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

Quasi-resonant diffusion of wave packets in one-dimensional disordered mosaic lattices

Ba Phi Nguyen [email protected] Department of Basic Sciences, Mientrung University of Civil Engineering, Tuy Hoa 620000, Vietnam Research Institute for Basic Sciences, Ajou University, Suwon 16499, Korea    Duy Khuong Phung Computing Fundamentals Department, FPT University, Hanoi 100000, Vietnam    Kihong Kim [email protected] Department of Physics, Ajou University, Suwon 16499, Korea School of Physics, Korea Institute for Advanced Study, Seoul 02455, Korea
Abstract

We investigate numerically the time evolution of wave packets incident on one-dimensional semi-infinite lattices with mosaic modulated random on-site potentials, which are characterized by the integer-valued modulation period κ\kappa and the disorder strength WW. For Gaussian wave packets with the central energy E0E_{0} and a small spectral width, we perform extensive numerical calculations of the disorder-averaged time-dependent reflectance, R(t)\langle R(t)\rangle, for various values of E0E_{0}, κ\kappa, and WW. We find that the long-time behavior of R(t)\langle R(t)\rangle obeys a power law of the form tγt^{-\gamma} in all cases. In the presence of the mosaic modulation, γ\gamma is equal to 2 for almost all values of E0E_{0}, implying the onset of the Anderson localization, while at a finite number of discrete values of E0E_{0} dependent on κ\kappa, γ\gamma approaches 3/2, implying the onset of the classical diffusion. This phenomenon is independent of the disorder strength and arises in a quasi-resonant manner such that γ\gamma varies rapidly from 3/2 to 2 in a narrow energy range as E0E_{0} varies away from the quasi-resonance values. We deduce a simple analytical formula for the quasi-resonance energies and provide an explanation of the delocalization phenomenon based on the interplay between randomness and band structure and the node structure of the wave functions. We explore the nature of the states at the quasi-resonance energies using a finite-size scaling analysis of the average participation ratio and find that the states are neither extended nor exponentially localized, but ciritical states.

I Introduction

Anderson localization of classical waves and quantum particles occurs due to the interference of multiply scattered waves in spatially random media. Since it was discovered theoretically over 60 years ago by Anderson, it has been studied extensively in many areas of physics [1, 2, 3]. Anderson localization arises universally for all kinds of waves and many aspects of the phenomenon have been explored and understood in detail [4, 5, 6, 7]. Nevertheless, there still exist features which are not fully understood and new aspects of localization continue to be discovered when new elements are included in the system [8, 9, 10, 11, 12, 13].

One of the prominent results of early theories of localization is that in the simplest one-dimensional (1D) and two-dimensional random systems, all eigenstates are exponentially localized even in the presence of infinitesimally weak disorder [14]. However, it has been found that this conclusion is not always true in more general random systems and there exist various situations where some states are not exponentially localized, but are either extended or critically localized. Representative examples include the cases where distinct kinds of impedance matching phenomena such as the Brewster anomaly [15, 16, 17, 18, 19, 20] and the Klein effect [21, 22, 23, 24] happen or the random potential is spatially correlated with short- or long-range correlations [25, 26, 27, 28, 29, 30, 31, 32, 33, 34]. For instance, it has been demonstrated both theoretically and experimentally that there exist extended states at two energy values in the 1D random dimer model which contains a special type of short-range correlated disorder [25, 26]. This model has been generalized to the case of random NN-mer systems and analytical expressions for the values of the resonant energies at which delocalization arises have been obtained [27, 28, 29, 30].

A discrete set of delocalized states can appear in short-range correlated random systems. In contrast, a continuum of delocalized states with sharp mobility edges was predicted theoretically and confirmed experimentally to appear in long-range correlated random systems in 1D [31, 32, 33, 34, 35, 36, 37]. The mobility edges, which are the energy values separating localized and extended states, can also be present in a wide range of quasiperiodic models in 1D [38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49]. Recently, an exactly solvable 1D model with multiple mobility edges called quasiperiodic mosaic lattice model has been proposed [43]. In that model, a quasiperiodic on-site potential exists only at periodically spaced sites, while the potential is constant at all other sites. The number and the positions of the mobility edges have been shown to depend sensitively on the modulation period κ\kappa.

The main aim of the present paper is to propose a new way of inducing delocalized states in 1D random systems. We consider a random version of the quasiperiodic mosaic lattice model, which we call disordered mosaic lattice model, where the on-site potential takes a random value only at equally spaced sites with a period κ\kappa. We study the transport and localization properties of the proposed model primarily by investigating the time evolution of wave packets incident on an effectively semi-infinite disordered mosaic lattice chain. More specifically, we calculate the time-dependent reflectance averaged over a large number of independent disorder configurations, R(t)\langle R(t)\rangle, for various values of the modulation period κ\kappa and the central energy of the wave packet, E0E_{0}. This approach based on the reflection geometry has substantial experimental advantages over those based on the transmission geometry.

We are especially interested in exploring the long-time scaling behavior of R(t)\langle R(t)\rangle, which obeys a power-law decay of the form tγt^{-\gamma} for all values of the parameters. From many previous researches, it has been solidly established that in the cases where the standard Anderson localization occurs, the exponent γ\gamma is equal to 2, while in those where the classical diffusion occurs, it is 3/2 [50, 51, 52, 53, 54, 55, 56, 57, 58]. From extensive numerical calculations, we will find that in the presence of the mosaic modulation, γ\gamma is equal to 2 for almost all values of E0E_{0}, while at a finite number of discrete values of E0E_{0} dependent on κ\kappa, γ\gamma approaches 3/2. In other words, although most eigenstates of the disordered mosaic lattice model are exponentially localized, there appear a finite number of discrete states that are not exponentially localized but display transport behavior characteristic of classical diffusion. We will also find that this phenomenon is independent of the disorder strength and occurs in a quasi-resonant manner. From the numerical results, we will deduce a simple analytical formula for the quasi-resonance energies at which delocalization occurs and provide an explanation of the phenomenon based on the interplay between randomness and band structure and the node structure of the wave functions. We also explore the nature of the states at the quasi-resonance energies using a finite-size scaling analysis of the average participation ratio and find that the states are neither extended nor exponentially localized, but ciritical states.

The rest of this paper is organized as follows. In Sec. II, we introduce the 1D disordered mosaic lattice model characterized by the time-independent Schrödinger equation within the nearest-neighbor tight-binding approximation. We also describe the numerical calculation method and the physical quantities of interest. In Sec. III, we present the numerical results and discuss the mechanism for the onset of the diffusive behavior. We also explore the nature of the states at the quasi-resonance energies using a finite-size scaling analysis of the average participation ratio. Finally, in Sec. IV, we conclude the paper.

