You are on page 1of 21

Fluctuation driven transitions in localized insulators:

Intermittent metallicity and path chaos precede delocalization


Valentina Ros∗
Université Paris-Saclay, CNRS, LPTMS, 91405, Orsay, France

Markus Müller†
Paul Scherrer Institut, CH-5232 Villigen PSI, Switzerland and
The Abdus Salam International Center for Theoretical Physics, Strada Costiera 11, 34151 Trieste, Italy
(Dated: March 12, 2021)
We study how interacting localized degrees of freedom are affected by slow thermal fluctuations
arXiv:2103.06806v1 [cond-mat.dis-nn] 11 Mar 2021

that change the effective local disorder. We compute the time-averaged (annealed) conductance in
the insulating regime and find three distinct insulating phases, separated by two sharp transitions.
The first occurs between the deep non-resonating insulator and a rarely-resonating insulator (or
intermittent metal). The average conductance is always dominated by rare temporal fluctuations.
However, in the intermittent metal, they are so strong that the system becomes metallic for an
exponentially small fraction of the time. A second transition occurs within that phase. At stronger
disorder, there is a single optimal path providing the dominant contribution to the conductance at
all times, but closer to delocalization, a transition to a phase with fluctuating paths occurs. This
last phase displays the quantum analogon of configurational chaos in glassy systems, in that thermal
fluctuations induce significant changes of the dominant decay channels. While in the insulator the
annealed conductance is strictly bigger than the conductance with typical, frozen disorder, we show
that the threshold to delocalization is insensitive to whether or not thermal fluctuations are admitted.
This rules out a potential bistability, at fixed disorder, of a localized phase with suppressed internal
fluctuations and a delocalized, internally fluctuating phase.

I. INTRODUCTION they are encountered with vanishing probability in the


thermodynamic limit. Likewise it is well-understood that
Strong Anderson localization and its interacting ana- temporally fluctuating disorder can induce delocaliza-
logue, Many-Body Localization (MBL), both arise due tion, even if at every instance of time the disorder real-
to the disorder-induced suppression of resonant cou- ization would be classified as localizing, if it were static.
plings between close-by states in configuration space. This phenomenon has been studied, e.g., in the context of
Strong disorder renders typical energy differences be- periodically modulated disordered Hamiltonians [12, 13]:
tween such configurations too large to be overcome by The slower the variations the higher is the danger to en-
the off-diagonal tunneling terms that connect them. The counter resonances over long enough time windows dur-
sparse remaining resonances turn out to be harmless (at ing which adiabatic state changes take place, which delo-
least in one dimension, in the case of many body sys- calize the system over the long run and restore ergodicity.
tems) and preserve localization, non-ergodicity and a re- This poses a fundamental problem when one tries to uti-
sistance that grows exponentially with system size. For lize disorder-induced many-body localization to protect
non-interacting quantum particles this has been under- topological order and anyon braiding at finite tempera-
stood already by Anderson [1], and it has been argued ture. [14]
at various levels of rigor for the many-body case over the
In systems with a conserved charge, a time depen-
last 15 years [2–4], taking up and pushing much further
dence of the disorder almost invariably increases the
an initial perturbative analysis by Fleishman and Ander-
time-averaged conductance associated to the charge. In-
son [5], cf. also [6–10] for recent reviews. MBL has been
deed, being an exponentially small quantity its time av-
argued to be the most robust route to achieve ergod-
erage is likely to be dominated by rare temporal fluctu-
icity breaking (without invoking spontaneous symmetry
ations during which the conductance increases exponen-
breaking), to suppress long range transport and to avoid
tially, even if only for an exponentially small fraction of
thermalization in isolated interacting quantum systems
the total time. Often it is the first exponential that wins,
[11].
as we will show with an explicit calculation in this work.
The above considerations hold, however, for typical re-
Likewise, the life time of excitations localized in the bulk
alizations of static disorder. It is well-known that rare
of a sample will likely be limited by such rare fluctuations,
correlated realizations of disorder can nevertheless ex-
too, rather than by the decay in the presence of a typical
hibit delocalization and finite transport, even though
disorder configuration. This then raises an interesting
question: Anderson’s criterion for the breakdown of lo-
calization requires that the decay rate to infinity turns
∗ valentina.ros@universite-paris-saclay.fr from being exponentially suppressed in the system size
† Markus.Mueller@psi.ch to becoming finite. Since it turns out that in the pres-
2

ence of fluctuations these decay rates depend on whether mal fluctuations shifted the localization phase boundary
one takes their annealed, i.e. time-averaged value, or towards stronger disorder. As we mentioned above, such
their quenched value in a static, typical configuration an effect would entail the possibility of an intermediate
one might surmise that the delocalization transition of range of disorder strength where two physically differ-
the quantum system actually depends on whether or not ent phases would be self-consistent and locally stable: (i)
the disorder is fluctuating. From this it is then a nat- a non-fluctuating, more strongly localized non-thermal
ural next step to ask what would happen if fluctuating phase with frozen effective potentials, and (ii) an ergodic
effective disorder arose not from an external source, but delocalized phase with thermally fluctuating effective po-
were generated internally, by the dynamics of the sys- tentials.
tem itself. If the delocalization transition indeed hinged Here we revisit this intriguing scenario. Our analysis
on the presence or absence of fluctuations, the instabil- shows, however, that such a bistability of localized and
ity would seem to depend on whether one approaches delocalized phases is impossible: while temporal fluctua-
the transition from the delocalized or the localized side. tions of local potentials definitely affect the spatial struc-
This would potentially allow for a region of bistability ture of localized excitations and their effective localiza-
where either phase would be self-consistent - a scenario tion length, the localization phase boundaries (or the as-
that would contrast with the currently favored scenario sociated crossovers) will be shown to be independent of
that the many-body delocalization transition occurs at a whether or not thermal fluctuations are included in the
unique, well-defined critical point. [15–17] analysis. This is so, even though the analytical contin-
However, our analysis will show that actually no such uation of the annealed (“fluctuating”) conductance from
coexistence regime exists. Nevertheless, we will find that the strong disorder regime actually would predict a de-
thermal fluctuations in an interacting insulator introduce localization occurring at stronger disorder than that of a
a rich phenomenology, including three distinct insulating disorder-quenched system. However, we will show that
phases. a phase transition inside the insulating regime, from a
The question about the effect of fluctuations is partic- deep, “non-resonating insulator” to a “rarely-resonating
ularly relevant in systems where the interactions have the insulator” or an “intermittent metal”, invalidates the an-
predominant role of tuning the effective disorder. This alytical continuation and thus, the prediction of a shifted
differs from their role in the canonical models of weakly delocalization transition. Those phases exhibit a qualita-
interacting disordered quantum particles that were stud- tively different behavior of their exponentially small av-
ied in the wake of MBL [2, 3]. In those cases the prime erage conductance, which is potentially observable [23].
role of interactions is to tune the number of scattering Indeed, in the rarely-resonating regime the conductance
channels that allow for long range transport. Here in- is dominated by rare thermal fluctuations that already
stead we are interested in systems of particles or spins, induce metallic behavior, whereas at a typical instant
where the interaction terms act mostly “classically”, in the sample still looks well insulating.
the sense that they commute with each other and with
In addition to this transition, we find that in di-
the disordered potential part of the Hamiltonian; a sim-
mensions d > 1 thermal fluctuations induce a fur-
ple example are density-density interactions of strongly
ther transition within the intermittent metallic phase.
localized electrons or Ising interactions of spins in random
That second transition is a quantum analogon of the
longitudinal fields. In such models, transport is mainly
freezing-unfreezing transition occurring in certain models
due to kinetic hopping or spin flip terms which compete
of glasses [24, 25]. While in static (or relatively weakly
with the interaction-induced potential landscape. The
fluctuating) disorder the conductance is dominated by
interaction terms thus strongly affect the effective local
the propagation along one (or very few) rather well-
energy spectrum that excitations encounter as they try
defined paths through the sample, this changes in the
to propagate. On one hand, thermal fluctuations of the
vicinity of delocalization, where the dominant path starts
degrees of freedom with which an excitation interacts
to fluctuate with thermal fluctuations of the effective lo-
(other spins or electrons) may generate effective local
cal disorder. This is an interesting analog of a well-known
fields that are more resonant with the considered exci-
phenomenon in glassy systems where the ground state
tation, enhancing small denominators and thus favoring
configuration often times changes in a chaotic manner
delocalization. In certain cases, on the other hand, the
with changes of the disorder potential. [26, 27]
interactions may even enhance the localization tendency
with increasing temperature, because thermal configura-
tional disorder translates into an increased width of the
disorder distribution [18–21].
The localization properties of one of the simplest re- II. MODEL AND LOCATOR APPROXIMATION
alizations of such a system was studied in Ref. 22 and
analyzed on a Bethe lattice for simplicity. Despite the
fact that the distribution of effective local fields (sampled As motivated above we consider models where the in-
over all sites) remains temperature-independent in that teractions predominantly shape the effective energy land-
model, the results suggested that the presence of ther- scape. We focus on spin systems on a lattice, with Hamil-
3

tonian of the generic form: heff


i evolve with time due to the fluctuations of the neigh-
X boring spins. If the latter fluctuate thermally, the prob-
H = Hcl ({σiz }) − J⊥ σi+ σj− + σi− σj+ ,

