψ(0)|Ψ(0)=1\langle\Psi(0)|\Psi(0)\rangle=1. γ\gamma is the lost rate which is inversely proportional to the life-time of the excitons. We define the escaping probability at location jj of the walker as

Ij=2γ0𝑑t|ψj(t)|2,I_{j}=2\gamma\int_{0}^{\infty}dt|\psi_{j}(t)|^{2}, (5)

which satisfies jIj=j2γ0𝑑t|ψj(t)|2=0𝑑tI(t)=1\sum_{j}I_{j}=\sum_{j}2\gamma\int_{0}^{\infty}dt|\psi_{j}(t)|^{2}=\int_{0}^{\infty}dtI(t)=1. IjI_{j} is the summation of luminescence intensity of all the exciton states and is the PL intensity profile.

The cooling of the excitons is another important physical factor in the exciton diffusion. With the increase of the time, the high-energy excitons relax to low-energy excitons and condense to their degenerate states. The condensed excitons in the confined well occupied the highly degenerate states can be further cooled before recombination. The cooling process can be well simulated by the imaginary time evolution. As we know, for a stationary state or eigenstate, if the time tt is artificially replaced by it-it, the wavefunction will exponential decay in its time evolution. According to the superposition principle, a state can be expanded as the linear superposition of all the eigenstates. In the imaginary time evaluation of the state, the components with the higher eigen energies have the faster decay rates which is equivalent to the relaxation of high energy components to the low energy components Bader et al. (2013). Although the energy levels are not well defined in the time-dependent non-Hermitian system, the energy levels have previously been observed in previous experiments Butov et al. (2002a, b); Snoke et al. (2002); Lai et al. (2004); High et al. (2009) and investigated theoretically Wilkes et al. (2010); Liu et al. (2006); Xu et al. (2012); Liu et al. (2014, 2009). We use the imaginary time-evolution of the Schrödinger equation in Eq. (1) to describe the cooling process.

In the following calculations, we solve the time-dependent Schrödinger equation (1) in a real time interval Δt\Delta t with the initial state Ψ(0)\Psi(0) to obtain the time evolution state Ψ(Δt)\Psi(\Delta t). Then the state Ψ(Δt)\Psi(\Delta t) is used as the initial state, we solve the time-dependent Schrödinger equation (1) in a imaginary time interval iΔti\Delta t^{\prime} to obtain the time evolution state Ψ(Δt+iΔt)\Psi(\Delta t+i\Delta t^{\prime}). At last, the wavefunction Ψ(Δt+iΔt)\Psi(\Delta t+i\Delta t^{\prime}) is normalized as Bader et al. (2013)

I(Δt+iΔt)=|Ψ(Δt+iΔt)|2=exp(2γΔt).\displaystyle I(\Delta t+i\Delta t^{\prime})=|\Psi(\Delta t+i\Delta t^{\prime})|^{2}=\exp\left(-2\gamma\Delta t\right). (6)

Repeating the above process 𝒩\mathcal{N} times until I(𝒩[Δt+iΔt])0I\left(\mathcal{N}\left[\Delta t+i\Delta t^{\prime}\right]\right)\sim 0, we get a serial of Ψ(n[Δt+iΔt])\Psi\left(n\left[\Delta t+i\Delta t^{\prime}\right]\right) to calculate Ij=2γn=0𝒩|Ψ(n[Δt+iΔt])|2I_{j}=2\gamma\sum_{n=0}^{\mathcal{N}}\left|\Psi\left(n\left[\Delta t+i\Delta t^{\prime}\right]\right)\right|^{2}. We define τ=Δt/Δt\tau=\Delta t^{\prime}/\Delta t. The larger τ\tau indicates the longer cooling time. So τ\tau is proportional to cooling speed.

It should be emphasized that the method above is effective since the normalization condition of the escape rate IjI_{j} in Eq. (5) guarantees the convergence of the calculations. Otherwise, without considering the time-dependent normalized condition, the propagation of a wave-pocket still meets the boundary and reflects under the open boundary condition. Even though the absorbing boundary condition is adopted, the boundary also have a great effect on the evolution of the wavefunction. In such a case, the interference fringes are formed by interference between incident and reflected waves and no convergent solution can be obtained.

III Numerical analysis

III.1 A wave pocket diffusion: without particle interactions

Refer to caption
Figure 2: The wave pocket diffusion without a periodic potential (left column) and with the static periodic potential (right column). The free diffusion of a wave pocket at different time without the dissipation γ=0\gamma=0 (a) and with the dissipation γ=0.01\gamma=0.01 (b). The diffusion of a wave pocket modulated by the static periodic potential (dashed line) at different time without the dissipation γ=0\gamma=0 (d) and with the dissipation γ=0.05\gamma=0.05 (e). IjI_{j} distribution for the different dissipation γ\gamma for the free diffusion (c) and for the diffusion modulated by the static periodic potential Vconv,jV_{\text{conv},j} (f). The parameters of the periodic potential in Eq. (3) are set as Aconv=1.6A_{\text{conv}}=1.6, ω=0\omega=0, and λ=64\lambda=64.

