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

Anomalous multifractality in quantum chains with strongly correlated disorder

Alexander Duthie [email protected] Physical and Theoretical Chemistry, Oxford University, South Parks Road, Oxford OX1 3QZ, United Kingdom    Sthitadhi Roy [email protected] Physical and Theoretical Chemistry, Oxford University, South Parks Road, Oxford OX1 3QZ, United Kingdom Rudolf Peierls Centre for Theoretical Physics, Clarendon Laboratory, Oxford University, Parks Road, Oxford OX1 3PU, United Kingdom    David E. Logan [email protected] Physical and Theoretical Chemistry, Oxford University, South Parks Road, Oxford OX1 3QZ, United Kingdom Department of Physics, Indian Institute of Science, Bangalore 560012, India
Abstract

We demonstrate numerically that a robust and unusual multifractal regime can emerge in a one-dimensional quantum chain with maximally correlated disorder, above a threshold disorder strength. This regime is preceded by a mixed and an extended regime at weaker disorder strengths, with the former hosting both extended and multifractal eigenstates. The multifractal states we find are markedly different from conventional multifractal states in their structure, as they reside approximately uniformly over a continuous segment of the chain, and the lengths of these segments scale non-trivially with system size. This anomalous nature also leaves imprints on dynamics. An initially localised wavepacket shows ballistic transport, in contrast to the slow, generally subdiffusive, transport commonly associated with multifractality. However, the timescale over which this ballistic transport persists again scales non-trivially with the system size.

Multifractal wavefunctions in quantum systems, which are neither extended nor localised, are characterised by anomalous statistics of their amplitudes [1]. While the effective volume occupied by such states grows unboundedly with system size, it is a vanishing fraction of the system volume; as such, they are often dubbed non-ergodic extended states. This is reflected in the scaling of the moments of the wavefunction amplitudes with system size. For a wavefunction ψ(x)\psi(x) defined on a discrete graph with LL sites,

x=1L|ψ(x)|2q{L(q1):extendedLDq(q1):multifractalL0:localised\displaystyle\sum_{x=1}^{L}|\psi(x)|^{2q}\sim\begin{cases}L^{-(q-1)}&:\mathrm{extended}\\ L^{-D_{q}(q-1)}&:\mathrm{multifractal}\\ L^{0}&:\mathrm{localised}\end{cases}\,

where 0<Dq<10<D_{q}<1 is the so-called multifractal dimension [1].

In the context of short-ranged disordered systems in finite-dimensions, multifractality is often a feature of critical points, such as Anderson transitions [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12] and quantum Hall plateau transitions [13, 14, 15], which are clearly fine-tuned points in parameter space. Multifractality is also realised, often robustly, in several long-ranged disordered hopping models, and fully connected random-matrix ensembles [16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29]. The presence of long-ranged physics, either emergently via diverging correlation lengths in the former, or explicitly via the structure of the models in the latter, unifies the two contexts. An interesting question thus arises: how can a robust multifractal phase be realised in a quantum system with inherently short-ranged interactions/hoppings? One possible avenue in a manifestly out-of-equilibrium setting is the time-periodic modulation of a quasiperiodic system with a mobility edge or a localisation transition [30, 31].

In this work, we demonstrate an alternative pathway to robust multifractality in an inherently short-ranged system, importantly in a time-independent Hamiltonian setting. The central ingredient is strong (in fact maximal, as clarified below) correlations in the disordered on-site potential of a one-dimensional chain. Interestingly, the origin and resultant structure of the multifractal states in such a system is markedly different from that of conventional multifractal states, which are associated with both rare large peaks and long polynomial tails of wavefunction amplitudes. In contrast, the multifractal states we find in this work reside over continuous segments of length \ell in the chain and crucially, within the segments, the wavefunction intensities are approximately uniform; |ψ(x)|21/|\psi(x)|^{2}\approx 1/\ell, and 0 elsewhere. Due to this structure, we refer to the states as tabletop states. The multifractality of the wavefunctions is then encoded in the scaling of these tabletop lengths with system size. The anomalous nature of the multifractal states also has implications for dynamics. Multifractality of eigenstates is often accompanied by slow dynamics [32, 33, 30, 34],111It should be noted that multifractality is not always a prerequisite for slow dynamics [49, 34].. However, in this case, we find ballistic spreading of a initially localised wavepacket, but over timescales that scale non-trivially with the system size.