II Theoretical model and method

II.1 Model

To describe 1D non-interacting spinless particle systems, we use the standard single-chain tight-binding model

J(ψn1+ψn+1)+εnψn=Eψn,\displaystyle J(\psi_{n-1}+\psi_{n+1})+\varepsilon_{n}\psi_{n}=E{\psi_{n}}, (1)

where ψn\psi_{n} and εn\varepsilon_{n} are the wave function amplitude and the on-site potential at the nn-th lattice site respectively. EE is the energy and JJ is the coupling strength between nearest-neighbor sites. From now on, we will measure all energy scales in units of JJ and set it equal to 1. In this study, we will investigate the transport and localization properties of a model which we call disordered mosaic lattice model. This model is defined by

εn={βn[W,W],n=mκV0,otherwise,\displaystyle\varepsilon_{n}=\left\{\begin{array}[]{l l}\beta_{n}\in[-W,W],&\quad\mbox{$n=m\kappa$}\\ V_{0},&\quad\mbox{otherwise}\end{array}\right., (4)

where the inlay parameter κ\kappa representing the period of the mosaic modulation is a fixed positive integer larger than 1 and mm is an integer running from 1 to NN. Then the total number of sites LL is equal to κN\kappa N. The on-site potential βn\beta_{n} at the mκm\kappa-th site is a random variable uniformly distributed in the interval [W,W][-W,W], where WW is the strength of disorder. In all other sites, the on-site potential takes a constant value of V0V_{0}. In [43], the authors have studied a quasiperiodic mosaic lattice model, where βn\beta_{n} is a quasiperiodic potential of Aubry-André-type [59]. Our model is different from that model in that βn\beta_{n} is random instead of quasiperiodic. We note that if βn\beta_{n} is a constant potential different from V0V_{0}, the model becomes perfectly periodic with period κ\kappa.

II.2 Method

Following the procedure given in [58], we first assume that a monochromatic wave of energy EE is incident from the left side of the disordered region and define the amplitudes of the incident, reflected, and transmitted waves AA, BB, and CC by

ψn={Aeiqn+Beiqn,n=1,2Ceiqn,n=L1,L,\displaystyle\psi_{n}=\left\{\begin{array}[]{l l}Ae^{iqn}+Be^{-iqn},&\quad\mbox{$n=1$},2\\ Ce^{iqn},&\quad\mbox{$n=L-1$},L\end{array}\right., (7)

where qq is related to EE by the free-space dispersion relation E=2cosqE=2\cos q. In the absence of dissipation, the law of energy conservation |B|2+|C|2=|A|2|B|^{2}+|C|^{2}=|A|^{2} should be satisfied. In order to solve Eq. (1) numerically, we first fix ψL1\psi_{L-1} to 1, then we obtain C=exp[iq(L1)]C=\exp[-iq(L-1)] and ψL=exp(iq)\psi_{L}=\exp(iq). Knowing the values of ψL\psi_{L} and ψL1\psi_{L-1}, we can solve Eq. (1) iteratively to obtain ψL2\psi_{L-2}, ψL3\psi_{L-3}, \cdots, ψ2\psi_{2}, ψ1\psi_{1}. Using the definition of AA and BB given in Eq. (7), we can express them in terms of ψ1\psi_{1} and ψ2\psi_{2}:

A=e2iqψ2ψ1eiq1e2iq,B=ψ1eiqψ21e2iq.\displaystyle A=e^{-2iq}\frac{\psi_{2}-\psi_{1}e^{-iq}}{1-e^{-2iq}},~{}B=\frac{\psi_{1}e^{iq}-\psi_{2}}{1-e^{-2iq}}. (8)

We are interested in the case where the length of the disordered region is sufficiently large. Then the large portion of the incident wave power will be reflected. Therefore we focus on the behavior of the reflection coefficient r~(E)\tilde{r}(E) and the reflectance R~(E)\tilde{R}(E) defined by

r~(E)=BA=e2iqψ1eiqψ2ψ2ψ1eiq,\displaystyle\tilde{r}(E)=\frac{B}{A}=e^{2iq}\frac{\psi_{1}e^{iq}-\psi_{2}}{\psi_{2}-\psi_{1}e^{-iq}},
R~(E)=|BA|2=|ψ2ψ1eiqψ2ψ1eiq|2.\displaystyle\tilde{R}(E)=\left|\frac{B}{A}\right|^{2}=\left|\frac{\psi_{2}-\psi_{1}e^{iq}}{\psi_{2}-\psi_{1}e^{-iq}}\right|^{2}. (9)

Next we consider a Gaussian wave packet characterized by the spectrum

f(E)=(2π)1/4σexp[(EE0)24σ2],\displaystyle f(E)=\frac{(2\pi)^{1/4}}{\sqrt{\sigma}}\exp\left[\ -\frac{(E-E_{0})^{2}}{4\sigma^{2}}\right], (10)

where E0E_{0} is the central energy of the wave packet and σ\sigma is its spectral width. f(E)f(E) satisfies

12π𝑑E[f(E)]2=1.\displaystyle\frac{1}{2\pi}\int_{-\infty}^{\infty}dE~{}\left[f(E)\right]^{2}=1. (11)

The time-dependent reflection coefficient r(t)r(t) of the incident Gaussian wave packet can be calculated using

r(t)=12π22𝑑Er~(E)f(E)exp(iEt),\displaystyle r(t)=\frac{1}{2\pi}\int_{-2}^{2}dE~{}\tilde{r}(E)f(E)\exp(iEt), (12)

where the range of integration is limited to 2E2-2\leq E\leq 2, since the reflection coefficient r~(E)\tilde{r}(E) can be defined only inside the band satisfying E=2cosqE=2\cos q. The disorder-averaged time-dependent reflectance R(t)\left\langle R(t)\right\rangle is obtained from

R(t)=|r(t)|2=1(2π)222𝑑E122𝑑E2r~(E1)r~(E2)f(E1)f(E2)exp[i(E1E2)t],\displaystyle\left\langle R(t)\right\rangle=\langle\left|r(t)\right|^{2}\rangle=\frac{1}{(2\pi)^{2}}\int_{-2}^{2}dE_{1}\int_{-2}^{2}dE_{2}~{}\langle\tilde{r}(E_{1})\tilde{r}^{\ast}(E_{2})\rangle f(E_{1})f^{\ast}(E_{2})\exp[i(E_{1}-E_{2})t], (13)

where \langle\cdots\rangle denotes averaging over a large number of different disorder configurations.