According to the above analysis, there are various sources for the coherent excitons. As a result, the initial distribution of the degenerate excitons can be considered as a wave pocket with multi-peaks. In order to understand the transport of the wave pocket, let us first study the simplest case of the diffusion of a wave pocket with one-peak. We solve the time-dependent nonlinear Schrödinger equation (1) in real-time only under the open boundary condition (ψ±L=0\psi_{\pm L}=0) for an initial Gauss wave pocket ψj(0)=(2κ/π)exp[κ(jj0)2]\psi_{j}(0)=\sqrt{(2\kappa/\pi)}\exp\left[-\kappa(j-j_{0})^{2}\right]. The results with the parameters j0=16j_{0}=-16, κ=0.006\kappa=0.006 and L=128L=128 are shown in Fig. 2. In the following discussions, the blue lines in all the figures indicate the shape of the initial wave pocket.

The physics of the wave pocket diffusion is very simple Zeng (2007); Sakurai (1994). For the Gauss wave pocket, its wave functions both in real space and momentum space are all of the Gauss type. According to the uncertainty relation ΔxΔp\Delta x\Delta p\approx\hbar, with the increase of time, the larger Δx\Delta x leads to the smaller Δp\Delta p, here Δx\Delta x and Δp\Delta p are the full width at half maximum of the Gauss wave pocket. If we approximately take vΔvv\approx\Delta v, the smaller Δp\Delta p leads to the lower the propagation speed. As a result, although the diffusion becomes slower and slower as the time increases, the wave finally arrives at the boundary and is reflected. Interference fringes are generated by the interference between reflected and incident waves. The above statements are true whether it is without the periodic potential in Fig. 2 (a) or with the static periodic potential in Fig. 2 (d). The difference is that more interference fringes appear in Fig. 2 (d) due to the interplay between the wave and the periodic potential. When we take the dissipation of the particles into account, the wave pocket decays in Fig. 2 (b) and (e) with the time goes on. The escape probability IjI_{j} in Eq. (5) is calculated and shown in Fig. 2 (c) and (f). As expected, the particle dissipation suppresses the wave pocket diffusion. From Fig. 2 (f), IjI_{j} can be deviated from the initial position due to the modulation of the external potential.

Refer to caption
Figure 3: The wave pocket diffusions without the cooling [black lines] and with the cooling [red lines] modulated by the static periodic potentials with the different intensities (a) Aconv=0A_{\text{conv}}=0, (b) Aconv=0.01A_{\text{conv}}=0.01 and (c) Aconv=0.4A_{\text{conv}}=0.4.

We next investigate the cooling effect on the wave pocket diffusion. The time-dependent nonlinear Schrödinger equation (1) is solved with the real-time and imaginary-time evolutions alternatively to obtain the escape rate IjI_{j} of the coherent excitons. The result of IjI_{j} is shown in Fig. 3. It can be seen that the cooling helps the wave pocket diffusion. We can understand the underlying physical picture as follow Zeng (2007); Sakurai (1994). A wave pocket in free space can be considered as the superposition of the plane waves with the different wave vectors. When the condensed Bosons are considered to be in a state of Gauss wave pocket in momentum space, their distribution in the Cartesian coordinate space is still Gauss type. With the time increases, the wave pocket will spread in coordinate space and will contract in momentum space. However, their waves are still Gauss types. With the temperature decreases, the low-energy probability of the Gauss wave pocket increases accordingly since the wave pocket in momentum space is modulated by the factor exp(βEk)\exp(-\beta E_{k}) (Ek=k2/2E_{k}=k^{2}/2). The narrower the wave pocket in momentum space, the wider wave pocket in coordinate space Zeng (2007); Sakurai (1994). As a result, the quantum mechanical effect is in opposition to the classical effect where the spreading of classical particle becomes slow with the decrease of the temperature.

By comparing the black lines and the red lines in Fig. 3 (b), it is found that the cooling also helps the wave pocket spread in a static periodic potential. The physical picture can also be understood in a similar way. In the static periodic potential, the eigenfunction of a particle is the Bloch wave which can be expanded in terms of the Wannier functions. Correspondingly, the energy bands come from the splitting of the degenerate energy levels. A wave pocket in periodic potential can be considered as a superposition of the Bloch waves with the different wave vectors. With time increasing and particle cooling, the particle probabilities in low-energy bands increase accordingly. The narrower the wave pocket in momentum space, the wider the wave pocket in coordinate space. As a result, the cooling enhances the diffusion of the wave pocket.

It is interesting to see in Fig. 3 (c) that the peaks of the IjI_{j} for τ>0\tau>0 are located at the crests of the periodic potential, which is in contrast to the case of τ=0\tau=0 where the peaks are located at the troughs. The reason can be understood as follows. According to the above analysis, with the lowering of the temperature, the Bosons tend to condense at k=0k=0, which results in the increasing of Δx\Delta x. In the large AconvA_{\text{conv}} case, the Wannier functions are the atomic orbital wavefunctions in a unit cell and the particles are confined in a single unit cell. The stronger the confinement, the larger the Δx\Delta x. The particles tend to move to the boundary of the unit cell to obtain the large Δx\Delta x which causes the dip of particle distribution at the center of the unit cell. This quantum behavior can be used to detect the exciton degeneracy.