Refer to caption
Figure 1: Schematic of the three regimes of the model Eq. 1, as a function of disorder strength. For small disorder, all states are extended, and for large enough disorder all states are multifractal. In between these two phases is a mixed regime, where the disorder-averaged spectrum hosts a mixture of extended and multifractal states.
Refer to caption
Figure 2: For weak and strong disorder strengths WW (top and bottom respectively), representative eigenstates are shown in blue and yellow. In the former, tabletops span the entire chain, whereas in the latter they span a sub-extensive segment. Black traces denote a realisation of the maximally correlated disorder potential (labels on right axis) wherein short distance fluctuations are visibly suppressed.

As a concrete model, we consider a disordered tight-binding Hamiltonian on a chain of length LL,

H=Wx=1Lϵx|xx|+x=1L1[|xx+1|+|x+1x|],H=W\sum_{x=1}^{L}\epsilon_{x}|x\rangle\langle x|+\sum_{x=1}^{L-1}\left[|x\rangle\langle x+1|+\ket{x+1}\bra{x}\right]\,, (1)

where WW denotes the disorder strength and the on-site potentials are drawn from a multivariate Gaussian distribution, ϵx𝒩(𝟎,𝐂)\vec{\epsilon_{x}}\sim\mathcal{N}(\mathbf{0},\mathbf{C}), with zero mean. The correlations in the potential are encoded in the covariance matrix 𝐂\mathbf{C}. We take the correlations in the potential to decay with distance as

C(r)[𝐂]x,x+r=ϵxϵx+r=f(r/L);f(0)=1.\displaystyle C(r)\equiv[\mathbf{C}]_{x,x+r}=\braket{\epsilon_{x}\epsilon_{x+r}}=f(r/L);\quad f(0)=1. (2)

The key point here is that the correlation is a function of r/Lr/L, which implies that in the thermodynamic limit C(r)1C(r)\to 1 for all sub-extensive rr, as limL(r/L)0\lim_{L\to\infty}(r/L)\to 0. In other words, the potentials of two sites a sub-extensive distance rr apart are completely slaved to each other in the thermodynamic limit. We refer to this as maximal correlations in the disorder [36, 37]. For specificity we choose f(r/L)=exp[r/(λL)]f(r/L)=\exp[-r/(\lambda L)], but emphasise that the specific functional form of ff is immaterial. In the following we set λ=1\lambda=1. Note that the limit of λ0\lambda\to 0 is singular, as there the model becomes the conventional 11D Anderson model with all eigenstates exponentially localised [38, 39].

The form of the correlations, (2), endows the disordered potential with an extensive lengthscale (λL\lambda L in this case), such that the potential fluctuations on sub-extensive scales are heavily suppressed and only those at extensive scales survive for large LL. This is already suggestive that the eigenstates can be extended over lengthscales which scale non-trivially with LL, resulting in multifractality. In fact, as demonstrated below, we find three distinct regimes as a function of WW. For sufficiently strong disorder a robust multifractal phase is found, where the average or typical tabletop lengths scale as LαL^{\alpha} with α<1\alpha<1; whereas for weak disorder we find α=1\alpha=1 for all eigenstates, indicating an extended phase. For a range of WW between these two regimes a mixed phase is found, where for a given energy some realisations host extended states and some multifractal. Establishing these robust multifractal and extended regimes with the intervening mixed regime in a model with maximally correlated disorder is the central result of this work, and is summarised in Fig. 1.

Before delving into a detailed analysis of the statistics of eigenstate tabletop lengths and the consequent multifractality, in Fig. 2 we show explicitly the tabletop nature of eigenstates. Operationally, we extract the tabletop edges for an eigenstate by scanning the chain for sites where the |ψx|2|\psi_{x}|^{2} jumps from zero (within numerical precision) and the tabletop length \ell is then simply the distance between the two such sites. It is evident from Fig. 2 that the eigenstates are approximately uniform over the tabletop segments. Hence it is natural to study the distribution of the tabletop lengths \ell, or equivalently of ~=/L\tilde{\ell}=\ell/L; denoting these distributions by PP_{\ell} and P~P_{\tilde{\ell}} respectively. Since we are interested in particular how \ell scales with system size LL, we define the exponents αm\alpha_{m} and αt\alpha_{t} from the mean and typical tabletop lengths as

mLαm;texp[log]Lαt,\displaystyle\ell_{m}\equiv\braket{\ell}\sim L^{\alpha_{m}};\quad\ell_{t}\equiv\exp[\braket{\log\ell}]\sim L^{\alpha_{t}}\,, (3)

where =𝑑P()\braket{\ell}=\int d\ell~{}\ell P_{\ell}(\ell) and similarly for log\braket{\log\ell}. In addition, we also define the exponent α=log/logL\alpha=\log\ell/\log L, and study its distribution which we denote by PαP_{\alpha}.