According to the previous theories of Anderson localization in the time domain [50, 52, 53, 54, 55, 58], the exponent characterizing the power-law decay of R(t)\left\langle R(t)\right\rangle in the long-time limit provides an important information on the state of the system. Specifically, it has been established that R(t)\left\langle R(t)\right\rangle decays as t2t^{-2} in the localized regime, while it decays as t3/2t^{-3/2} in the regime of classical diffusion. Finally, we point out that studying Anderson localization based on the measurements in the reflection geometry has substantial experimental advantages over more conventional approaches in the transmission geometry, in that the reflectance is often more easily measurable than the transmittance and there exist situations where measurements in the transmission mode are not possible. These advantages are especially relevant to the fields such as optics, acoustics, and seismology [60, 61, 62, 63].

III Numerical results and discussion

III.1 Numerical results

Refer to caption
Figure 1: Ln–ln plot of the disorder-averaged time-dependent reflectance R(t)\langle R(t)\rangle versus time tt for various values of E0E_{0} when κ=1\kappa=1, W=2W=2, σ=0.05\sigma=0.05, V0=0V_{0}=0, and L=1000L=1000. The time dependence of R(t)\langle R(t)\rangle obeys an inverse-square law of the form R(t)t2\langle R(t)\rangle\propto t^{-2} in the long-time limit.
Refer to caption
Refer to caption
Refer to caption
Figure 2: Ln–ln plots of R(t)\langle R(t)\rangle versus tt for various values of E0E_{0} when κ=2\kappa=2 and (a) W=1W=1, (b) W=2W=2, and (c) W=4W=4. σ\sigma and V0V_{0} are fixed to 0.05 and 0 respectively and LL is chosen to be 1300 in (a) and 1000 in (b) and (c). The curve for E0=0E_{0}=0 that is highlighted with a thick line scales as t3/2t^{-3/2} in the long-time limit in contrast with other curves that scale as t2t^{-2}.

In most of our calculations, we have set V0=0V_{0}=0 and computed R(t)\langle R(t)\rangle by averaging over 10,000 distinct random configurations of εn\varepsilon_{n}. The step size in energy, ΔE\Delta E, is 10310^{-3}. We have chosen the system size LL to be sufficiently large so that the transmission is negligible and the system can be considered an effectively semi-infinite medium.

Our main goal is to study the dynamics of wave packet propagation in 1D disordered mosaic lattices with κ2\kappa\geq 2. However, it is instructive to first consider the κ=1\kappa=1 case corresponding to the ordinary Anderson model to clarify the effects of the mosaic modulation. In a recent study, a numerical analysis of the time-dependent reflectance for the 1D Anderson model has been presented [58]. It has been reported that for any W[1,4]W\in[1,4] and E0[2+2σ,22σ]E_{0}\in[-2+2\sigma,2-2\sigma], the time dependence of R(t)\langle R(t)\rangle obeys an inverse-square law of the form R(t)t2\langle R(t)\rangle\propto t^{-2} in the long-time limit. This implies that all the incident wave packets exhibit an Anderson localization behavior. It has also been shown that this behavior is independent of the spectral shape of the wave packet or its spectral width σ\sigma as long as it is as small as 0.05 and 1. In the present work, we consider Gaussian wave packets with a fixed spectral width of σ=0.05\sigma=0.05. In Fig. 1, we show our calculations for the time decay of R(t)\langle R(t)\rangle when κ=1\kappa=1, W=2W=2, and the system size L=1000L=1000, which agree well with [58].

We are interested in exploring the transport behavior when the random mosaic modulation defined by Eq. (4) is introduced into the on-site potential of a lattice model. In Fig. 2, we set κ=2\kappa=2 and plot the time evolution of R(t)\langle R(t)\rangle for various values of E0E_{0} when W=1W=1, 2, and 4. We find that for all values of E0E_{0}, R(t)\langle R(t)\rangle shows a power-law decay of the form R(t)tγ\langle R(t)\rangle\propto t^{-\gamma} in the long-time limit. The exponent γ\gamma is equal to 2 for almost all E0E_{0} values except for a narrow region close to E0=0E_{0}=0. In contrast, the value of γ\gamma at E0=0E_{0}=0 is equal to 3/23/2 with a good approximation and is markedly different from the Anderson localization behavior shown for other E0E_{0} values. In addition, the value of R(t)\langle R(t)\rangle for E0=0E_{0}=0 is noticeably larger than those for other E0E_{0} values when lnt\ln t is sufficiently large, as can be seen from the curves drawn with thick lines in Fig. 2. This behavior occurs for all values of the disorder parameter WW considered here, as long as the system length LL is sufficiently large such that the effect due to the leakage of the wave packet into the transmitted region is negligible. When the disorder is weak, we need to use a larger LL to satisfy such a condition.

Refer to caption
Refer to caption
Refer to caption
Figure 3: Ln–ln plots of R(t)\langle R(t)\rangle versus tt for various values of E0E_{0} when κ=3\kappa=3 and (a) W=1W=1, (b) W=2W=2, and (c) W=4W=4. σ\sigma, V0V_{0}, and LL are fixed to 0.05, 0, and 1050 respectively. The curve for E0=1E_{0}=1 that is highlighted with a thick line scales as t3/2t^{-3/2} in the long-time limit in contrast with other curves that scale as t2t^{-2}.
Refer to caption
Refer to caption
Refer to caption
Figure 4: Ln–ln plots of R(t)\langle R(t)\rangle versus tt for various values of E0E_{0} when κ=4\kappa=4 and (a) W=1W=1, (b) W=2W=2, and (c) W=4W=4. σ\sigma, V0V_{0}, and LL are fixed to 0.05, 0, and 1200 respectively. The curves for E0=0E_{0}=0 and 1.4 (2\approx\sqrt{2}) scale as t3/2t^{-3/2} in the long-time limit in contrast with other curves that scale as t2t^{-2}.
Refer to caption
Refer to caption
Refer to caption
Figure 5: The power-law exponent γ\gamma obtained by fitting the numerical results in the long-time region where 5<lnt<85<\ln t<8 plotted versus E0E_{0} when W=2W=2, 3, 4, 5 and (a) κ=2\kappa=2, (b) κ=3\kappa=3, and (c) κ=4\kappa=4. σ\sigma and V0V_{0} are fixed to 0.05 and 0 respectively and LL is 1000 in (a), 1050 in (b), and 1200 in (c). The two dashed lines correspond to γ=3/2\gamma=3/2 and 22 which are expected for diffusive and localized wave packets respectively.