Refer to caption
Figure 4: The wave pocket diffusions in a moving lattice without considering the cooling (a) and considering the cooling (b). IjI_{j} of the two cases of (a) and (b) are given in (c). (e) and (g) show IjI_{j} with the different moving lattice intensity AconvA_{\text{conv}} and moving speed vconvv_{\text{conv}}. The cooling speed τ\tau, the lattice intensity AconvA_{\text{conv}} and the lattice velocity vconvv_{\text{conv}} dependencies of the transport distant M1M_{1} are shown in (d), (f) and (h) respectively. The parameters used in all figures are: AconvA_{\text{conv}} = 1.6, vconvv_{\text{conv}} = 1.48, γ=0.05\gamma=0.05.

The wave pocket diffusion in a moving lattice is shown in Fig. 4. The moving lattice modifies the shape of the Gauss wave pocket. With the increase of the time, the wave pocket follows the moving lattice while its height decreases obviously due to the finite lifetime of the exciton [Fig. 4 (a) and (b)]. The exciton cooling (τ=0.5\tau=0.5) in Fig. 4 (b) dramatically increases the transport distance in comparison with the case without the cooling (τ=0\tau=0) in Fig. 4 (a). The escaping probability IjI_{j} corresponding to the case in Fig. 4 (a) and (b) is presented in Fig. 4 (c). The average transport distance M1M_{1} increases with the cooling parameter τ\tau and then tends to saturation as is shown in Fig. 4 (d). We can conclude that the cooling is the key factor to the exciton transport via the conveyer.

The reason can be understood from the case of the low velocity and the high AconvA_{\text{conv}} of the moving periodic potential. In such case, the Bloch theory can be assumed to be correct approximately and the Wannier functions can be approximated as the atom-orbital functions (particle wave functions in unit cell). The overlap of the neighbouring low orbital functions is less than that of the high orbital functions. The little overlap of the neighbouring low orbital functions leads to the little tunneling between unit cells. With the cooling, the excitons tend to occupy the low energy bands. The occupation probability of the low orbital state increases with the cooling parameters τ\tau. With increase of the cooling further, most part of the particles are in the low orbital state. As a result, the particles follow the moving lattice. So the change of the average transport distance M1M_{1} obvious increase with τ\tau as is shown in Fig. 4 (d).

The tunneling between unit cells decreases with the increase of the lattice height AconvA_{\text{conv}}. In the high AconvA_{\text{conv}} case, the tunneling between unit cells is inconspicuous. So the transport distance increases with the AconvA_{\text{conv}} and then tends to saturation as is shown in Fig. 4 (e). In the low lattice speed case, the average transport distance M1M_{1} increases with the lattice speed as is shown in Fig. 4 (h), which indicates the exciton transport follows the moving lattices. In the high lattice speed case, the energy band theory breaks down. It is definitely that the tunneling between the unit cells decreases with the increase of the lattice speed which is adverse to the particles transport. We therefore argue that the particles follow the moving lattices only in the low velocity and high lattice amplitude cases.

III.2 A wave pocket diffusion: with particle interactions

Refer to caption
Figure 5: The wave pocket diffusions of the weak repulsive particles. The following parameters are used in the numerical analysis: Aconv=0.8A_{\text{conv}}=0.8, vconv=2.08v_{\text{conv}}=2.08, τ=0.5\tau=0.5 and γ=0.05\gamma=0.05. For the pure repulsive case, the interaction parameter gg-dependent effective potential g|ψj|2g|\psi_{j}|^{2}, the distribution patterns IjI_{j} with different gg and transport distance M1M_{1} as a function of gg are given in (a), (b) and (c) respectively. For the complex interaction g1|ψj|2+g2|ψj|4-g_{1}|\psi_{j}|^{2}+g_{2}|\psi_{j}|^{4} case, we set g2=ϵg12g_{2}=\epsilon g_{1}^{2} and ϵ=8\epsilon=8. The interaction parameter g1g_{1}-dependent effective potential, the distribution patterns IjI_{j} with the different g1g_{1} and transport distance M1M_{1} as a function of g1g_{1} are given in (d), (e) and (f) respectively.

In general, the excitons are considered to have the weak repulsive interactions. We study the exciton transport with this kind of interactions in Figs. 5 (a), (b) and (c). The interaction g|Ψj|2g|\Psi_{j}|^{2} acts as the effective potential as is shown in Fig. 5 (a) with a different strength gg. IjI_{j} with the different gg is shown in Fig. 5 (b). It indicates that the interactions have a significant effect on the particle transport. M1M_{1} as a function of gg is shown in Fig. 5 (c). Two different effects affect the particle transport for the repulsive interactions. One is that the repulsive interactions always favor the particle spreading. The other is that the repulsive interactions modify the lattice intensity equivalently, which is detrimental to particle transport. So the particle transport M1M_{1} is a non-monotonic function of gg.