Refer to caption
Figure 3: (a)-(c): Main panels show distributions P~(~)P_{\tilde{\ell}}(\tilde{\ell}) of ~=/L\tilde{\ell}=\ell/L for representative disorder strengths in the extended (W=0.5W=0.5), mixed (W=2W=2), and multifractal (W=8W=8) regimes, for different system sizes LL (indicated in panel (f)). Insets show the corresponding P()P_{\ell}(\ell) distributions. (d)-(f): The distribution Pα(α)P_{\alpha}(\alpha) for the corresponding WW values. Statistics are obtained over 2000020000 disorder realisations.

We turn now to numerical results, which unless stated otherwise refer to band centre states (results for other energies remain qualitatively the same). For weak disorder, we find that the tabletops span not only an extensive segment of the chain but the entirety of it, such that P()=δ(L)P_{\ell}(\ell)=\delta(\ell-L) or equivalently Pα(α)=δ(α1)P_{\alpha}(\alpha)=\delta(\alpha-1) [Figs. 3(a) and 3(d)]. This is the extended regime indicated in Fig. 1. On increasing WW, the distribution P()P_{\ell}(\ell) develops finite weight at sub-extensive \ell while retaining the rest of the weight at extensive \ell [Figs. 3(b) and 3(e)]. This is the mixed regime referred to in Fig. 1. Note that it is important to distinguish the multifractal states with sub-extensive \ell from those that occupy an extensive =aL\ell=aL with a1a\leq 1, and hence confirm the presence of the former. That this is indeed the case is evidenced in Fig. 3(b) where the weight of the distribution P~(~)P_{\tilde{\ell}}(\tilde{\ell}) at ~=0\tilde{\ell}=0 grows with increasing LL. Further clear evidence for the mixed regime is also seen in Fig. 3(e) by the fact that Pα(α)P_{\alpha}(\alpha) retains weight at both α=1\alpha=1 and α<1\alpha<1 as LL\to\infty. Finally, at strong disorder, the system enters a regime where all states in the spectrum are multifractal. As shown in Fig. 3(c), for finite ~\tilde{\ell} the weight in P~(~)P_{\tilde{\ell}}(\tilde{\ell}) ultimately decays with increasing LL, suggesting that there exist no states with extensive tabletop lengths. That all states are indeed multifractal is further confirmed by the distribution Pα(α)P_{\alpha}(\alpha) (which is well converged with LL) having support strictly on 0<α<10<\alpha<1 as shown in Fig. 3(f). The vanishing of the weight of Pα(α)P_{\alpha}(\alpha) at α=1\alpha=1 and α=0\alpha=0 implies respectively the absence of extended and localised states. We add that this three-phase picture is also consistent with results for the scaling of transmittances with LL, obtained via a transfer matrix calculation [40].

Refer to caption
Figure 4: Left: Exponents αm\alpha_{m} (circles) and αt\alpha_{t} (triangles), defined in Eq. 3, as a function of disorder strength WW. Right: Fits of m\ell_{m} and t\ell_{t} versus LL on logarithmic axes, used to extract the exponents. Data shown for W=0.5W=0.5, 22, 33, and 44.

Having established the presence of a multifractal regime, along with an extended and a mixed regime, based on the distributions P~P_{\tilde{\ell}} and PαP_{\alpha}, we next present results for the scaling of mean and typical tabletop lengths with LL. In particular, Fig. 4 shows results for αm\alpha_{m} and αt\alpha_{t} (defined in Eq. 3) as a function of WW. For weak disorder we find both αm=1\alpha_{m}=1 and αt=1\alpha_{t}=1, consistent with the presence solely of extended states. On increasing WW and entering the mixed regime, αt\alpha_{t} and αm\alpha_{m} decrease from 1, indicating the emergence of multifractal states in the spectrum. Note that on entering the mixed regime, αm\alpha_{m} deviates from 1 less markedly than αt\alpha_{t}. This is natural, as in the presence of both extended and multifractal states the mean m\ell_{m} is dominated by the extended states with L\ell\sim L, whence αm\alpha_{m} is closer to 1 than αt\alpha_{t}. Finally, on increasing WW further into the regime where the spectrum has solely multifractal states, αm\alpha_{m} and αt<αm\alpha_{t}<\alpha_{m} continue to decrease monotonically.