(1) ability to see a field hi at a given site i is given by
hi,ji  
X Y e−βj σjz X
where Hcl is a function of the classical Ising spin vari- P (hi |~i )= δ hi −i − Jij σjz .
ables σjz only, and thus only contains mutually commut- z
2cosh(βj )
{σj }|j∈∂i j∈∂i j∈∂i
ing terms. Quantum fluctuations and dynamics arise (4)
only through the spin-flip term with amplitude J⊥ . The The on-site distribution depends on the random fields at
notation hi, ji indicates that the sites i, j are nearest- the site i and at the neighboring sites j ∈ ∂i, which we
neighbors in the lattice. collectively denote by ~i . The distribution of local fields
The effective local field seen by spin i is sampled aver all sites reads
∂Hcl
 
N N
heff
i =− , (2) 1 X 1 X X
∂σiz P (h)= δ(h−heff
i )= δ h−i − Jij σjz ,
N i=1 N i=1
j∈∂i
which for pairwise Ising interactions takes the form (5)
X where the {σjz } realize a typical classical configura-
heff
i = i + Jij σjz ≡ heff
i (i , ~
σ∂i ) , (3)
tion sampled from the Gibbs ensemble at temperature
j∈∂i
T ≡ β −1 . This distribution in general depends on the
where ~σ∂i denotes the configurations σjz of the spins in temperature. However,
P 2 here we restrict ourselves to tem-
the neighborhood ∂i of σiz , see Fig. 1. We assume peratures T  j Jij /W , in such a way that we can
the random fields i to be independent on every site, neglect the polarizing influence of spin i on its neighbors.
with an identical distribution f () with width W . Mod- We also assume statistical time-reversal symmetry, im-
els of the form (1) arise rather ubiquitously in the the- plying that the distribution of local fields f () is even.
ory of disordered quantum magnets [28, 29], quantum In this case the probability of finding neighboring spins
Coulomb glasses [30–33], disordered supersolids [34–36], up or down, averaged over all those spins, is always
cold atomic systems [37], or disordered superconductors z
e−βσ
Z
z 1
[38, 39]. p(σ ) = d f ()= , (6)
2 cosh(β) 2
independent of the temperature. It follows that the dis-
tribution of local fields P (h) in Eq. (5) is T -independent
as well. While it would not be difficult to take correlation
effects at lower temperatures into account, our simplifi-
cation eliminates a potential temperature dependence of
the decay rates that arises trivially from a T -dependence
of the local field distribution. It thereby allows us to
focus on the effect of thermal fluctuations only. A case
where correlation effects are strong and P (h) does depend
significantly on temperature at low T is instead analyzed
in Ref. 40.
Notice that the above definition of the distribution
P (h) makes sense even without a thermal average and
even if the system were completely frozen, realizing a sin-
gle classical configuration sampled from the Gibbs distri-
bution, but without invoking a dynamic averaging. This
is because the average over the entire sample ensures that
all possible local configurations are sampled and repre-
FIG. 1. Schematic representation of the set-ups discussed in
sented in the sum (5). This makes a thermal average
this work. The top figure shows sites s belonging to a path
of length R, each site being connected to other N “environ-
superfluous, since P (h) is self-averaging.
mental” spins (here N = 2) belonging to the neighborhood
∂s. The bottom figure shows in red one out of exponentially
many paths of length R on a tree with branching number Locator approximation for the decay rate
k = 2. Each site s of the tree is connected to N  1 environ-
mental spins in ∂s. To characterize localization for the model (1), we con-
sider the decay rate of a local (spin-flip) excitation cre-
The nearly classical variables σiz have a dynamics in- ated at a site 0 in the bulk of the lattice. In a localized
duced by the transverse term, and thus the local fields system, the decay rate vanishes in the thermodynamic
4

limit, even if one couples the system to a bath at the dynamic variables that are allowed to fluctuate at each
boundary. In the tree approximation in which loops are site with their thermal probabilities 2 . In the latter case,
neglected, approximate expressions for the decay rate Γ one can have rare fluctuations in the environmental spins
(via the boundary) can be derived, based on the lineariza- that give rise to local fields that are more frequently close
tion [39, 41–45] of recursion equations for the imaginary to ω than in a typical thermal configuration. In an opti-
parts of Green’s functions. For Hamiltonians such as in mized fluctuation of environmental spins, the abundance
Eq. (1), in appropriate units they take the form: of small energy denominators is higher, which opens a
more efficient decay channel for the excitation than a
2
X Y J⊥ typical configuration [22]. Since the probability for a
ΓR [~, ~σ∂ ; ω] = , (7)
[ω − heff
s (s , ~σ∂s )]2 small deviation of denominators from a typical thermal
P:0→BR s∈P
distribution only decreases with the square of the devia-
where the sum is over the exponentially many lattice tion, while its effect on the decay rate is linear, we gener-
paths P that connect the bulk site 0 to the boundary ally expect that fluctuations enhance the decay. In other
BR , assumed to be at lattice distance R from site 0. By words, the annealed decay rate is expected to be strictly
∂ = ∪∂s we denote the set of all spins that are neigh- bigger than its quenched counterpart, except possibly at
bors of sites s on any such path P. Different paths in the localization transition. This will be confirmed by our
the sum contribute with amplitudes that are a product explicit calculations below.
of locators, one for each site. The denominators essen- To investigate the effect of such fluctuations, we de-
tially correspond to the mismatch between the energy of scribe the liquid environment by computing the annealed
the propagating excitation ω and the effective field at average (which is essentially equivalent to the time-
the site. This expression is obtained within the so-called average) of the decay rates associated to each path in
forward scattering approximation, that corresponds to (7), obtaining:
taking the leading order in the quantum fluctuations J⊥ . X Z Y
(
Y J2
)
In particular, within this approximation, self-energy cor- ΓR [~;ω]= dhs P (hs |~s )min 1, ⊥
,
rections to the denominators are neglected [41, 44, 46]. [ω−hs ]2
P:0→BR s∈P s∈P
If the lattice is locally tree-like, the spins ~σ∂i with which (8)
the excitation interacts are different from site to site, and where the field distribution P (hs |~s ) is averaged over
therefore the locators at different sites are statistically in- the thermal distribution of configurations of neighboring
dependent. When having in mind more general lattices spins. Note that we have to be cautious when averag-
which are not tree-like, we will make this approximation, ing the decay rate: While within the insulating phase it
too. 1 typically decreases exponentially with the path length,
The excitation is considered to be localized whenever it might happen that on paths with rare configurations,
the typical value of ΓR decays exponentially in R with a where small denominators are more frequent, the prod-
positive spatial decay constant γ > 0, hlog ΓR i = −Rγ + uct of locators becomes exponentially large. This is ob-
o(R) (see below for a precise definition of the average). viously an unphysical artifact which arises from our for-
The vanishing of this constant, γ = 0, can thus be taken ward approximation and its neglect of self-energy correc-
as signature for the onset of a delocalized phase. tions. Those would introduce correlations between the
local fields and in particular suppress the effect of small
denominators, ensuring that the decay rate never grows
III. FLUCTUATION-ENHANCED DECAY RATE exponentially with distance. Indeed, from physical con-
siderations, the decay rate can at best become of order
The amplitude of each path in the sum (7) is affected O(1). In Eq. (8) we have remedied this artifact of our ap-
by the interactions with the neighboring spins, as it de- proximation by introducing an upper cutoff of 1 on the
pends explicitly on the dynamical variables ~σ∂ . We refer locator product.
to these variables as the “neighboring” or “environmen- To obtain a meaningful decay rate, we still need to
tal” spins in the following. It is natural to expect that specify how to average over the random fields i that
different values of the decay rate are obtained depending enter the above calculation. In the case of a liquid en-
on whether these variables are treated as frozen in a typ- vironment, the decay rate should first be averaged over
ical configuration at inverse temperature β, or as liquid the annealed environmental spin variables, as described
above; to obtain the typical decay rate, the resulting Γ0
should then be logarithmically averaged over the local
1 Notice that in single particle problems the effect of self-energies
can be accounted for approximately, by imposing a lower bound
to the energy denominators of the locators as suggested already 2 The difference between frozen and liquid environmental spins is
in Ref. [1]. This preserves the statistical independence of lo- particularly relevant for models where the local fields along a
cators along the path. We implement these corrections in the given path depend mostly on environmental spins off the path,
calculations in Appendix B, the results of which are shown in rather than a fixed local random field. This is usually the case,
Sec. V. for lattices with a large connectivity.
5