We have phenomenally proposed the two-body attractive and three repulsive interactions to understand the exciton patterns. To study how the complex interactions modify the exciton transport, we show the result with the effective interaction potential g1|ψj|2+g2|ψj|4-g_{1}|\psi_{j}|^{2}+g_{2}|\psi_{j}|^{4} in Fig. 5 (d). Here g2=ϵg12g_{2}=\epsilon g_{1}^{2} with ϵ=8\epsilon=8. For the large g1g_{1} cases, the excitons are in repulsive interactions as that in Fig. 5 (a). So IjI_{j} in Fig. 5 (e) and M1M_{1} in Fig. 5 (f) are similar to those of Fig. 5 (b) and M1M_{1} in Fig. 5 (c) respectively.

Refer to caption
Figure 6: The wave pocket diffusions of the attractive particles. The following parameters are used in the numerical analysis: Aconv=0.2A_{\text{conv}}=0.2, vconv=2.08v_{\text{conv}}=2.08, τ=0.5\tau=0.5 and γ=0.05\gamma=0.05. For the pure attractive case, the interaction parameter gg-dependent effective potential g|ψj|2-g|\psi_{j}|^{2}, the distribution patterns IjI_{j} with different gg and transport distance M1M_{1} as a function of gg are given in (a), (b) and (c) respectively. For the complex interaction g1|ψj|2+g2|ψj|4-g_{1}|\psi_{j}|^{2}+g_{2}|\psi_{j}|^{4} in the weak attractive interaction region, we set g2=ϵg12g_{2}=\epsilon g_{1}^{2} and ϵ=8\epsilon=8. The interaction parameter g1g_{1}-dependent effective potential, the distribution patterns IjI_{j} with the different g1g_{1} and transport distance M1M_{1} as a function of g1g_{1} are given in (d), (e) and (f) respectively.

In the low exciton density case, g1|ψj|2+g2|ψj|4-g_{1}|\psi_{j}|^{2}+g_{2}|\psi_{j}|^{4} is in the attractive interaction region. It is interesting to study how the attractive interactions g|ψj|2-g|\psi_{j}|^{2} and g1|ψj|2+g2|ψj|4-g_{1}|\psi_{j}|^{2}+g_{2}|\psi_{j}|^{4} modify the particle transport. We present the effective attractive interactions in Figs. 6 (a) and (d). In order to study the effects of the weak interaction more clearly, a small lattice intensity Aconv=0.2A_{\text{conv}}=0.2 is adopted. This is in contrast to the repulsive case shown in Figs. 5 (a) and (d) [Aconv=0.8A_{\text{conv}}=0.8]. In the pure attractive case g|ψj|2-g|\psi_{j}|^{2}, a wave pocket will collapse in the time evolution as is shown in Fig. 6 (b). However, the finite exciton lifetime prevents the pocket from collapsing further.

The attractive interactions also have two effects on particle transport. One is that the attractive interactions always hinder the particle spread. The other is that the increase of the effective lattice intensity facilitates the particle transport. So the particle transport M1M_{1} is also a non-monotonic function of gg [see Fig. 6 (c)]. In the small g1<4g_{1}<4 case, the weak complex interactions are in the attractive region, which also modify the exciton transport in Fig. 6 (e) and (f). However, M1M_{1} shows a monotonic variation as a function of g1g_{1}.

III.3 The disorder effects

The data of the exciton transport distance M1M_{1} via conveyer as a function of the conveyer amplitude AconvA_{\text{conv}} are given in Fig.1 (g). It shows that the exciton cloud extension M1M_{1} is not affected by the motion of a shallow conveyer. Across the transition point however, the exciton cloud starts to follow the conveyer, and M1M_{1} increases with the increase of AconvA_{\text{conv}}. At higher conveyer amplitude, M1M_{1} tends to saturation. As we know, the random disorders or impurities exist inevitably in the coupled quantum well grown with molecular beam epitaxy. The destructive interference of scattered waves due to the strong disorders leads to the Anderson localizations Anderson (1958); Abrahams (2010). As a result, the experimental data is naturally explained as the dynamical localization-delocalization transition. The dynamical localization was found in two-band system driven by the DC-AC electric field, in which the Rabi oscillation is quenched under the certain ratio of Bloch frequency and AC frequency Zhao et al. (1996); Liu and Ma (2003). An interesting issue here is whether the data can really be explained by the dynamical localization or not.

The Anderson localizations are generally studied with indirect method where the random disorders are simulated by a quasiperiodic on-site modulation Cai et al. (2013); Sun and Liu (2023). The quasiperiodic potential is set to be site dependent, i.e.

Vdis,j=2Adiscos(2παj),V_{\text{dis},j}=2A_{\text{dis}}\cos(2\pi\alpha j),

with AdisA_{\text{dis}} being the strength, α\alpha being an irrational number which is used to characterize the quasiperiodicity. The irrational number is usually taken as the value of the inverse of golden ratio [α=(51)/2\alpha=(\sqrt{5}-1)/2]. The value is closely related to the Fibonacci number which is defined by Fn/Fn+1F_{n}/F_{n+1} where the Fibonacci sequence of numbers FnF_{n} is defined using the recursive relation with the seed values F0=0F_{0}=0 , F1=1F_{1}=1 and Fn=Fn1+Fn2F_{n}=F_{n-1}+F_{n-2} Kohmoto (1983); Sun and Liu (2023). The merit of using Fibonacci numbers is that the quasiperiodic potential Vdis(j+Fn+1)=Vdis(j)V_{\text{dis}}\left(j+F_{n+1}\right)=V_{\text{dis}}\left(j\right) becomes periodic.