We close our analysis of the multifractal statistics of tabletop eigenstates with results of a standard probe of multifractality, the generalised inverse participation ratios (IPR), defined as q(ψ)=x=1L|ψ(x)|2q\mathcal{I}_{q}(\psi)=\sum_{x=1}^{L}|\psi(x)|^{2q}. We will be interested in the scaling with LL of both the mean and typical IPR,

q,m=qLτq,m,q,t=exp[logq]Lτq,t.\displaystyle\mathcal{I}_{q,m}=\braket{\mathcal{I}_{q}}\sim L^{-\tau_{q,{m}}},~{}~{}\mathcal{I}_{q,t}=\exp[\braket{\log\mathcal{I}_{q}}]\sim L^{-\tau_{q,{t}}}\,. (4)

Extended states have τq,m/t=q1\tau_{q,{m/t}}=q-1, while for exponentially localised states τq,m/t=0\tau_{q,{m/t}}=0 for q>0q>0. An intermediate behaviour for τq\tau_{q} indicates multifractal states [1]. Fig. 5(a) shows results for τq,m\tau_{q,{m}} and τq,t\tau_{q,{t}}, for representative WW values in each of the three regimes. For the weakest disorder, which lies in the extended regime, we indeed find τq,m/t=q1\tau_{q,{m/t}}=q-1. The presence of multifractal states upon increasing WW is borne out by 0<τq,m/t<q10<\tau_{q,{m/t}}<q-1 for q>1q>1. In Fig. 5(b), we focus on τ2,m/t\tau_{2,m/t} as a function of WW. Note that on increasing WW and entering the mixed regime, τ2,m\tau_{2,{m}} deviates from the ergodic value of 1 more markedly than τ2,t\tau_{2,t}, reflecting the fact that the mean IPR is dominated by the small fraction of multifractal states which have qualitatively larger IPRs than the extended ones.

Refer to caption
Figure 5: (a) Exponents τq,m\tau_{q,{m}} (circles) and τq,t\tau_{q,t} (triangles), defined in Eq. 4, for representative WW in the three regimes. Black and red dashed lines respectively denote extended and localised behaviour. (b) Exponents τ2,m\tau_{2,{m}} and τ2,t\tau_{2,{t}} as a function of WW. Inset shows representative plots of 2\braket{\mathcal{I}_{2}} versus LL, used to extract τ2,m\tau_{2,m}, for W=0,1,2,3,4,5W=0,1,2,3,4,5.

An explanatory comment is in order regarding the lack of energy resolution between extended and multifractal states in the mixed regime. The density of states (DoS) for the model (1) with the correlated disorder (2) can be shown to fluctuate strongly across disorder realisations [40]. For each realisation we choose to refer energies relative to the centre of the spectrum (tantamount to HHTr[H]H\to H-\mathrm{Tr}[H]). However, the fluctuations in all higher moments of the DoS also remain finite in the thermodynamic limit. As a result, whether an eigenstate at a given energy (relative to the centre of the spectrum) is extended or multifractal depends on the specific disorder realisation (though for any given realisation, multifractal and extended states do not of course coexist at the same energy). Averaging over disorder realisations therefore smears out any energy resolution in the mixed regime, thereby precluding the traditional notion of a mobility edge in the averaged DoS.

So far, we have focussed on ‘static’ properties of the model. Since multifractality often goes hand in hand with slow dynamics [32, 33, 30, 34], it is worth asking what imprint the anomalous multifractal states leave on the dynamics in the present case. In order to answer this question, we consider the spreading of an initially localised wavepacket, in particular its second moment defined as

X2(t)=x=1Lx2|ψ(x,t)|2[x=1Lx|ψ(x,t)|2]2,\displaystyle X^{2}(t)=\Big{\langle}\sum_{x=1}^{L}x^{2}|\psi(x,t)|^{2}-\big{[}\sum_{x=1}^{L}x|\psi(x,t)|^{2}\big{]}^{2}\Big{\rangle}\,, (5)

with ψ(x,t=0)=δx,L/2\psi(x,t=0)=\delta_{x,L/2}. The simplest expected behaviour for X2(t)X^{2}(t) is