fields i . Notice that in this case the decay rate depends We caution that in this simple 1d case the localization-
on the local random fields i via the energy denominators, delocalization transition predicted by γ = 0 should be
as well as via the distribution P (hs |~s ) in (4), since the taken with a grain of salt, since it is well known that for
Boltzmann weight of the neighboring spins σjz , depends single particle Anderson localization in 1d, there is never
on the local fields j . In contrast, when the environment a genuine delocalization. In that case γ = 0 only marks
is treated as frozen (anticipating a fully localized phase) the crossover to the weak localization regime, where the
the average of log Γ should be taken over local fields and localization length becomes large, even though it does
the configuration of neighboring spins, since the latter are not diverge due to weak localization effects. For many-
quenched during the decay time. The resulting distribu- body problems however, the latter are usually inessential
tion is then simply given by P (h) in (5). This leads us and γ = 0 can still be taken as a reasonable estimate
to define the following two spatial decay constants γF,L for the transition. Despite its simplicity this example al-
that characterize the decay rates in frozen and liquid en- ready exhibits the transition (within the insulating phase
vironment: of a system in a fluctuating environment) between the
Z Y ! “non-resonating” and the “rarely-resonating” insulator.
log ΓR The latter fluctuates into becoming metallic for a small
γF = − min lim dhi P (hi ) ,0 ,
R→∞ R fraction of the time. Technically, it is characterized by
i
Z Y ! (9) dominant environmental configurations that saturate the
log ΓR bound in Eq. (8), while in the non-resonating insulator
γL = − min lim di f (i ) ,0 ,
R→∞ R the dominant fluctuations still have exponentially small
i
decay rates, so that the bound in Eq. (8) remains irrele-
where i runs over all lattice sites, and the subscripts F vant.
and L stand for “Frozen” and “Liquid”, respectively. The
expression for ΓR entering in the definition of γF is given
by (7) with the notation heff
s → hi , while ΓR in the def- Frozen vs annealed decay rate
inition of γL is given in (8). As usual, the convexity of
the logarithm implies γL ≤ γF . This is in line with the We start by deriving explicit expressions for the spatial
physical expectation that a fluctuating environment of decay constants γF,L . We focus on frequencies in the
neighboring spins can increase the abundance of small middle of the range of local excitations, choosing ω = 0
denominators. for simplicity and dropping the dependence on ω from
now on. The frozen constant γF , when positive, is readily
computed as:
Coexistence of frozen and liquid phases?
Z 2
J⊥
As discussed above, delocalization happens when γ = γF = − dh P (h) log . (10)
h
0. From the fact that the annealed decay rate is always
bigger or equal to its frozen counterpart, one might ex- Recall that the distribution of local fields P (h) does not
pect that at a given temperature and at fixed W , the depend on temperature, and therefore γF does neither.
(c)
critical value of the transverse field, J⊥ , at which γ = 0 In contrast, the liquid decay rate does depend on tem-
depends on whether the environment fluctuates or not, perature. To compute it, we re-write:
and that it might be smaller for a fluctuating environ- Z ∞
(c) (c)
ment than for a frozen environment, J⊥,L < J⊥,F . If ΓR [~] = dx min e−Rx , 1 PR (x|~),

(11)
that happened, it would suggest a regime of coexistence −∞
(c) ∗(c)
J⊥,L < J⊥ < J⊥,F , in which both assumptions, a frozen,
localized phase or a liquid, delocalized phase are self- where PR (x|~) is the probability (over the thermal config-
consistent. However, this scenario will be ruled out be- urations of neighboring spins) of finding a path amplitude
(c) (c)
low, as we show that in fact J⊥,L = J⊥,F , since the an- that decays with spatial decay constant x. Formally it
nealed and frozen averages become equal at criticality. equals to:

R
Z Y R !
2 X J⊥
PR (x|~) = dhs P (hs |~s )δ log + x .
IV. ONE DIMENSION - A SINGLE DECAY s=1
R s=1 hs
PATH (12)
We show in (18) below that when evaluated for a typical
To illustrate the phenomenology of annealed and realization ~typ of the quenched random fields on path
frozen decay rates in the simplest possible framework we sites and on their neighbors, this probability can be re-
consider first the case in which only one decay channel written as:
is accessible to the excitation, meaning that the sums
in (7) and (8) reduce to a single path, see Fig. 1 (top). PR (x|~typ ) = e−RΣ(x)+o(R) . (13)
6

With this one readily obtains the large R limit of Γ, After this digression, let us now return to evaluating
Z ∞ the annealed spatial decay rate in a fluctuating environ-
ΓR [~typ ] = dx min e−Rx , 1 e−RΣ(x)+o(R) , (14)
 ment, which we expect to be smaller than the one cor-
−∞ responding to typical configurations of the environment.
The configurations of ~σ∂i that dominate the annealed av-
via a saddle point calculation. erage are usually exponentially rare (Σ > 0). As a matter
To obtain the probability of a spatial decay rate we of fact, implementing the constraint in (8) we find:
represent the constraint in (12) by an integral over an
0 ∞
auxiliary variable ξ, which leads to:
Z Z
0
ΓR [~typ ]= dxe−RΣ(x)+o(R) + dxe−R[x+Σ(x)]+o(R) .
R ∞ −∞ 0
Z
PR (x|~) = dξ eR[iξx+Φ̃R (iξ,~)] (15) (22)
2π −∞
Given that we restrict to the regime in which xtyp >
with: 0, the first integrand in (22) assumes its maximum at
R Z  a value of x outside the integration domain, and thus
1 X J
2z log h⊥ the integral is dominated by the contribution from the
Φ̃R (z,~) = log dhs P (hs |~s )e s . (16)

R s=1 boundary x = 0. The second term is dominated by the
point xSP that solves the saddle-point equation Σ0 = −1
The probability (15) depends on the field realization ~ or:
and is in general not self-averaging, being exponentially
small in R; however, the quantity (16) is intensive. Its ξ ∗ (xSP ) = 1, (23)
typical value equals
as follows from (19) and (20). The solution reads:
Z Y
Φ(z) = [di f (i )]Φ̃R (z,~) =
D 2 2 E
Z log Jh⊥ Jh⊥
i
(17) xSP = − d~s p(~s ) D 2 E . (24)
J⊥
Z Z
J