In Fig. 3, we set κ=3\kappa=3 and plot the time evolution of R(t)\langle R(t)\rangle for the same values of E0E_{0} and WW as in Fig. 2. Similarly to the κ=2\kappa=2 case, the localization behavior takes place for most values of E0E_{0}, while the diffusive behavior appears only near a special value of E0E_{0}. In contrast to the previous case, however, the diffusive behavior with γ=3/2\gamma=3/2 is observed at E0=1E_{0}=1. In addition, we have verified numerically that a similar diffusive behavior also occurs at E0=1E_{0}=-1, though it has not been shown here explicitly.

Next, in Fig. 4, we consider the κ=4\kappa=4 case. The overall long-time behavior is similar to the κ=2\kappa=2 and 3 cases, but the diffusive behavior with γ=3/2\gamma=3/2 occurs at two different values E0=0E_{0}=0 and 1.4 (2\approx\sqrt{2}) in the present case. In addition, a similar behavior is also observed at E0=2E_{0}=-\sqrt{2}. Since the diffusive long-time behavior appears only in a narrow range of E0E_{0} values, it can be considered as a kind of quasi-resonance. Combining the results obtained for κ=2\kappa=2, 3, and 4, we deduce that the quasi-resonances occur symmetrically with respect to E0=0E_{0}=0 and their total number is equal to κ1\kappa-1. We also find that this behavior is unaffected by the disorder strength when W1W\gtrsim 1.

In order to examine the quasi-resonant nature of the diffusive long-time behavior more clearly, we plot the exponent γ\gamma obtained by fitting the numerical results in the long-time region where 5<lnt<85<\ln t<8 versus E0E_{0}, when W=2W=2, 3, 4, 5 and κ=2\kappa=2, 3, 4 in Fig. 5. We point out that there is a mirror symmetry in a statistical sense with respect to E0=0E_{0}=0, though the region with E0<0E_{0}<0 is not shown here. Numerical results for many E0E_{0} values around the sharp dips at E0=0E_{0}=0, 11, and 2\sqrt{2} in addition to those shown in Figs. 2, 3, and 4 have been used in Fig. 5. As E0E_{0} varies away from the values at the dips, γ\gamma increases rapidly from 1.5 to 2. The half-widths (full widths at half maximum) of all the dips are similar and roughly equal to 0.15. In order to show the positions of the quasi-resonances more clearly, some values of E0E_{0} versus γ\gamma around the sharp dips are listed in Table 1.

Table 1: List of the values of E0E_{0} versus γ\gamma around the sharp dips at E0=1E_{0}=1 in Fig. 5(b) and at E0=1.4E_{0}=1.4 in Fig. 5(c). The minimum value of γ\gamma approaches 3/23/2 at E0=1E_{0}=1 when κ=3\kappa=3 and at E0=1.42E_{0}=1.42 (2\approx\sqrt{2}) when κ=4\kappa=4.
            κ=3\kappa=3                     κ=4\kappa=4
E0E_{0} γ\gamma E0E_{0} γ\gamma
0.90 1.778 1.30 1.822
0.92 1.710 1.32 1.756
0.94 1.636 1.34 1.680
0.96 1.577 1.36 1.610
0.98 1.534 1.38 1.554
1.00 1.521 1.40 1.527
1.02 1.535 1.42 1.510
1.04 1.577 1.44 1.536
1.06 1.640 1.46 1.580
1.08 1.715 1.48 1.638
1.10 1.786 1.50 1.710
Refer to caption
Figure 6: Ln–ln plots of R(t)\langle R(t)\rangle versus tt for several values of κ\kappa when E0=0E_{0}=0 and W=4W=4. σ\sigma and V0V_{0} are fixed to 0.05 and 0 respectively. The curves for even values of κ\kappa scale as t3/2t^{-3/2} while those for odd values of κ\kappa scale as t2t^{-2} in the long-time limit.

In the rest of this subsection, we present some additional results which are useful in deducing a general formula for the quasi-resonant energy values. In Fig. 6, we plot R(t)\langle R(t)\rangle versus tt at E0=0E_{0}=0 when W=4W=4, κ=2\kappa=2, 3, 4, 5, and 6, and V0=0V_{0}=0. We find that the behavior of the incident wave packet at E0=0E_{0}=0 depends on the parity of κ\kappa. When κ\kappa is odd, the wave packet exhibits Anderson localization, while, when κ\kappa is even, it does a diffusive behavior.

Finally, in Fig. 7, we compare the long-time scaling behavior at E0=0E_{0}=0 and 0.2, when V0=0V_{0}=0 and V0=0.2V_{0}=0.2 for κ=2\kappa=2, 4, and 6. A diffusive behavior is observed at E0=0E_{0}=0 when V0=0V_{0}=0, while it is found at E0=V0E_{0}=V_{0} when V0=0.2V_{0}=0.2. Therefore the quasi-resonance energy is found to depend on the value of V0V_{0} as well as κ\kappa.

Refer to caption
Refer to caption
Figure 7: Ln–ln plots of R(t)\langle R(t)\rangle versus tt for κ=2,\kappa=2, 4, 6, E0=0E_{0}=0, 0.2, σ=0.5\sigma=0.5, and W=4W=4, when (a) V0=0V_{0}=0 and (b) V0=0.2V_{0}=0.2. In (a), the curves for E0=0E_{0}=0 scale as t3/2t^{-3/2} and those for E0=0.2E_{0}=0.2 scale as t2t^{-2} in the long-time limit. In (b), the opposite behavior is observed.
Refer to caption
Figure 8: Band structure of a periodic mosaic lattice model when κ=4\kappa=4, β=10\beta=10, and V0=0V_{0}=0. The gray region denotes the forbidden bands. The dispersion relation between EE and (4/π)q(4/\pi)q is plotted in red inside the allowed bands. The energy values at the upper edges of the allowed bands are 2\sqrt{2}, 0, and 2-\sqrt{2}, which agree precisely with Eq. (14).

III.2 Quasi-resonant diffusion of wave packets

The numerical results presented above clearly show that a wave packet propagating inside a disordered mosaic lattice displays a diffusive behavior for some special values of the wave packet’s central energy in a quasi-resonant manner, although it exhibits Anderson localization for all the other values of the central energy. In other words, mosaic modulation of the lattice potential causes the usual exponential Anderson localization to be destroyed at special discrete values of the energy dependent on the modulation period. Thus we find that the periodic mosaic modulation is a feature that can induce a new type of delocaliztion in low-dimensional disordered systems.