X2(t){t2z;ttL2β;tt,\displaystyle X^{2}(t)\sim\begin{cases}t^{2z};\quad&t\ll t_{\ast}\\ L^{2\beta};\quad&t\gg t_{\ast}\end{cases}\,, (6)

with crossover scale tLβ/zt_{\ast}\sim L^{\beta/z}. For example, z,β=1z,\beta=1 would correspond to ballistic transport until the wavepacket spans the entire system; while for localised states z,β=0z,\beta=0, indicating the absence of transport. Conventionally, multifractal states lead to 0<z,β<10<z,\beta<1 [32, 33, 30, 34].

Refer to caption
Figure 6: Dynamics of an initially localised wavepacket in the multifractal regime, for W=6W=6. Left: Ballistic transport of the wavepacket as X2(t)X^{2}(t) (defined in Eq. 5) initially grows as t2t^{2}. Right: The saturation value, X2(t)L2βX^{2}(t\to\infty)\sim L^{2\beta} plotted as a function of LL, with β0.85\beta\simeq 0.85. The inset shows the scale-collapsed form X2(t)=L2βg(t/Lβ)X^{2}(t)=L^{2\beta}g(t/L^{\beta}), with g(y)y2g(y)\sim y^{2} for y1y\ll 1.

Fig. 6 shows results for the maximally correlated model in the pure multifractal phase. Remarkably, the transport is ballistic with z=1z=1 but the saturation of X2(t)X^{2}(t) scales as L2βL^{2\beta} with exponent β<1\beta<1. In fact, the inset in Fig. 6 shows the scale-collapsed behaviour X2(t)=L2βg(t/Lβ)X^{2}(t)=L^{2\beta}g(t/L^{\beta}) with g(y)y2g(y)\sim y^{2} for y1y\ll 1, which confirms the ballistic spreading and sub-extensive saturation of the wavepacket. We now show how a toy model of the anomalous multifractal states qualitatively rationalises the above dynamics.

For simplicity, let us assume that in a given realisation the initially localised wavepacket lies at the centre of a tabletop segment of length \ell. Approximating the eigenstates therein as fully extended over the segment, the wavepacket spreads as a truncated Gaussian

|ψ(x,t)|2=Θ(2|x|)ex2/t2/(tπErf(/2t)),\displaystyle|\psi(x,t)|^{2}=\Theta(\ell-2|x|)e^{-x^{2}/t^{2}}/\big{(}t\sqrt{\pi}~{}\mathrm{Erf}(\ell/2t)\big{)}\,, (7)

where we have absorbed the velocity into the units of time. X2(t)X^{2}(t) for the wavepacket (7) behaves in time as

X2(t)=12t2[1e2/4t2(/t)/(πErf(/2t))].\displaystyle X_{\ell}^{2}(t)=\tfrac{1}{2}t^{2}[1-e^{-\ell^{2}/4t^{2}}(\ell/t)/\big{(}\sqrt{\pi}~{}\mathrm{Erf}(\ell/2t)\big{)}]\,. (8)

Averaging over disorder is equivalent to integrating over the probability distribution of the initial site residing in a tabletop of length \ell, denoted by Φ(,L)\Phi_{\ell}(\ell,L). Thus

X2(t)=12t2[12π𝑑Φ(,L)e2/4t22tErf(/2t)].\displaystyle X^{2}(t)=\tfrac{1}{2}t^{2}\left[1-\frac{2}{\sqrt{\pi}}\int d\ell~{}\Phi_{\ell}(\ell,L)\frac{\ell e^{-\ell^{2}/4t^{2}}}{2t~{}\mathrm{Erf}(\ell/2t)}\right]\,. (9)

For timescales much smaller than a typical tabletop length, the second term in Eq. 9 is negligibly small and the ballistic behaviour X2(t)t2X^{2}(t)\sim t^{2} is recovered. In the opposite limit where tt is much larger than typical tabletop lengths (such that /2t1\ell/2t\ll 1 in (9)), the wavepacket spreading saturates and X2𝑑Φ(,L)2X^{2}\simeq\int d\ell~{}\Phi_{\ell}(\ell,L)\ell^{2}, which one expects to scale non-trivially with LL due to that of Φ(,L)\Phi_{\ell}(\ell,L).