To simplify the discussion, the particle interaction, lifetime and cooling are not taken into account. The Hamilton in Eq. (II) can be written as

H(t)=j[t~(|ψjψj+1|+h.c.)+𝒱j(t)|ψjψj|],H\left(t\right)=\sum_{j}\left[\tilde{t}\left(\left|\psi_{j}\right\rangle\left\langle\psi_{j+1}\right|+\text{h.c.}\right)+\mathcal{V}_{j}\left(t\right)\left|\psi_{j}\right\rangle\left\langle\psi_{j}\right|\right], (7)

here 𝒱j(t)=Vconv,j(t)+Vdis,j\mathcal{V}_{j}\left(t\right)=V_{\text{conv},j}(t)+V_{\text{dis},j}, which is a spatial periodic λFn+1\lambda F_{n+1} and time periodic TT function 𝒱j(t)=𝒱j(t+T)=𝒱j+λFn+1(t)\mathcal{V}_{j}\left(t\right)=\mathcal{V}_{j}\left(t+T\right)=\mathcal{V}_{j+\lambda F_{n+1}}\left(t\right) here λ\lambda is an integer. The time dependence of the Hamiltonian (7) leads to no stationary states in the system. However, the Hamiltonian (7) has the time periodicity. We can write the state as

|ψj(t)=eiεt|cj(t),\left|\psi_{j}\left(t\right)\right\rangle=e^{-i\varepsilon t}\left|c_{j}\left(t\right)\right\rangle,

where ε\varepsilon is the quasienergy and |cj(t)=|cj(t+T)\left|c_{j}\left(t\right)\right\rangle=\left|c_{j}\left(t+T\right)\right\rangle is periodic function which satisfies the time-dependent equation

[H(t)it]|cj(t)=ε|cj(t)\left[H\left(t\right)-i\partial_{t}\right]\left|c_{j}\left(t\right)\right\rangle=\varepsilon\left|c_{j}\left(t\right)\right\rangle

according to the Floquet theorem Grifoni and Hänggi (1998); Liu and Ma (2003). As eiε(t+T)=eiεte^{i\varepsilon\left(t+T\right)}=e^{i\varepsilon t}, it requires the quasienergy satisfies εl=2πlT=ωl\varepsilon_{l}=\frac{2\pi l}{T}=\omega l where ll is an integer. To ensure the consistency between the Floquet theorem and the Bloch theorem in the form, the period is assumed to be T=NaT=Na and the quasienergy is confined in the first time Brillouin zone ]π/a,π/a]\left]-{\pi}/{a},{\pi}/{a}\right], with aa being the time unit. ]π/a,π/a]\left]-{\pi}/{a},{\pi}/{a}\right] means an excluded value π/a-{\pi}/{a} on the left and included π/a{\pi}/{a} on the right. As a result, ll can be taken as ]N/2,N/2]\left]-{N}/{2},{N}/{2}\right]. From the time-dependent nonlinear Schrödinger equation, the Hamiltonian in Eq. (7) can be transformed to a tight-binding Floquet operator in the second quantization form

(t)\displaystyle\mathcal{H}\left(t\right) =\displaystyle= j[t~(cj(t)cj+1(t)+h.c.)\displaystyle\sum_{j}\left[\tilde{t}\left(c_{j}^{{\dagger}}\left(t\right)c_{j+1}\left(t\right)+\text{h.c.}\right)\right.
+cj(t)(𝒱j(t)it)cj(t)].\displaystyle+\left.c_{j}^{{\dagger}}\left(t\right)\left(\mathcal{V}_{j}\left(t\right)-i\partial_{t}\right)c_{j}\left(t\right)\right].

After the Fourier transform

cj(t)\displaystyle c_{j}\left(t\right) =\displaystyle= N1/2n=N/2N/2ceiωntj,n,\displaystyle N^{-1/2}\sum_{n=-N/2}^{N/2}c{}_{j,n}e^{i\omega nt},
cj(t)\displaystyle c_{j}^{\dagger}\left(t\right) =\displaystyle= N1/2n=N/2N/2ceiωntj,n,\displaystyle N^{-1/2}\sum_{n=-N/2}^{N/2}c{}_{j,n}^{\dagger}e^{-i\omega nt},

and the inner product (t)=1T0T(t)𝑑t=1Nn=N/2N/2(n)\left\langle\left\langle\mathcal{H}\left(t\right)\right\rangle\right\rangle=\frac{1}{T}\int_{0}^{T}\mathcal{H}\left(t\right)dt=\frac{1}{N}\sum_{n=-N/2}^{N/2}\mathcal{H}\left(n\right), the time-dependent Floquet operator (t)\mathcal{H}\left(t\right) becomes Grifoni and Hänggi (1998); Liu (2021)

\displaystyle\mathcal{H} =\displaystyle\mathcal{=} j,n{(t~ccj,n+j+1,nAconvei2πηjccj,n1+j,nh.c.)\displaystyle\sum_{j,n}\left\{\left(\tilde{t}c{}_{j,n}^{\dagger}c{}_{j+1,n}+A_{\text{conv}}e^{-i2\pi\eta j}c{}_{j,n-1}^{\dagger}c{}_{j,n}+\text{h.c.}\right)\right. (9)
+[Adiscos(2παj)nω]ccj,n}j,n.\displaystyle+\left.\left[A_{\text{dis}}\cos\left(2\pi\alpha j\right)-n\omega\right]c{}_{j,n}^{\dagger}c{}_{j,n}\right\}.