2z log h⊥
d~s p(~s ) log dhs P (hs |~s )e s , h

Q This value belongs to the integration domain as long


where p(~s ) = f (s ) u∈∂s f (u ). Therefore the proba- as xSP > 0. In
 this case one finds ΓR =
bility (15) evaluated on typical realizations ~typ , after a exp −γLnaive R + o(R) with
rotation in the complex plane, can be written as * +
J ⊥ 2
Z
i∞ naive
γL = −Φ(1) = − d~s p(~s ) log . (25)
Z
PR (x|~typ ) = cR dξ eR[ξx+Φ(ξ)] ≡ e−RΣ(x)+o(R) h
−i∞
(18) However, once xSP saturates to zero, the boundary value
where cR is a constant that is sub-exponential in R. For x = 0 dominates the second integral, implying ΓR =
large R, the leading order term Σ(x) can be obtained exp (−γLres R + o(R)) with
from a saddle point calculation as:
γLres = Σ(x = 0). (26)
Σ(x) = −[ξ ∗ (x)x + Φ(ξ ∗ (x))] (19)
Therefore, if positive, the liquid decay constant reads
where ξ ∗ (x) is defined as the inverse function of:
D 2 2ξ E γL = Θ(xSP )γLnaive + Θ(−xSP )γLres . (27)
0
Z log Jh⊥ Jh⊥
x(ξ) = −Φ (ξ) = − d~s p(~s ) D 2ξ E , (20)
J⊥ The fluctuations-induced transition in the insulator
h

R
with the notation h. . . i = dhP (h|~s )(. . . ). Within this The behavior of the spatial decay constants γF in (10)
framework the typical decay constant (as in a frozen ther- and γL in (27) is shown in Fig. 2 for a system with Hamil-
mal configuration of the environment) is recovered set- tonian (1) defined on a path with N = 2 environmental
ting ξ ∗ = 0, with which one indeed finds spins attached to each site as in Fig. 1 (top). The fields
i are taken to be uniformly distributed in the interval
xtyp = x(ξ ∗ = 0) = γF (21) [−W/2, W/2], and the interactions Jij ≡ Jz are constant.
The analysis of this case of a single decay channel leads
upon comparing (20) and (10). Further, from (19) and to the following conclusions:
using the identity Φ(0) = 0 (cf. Eq. (17)) one confirms
that Σ(xtyp ) = 0 for this decay rate, as it must be for the (i) Thermal fluctuations do affect spatial localization.
logarithm of the probability of typical configurations. The liquid decay rate is always smaller than the
7

increase, the dominating spin configurations along


a path become such that the far distance tunnel-
ing amplitude reaches values O(1), which does not
decay exponentially with R. This is the “rarely-
resonating” regime. However, the system still re-
mains an insulator since the rate Γ is now pro-
portional to the probability of occurrence of such
configurations through fluctuations, which is ex-
ponentially small in the path length and given by
∼ e−RΣ(0) with Σ(0) = Φ(ξ ∗ (xSP = 0)).

(iii) Thermal fluctuations do not affect the transition


point out of the insulator. The vanishing of the
decay constant, γ = 0, occurs at exactly the same
value of the parameters in both the frozen and
the liquid case: for γL = minx≥0 {x + Σ(x)} to
FIG. 2. Comparison between the decay constants γF (blue)
be zero, one indeed needs that both x = 0 and
and γL (red/purple), frozen and liquid environments, respec-
tively, for a chain (or a single path) each spin of which is cou- Σ(x = 0) = 0, cf. Eq. (26). Since Σ(x) vanishes for
pled to N = 2 environmental spins. Parameters are W = 1, x = xtyp (by definition), this implies together with
β = 0 and Jz = .26. The red dot marks the transition be- Eq. (21) that at these parameters γF = xtyp = 0 as
tween the non-resonating and the rarely-resonating insulator well. Exactly at the transition, the configurations
of a system with a liquid environment. It occurs when the of the environmental spins that dominate the av-
dominating fluctuations give rise to spatial decay constants erage conductance and lead to path weight of am-
xSP = 0. In the non-resonating insulator the total liquid plitude ∼ 1 become typical. Therefore they con-
decay constant is larger than xSP = 0 by an amount Σ(xSP ), tribute to the quenched conductance and the two
because the dominant fluctuations are exponentially rare tem- conductances become equal.
poral fluctuations. The black dot marks the “delocalization”
(c)
crossover J⊥ at which both decay constants become zero. At
this point the system turns metallic, irrespective of the nature
In conclusion, even though the fluctuations of neigh-
of the environment. The dashed red line shows the analytic boring spins can indeed open more favorable decay chan-
continuation of the curve xSP + Σ(xSP ) into the region where nels for the decaying excitations, they do not shift the
the decay rate of rare thermal fluctuations, naively evaluate, boundary of the localized phase. We note that in order
would take unphysical negative values xSP < 0. to reach this last conclusion it was crucial to impose the
physical bound prohibiting exponentially large, unphysi-
cal path amplitudes: Failing to do so, one would miss the
frozen one. Indeed, for generic values of the pa- transition in (ii) and would erroneously conclude that the
rameters it holds that: two spatial decay constants γL and γF vanish at different
points in parameter space (see the dashed red line in Fig.
xSP + Σ(xSP ) ≤ xtyp , (28) 2).

where xtyp is defined by Σ(xtyp ) = 0. Thus,


γL ≤ γF , consistent with the expectation that an V. OPTIMIZING OVER MANY PATHS
annealed average over the environmental spins is
in general dominated by configurations that are
more resonant with the decaying excitation, but We now extend these results to the case of an expo-
are too rare to contribute to the typical spatial nentially large number of paths as is relevant for higher
rate γF (as they have exponentially small probabil- dimensions d > 1. In particular, we will confirm the per-
ity, Σ(xSP ) > 0). Fluctuations of the environmen- sistence of the transition between a non-resonating and
tal spins thus lead to a larger effective localization a rarely-resonating insulator in the presence of a liquid
length of local excitations. environment. However, in addition, we identify another
transition within the rarely resonating phase, which is the
(ii) Within the insulator, a transition occurs in the an- analogue of the glass transition in the related classical
nealed conductance. The spatial decay constant γL problem of a directed polymer in a random medium: On
exhibits a non-analyticity when xSP reaches zero. the more insulating side of the rarely resonating phase,
For weak hopping terms (or quantum fluctuations), the dominant path that occasionally becomes metallic
xSP > 0. This means that the thermal configura- lives is a fixed, frozen optimal path. However, upon
tions that contribute dominantly to the decay rate approaching the metallic phase, there are exponentially
give rise to a path weight exp(−xSP R) that is it- many paths that occasionally become metallic through
self exponentially small in R, and resonances are thermal fluctuations. In other words, the optimal paths
strongly suppressed. As the quantum fluctuations itself starts to fluctuate in space.
8

In order to simplify the analytical treatment, we con- the calculations, we assume the distribution f () to be
sider a Bethe lattice of branching number k, where addi- Gaussian as well, with standard deviation W ,
tionally each site s has its own N environment spins σjz
2
with j ∈ ∂s, see Fig. 1 (bottom). Each of these √ spins e− 2W 2
interacts with σsz through a coupling Jsj ≡ Jz / N , and f () = √ . (35)
2πW 2
sits in a field j drawn independently for every site from
the distribution f (). The quenched fields s along the Then the distribution of (33),
path are also random and independent, with identical
distribution f (). We assume N to be very large, so that (Es −s )2
Z −
e 2vM
(36)
the field transmitted to a site s by the environmental ρ(Es ) = ds f (s ) √ ,
spins, 2πvM

Jz X z is itself Gaussian, and:


henv
s =√ σj , (29)
N j∈∂s 
h2s

Z exp − 2(J 2 +W 2)
z
is a fluctuating Gaussian variable. The σjz are thermally P (hs ) = dEs ρ(Es ) P (hs |Es ) = p
2π(Jz2 + W 2 )
fluctuating and have the probability distribution p(σjz ) =
(37)
(1 + mj σjz )/2 where mj = tanh(βj ). Thus the henv
s have
is independent of temperature as it has to be, see the
means Ms and variances Vs (with respect to the thermal
discussion around Eq. (5).
fluctuations) that depend explicitly on the random fields
~∂s , and read:
N Biased distribution of quenched fields along the
Jz X optimal path
Ms ≡ M (~∂s ) = henv
s = √ mj ,
N j=1
N Let us now discuss how to account for the presence of
2 Jz2 X
Vs ≡ V (~∂s ) = (henv
s )
2 − henv
s = (1 − m2j ), multiple paths. The main difficulty in this case consists
N j=1 in simultaneously taking into account the constraint on
(30) the liquid decay constant being non-negative, (8), and
the statistics of the quenched fields ~. As an exponen-
where the line denotes the average over thermal fluctu- tially large number of decay channels is available to the
ations of the σjz in their fixed random fields. Both Ms system, some of them have highly atypical distributions
and Vs are random variables; however, Vs has negligible of quenched fields ~ along the path and its environmental
fluctuations and tends to the fixed value sites. Those can lead to large deviations of path ampli-
Z tudes. If the fluctuations of the local fields are sufficiently
Vs → Vβ ≡ Jz d f ()[1 − tanh2 (β)].
2
(31) strong, these large deviations affect the statistical behav-
ior of sums of the form (7) and (8), in that they become
In contrast, the mean Ms fluctuates from site to site ac- dominated by a single (or at best a few) optimal path.
cording to a Gaussian distribution with zero mean and Let us first discuss the non-interacting limit of the
variance: problem, where the environmental spins are absent. In
Z this case it is known that in the whole localized phase
vM = Jz2 df () tanh2 (β) = Jz2 − Vβ . (32) sums of the form (7) are dominated by one single (or
at best a few) contribution(s) or “optimal path(s)”
Given a fixed configuration ~s = (s ,~∂s ) of the quenched [1, 41, 47], associated to a particularly large amplitude.
local fields on the given sites s and of the N neighboring In the non-interacting setting, on a Bethe lattice, this
ones j ∈ ∂s, the probability (over the thermal fluctua- can be rationalized straightforwardly via a mapping to
tions) to find an effective local field hs at the site s is a directed polymer, by identifying the sum Γ with the
therefore a Gaussian that depends on s and ~∂s only partition function of the polymer anchored at the origin
through the combination: of the lattice [24]. This mapping allows one to compute
the spatial decay constant γ as the free energy density
Es (~s ) = s + M (~∂s ). (33) of the polymer. Under this mapping the domination of
Γ by one (or few) optimal path(s) is the equivalent of
reading, the frozen (or glassy) phase of the polymer: the local-

−Es ]2
 ized phase for a non-interacting problem always maps to
exp − [hs2V β the glassy phase of the corresponding polymer problem.
P (hs |Es ) = . (34)
Delocalization occurs when the amplitude of the optimal
p
2πVβ
path becomes of order O(1), which means that exactly at
This corresponds to (4) and where we replaced the lo- the transition, among the exponentially many paths one
cal variance with the averaged one. To further simplify typically finds just one (or very few) that allow a decay
9

to the boundary. 3 Thanks to this mapping, the frozen frozen phase with essentially one optimal path. More
path nature of the localized phase is straightforward to generally it can be obtained by exploiting the directed
see, though one has to remember that the mapping relies polymer analogy, which we recall in Appendix A. The
on the forward approximation [46]. However, the frozen decay rate is obtained from the analog of the “replicated
path nature remains robust when self-energy corrections free energy”:
are added, as follows from rigorous results [47, 48]. In 1
 Z J 2η 

contrast, an interacting system with a fluctuating envi- ψF (η) = − log k dh P (h) , (40)
η h
ronment does not necessarily have to be in a frozen path
phase, as we will see below. with the distribution of fields P (h) as defined in (37).
As we recall in Appendix A, the expression for the non- The replicated free energy ψF (η) should be maximized
interacting decay constant γ on a Bethe lattice, as de- over 0 ≤ η ≤ 1. More precisely, denoting η ∗ the argu-
rived in [24], can be found from a convenient variational ment which maximizes ψF (40), one obtains for the decay
formulation [45]: the amplitude of the dominating path constant
is distributed like the maximum among k R path ampli-
(
ψF (η ∗ ) if η ∗ ≤ 1
tudes with negligible mutual correlations, which reflects γF = (41)
that the correlations among different paths on a Bethe ψF (1) if η ∗ > 1.
lattice do not matter for the problem at hand. The domi- The first case applies in the frozen phase where the
nating path will host a rather biased distribution of local sum over paths is dominated by few contributions
fields, ρopt opt
k (). The optimal distribution ρk () along the (sub-exponentially many in number). We discuss the
optimal path can be determined by maximizing the ex- detailed calculation of this function in Appendix B.
pression of the decay constant, while making sure that Here, we simply remark that in order for (40) to
the probability of observing such a distribution on a ran- be defined for η ≥ 1/2, one needs to regularize the
domly picked path is at least O(k −R ), which ensures that divergence arising due to small field denominators.
such a path indeed typically occurs on a Cayley tree of Physically this is necessary since small denominators
depth R. The logarithm of the probability Pρ of finding a need to be resummed, which effectively cuts of the effect
path of length R with such a biased distribution is mea- of too small denominators. In practice it suffices to
sured by the relative decrease in entropy, or Kullback- impose a cutoff to the integration, which regularizes
Leibler divergence, of ρopt
k () as compared to the typical, the singularity from small denominators h ≈ 0 [1], see
unbiased distribution ρ() of local quenched fields: Appendix B for details. For a frozen environment we
! find that η ∗ < 1 in the whole localized phase, exactly as
ρopt
Z
opt k () in a non-interacting problem. This is expected, since the
log Pρ = −R ρk () log d. (38)
ρ() interactions enter the formalism only via the distribution
of the effective local fields.
Therefore:
(Z 2 ) We point out that it follows from the temperature
J opt independence of P (h) (cf. Eq. (37)) that the decay
γ= min log ρk ()d , (39)
ρopt
log Pρ  constant γF is independent of T . This implies that also
k : R =− log(k)
the location where it vanishes, i.e., the delocalization
see Eqs. A8 and (A13) in Appendix A. transition in a frozen environment, is temperature
This reasoning can straightforwardly be extended to independent. Since we will show later on that the
the computation of the frozen decay constant γF in the delocalization transition is independent of whether or
presence of environmental spins. Moreover, it will also not fluctuations are included, we draw the non-trivial
allow us to derive compact expressions for the liquid de- conclusion that the transition to the metallic regime
cay constant γL , and to account for the bound on the is entirely temperature independent in the model we
path amplitude as we did in the case of a single path. consider here.

Liquid environment - Let us now turn to the more


Directed polymers in frozen vs liquid environments complex calculation of the decay constant in a liquid,
fluctuating environment. Let us first assume that the
Frozen environment - The decay rate in a frozen en- sum (7) is dominated by an optimal path, with a config-
vironment can be obtained following the recipe given in uration of fields ~s and thus of effective on-site fields Es
the previous section, which gives the correct result in a along the path. We call ρoptk (Es ) the probability density
describing the frequency with which the effective field Es
is encountered along that path, which of course differs
from the distribution across the whole system,
3 2
Es
There is no transition in the equivalent directed polymer prob- − 2(v 2
lem, where the delocalization point merely maps to a point where e M +W )
(42)
the free energy density (the analogue of γ) vanishes.
ρopt
k (Es ) 6= ρ(Es ) = p .
2π(vM + W 2 )
10

The optimal path amplitude depends not only on the where


local fields s at each site of the path, but via the fluctu- 2 2 !
ρopt (E)
Z Z
ating averages Ms also on the fields on the neighboring J⊥ J⊥
xSP = − dE k dh P (h|E) log
sites ∂s. Let us denote by E~ the collection of effective ω1 (E) h h
fields Es along the optimal path. (51)
We proceed by determining the decay constant γL [ρopt k ]
along this single optimal path as described in Sec. IV, and the two constants are given by
and subsequently determine the biased distribution (42)
by optimizing the resulting constant over ρopt
Z
under the
k γLnaive [ρopt
k ] = −Φ[1; ρopt
k ] = − dE ρopt
k (E) log[ω1 (E)]
constraint:
! (52)
ρopt
Z
opt k (E) and
dE ρk (E) log = log k. (43)
ρ(E)
γLres [ρopt opt ∗ opt
k ] = Σ[x = 0; ρk ] = −Φ[ξ ; ρk ]. (53)
The analogue of (12) describing the probability to en-
counter a thermal fluctuation giving rise to the decay It is straightforward to check 4 from these two expressions
constant x along this path now takes the form: that Eq. (50) can be compactly rewritten as
R !
γL [ρopt opt
Z Y
2 X J⊥