Finally, we provide a simple heuristic argument for the physics underlying the emergence of the multifractal regime. Note that with the form of the correlations, C(r)=er/LC(r)=e^{-r/L}, the sequence of site energies ,ϵi,ϵi+1,\dots,\epsilon_{i},\epsilon_{i+1},\dots is readily shown to be a martingale. Thus, given ϵi\epsilon_{i}, ϵi+1\epsilon_{i+1} is a Gaussian random number with mean e1/Lϵie^{-1/L}\epsilon_{i} and variance 2/L\simeq 2/L. This leads to the site energies of nearby sites being very close to each other with high probability; to which we attribute the approximately uniform |ψ(x)|2|\psi(x)|^{2} across the tabletop segment. However, since the conditional distribution P(ϵi+1|ϵi)P(\epsilon_{i+1}|\epsilon_{i}) is a Gaussian, it has unbounded support. As a result, with very low probability, there can be neighbouring sites where the site energies are wildly different. The break in the tabletops may be attributed to such rare regions. Moreover, as the sequence of site energies is a martingale, the probability of a tabletop terminating at a site is independent of the energies of the sites prior to it. It is then intuitively natural to regard the tabletop breaks as a Poisson process. This in turn implies P()eγP_{\ell}(\ell)\sim e^{-\gamma\ell} where γ\gamma is the density of tabletop breaks, which decays with increasing LL 222As each tabletop state has two breaks, and the density of tabletops of length Lα\ell\sim L^{\alpha} is Lα\sim L^{-\alpha}.. An exponential P()P_{\ell}(\ell) with a rate that decays with LL is indeed consistent with the numerical results shown in Fig. 3(c) in the strong disorder regime. Obviously, however, obtaining the precise multifractal exponents would require a theory which takes into account the fluctuations of |ψ(x)|2|\psi(x)|^{2} within the tabletop length.

In summary, we demonstrated numerically that a one-dimensional quantum chain (1), with maximally correlated disorder (2), hosts a robust multifractal regime. These multifractal eigenstates are strikingly unusual. They reside approximately uniformly over segments of the chain whose lengths scale non-trivially with LL, whence the multifractal statistics arise. Such an anomalous structure also leaves imprints on the dynamics – in the multifractal phase an initially localised wavepacket spreads ballistically, but over timescales that scale as LβL^{\beta} with β<1\beta<1, beyond which the expansion saturates.

We note that the behaviour found here differs radically 333Perhaps expectedly, given that the Hausdorff dimension of any tree with K2K\geq 2 is infinite, while that of a 1D chain is unity. from the same model considered on a tree with connectivity K2K\geq 2 [37], which hosts localised states but not a multifractal phase. The present work is also very different from earlier studies [43, 44] of 1D Anderson localisation with long-ranged, power-law (and hence scale-free) disorder correlations. These models too do not host a robust multifractal regime. This suggests that the system-size dependent scale introduced by the correlations (2), along with the martingale nature of the site-energies, is responsible for the robust multifractal regime in the 1D model with maximally correlated disorder.

Finally, it is interesting to note that the effective Fock-space disorder in a many-body localised system is likewise maximally correlated [36], and the many-body localised eigenstates indeed exhibit multifractality on the Fock space [45, 46, 47, 48]. Whether a concrete connection between maximal disorder correlations and multifractality exists, and if so under what conditions, remains an open question.

Acknowledgements.
We thank Ivan Khaymovich for helpful comments on the manuscript. This work was supported in part by EPSRC, under Grant No. EP/L015722/1 for the TMCS Centre for Doctoral Training, and Grant No. EP/S020527/1.