From our numerical results given in the previous subsection, we can easily deduce that the central energy at the quasi-resonances, ERE_{R}, is given precisely by

ER=V0+2cos(πκn)(n=1,2,,κ1).\displaystyle E_{R}=V_{0}+2\cos\left(\frac{\pi}{\kappa}n\right)~{}~{}(n=1,2,\cdots,\kappa-1). (14)

The number of ERE_{R} is equal to κ1\kappa-1. ERE_{R} is distributed symmetrically with respect to V0V_{0} and ER=V0E_{R}=V_{0} is included only when κ\kappa is even. For several small values of κ\kappa, we obtain

ER={V0for κ=2V0±1for κ=3V0,V0±2for κ=4V0±512,V0±5+12for κ=5V0,V0±1,V0±3for κ=6.\displaystyle E_{R}=\left\{\begin{array}[]{l l}V_{0}&\quad\mbox{for $\kappa=2$}\\ V_{0}\pm 1&\quad\mbox{for $\kappa=3$}\\ V_{0},V_{0}\pm\sqrt{2}&\quad\mbox{for $\kappa=4$}\\ V_{0}\pm\frac{\sqrt{5}-1}{2},V_{0}\pm\frac{\sqrt{5}+1}{2}&\quad\mbox{for $\kappa=5$}\\ V_{0},V_{0}\pm 1,V_{0}\pm\sqrt{3}&\quad\mbox{for $\kappa=6$}\\ \cdots\end{array}\right.. (21)

The analytical formula for the quasi-resonance energy ERE_{R} appears to agree with the formula obtained for the resonance energy [e.g., Eq. (16) of Ref. 29] where the delocalization occurs in 1D binary random NN-mer models with N=κN=\kappa. However, this agreement is only coincidental and superficial. In the binary random NN-mer model, the resonance energy has been obtained from the condition that the transmittance through a single NN-mer is unity, and therefore the transmittance at the resonance energies is identically equal to 1 and the corresponding states are completely extended [27, 28, 29, 30]. In contrast, at our quasi-resonance energies, the transmittance is not equal to 1 and the states are not extended but critical states, as will be explained in Sec. III.3. In the binary random NN-mer model of Ref. 29, the on-site potential can take one of the two values 0 and V0V_{0}. The problem becomes nonrandom and trivial if V0=0V_{0}=0 and so we need to exclude that case. In order to have extended states, it is also necessary to have the condition that |ER|2|E_{R}|\leq 2, which gives another constraint for V0V_{0} [29]. On the contrary, in our model, V0V_{0} is just a background on-site potential and can take any arbitrary value including zero. We also point out that the cosine form for the resonance energy of the NN-mer model arises only when the on-site potentials at all the NN sites of an NN-mer are the same. Even if they are not uniform, there can still exist resonance energies, but they will not be given by a simple cosine form.

We pay close attention to the results that the dependence of γ\gamma on E0E_{0} shown in Fig. 5 is insensitive to the strength of disorder WW. In our model, the random potential is present only at periodically spaced sites n=κmn=\kappa m (m=1,2,3,m=1,2,3,\cdots). In order for the numerical results to be insensitive to the disorder strength, it is natural to assume that the wave-function amplitudes take very small values close to zero at those sites. That is, the wave function has nodes there. This implies that the eigenfunctions have a periodicity with the wavelength λ\lambda satisfying

κ=λ2n(n=1,2,3,).\displaystyle\kappa=\frac{\lambda}{2}n~{}~{}~{}~{}(n=1,2,3,\cdots). (22)

For such eigenfunctions, the energy eigenvalues should be very close to the values obtained in the disorder-free case, which are given by E=V0+2cosqE=V_{0}+2\cos q. Since the wave vector qq is related to λ\lambda by q=2π/λq=2\pi/\lambda, we can obtain Eq. (14) in a straightforward way. This argument strongly suggests that the states at the energies ERE_{R} are rather insensitive to the disorder and therefore are not standard exponentially localized states. Our conjecture that the wave functions at the quasi-resonance energies have nodes at n=κmn=\kappa m will be confirmed later in Fig. 9 in Sec. III.3.

It is highly instructive to consider a related periodic model, which we may call periodic mosaic lattice model, defined by Eq. (4), but with βn\beta_{n} replaced by a constant β\beta. Since this model is strictly periodic with a period κ\kappa, it contains alternating forbidden and allowed bands. In Fig. 8, we show its band structure when κ=4\kappa=4, β=10\beta=10, and V0=0V_{0}=0. The gray-colored region denotes the forbidden bands. The dispersion relation between EE and (4/π)q(4/\pi)q is plotted in red inside the allowed bands. We have found that the energy values at the upper edges of the allowed bands are precisely given by Eq. (14) for all values of κ\kappa, β\beta, and V0V_{0}. As the potential strength β\beta increases to large values, the widths of the allowed bands become very small. In the large β\beta limit, the allowed bands consist of infinitesimally narrow regions at the energies given by Eq. (14). If we compare this limiting case with the disordered mosaic lattice model for large WW, we find that the forbidden bands and the narrow allowed bands at ERE_{R} in the periodic case are replaced respectively by exponentially localized states and diffusive states in the random case.

Refer to caption
Figure 9: Average participation ratio P(E)\langle P(E)\rangle of a disordered mosaic lattice plotted versus (a) energy EE for L=1000L=1000 and (b) lattice size LL for E=1E=1, when κ=3\kappa=3 and W=1W=1, 2. P(E)\langle P(E)\rangle is obtained by averaging over the eigenstates within the interval ΔE=0.1\Delta E=0.1 around EE and over 10001000 distinct disorder realizations. Two typical wave functions corresponding to the eigenstates with (c) E=0E=0 and (d) E=1E=1 for a single disorder configuration plotted versus site index nn when κ=3\kappa=3, L=1000L=1000, and W=5W=5. Inset of (d) shows an expanded plot of the wave function in the range of 0<n<300<n<30.

III.3 Nature of the states at the quasi-resonance energies

In order to better understand the nature of the states at the quasi-resonance energies, we digress briefly from the study of the time-dependent reflectance to consider the spectral properties of a finite disordered mosaic lattice. Specifically we calculate the participation ratio P(Ek)P(E_{k}) for the kk-th eigenstate with the energy eigenvalue EkE_{k}, which is defined by [64]

P(Ek)=(n|ψn(k)|2)2n|ψn(k)|4,\displaystyle P\left(E_{k}\right)=\frac{\left(\sum_{n}\left|\psi_{n}^{(k)}\right|^{2}\right)^{2}}{\sum_{n}\left|\psi_{n}^{(k)}\right|^{4}}, (23)

where ψn(k)\psi_{n}^{(k)} is the value of the kk-th eigenfunction at the site nn (n=1,2,,Ln=1,2,\cdots,L). For a finite lattice, P(Ek)P(E_{k}) gives approximately the number of sites over which the kk-th eigenfunction is extended. To study the large-LL scaling behavior of the participation ratio in a disordered system, it is convenient to introduce a double-averaged quantity P(E)\langle P(E)\rangle, where P(E)P(E) is obtained by averaging over all eigenstates within a narrow interval ΔE\Delta E around EE and \langle\cdots\rangle denotes averaging over a large number of independent disorder configurations. When LL is sufficiently large, we find a power-law scaling behavior of the form

P(E)Lx\displaystyle\langle P(E)\rangle\propto L^{x} (24)

with the scaling exponent xx. It has been well-established that for extended states, the scaling exponent xx is equal to 1 and P(E)\langle P(E)\rangle increases linearly with increasing the lattice size LL. On the contrary, for (exponentially) localized states, the exponent xx is zero, that is, P(E)\langle P(E)\rangle does not depend on LL and converges to a constant value as LL\to\infty. For critical states at the boundary between extended and localized states, the exponent should be in the intermediate range of 0<x<10<x<1. Thus the finite-size scaling analysis of P(E)\langle P(E)\rangle provides a very useful information about the nature of the states.