~
PR (x|E)= dhs P (hs |Es )δ log +x . (44) k ] = max0≤ξ≤1 −Φ[ξ; ρk ] . (54)
s
R s=1 hs
These equations express the fact that the decay constant
opt
Let us denote with E~typ a realization of fields along the γL corresponds to the naive annealed average γLnaive , un-
path, that is typical with respect to the unknown distri- less the expression for the latter is dominated by unphys-
bution (42). We have ical, exponentially growing contributions reflected by a
negative growth rate xSP < 0 at the saddle point. In
opt
opt
PR (x|E~typ ) = e−RΣ[x;ρk ]+o(R) , (45) that case the correct decay constant is given by (53),
which encodes the probability to encounter a thermal
where Σ[x; ρopt
k ] is now a functional of the biased density fluctuation that produces a metallic conduction along
opt the path. The logarithm of this probability is Σ[x =
ρk , defined as:
0; ρopt ∗ opt ∗
k ] = −Φ[ξ ; ρk ], cf. (46), where ξ is the solution
Σ[x; ρopt ∗ ∗ opt

k ] = − ξ x + Φ[ξ ; ρk ] . (46) of −dΦ/dξ = x = 0, or:
Here 2ξ∗ 2 !
ρopt
Z Z
k (E)
J⊥ J⊥
dE dh P (h|E) log = 0.
Z "Z 2ξ #
J⊥ ωξ∗ (E) h h
Φ[ξ; ρopt
k ] = dE ρopt
k (E) log dh P (h|E) ,
h (55)
(47) It remains to find the optimal biased density ρopt k . Our
analysis below closely parallels the calculation in App. A
and ξ ∗ = ξ ∗ [x; ρopt
k ] is obtained by inverting the saddle for the non-interacting case. We optimize the decay rate
point condition x = −dΦ[ξ; ρoptk ]/dξ, which reads explic- subject to the Kullback-Leibler constraint (43) and the
itly: normalization constraint on ρopt k . To do so we define the
2ξ 2 ! functional:
ρopt
Z Z
k (E)
J⊥ J⊥
x = − dE dh P (h|E) log Z
ωξ (E) h h L[ρk ; ξ, µ1 , µ2 ] = − dE ρopt
opt
k (E) log[ωξ (E)]+
(48) "Z ! #
with opt ρopt
k (E)
µ1 dE ρk (E) log − log k + (56)
Z 2ξ
J⊥ ρ(E)
ωξ (E) = dh P (h|E) . (49) Z 
h
µ2 dE ρopt
k (E) − 1 .
In this notation, we leave the dependence on the cou-
plings J⊥ and Vβ is implicit. Similarly as in the frozen
case, we use a cut-off around small fields to regularize
the integral (49) for ξ ≥ 1/2, see Appendix (B) for de- 4 This follows from the fact that for fixed ρopt
k , the function x(ξ)
tails. Following the same steps as in Sec. IV we ob-
defined in Eq. (48) is monotonically decreasing with ξ: ∂x/∂ξ ≤
tain for the decay constant γL , defined via the decay 0, as can be verified with simple convexity arguments along the
opt
exp(−γL R) ∼ dx min e−xR , 1 PR (x|E~typ
R 
): lines of those presented in Appendix C. Therefore, if x(ξ) is found
to be negative at ξ = 1, the equality x = 0 will be attained at a
γL [ρopt
k ]=Θ(xSP )γL [ρk ]+Θ(−xSP )γLres [ρopt
naive opt
k ], (50)
value of ξ < 1.
11

The first line is −Φ[ξ; ρopt ∗


k ] with ξ = ξ < 1 in the res- the saddle point decay rate is positive, xSP > 0,
onating phase or ξ = 1 in the non-resonating phase, as and thus physical. xSP > 0 implies ∂ΨL /∂ξ > 1,
can be seen from Eqs. (47,52, 53). For any value of ξ, we which guarantees that ΨL (µ∗ , 1) is indeed a maxi-
find that the normalized solution of mum on the domain 0 ≤ µ, ξ ≤ 1. When this holds
true, it turns out that the maximizing µ∗ always
δL[ρopt
k ; ξ, µ1 , µ2 ] satisfies µ∗ < 1. This follows from a simple convex-
=0 (57)
δ ρopt
k ity argument given in Appendix C. We recall that
µ∗ < 1 indicates that in this phase the decay rate is
reads dominated by a sub-exponential number of paths.
ρ(E) [ωξ (E)]µ Since ξ = 1 we further know that each of those
ρopt
k (E) =
R , (58) contributes with a strictly exponentially decaying
dEρ(E) [ωξ (E)]µ
term (xSP > 0). This regime thus corresponds to
where we have substituted µ ≡ 1/µ1 . Injecting this form the deep, non-resonating insulator phase. The total
into the functional L we obtain the function spatial decay constant is obtained substituting the
 Z  optimal distribution into (52) for the annealed ther-
1 mal average along the optimal path, which yields
ΨL (µ, ξ) = − log k dEρ(E) [ωξ ]µ , (59)
µ
γL = γLnaive [ρopt ∗ ∗
k (µ = µ )] = ΨL (µ , 1). (62)
which still needs to be extremized with respect to µ,
whereby it turns out that the relevant extremum is al- (b) If the assumption ξ = 1 and maximizing over µ
ways a maximum. Notice that in this expression µ plays leads to the physically inconsistent xSP < 0 in (51),
the same role as the parameter η in the decay rate (40) we know that ΨL assumes its maximum for ξ < 1.
in a frozen environment. Let µ∗ the argument that Assuming ξ < 1 in turn means that the dominant
maximizes ΨL , i.e., that solves: thermal fluctuations constrained in such a way that
we make sure that the dominant path amplitudes
∂ just reach 1 (or xSP = 0). Such a regime corre-
ΨL (µ, ξ) = 0. (60)
∂µ sponds to a rarely-resonant insulator. In that case
the physical bound on the path amplitude should
Note that imposing the constraint (43) on the maxi- be implemented by solving simultaneously the two
mal Kullback-Leibler divergence only makes sense if the saddle point equations:
decay rate is maximized by a single path (or at most a
sub-exponential number of paths). This is the case if and ∂ ∂
ΨL (µ, ξ) =0= ΨL (µ, ξ) , (63)

only if one finds µ∗ < 1. Otherwise one should impose no ∂µ ∗
µ ,ξ ∗ ∂ξ µ∗ ,ξ ∗
constraint, which is equivalent to setting µ = 1, in perfect where the second equation ensures that x = 0, see
analogy to the non-interacting case. At the point where (55). As long as the solution to both equations
the solution of Eq. (60) reaches µ∗ = 1, the quenched yields µ∗ < 1, there is a single optimal path, which
average over the random fields E becomes exactly equal does not change under thermal fluctuations. Rare
to the annealed average, as one can see by setting µ = 1 thermal fluctuations on this path turn it metal-
in (59). This corresponds to a melting transition in the lic, which dominates the decay. This regime cor-
associated polymer problem. While the Lagrange param- responds to a path-frozen, rarely resonating insula-
eter µ controls the number of paths that contribute, we tor.
recall that the remaining parameter ξ is associated with
constraining the sampling of thermal fluctuations that Once the maximizing µ∗ approaches µ∗ = 1, the
lead to potentially negative decay constants. The ex- system undergoes a transition to a non-frozen phase
pression for ΨL still has to be maximized with respect to in terms of the dominating decay paths. In this
ξ over the interval 0 < ξ ≤ 1. We recall that a maximum regime, one has to set µ = 1, while
ξ ∗ is determined

with ξ < 1 signals that the dominant thermal fluctua- as the solution of ∂ξ ΨL (1, ξ) ∗ = 0. One then

ξ
tions turn the system temporarily metallic on the given obtains:
path, indicating a rarely resonating insulator phase. (
The above formalism suggests the following algorithm ΨL (µ∗ , ξ ∗ ) if µ∗ ≤ 1 path-frozen
to calculate the decay constant γL : γL = γLres = (64)
ΨL (1, ξ ∗ ) if µ∗ > 1 unfrozen .
(a) One first assumes ξ = 1 and determines µ∗ from This yields the decay constant for the rarely-
∂ resonating insulator.
ΨL (µ, 1) ∗ = 0. (61)

∂µ µ It is straightforward to deduce from the above that this
procedure is equivalent to the double maximization of ΨL
The optimal distribution of E is then given by over a compact interval:
Eq. (58) with µ = µ∗ and ξ = 1. Upon substi-
tuting this in Eq. (51), one then checks whether γL = max0≤µ,ξ≤1 ΨL (µ, ξ). (65)
12