It is obviously that several sub-bands forming for different nn of the model (9). For a given nn, the model (9) can be mapped to Harper model which is the two-dimensional electron gas in a magnetic field if η\eta is regarded as a magnetic flux penetrating the unit cell Harper (1955).

Under the Floquet ansatz cj(t)cj(t)=nccj,nj,n\left\langle\left\langle c_{j}^{{\dagger}}\left(t\right)c_{j}\left(t\right)\right\rangle\right\rangle=\sum_{n}c{}_{j,n}^{\dagger}c{}_{j,n}, we can define a time-average inverse of the participation ratio (TMIRP) of the normalized eigenstate |cj,nl\left|c{}_{j,n}\right\rangle_{l} corresponding to the eigenvalue εl\varepsilon_{l},

TMIPRl=j,n|c|j,ncj,nll|4.\text{TMIPR}_{l}=\sum\limits_{j,n}\left|{}_{l}\left\langle c{}_{j,n}^{\dagger}|c{}_{j,n}\right\rangle_{l}\right|^{4}.

The localization of the whole system can be characterized by the average of TMIPR

TMIPR¯=lTMIPRl/L.\overline{\text{TMIPR}}=\sum_{l}{\text{TMIPR}_{l}}/L.

For a delocalizaton phase of the system, TMIPR¯\overline{\text{TMIPR}} is of the order 1/L1/L, whereas it approaches 11 in the localized phase.

Refer to caption
Figure 7: The TMIPR¯\overline{\text{TMIPR}}^{\prime}s dependence on AdisA_{\text{dis}} for different AconvA_{\text{conv}}. The data have been expanded three and six times for the cases of Aconv=0.6A_{\text{conv}}=0.6 and 1.21.2. The green dashed line is the slop of black line and is used to show the phase transition occurring at Adis=1A_{\text{dis}}=1 for the AAH model (Aconv=0A_{\text{conv}}=0). The parameters used in the numerical calculation are L=N=65L=N=65, α=0.2\alpha=0.2 and ω=0.5\omega=0.5.

We diagonalize the time-mean tight-binding Hamiltonian (9) to get the eigenstates |cj,nl\left|c{}_{j,n}\right\rangle_{l} for calculating TMIPR¯\overline{\text{TMIPR}}. The result is shown in Fig. 7. In the case of Aconv=0A_{\text{conv}}=0, the Hamiltonian (7) is reduced to the Aubry-André -Harper (AAH) model Aubry and André (1980); Harper (1955). The AdisA_{\text{dis}} dependence of TMIPR¯\overline{\text{TMIPR}} and its first derivative (TMIPR¯)=d(TMIPR)¯/d(Adis)\left(\overline{\text{TMIPR}}\right)^{\prime}=d\overline{\left(\text{TMIPR}\right)}/d\left(A_{\text{dis}}\right) are shown in Fig. 7 by the black line and the green dashed line, respectively. A peak is found at Adis=1A_{\text{dis}}=1 which is consistent with that of the AAH model where the Anderson localization transition occurs at Adis=t~A_{\text{dis}}=\tilde{t}. It indicates the effectiveness of the time-mean method. We further calculate the AdisA_{\text{dis}} dependence of TMIPR¯\overline{\text{TMIPR}} for different AconvA_{\text{conv}}. Similar to the above research, for a wave pocket driven by the moving lattices, the wave pocket spreads over time, and tends to delocalization. For a wave pocket driven by the random lattice, it tends to localize. As a result, the interplay of the two lattices leads to suppression of the TMIPR¯\overline{\text{TMIPR}} with the increase of AconvA_{\text{conv}}. As we know, when there is a competition between localization and delocalization in particular, a localization-delocalization transition usually occurs. It is surprising that no visible localization-delocalization transition is found in the AdisA_{\text{dis}} dependence of TMIPR¯\overline{\text{TMIPR}} in Eq. (7) for the case of Aconv0.6A_{\text{conv}}\geq 0.6.