References

  • Evers and Mirlin [2008] F. Evers and A. D. Mirlin, Anderson transitions, Rev. Mod. Phys. 80, 1355 (2008).
  • Wegner [1980] F. Wegner, Inverse participation ratio in 2+ϵ\epsilon dimensions, Z. Phys. B 36, 209 (1980).
  • Castellani and Peliti [1986] C. Castellani and L. Peliti, Journal of Physics A: Mathematical and General 19, L429 (1986).
  • Chalker [1988] J. T. Chalker, Scaling and correlations at a mobility edge in two dimensions, J. Phys. C: Solid State Phys. 21, L119 (1988).
  • Chalker [1990] J. T. Chalker, Scaling and eigenfunction correlations near a mobility edge, Physica A 167, 253 (1990).
  • Bauer et al. [1990] J. Bauer, T.-M. Chang, and J. L. Skinner, Correlation length and inverse-participation-ratio exponents and multifractal structure for Anderson localization, Phys. Rev. B 42, 8121 (1990).
  • Schreiber and Grussbach [1991] M. Schreiber and H. Grussbach, Multifractal wave functions at the Anderson transition, Phys. Rev. Lett. 67, 607 (1991).
  • Janssen [1994] M. Janssen, Multifractal analysis of broadly-distributed observables at criticality, Int. J. Mod. Phys. B 08, 943 (1994).
  • Chalker et al. [1996] J. T. Chalker, V. E. Kravtsov, and I. V. Lerner, Spectral rigidity and eigenfunction correlations at the Anderson transition, JETP Letters 64, 386 (1996).
  • Mirlin and Evers [2000] A. D. Mirlin and F. Evers, Multifractality and critical fluctuations at the Anderson transition, Phys. Rev. B 62, 7920 (2000).
  • Evers and Mirlin [2000] F. Evers and A. D. Mirlin, Fluctuations of the inverse participation ratio at the Anderson transition, Phys. Rev. Lett. 84, 3690 (2000).
  • Cuevas et al. [2001a] E. Cuevas, M. Ortuño, V. Gasparian, and A. Pérez-Garrido, Fluctuations of the correlation dimension at metal-insulator transitions, Phys. Rev. Lett. 88, 016401 (2001a).
  • Chalker and Coddington [1988] J. T. Chalker and P. D. Coddington, Percolation, quantum tunnelling and the integer Hall effect, J. Phys. C: Solid State Phys. 21, 2665 (1988).
  • Huckestein [1995] B. Huckestein, Scaling theory of the integer quantum Hall effect, Rev. Mod. Phys. 67, 357 (1995).
  • Evers et al. [2001] F. Evers, A. Mildenberger, and A. D. Mirlin, Multifractality of wave functions at the quantum Hall transition revisited, Phys. Rev. B 64, 241303 (2001).
  • Levitov [1990] L. S. Levitov, Delocalization of vibrational modes caused by electric dipole interaction, Phys. Rev. Lett. 64, 547 (1990).
  • Mirlin et al. [1996] A. D. Mirlin, Y. V. Fyodorov, F.-M. Dittes, J. Quezada, and T. H. Seligman, Transition from localized to extended eigenstates in the ensemble of power-law random banded matrices, Phys. Rev. E 54, 3221 (1996).
  • Parshin and Schober [1998] D. A. Parshin and H. R. Schober, Multifractal structure of eigenstates in the Anderson model with long-range off-diagonal disorder, Phys. Rev. B 57, 10232 (1998).
  • Levitov [1999] L. S. Levitov, Critical Hamiltonians with long range hopping, Annalen der Physik 8, 697 (1999).
  • Varga and Braun [2000] I. Varga and D. Braun, Critical statistics in a power-law random-banded matrix ensemble, Phys. Rev. B 61, R11859 (2000).
  • Cuevas et al. [2001b] E. Cuevas, V. Gasparian, and M. Ortuño, Anomalously large critical regions in power-law random matrix ensembles, Phys. Rev. Lett. 87, 056601 (2001b).
  • Cuevas [2003] E. Cuevas, Multifractality of Hamiltonians with power-law transfer terms, Phys. Rev. B 68, 184206 (2003).
  • Monthus and Garel [2010] C. Monthus and T. Garel, The Anderson localization transition with long-ranged hoppings: analysis of the strong multifractality regime in terms of weighted Lévy sums, J. Stat. Mech. , P09015 (2010).
  • Kravtsov et al. [2015] V. E. Kravtsov, I. M. Khaymovich, E. Cuevas, and M. Amini, A random matrix model with localization and ergodic transitions, New Journal of Physics 17, 122002 (2015).
  • Deng et al. [2019] X. Deng, S. Ray, S. Sinha, G. V. Shlyapnikov, and L. Santos, One-dimensional quasicrystals with power-law hopping, Phys. Rev. Lett. 123, 025301 (2019).