Let us briefly discuss the resulting expressions for the


spatial decay constants. Comparing (59) with the ex-
pression (40) for the frozen case, we see that ΨL (µ, ξ)
plays the role of the replicated free energy of a directed
polymer, but now for a liquid environment. While in the
frozen case the thermal realization of the environmental
spins ~σ∂z and the random local fields ~ are treated on the
same footing, entering as quenched disorder into the dis-
tribution P (h)), in a liquid environment the thermal fluc-
tuations are fast and averaged over first. This leads to the
modified locator ωξ in (49), which takes the place of the
simpler locator (J⊥ /h)2 of the frozen problem. The aver-
ages over the random local fields and over the configura-
tion of environmental spins, respectively, are associated
with the two distinct parameters, µ and ξ. The param-
eter µ controls whether the number of paths that domi-
nate the decay rate is small (in the path-frozen regime,
µ∗ < 1) or exponentially large (in the path-molten in-
sulator µ = 1). The parameter ξ instead is tuned such
as to control the path amplitudes and prevent unphysi-
cal, exponentially growing contributions to the decay rate
arising in the rarely-resonating insulator (which requires
a non-trivial value ξ ∗ < 1).
These parameters are thus seen to take rather different
roles. While at first sight it may look as if the transition
from a rarely-resonating to a non-resonating insulator, as
signaled by the crossing of ξ = 1, corresponds to some
kind of glass transition, due to its formal resemblance to
freezing transitions in replica theory, this is actually not
the case. At a glass transition, the number of metastable
configurational valleys contributing significantly to the FIG. 3. Top. Decay constants γF,L in a frozen and liquid
total free energy valleys shrinks from exponentially many environment, respectively, on a Bethe lattice with branching
to O(1). In our case, the role of configurational valleys of number k = 2, while every lattice spin couples to a sepa-
a directed polymer is assumed by distinct decay paths, rate environment made of N  1 spins. We assume infinite
and the corresponding unfreezing transition is indicated temperature, β = 0. The decay constants are plotted as a
by µ∗ reaching 1. In contrast, ξ ∗ becoming smaller than function
√ 2 of the ratio between the effective disorder strength
one indicates the vanishing of the dominant decay rate W + Jz2 , which controls the width of the distribution P (h)
of the local fields, and the amplitude of quantum fluctuations
observed in rare thermal fluctuations, xSP = 0, but it
J⊥ . Localization is always stronger in a frozen environment,
does not indicate the vanishing of the logarithm of the as seen by the strict inequality γF > γ√L which holds all the
associated probability Σ(xSP ) (which is what one might way to the delocalization transition at W 2 + Jz2 /J⊥ = 5.54.
expect from a freezing transition). Indeed, Σ(xSP ) is still Close to the metallic phase the insulator with liquid environ-
strictly positive at the transition to the resonating insu- ment is in a rarely resonating phase. As the effective disor-
lator. der
√ increases, γL undergoes first a path-freezing transition at
W 2 + Jz2 /J⊥ ≈ 15.01, and √ subsequently a transition to a
non-resonating insulator at W 2 + Jz2 /J⊥ = 31.23.
Bottom. Evolution of the parameter η ∗ that extremizes
VI. THREE INSULATING PHASES BROUGHT
ΨF (η) for a frozen environment, and of the parameters µ∗ , ξ ∗
ABOUT BY THERMAL FLUCTUATIONS
maximizing the liquid functional ΨL capturing the decay rate
in a liquid environment. A non-resonating insulator is identi-
In Fig. 3 (Top) we plot the decay constants γF , γL for fied by ξ ∗ = 1, while the path un-frozen insulator has µ∗ = 1.
a Bethe lattice with branching number k = 2 (each site These two phases are always separated by an intermediate
having k + 1 = 3 neighbors on the lattice) as a function phase with non-trivial µ∗ , ξ ∗ < 1 corresponding to a rarely
of disorder. Fig. 3 (Bottom) shows the corresponding resonating, but path-frozen insulator. At the delocalization
evolution of the liquid parameters µ, ξ, as well as the transition, ξ ∗ = η ∗ .
parameter η in the frozen case, as defined in the previous
section. We refer to Appendix B for details about the that:
computation. While a frozen environment only gives rise
to a single insulating phase, the situation of a fluctuating (i) Upon decreasing the disorder or increasing the
environment is much richer. From the plots one can see quantum fluctuations, the transition from a non-
13

resonating to a rarely-resonating insulator is pre- where P (h) is the unconstrained distribution defined in
served from the 1d situation discussed in Sec- Eq.(37). Here ξ ∗ is fixed by the condition:
tion IV, even though now there are exponentially
many paths available for decay. Z 2 2ξ∗
J⊥ J⊥
(ii) The liquid decay rate undergoes a further transi- dh P (h) log = 0. (67)
h h
tion within the rarely-resonating insulator phase.
It can be identified with an unfreezing transition of
the corresponding directed polymer problem. Ob- Assuming ξ ∗ > 0 we can divide each of these two van-
viously, such a configurational unfreezing cannot ishing quantities by some strictly positive quantities to
occur in a 1d setting with only a single decay path obtain the equation:
available. This transition between a path-frozen
and an unfrozen regime always occurs within the  2ξ∗ 
rarely-resonating regime, as we prove in Appendix log k dh P (h) Jh⊥
R 2 2ξ∗
dh P (h) log Jh⊥ Jh⊥
R
C. In contrast, in a system with a frozen environ- = ,
2ξ∗
ment the dominant decay always occurs along the ξ∗ R
dh P (h) Jh⊥
[49] same or the same few paths. It is the thermal (68)
fluctuations of environmental spins and the related which is equivalent to the condition for a maximum of
changes in local fields that make the dominant de- the frozen decay functional ΨF ,
cay paths of a system with liquid environment fluc-
tuate. This path melting always takes place in a
boundary regime adjacent to the transition to the d
η2 ΨF (η) ∗ = 0. (69)

metal. We will come back to the properties of this dη η=ξ
phase in Sec. VII
(iii) The delocalization transition occurs at the same This means that when the liquid decay vanishes, γL = 0,
value of parameters, whether a frozen or a liquid en- the extremizers ξ ∗ for the liquid case and µ∗ for the frozen
vironment are considered. This remains unchanged case coincide, ξ ∗ = η ∗ . This is seen in the explicit so-
from the 1d case of a single path. The transition lution of Fig. 3 (Bottom) . The frozen decay is given
always occurs out of the path-unfrozen, rarely res- by Eq. (40). By virtue of Eq. (67) it vanishes as well,
onating phase, where the parameter µ = 1 signals γF = 0, which completes our proof. We conclude that
the contribution of exponentially many paths. This the sequence of transitions shown in Fig. 3 and the co-
is proven in Appendix C. As we argued earlier, ev- incidence of liquid and frozen delocalization hold for any
erywhere within the insulator the strict inequality choice of model parameters (such as the connectivity k,
γF > γL holds. e.g.). We recall that our treatment of the decay is only
approximate in that it is essentially a forward approxima-
We will discuss how the different insulating phases could tion where small denominators are suitably regularized.
be distinguished by physical observables in Sec. VII. This captures only approximately the (anti)correlations
between the locators at subsequent sites along the path
that are induced by the self-energy corrections [50]. How-
No coexistence of frozen and liquid phases ever, despite this approximation, we find that at the lo-
calization transition the values of η ∗ = ξ ∗ come close to
Let us now address the question of the possibility of the exact value of 1/2, which follows from a symmetry
a coexistence of frozen and liquid phases in this class of argument [41].
models. With the above results in hand we can now show The physical reason for the coincidence of the liquid
explicitly that this is excluded since the two decay rates and the frozen delocalization transitions, γL = γF = 0
γF , γL vanish exactly at the same value of the hopping is relatively straightforward. For any given thermal real-
and disorder parameters. To prove this, since γF ≥ γL ≥ ization of the neighbor spin configuration, the local fields
0, it suffices to show that γL = 0 necessarily implies are as in a frozen configuration. For static configurations
γF = 0. As we show in Appendix C, in the presence it is known that on a Cayley tree there is an optimal
of a liquid environment the transition to delocalization path that dominates the decay. The thermally fluctuat-
always occurs inside a path-unfrozen phase, where the ing problem can have a non-exponential decay rate only,
parameter µ takes the value µ = 1. Setting µ = 1 in (59) if a typical thermal realization of neighbor spin config-
the liquid delocalization transition (γL = 0) requires urations gives rise to non-exponential decay, otherwise
 Z  non-exponential decay would only occur in exponentially
γL = − log k dEρ(E) ωξ ∗ rare fluctuations. However, the requirement that typical
(66) configurations give rise to non-exponential decay is ex-
 Z J 2ξ∗ 
⊥ actly the same as that for delocalization within a frozen
= − log k dh P (h) = 0,
h environment.
14