The localization-delocalization transition can also be captured by the asymptotic behavior over long time of the second-order moment of position operator σ2(t)\sigma^{2}(t) defined by the wave spreading Wilkinson and Austin (1994); Jeon et al. (1998); Longhi (2021)

σ2(t)=jcj(t)|j2|cj(t),\sigma^{2}(t)=\sum_{j}\left\langle c_{j}(t)\right|j^{2}\left|c_{j}(t)\right\rangle,

with the initial localized state |cj(0)\left|c_{j}(0)\right\rangle=δj.0=\delta_{j.0}. It gives the opportunity to further check the non-existence of the localization-delocalization transition. The asymptotic spreading of σ2(t)\sigma^{2}(t) is described by the power law, i.e. σ2(t)t2δ\sigma^{2}(t)\sim t^{2\delta} where δ\delta is the diffusion exponent. In terms of dynamical behavior of a wave pocket, measured by the exponent δ=δ(Adis)\delta=\delta(A_{\text{dis}}), the phase transition is discontinuous since δ(Adis)=1\delta(A_{\text{dis}})=1 for AdisA_{\text{dis}} <t~<\tilde{t} (ballistic transport), δ(Adis)1/2\delta(A_{\text{dis}})\simeq 1/2 at the critical point Adis=t~A_{\text{dis}}=\tilde{t} (almost diffusive transport), and δ(Adis)=0\delta(A_{\text{dis}})=0 in the localized phase Adis>t~A_{\text{dis}}>\tilde{t} (dynamical localization). When the spreading velocity v(Adis)σ(t)/tv(A_{\text{dis}})\sim\sigma(t)/t is further defined, the phase transition turns out to be smooth, with v(Adis)v(A_{\text{dis}}) being continuous function of potential amplitude AdisA_{\text{dis}} and v(Adis)=0v(A_{\text{dis}})=0 for Adis0.6A_{\text{dis}}\geq 0.6.

We still use the above method to characterize the dynamical localization. We solve the time-dependent nonlinear Schrödinger equation of the Hamiltonian (7) with the initial localized state |cj(0)\left|c_{j}(0)\right\rangle =δj.0=\delta_{j.0} to obtain |cj(t)\left|c_{j}(t)\right\rangle and v(Adis)v(A_{\text{dis}}). The dependence of v(Adis)v(A_{\text{dis}}) on AdisA_{\text{dis}} is shown in Fig. 8. For the case of Adis=0A_{\text{dis}}=0, the Hamiltonian in Eq. (7) is reduced to the AAH model. The dependence of Aconv{A_{\text{conv}}} on v(Adis)v(A_{\text{dis}}) [black line] shows a transition at Adis=1A_{\text{dis}}=1(t~)(\tilde{t}) obviously. With the applying of the moving lattices, either increasing its intensity AconvA_{\text{conv}} [vconvv_{\text{conv}} unchanged] in Fig. 8 (a) or increasing its velocity vconvv_{\text{conv}} [AconvA_{\text{conv}} unchanged] in Fig. 8 (b), the spreading velocity v(Adis)v(A_{\text{dis}}) is suppressed. The most obvious feature is the disappearance of the delocalization-localization transition. We also calculate the conveyer amplitude Aconv{A_{\text{conv}}}^{\prime}s dependence on M1M_{1} for the different disorder intensity AdisA_{\text{dis}} [vconv=2.02v_{\text{conv}}=2.02] in Fig.8 (c) and for the different conveyer velocity vconvv_{\text{conv}} [Adis=1.6A_{\text{dis}}=1.6] (d). Recalling the behavior of TMIPR¯\overline{\text{TMIPR}} in Fig. 7, we therefore argue that the moving lattice breaks the Anderson localization transition.

Refer to caption
Figure 8: vv as a function of the disorder amplitude AdisA_{\text{dis}} for the largest propagation time t=150t=150 where in (a) the speed of the moving lattice is set to be vconv=1.48v_{\text{conv}}=1.48 and in (b) the disorder amplitude is set to be Aconv=1.6A_{\text{conv}}=1.6 respectively. M1{M_{1}}^{\prime}s dependence on the conveyer amplitude AconvA_{\text{conv}} for the different disorder intensity AconvA_{\text{conv}} [vconv=2.02v_{\text{conv}}=2.02] (c) and for the different conveyer speed vconvv_{\text{conv}} [Aconv=1.6A_{\text{conv}}=1.6] (d). The other parameters γ=0.05\gamma=0.05 and τ=0.5\tau=0.5.

III.4 Comparing the theory with the experiments

After detailed investigations of the various effects on the diffusion of a wave pocket in the moving lattice, we are ready to study the experimental data. According to the discussions in sec. III.3, the plateaus can not be explained by delocalization-localization transition due to the disorders, even though the disorders suppress the wave-pocket transport. The disorders are firstly neglected in following numerical calculations. In addition, as the coherent excitons have different origins, the wave function of the coherent excitons was assumed with two peaks initially. The motivation of the two peaks comes from the two rings reported in Refs. Butov et al. (2002a, b). We solve the time-dependent nonlinear Schrödinger equation (1) with the initial wave function [black line in Fig. 9] to obtain IjI_{j} [red line in Fig. 9 (a)]. The appearance of the four peaks is basically consistent with the interference fringe.