In Figs. 9(a) and 9(b), we show the results of numerical calculations of P(E)\langle P(E)\rangle obtained when κ=3\kappa=3, ΔE=0.1\Delta E=0.1, W=1W=1, 2, and the number of disorder configurations is 1000. In Fig. 9(a), we find that the average participation ratio for L=1000L=1000 has pronounced peaks at the quasi-resonance energies E=±1E=\pm 1, while it remains small at other energies. We have confirmed that for the value of EE away from ±1\pm 1, P(E)\langle P(E)\rangle approaches a constant as LL increases to large values, implying that xx is zero and the states are exponentially localized. In contrast, at E=±1E=\pm 1, xx is neither 0 nor 1 as demonstrated for E=1E=1 in Fig. 9(b). From fitting the data to Eq. (24), we obtain x0.25x\approx 0.25 for W=1W=1 and x0.23x\approx 0.23 for W=2W=2. These results clearly demonstrate that the states at the quasi-resonance energies are neither extended nor exponentially localized states. We notice that they closely resemble the critical states observed at Anderson localisation-delocalisation and quantum Hall plateau transitions [3, 65]. They are totally different from the resonant states appearing in the binary random NN-mer model, where such states have been found to be completely extended.

In Figs. 9(c) and 9(d), we illustrate the spatial distributions of the wave function amplitudes obtained for a single disorder configuration when κ=3\kappa=3, L=1000L=1000, W=5W=5, and E=0E=0, 1. When the energy is away from the quasi-resonance energies, we find that the wave function is strongly localized as in Fig. 9(c). At the quasi-resonance energies, however, the wave function has a distinct spatial distribution with several disjointed occupied regions as in Fig. 9(d). This is a unique characteristic frequently observed in critical or multifractal states. We also find that the wave function amplitudes are very close to zero at all the sites satisfying n=mκn=m\kappa (m=1,2,m=1,2,\cdots) as shown in the inset of Fig. 9(d), which is fully consistent with our conjecture that there are wave function nodes at such sites. This node structure is absent for localized wave functions and occurs only at the quasi-resonance energies. It leads to the behavior that when WW is sufficiently large, the quasi-resonant states are insensitive to the strength of disorder.

From separate calculations for the eigenfunctions of the periodic mosaic model, we have confirmed that they also have nodes at n=mκn=m\kappa (m=1,2,m=1,2,\cdots) only for the energies given by Eq. (14). Let us suppose that starting from the periodic mosaic model, we turn on the weak disorder at the sites n=mκn=m\kappa (m=1,2,m=1,2,\cdots) and increase its strength gradually. The special node structure of the wave function we have discussed so far will be maintained at the quasi-resonance energies regardless of the strength of disorder. Therefore the node structure of the wave functions is a crucial feature that connects the periodic and disordered mosaic lattice models and induces the critical states at the quasi-resonance energies. A more systematic and comprehensive analysis about the nature of the quasi-resonant states in the framework of the spectral problem will be presented elsewhere.

IV Conclusion

In this paper, we have studied numerically the time evolution of Gaussian wave packets in an effectively semi-infinite disordered mosaic lattice chain where the on-site potential takes a random value only at periodically spaced sites. We have performed extensive numerical calculations of the disorder-averaged time-dependent reflectance, R(t)\langle R(t)\rangle, for various values of the wave packet’s central energy E0E_{0}, the modulation period κ\kappa, and the disorder strength WW. We have found that the long-time behavior of R(t)\langle R(t)\rangle obeys a power-law decay of the form tγt^{-\gamma} in all cases. In the absence of the mosaic modulation (i.e., κ=1\kappa=1), the exponent γ\gamma is equal to 2 regardless of the parameters, implying the onset of the standard Anderson localization. When the mosaic modulation is turned on (i.e., κ2\kappa\geq 2), γ\gamma is still equal to 2 for almost all values of E0E_{0}, while at a finite number (equal to κ1\kappa-1) of discrete values of E0E_{0} dependent on κ\kappa, γ\gamma approaches 3/2, implying the onset of the classical diffusion. We have found that this phenomenon is independent of WW as long as it is sufficiently large and occurs in a quasi-resonant manner such that γ\gamma varies rapidly from 3/2 to 2 in a narrow energy range as E0E_{0} varies away from the quasi-resonance values. We have deduced a simple analytical formula for the quasi-resonance energies and provided an explanation of this novel delocalization phenomenon based on the interplay between randomness and band structure and the node structure of the wave functions. We have also explored the nature of the states at the quasi-resonance energies using a finite-size scaling analysis of the average participation ratio and found that the states are neither extended nor exponentially localized, but ciritical states. The model proposed here can be readily realized experimentally using various physical systems, which include coupled optical waveguide arrays, synthetic photonic lattices, and ultracold atoms.

In the future research, it is desirable to perform a comprehensive spectral analysis of the participation ratio, the logarithmic transmittance, and the localization length to better elucidate the true nature of the states at the quasi-resonance energies as well as to test the theoretical predictions experimentally. Another interesting direction of research is to explore the spreading dynamics of a wave packet launched at the center of a long disordered mosaic lattice, which is expected to provide an understanding of the states at the quasi-resonance energies from a different perspective.

Acknowledgements.
B.P.N. would like to thank Felix Izrailev for carefully reading a draft version of the manuscript and providing valuable comments. We also appreciate greatly very helpful comments and suggestions by an anonymous referee and Seulong Kim. This research was supported through a National Research Foundation of Korea Grant (NRF-2022R1F1A1074463) funded by the Korean Government. It was also supported by the Basic Science Research Program funded by the Ministry of Education (2021R1A6A1A10044950) and by the Global Frontier Program (2014M3A6B3063708).