  • Monthus [2019] C. Monthus, Multifractality in the generalized Aubry–André quasiperiodic localization model with power-law hoppings or power-law Fourier coefficients, Fractals 27, 1950007 (2019).
  • Khaymovich et al. [2020] I. M. Khaymovich, V. E. Kravtsov, B. L. Altshuler, and L. B. Ioffe, Fragile extended phases in the log-normal Rosenzweig-Porter model, Phys. Rev. Research 2, 043346 (2020).
  • Kravtsov et al. [2020] V. E. Kravtsov, I. M. Khaymovich, B. L. Altshuler, and L. B. Ioffe, Localization transition on the random regular graph as an unstable tricritical point in a log-normal rosenzweig-porter random matrix ensemble (2020), arXiv:2002.02979 [cond-mat.dis-nn] .
  • Biroli and Tarzia [2021] G. Biroli and M. Tarzia, Lévy-Rosenzweig-Porter random matrix ensemble, Phys. Rev. B 103, 104205 (2021).
  • Roy et al. [2018] S. Roy, I. M. Khaymovich, A. Das, and R. Moessner, Multifractality without fine-tuning in a Floquet quasiperiodic chain, SciPost Phys. 4, 25 (2018).
  • Sarkar et al. [2021] M. Sarkar, R. Ghosh, A. Sen, and K. Sengupta, Mobility edge and multifractality in a periodically driven Aubry-André model, Phys. Rev. B 103, 184309 (2021).
  • De Tomasi et al. [2016] G. De Tomasi, S. Roy, and S. Bera, Generalized Dyson model: Nature of the zero mode and its implication in dynamics, Phys. Rev. B 94, 144202 (2016).
  • Monthus [2017] C. Monthus, Multifractality of eigenstates in the delocalized non-ergodic phase of some random matrix models: Wigner–Weisskopf approach, J. Phys. A: Math. Theor. 50, 295101 (2017).
  • Khaymovich and Kravtsov [2021] I. M. Khaymovich and V. E. Kravtsov, Dynamical phases in a “multifractal” Rosenzweig-Porter model, SciPost Phys. 11, 45 (2021).
  • Note [1] It should be noted that multifractality is not always a prerequisite for slow dynamics [49, 34].
  • Roy and Logan [2020a] S. Roy and D. E. Logan, Fock-space correlations and the origins of many-body localization, Phys. Rev. B 101, 134202 (2020a).
  • Roy and Logan [2020b] S. Roy and D. E. Logan, Localization on certain graphs with strongly correlated disorder, Phys. Rev. Lett. 125, 250402 (2020b).
  • Anderson [1958] P. W. Anderson, Absence of diffusion in certain random lattices, Phys. Rev. 109, 1492 (1958).
  • Mott and Twose [1961] N. F. Mott and W. D. Twose, The theory of impurity conduction, Advances in Physics 10, 107 (1961).
  • [40] See supplementary material at [URL] for (I) results on transmittance and (II) derivation of the lack of self-averaging in the density of states.
  • Note [2] As each tabletop state has two breaks, and the density of tabletops of length Lα\ell\sim L^{\alpha} is Lα\sim L^{-\alpha}.
  • Note [3] Perhaps expectedly, given that the Hausdorff dimension of any tree with K2K\geq 2 is infinite, while that of a 1D chain is unity.
  • Izrailev and Krokhin [1999] 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).
  • Croy et al. [2011] A. Croy, P. Cain, and M. Schreiber, Anderson localization in 1d systems with correlated disorder, Eur. Phys. J. B 82, 107 (2011).
  • De Luca and Scardicchio [2013] A. De Luca and A. Scardicchio, Ergodicity breaking in a model showing many-body localization, Europhys. Lett. 101, 37003 (2013).
  • Macé et al. [2019] N. Macé, F. Alet, and N. Laflorencie, Multifractal scalings across the many-body localization transition, Phys. Rev. Lett. 123, 180601 (2019).
  • De Tomasi et al. [2021] G. De Tomasi, I. M. Khaymovich, F. Pollmann, and S. Warzel, Rare thermal bubbles at the many-body localization transition from the Fock space point of view, Phys. Rev. B 104, 024202 (2021).
  • Roy and Logan [2021] S. Roy and D. E. Logan, Fock-space anatomy of eigenstates across the many-body localization transition, Phys. Rev. B 104, 174201 (2021).
  • De Tomasi et al. [2020] G. De Tomasi, S. Bera, A. Scardicchio, and I. M. Khaymovich, Subdiffusion in the anderson model on the random regular graph, Phys. Rev. B 101, 100201 (2020).

See pages 1 of SM See pages 2 of SM