Refer to caption
Figure 9: (a) Red line: the PL intensity profile IjI_{j} . Black line: the initial wave-pocket. Green-dashed line: the initial conveyer. The parameters vconv=1.48v_{\text{conv}}=1.48 and Aconv=1.6A_{\text{conv}}=1.6 are used in the calculations. (b) M1M_{1} as a function of the conveyer amplitude AconvA_{\text{conv}} for the different conveyer velocity vconv=v_{\text{conv}}= 1.48, 3.29, 4.03 and 6.25 respectively. (b1) Aturn-onA_{\text{turn-on}} versus the conveyer velocity vconvv_{\text{conv}}. (c) M1M_{1} as a function of the interaction parameter AconvA_{\text{conv}}. (d) M1M_{1} as a function of the interaction parameter gg with the parameters vconv=4.03v_{\text{conv}}=4.03, Aconv=2.5A_{\text{conv}}=2.5. The other parameters in these calculations are set as g=1.2g=1.2,τ=0.5\tau=0.5 and γ=0.05\gamma=0.05.

The conveyer amplitude Aconv{A_{\text{conv}}}^{\prime}s dependence on M1M_{1} is shown in Fig. 9 (b) for different vconvv_{\text{conv}}. It indicates that the transport distance M1M_{1} increases with the conveyer amplitude AconvA_{\text{conv}}. However, the exciton transport is less efficient for higher velocity. We also study the low-velocity case [green line] and find that the exciton transport is also less efficient for lower velocity. It indicates that the effective transport occurs at moderate conveyer speed.

Although no disorders are involved in the potential, the plateau can still be found in the low moving lattice amplitude AconvA_{\text{conv}}. In particular, the plateau width increases with the conveyer velocity vconvv_{\text{conv}} [see Fig. 9 (b1)]. This is consistent with the experimental data [given by black points in Fig. 1 (g)] where Aturn-onA_{\text{turn-on}} [defined as AconvA_{\text{conv}} at the line intersection] increases with vconvv_{\text{conv}}. The result with the disorders is shown in Fig. 9 (c), it is interesting to see that the plateau is destroyed with the increase the disorder intensity AdisA_{\text{dis}}. We therefore believe that the plateau cannot be caused by the dynamical localization-delocalization transitions due to the disorders. We infer that the experimental sample is very clean. Their random potential strength is very weak, even if impurities exist.

According to the discussion in sec. III.2, there is no difference between g|ψj|2g|\psi_{j}|^{2} and g1|ψj|2+g2|ψj|4-g_{1}|\psi_{j}|^{2}+g_{2}|\psi_{j}|^{4} in describing the exciton interactions except for the case of the very low particle density. We use the interaction g|ψj|2g|\psi_{j}|^{2} to calculate the gg dependence of M1M_{1} as is shown in Fig. 9 (d). The transport distance M1M_{1} increases firstly and then decreases with gg. The behavior is similar to the experimental data in Fig. 1 (h) where the exciton transport via conveyer M1M_{1} increases firstly and then decreases with the excitation power log(Pex)\log(P_{\text{ex}}). As gNg\propto N when the wave function ψj\psi_{j} is normalized, it indicates that the coherent exciton number Nlog(Pex)N\propto\log(P_{\text{ex}}). We therefore infer that it is less efficient to increase the coherent exciton number by increasing the laser power PexP_{\text{ex}}. The cooling speed and long lifetime are still the key factors.

IV Summary

In summary, we have investigated the experimental data of electrostatic conveyor belt for excitons to understand the dynamical behaviors. We found that the formation of exciton patterns come from both the spatially-separated hot excitons and cooled excitons. The hot excitons can be regarded as classical particles whose transport can be well described by the classical diffusion equation. However, the cooled excitons are the coherent Bosons which must be described by the Schrödinger equation. The studies capture the nature of the exciton diffusion in conveyor belt, i. e. the excitons are cooled down during the transport. In particular, the excitons are in highly degenerate states far from the laser spot, and their temperature becomes lower and lower. This is why the method of the real-time and imaginary-time evolution of the Schrödinger equation can give a good account of the spatial separation patterns. Even though some discrepancies exist between the theory and experiment, the numerical results are expected to be realized in the further.

By calculating the distribution of the escape probability, we get the PL stripes and the transport distance that are consistent with the experiment. We find the cooling speed is the key factor to the transport distance. Based on the comparison between the calculation and the experimental data of transport distance, a function of density, we find increasing the excitation power is not an effective approach to increase the coherent excitons. We also find that the disorders failed to induce the dynamical localization-delocalization transition in the moving lattices. As a result, we infer the sample is basically impurity free.

In order to realize the controllable exciton transport in the moving lattices, according to our studies, two priority research directions are obvious. Experimental study should still focus on finding the ways of the long lifetime and the fast cooling speed of the excitons. Theoretically, designing achievable moving lattice to realize the dynamical localization-delocalization transition may be a new research direction. After finishing the manuscript, we are aware of the experiment of transport and localization of indirect excitons in a van der Waals MoSe2/WSe2 heterostructure Fowler-Gerace et al. (2023). Whether the theory present in the manuscript can explain the data or not is also worthy of further theoretical study.

Acknowledgements.
This work was supported by Hebei Provincial Natural Science Foundation of China (Grant No. A2010001116, A2012203174, and D2010001150), and National Natural Science Foundation of China (Grant Nos. 10974169, 11174115 and 10934008).

References