References

  • [1] P. W. Anderson, Absence of diffusion in certain random lattices, Phys. Rev. 109, 1492 (1958).
  • [2] P. A. Lee and T. V. Ramakrishnan, Disordered electronic systems, Rev. Mod. Phys. 57, 287 (1985).
  • [3] F. Evers and A. D. Mirlin, Anderson transitions, Rev. Mod. Phys. 80, 1355 (2008).
  • [4] H. Hu, A. Strybulevych, J. H. Page, S. E. Skipetrov, and B. A. van Tiggelen, Localization of ultrasound in a three-dimensional elastic network, Nat. Phys. 4, 945 (2008).
  • [5] G. Modugno, Anderson localization in Bose-Einstein condensates, Rep. Prog. Phys. 73, 102401 (2010).
  • [6] S. A. Gredeskul, Y. S. Kivshar, A. A. Asatryan, K. Y. Bliokh, Y. P. Bliokh, V. D. Freilikher, and I. V. Shadrivov, Anderson localization in metamaterials and other complex media, Low Temp. Phys. 38, 570 (2012).
  • [7] M. Segev, Y. Silberberg, and D. N. Christodoulides, Anderson localization of light, Nat. Photon. 7, 197 (2013).
  • [8] R. R. Naraghi and A. Dogariu, Phase transitions in diffusion of light, Phys. Rev. Lett. 117, 263901 (2016).
  • [9] D. Leykam, J. D. Bodyfelt, A. S. Desyatnikov, and S. Flach, Localization of weakly disordered flat band states, Eur. Phys. J. B 90, 1 (2017).
  • [10] S. Derevyanko, Anderson localization of a one-dimensional quantum walker, Sci. Rep. 8, 1795 (2018).
  • [11] Y. Sharabi, H. H. Sheinfux, Y. Sagi, G. Eisenstein, and M. Segev, Self-induced diffusion in disordered nonlinear photonic media, Phys. Rev. Lett. 121, 233901 (2018).
  • [12] B. P. Nguyen, T. K. T. Lieu, and K. Kim, Numerical study of the transverse localization of waves in one-dimensional lattices with randomly distributed gain and loss: effect of disorder correlations, Waves Random Complex Media 32, 390 (2022).
  • [13] P. Pelletier, D. Delande, V. Josse, A. Aspect, S. Mayboroda, D. N. Arnold, and M. Filoche, Spectral functions and localization-landscape theory in speckle potentials, Phys. Rev. A 105, 023314 (2022).
  • [14] E. Abrahams, P. W. Anderson, D. C. Licciardello and T. V. Ramakrishnan, Scaling theory of localization: Absence of quantum diffusion in two dimensions, Phys. Rev. Lett. 42, 673 (1979).
  • [15] J. E. Sipe, P. Sheng, B. S. White, and M. H. Cohen, Brewster anomalies: A polarization-induced delocalization effect, Phys. Rev. Lett. 60, 108 (1988).
  • [16] K. Kim, F. Rotermund, D.-H. Lee, and H. Lim, Propagation of pp-polarized electromagnetic waves obliquely incident on stratified random media: Random phase approximation, Waves Random Complex Media 17, 43 (2007).
  • [17] K. J. Lee and K. Kim, Universal shift of the Brewster angle and disorder-enhanced delocalization of pp waves in stratified random media, Opt. Express 19, 20817 (2011).
  • [18] T. M. Jordan, J. C. Partridge, and N. W. Roberts, Suppression of Brewster delocalization anomalies in an alternating isotropic-birefringent random layered medium, Phys. Rev. B 88, 041105(R) (2013).
  • [19] K. Kim, Anderson localization of electromagnetic waves in randomly-stratified magnetodielectric media with uniform impedance, Opt. Express 23, 14520 (2015).
  • [20] K. Kim and S. Kim, Anderson localization and Brewster anomaly of electromagnetic waves in randomly-stratified anisotropic media, Mater. Res. Express 6, 085803(2019).
  • [21] Q. Zhao, J. Gong, and C. A. Müller, Localization behavior of Dirac particles in disordered graphene superlattices, Phys. Rev. B 85, 104201 (2012).
  • [22] A. Fang, Z. Q. Zhang, S. G. Louie, and C. T. Chan, Anomalous Anderson localization behaviors in disordered pseudospin systems, Proc. Natl. Acad. Sci. U.S.A. 114, 4087 (2017).
  • [23] S. Kim and K. Kim, Anderson localization and delocalization of massless two-dimensional Dirac electrons in random one-dimensional scalar and vector potentials, Phys. Rev. B 99, 014205 (2019).
  • [24] S. Kim and K. Kim, Anderson localization of two-dimensional massless pseudospin-1 Dirac particles in a correlated random one-dimensional scalar potential, Phys. Rev. B 100, 104201 (2019).
  • [25] D. H. Dunlap, H.-L. Wu, and P. W. Phillips, Absence of localization in a random-dimer model, Phys. Rev. Lett. 65, 88 (1990).
  • [26] V. Bellani, E. Diez, R. Hey, L. Toni, L. Tarricone, G. B. Parravicini, F. Domínguez-Adame, and R. Gómez-Alcalá, Experimental evidence of delocalized states in random dimer superlattices, Phys. Rev. Lett. 82, 2159 (1999).
  • [27] H.-L. Wu, W. Goff, and P. W. Phillips, Insulator-metal transitions in random lattices containing symmetrical defects, Phys. Rev. B 45, 1623 (1992).
  • [28] F. M. Izrailev, T. Kottos, and G. P. Tsironis, Hamiltonian map approach to resonant states in paired correlated binary alloys, Phys. Rev. B 52, 3274 (1995).
  • [29] A. Kosior, J. Major, M. Płodzień, and J. Zakrzewski, Role of correlations and off-diagonal terms in binary disordered one-dimensional systems, Acta Phys. Pol. 128, 1002 (2015).
  • [30] J. Major, Extended states in disordered one-dimensional systems in the presence of the generalized NN-mer correlations, Phys. Rev. A 94, 053613 (2016).
  • [31] F. A. B. F. de Moura and M. L. Lyra, Delocalization in the 1D Anderson model with long-range correlated disorder, Phys. Rev. Lett. 81, 3735 (1998).
  • [32] F. M. Izrailev and A. A. Krokhin, Localization and the mobility edge in one-dimensional potentials with correlated disorder, Phys. Rev. Lett. 82, 4062 (1999).
  • [33] J. W. Kantelhardt, S. Russ, A. Bunde, S. Havlin, and I. Webman, Comment on “Delocalization in the 1D Anderson model with long-range correlated disorder”, Phys. Rev. Lett. 84, 198 (2000).
  • [34] U. Kuhl, F. M. Izrailev and A. A. Krokhin, Experimental observation of the mobility edge in a waveguide with correlated disorder, Appl. Phys. Lett. 77, 633 (2000).
  • [35] L. Sanchez-Palencia, D. Clement, P. Lugan, P. Bouyer, G.V. Shlyapnikov, and A. Aspect, Anderson localization of expanding Bose-Einstein condensates in random potentials, Phys. Rev. Lett. 98, 210401 (2007).
  • [36] M. Piraud, P. Lugan, P. Bouyer, A. Aspect, and L. Sanchez-Palencia, Localization of a matter wave packet in a disordered potential, Phys. Rev. A 83, 031603(R) (2011).
  • [37] J. Billy, V. Josse, Z. Zuo, A. Bernard, B. Hambrecht, P. Lugan, D. Clement, L. Sanchez-Palencia, P. Bouyer, and A. Aspect, Direct observation of Anderson localization of matter-waves in a controlled disorder, Nature 453, 891 (2008).
  • [38] J. Biddle and S. Das Sarma, Predicted mobility edges in one-dimensional incommensurate optical lattices: An exactly solvable model of Anderson localization, Phys. Rev. Lett. 104, 070601 (2010).
  • [39] A. Purkayastha, A. Dhar, and M. Kulkarni, Nonequilibrium phase diagram of a one-dimensional quasiperiodic system with a single-particle mobility edge, Phys. Rev. B 96, 180204(R) (2017).
  • [40] H. P. Lüschen, S. Scherg, T. Kohlert, M. Schreiber, P. Bordia, X. Li, S. Das Sarma, and I. Bloch, Single-particle mobility edge in a one-dimensional quasiperiodic optical lattice, Phys. Rev. Lett. 120, 160404 (2018).
  • [41] F. A. An, E. J. Meier, and B. Gadway, Engineering a flux-dependent mobility edge in disordered zigzag chains, Phys. Rev. X 8, 031045 (2018).
  • [42] H. Yao, H. Khouldi, L. Bresque, and L. Sanchez-Palencia, Critical behavior and fractality in shallow one-dimensional quasi-periodic potentials, Phys. Rev. Lett. 123, 070405 (2019).
  • [43] Y. Wang, X. Xia, L. Zhang, H. Ya, S. Chen, J. You, Q. Zhou and X.-J. Liu, One-dimensional quasiperiodic mosaic lattice with exact mobility edges, Phys. Rev. Lett. 125, 196604 (2020).
  • [44] Q.-B. Zeng, R. Lü, and L. You, Topological superconductors in one-dimensional mosaic lattices, EPL 135, 17003 (2021).
  • [45] Y. Liu, Y. Wang, X.-J. Liu, Q. Zhou, and S. Chen, Exact mobility edges, PTPT-symmetry breaking, and skin effect in one-dimensional non-Hermitian quasicrystals, Phys. Rev. B 103, 014203 (2021).
  • [46] Q.-B. Zeng and R. Lü, Topological phases and Anderson localization in off-diagonal mosaic lattices, Phys. Rev. B 104, 064203 (2021).
  • [47] L.-Y. Gong, H. Lu, and W.-W. Cheng, Exact mobility edges in 1D mosaic lattices inlaid with slowly varying potentials, Adv. Theory Simul. 4, 2100135 (2021).
  • [48] Z.-H. Wang, F. Xu, L. Li, D.-H. Xu, and B. Wang, Topological superconductors and exact mobility edges in non-Hermitian quasicrystals, Phys. Rev. B 105, 024514 (2022).
  • [49] D. Dwiputra and F.P. Zen, Single-particle mobility edge without disorder, Phys. Rev. B 105, L081110 (2022).
  • [50] B. White, P. Sheng, Z. Q. Zhang, and G. Papanicolaou, Wave localization characteristics in the time domain, Phys. Rev. Lett. 59, 1918 (1987).
  • [51] F. J. P. Schuurmans, M. Megens, D. Vanmaekelbergh, and A. Lagendijk, Light scattering near the localization transition in macroporous GaP networks, Phys. Rev. Lett. 83, 2183 (1999).
  • [52] M. Titov and C. W. J. Beenakker, Signature of wave localization in the time dependence of a reflected pulse, Phys. Rev. Lett. 85, 3388 (2000).
  • [53] P. M. Johnson, A. Imhof, B. P. J. Bret, J. G. Rivas, and A. Lagendijk, Time-resolved pulse propagation in a strongly scattering material, Phys. Rev. E 68, 016604 (2003).
  • [54] S. E. Skipetrov and B. A. van Tiggelen, Dynamics of weakly localized waves, Phys. Rev. Lett. 92, 113901 (2004).
  • [55] S. E. Skipetrov and B. A. van Tiggelen, Dynamics of Anderson localization in open 3D media, Phys. Rev. Lett. 96, 043902 (2006).
  • [56] K. M. Douglass, S. John, T. Suezaki, G. A. Ozin, and A. Dogariu, Anomalous flow of light near a photonic crystal pseudo-gap, Opt. Express 19, 25320 (2011).
  • [57] A. Aubry, L. A. Cobus, S. E. Skipetrov, B. A. van Tiggelen, A. Derode, and J. H. Page, Recurrent scattering and memory effect at the Anderson localization transition, Phys. Rev. Lett. 112, 043903 (2014).
  • [58] S. E. Skipetrov and A. Sinha, Time-dependent reflection at the localization transition, Phys. Rev. B 97, 104202 (2018).
  • [59] S. Aubry and G. André, Analyticity breaking and Anderson localization in incommensurate lattices, Ann. Isr. Phys. Soc. 3, 133 (1980).
  • [60] T. Durduran, R. Choe, W. B. Baker, and A. G. Yodh, Diffuse optics for tissue monitoring and tomography, Rep. Prog. Phys. 73, 076701 (2010).
  • [61] D. A. Boas and A. K. Dunn, Laser speckle contrast imaging in biomedical optics, J. Biomed. Opt. 15, 011109 (2010).
  • [62] S. Shahjahan, A. Aubry, F. Rupin, B. Chassignole, and A. Derode, A random matrix approach to detect defects in a strongly scattering polycrystal: How the memory effect can help overcome multiple scattering, Appl. Phys. Lett. 104, 234105 (2014).
  • [63] Diffuse Waves in Complex Media, edited by J.-P. Fouque (Kluwer, Dordrecht, 1999).
  • [64] D. J. Thouless, Electrons in disordered systems and the theory of localization, Phys. Rep. 13, 93 (1974).
  • [65] C. Castellani and L. Peliti, Multifractal wavefunction at the localisation threshold, J. Phys. A: Math. Gen. 19, L429 (1986).