VII. PHENOMENOLOGY OF THE conductance across a mesoscopic sample, if direct tunnel-


INSULATING PHASES ing across the sample is relevant. In path-frozen phases
the conductance is essentially dominated by a unique op-
Fluctuating decay paths: path chaos timal path connecting the two leads. (In finite dimen-
sions, other than on a Bethe lattice, such an optimal
Our analysis in a liquid environment reveals the fact path is only defined up to small, spatially local fluctua-
that thermal fluctuations of the neighbor spin configu- tions.) Accordingly, the conductance is strongly suscep-
rations can favor different paths to become the instan- tible to disturbances that alter the local fields on that
taneously optimal decay channels. In other words, in path. Such local perturbations could be produced, e.g.,
the course of time the dominating decay path fluctuates, by an atomically sized tunneling tip in the close vicinity
because the effective local field distribution fluctuates. of the surface of a 2D sample. By scanning the sample
However, this happens only close enough to the delo- surface laterally (parallel to the lead interfaces) one may
calization transition, in the phase that we dubbed the expect a sudden spike or drop in conductance as the tip
path-unfrozen phase. This is a close analogue of a phe- approaches the dominating conductance channel. In an
nomenon observed in numerical studies of Anderson lo- un-frozen, path-fluctuating insulator instead, the domi-
calization, where optimal decay paths were found to be nant path is constantly fluctuating itself and any static
very sensitive to changes in the disorder potential. [49] In local disturbance has little effect on the conductance.
the closely related classical problem of a directed poly- Local disturbances in the path-frozen phase of Ander-
mer in random media, and more generally in spin glasses son insulators can generally modify the localization prop-
and similar glassy systems, a related kind of effect is erties of excitations, in case they happen to affect the
known under the notion of “chaos” [26, 27]: The opti- optimal propagation channel of a relevant single parti-
mal (ground state) configuration of the glassy system is cle wavefunction. It has recently been shown in single
very sensitive to changes in the disorder realization, or particle problems that indeed optimal decay paths can
to control parameters such as the temperature, such that abruptly change as the potential landscape is altered [49],
sudden jumps of the ground state configuration as a func- and analogies to closely related shocks and avalanches in
tion of those parameters may occur, provided the system glassy directed polymer problems have been drawn. It
has enough time to equilibrate. might be interesting to study the continuously occurring
The last point constitutes an important difference “wavefunction shocks” due to thermal fluctuations and
between quantum localization problems and classical the associated spatial range of paths that contribute to
glasses, however. The way these two classes of systems the conductance.
explore the available configuration space (i.e., the various Probing the transition between the rarely-resonating
paths for our localization problem) is fundamentally dif- and the non-resonating regimes is more subtle. Indeed,
ferent. A glassy system usually has to overcome huge it requires to determine whether the exponentially rare
energy barriers in configuration space to settle into a fluctuations that dominate the average (annealed) con-
new low energy configuration. This may take extremely ductance correspond to a decay rate that is itself expo-
long time, and thus often results in the system falling nentially small or whether it is O(1), meaning that the
out of equilibrium, displaying aging, etc. This renders sample is in some sense an intermittent metal. The oc-
the observation of equilibrium chaos in glasses very chal- currence of a metallic-like fluctuation, even just over brief
lenging. The analogue of chaos in quantum localization time windows, opens at least in principle the possibility
problems instead is probed very differently. By its very for non-negligible, sample-spanning coherence effects in
nature, quantum mechanics probes all available decay the conductance. In particular, Fabry-Pérot-like oscilla-
paths simultaneously and instantaneously, like in a tun- tions of conductance as a function of the distance between
neling problem with parallel tunneling channels. In par- two reflecting barriers inside the sample or at its bound-
ticular, the system usually retains no memory of optimal aries, can occur with significant amplitude only if, at least
paths that were favored by past disorder configurations for some instants of time, the transmission from barrier
that arise in the course of fluctuations. Insofar the decay to barrier is quasi-metallic and multiply reflected waves
channels react instantaneously to the fluctuations in the can potentially interfere with each other. In contrast,
disorder realization and thus offer an interesting exper- more local quantum interference effects, such as weak lo-
imental route to studying chaos phenomena, which are calization due to a sample-threading magnetic flux and
much harder to probe in their thermodynamic analogues. the ensuing alteration of conductance would hardly allow
to distinguish between a non-resonating insulator and an
intermittent metal. Indeed, both regimes are effected by
Experimental distinction of insulating phases magnetic fields. In fact, strong insulators are usually dis-
proportionately more affected than weak ones, since local
It is interesting to discuss how the three insulating interference effects are exponentially amplified in strong
phases could be distinguished by experimental observ- insulators. [44, 51, 52]
ables. For electrons, or other excitations carrying a con- The transition from non-resonating insulator to inter-
served charge, the decay rates are related to the system’s mittent metal also shows in a kink in the exponential de-
15

cay of the conductance with sample length. Likewise one curs and a rarely-resonating, but path-unfrozen insulator
may expect the conductance noise to bear a signature of emerges. In this least insulating phase an exponentially
the transition, since the conductance fluctuations in the large number of paths contribute to the time-averaged
resonating insulator start to be more strongly bounded decay (we recall that the paths are well defined in a Cay-
from above. ley tree approximation to real space, while in real space
spatially local fluctuations around a dominant decay path
will contribute as well). It is still true that at every in-
VIII. DISCUSSION AND CONCLUSIONS stant of time, if one were to freeze the configuration of
the environmental spins, one path essentially dominates
the decay. However, this dominant path is now strongly
We have analyzed the decay rate of local excita-
sensitive to the thermally fluctuating effective local fields,
tions interacting with thermally fluctuating environmen-
in an analogous manner as ground states of glassy sys-
tal spins. One of our driving questions was to under-
tems can change substantially with small modifications
stand whether internally generated fluctuations are able
to the couplings or thermodynamic parameters. Here,
to shift the boundary of stability of the localized phase,
we can prove for the model of a Cayley tree with neigh-
a scenario that would entail the possibility of a coex-
boring spins that such a path-chaotic phase generically
istence regime between a delocalized and a (possibly
exists close to the delocalization transition. As the lat-
metastable, but long-lived) localized phase in the vicin-
ter is reached, the decay rate becomes order O(1) and
ity of the localization-delocalization transition. It would
independent of the system size. As mentioned above, its
also open the possibility to tune delocalization by tem-
location is independent of whether thermal fluctuations
perature, via a mechanism that is different from the one
are accounted for or whether the environment is taken to
usually discussed in the context of MBL [2, 3], an idea
be frozen.
originally raised in Ref. [22].
Our analysis has indeed confirmed that the annealed Note that despite some similarities, the transition asso-
rate of decay (averaging over thermal fluctuations of the ciated with the unfreezing of the dominant path is differ-
neighboring spins) is always bigger than the quenched ent in nature from the crossover, or putative transition,
decay rate (that of a typical, frozen environmental con- between a fully ergodic phase and non-ergodic delocal-
figuration) because the fluctuation average is dominated ized phase, which has been controversially discussed, es-
by rare thermal configurations. However, the difference pecially for non-interacting problems on Bethe lattices
between the two rates diminishes upon approaching the or Cayley trees [50, 53–61]. Indeed, the transition we
metallic regime and disappears exactly at the delocaliza- have identified in this work occurs within the insulating
tion transition. The location of that transition is there- phase, even though it requires interactions with liquid-
fore independent on whether the fluctuations are taken like fluctuating degrees of freedom. This transition en-
into account or not, and the putative coexistence or bista- tails a non-analytic behavior of the spatial decay constant
bility scenario is ruled out. γL governing the decay rates of excitations in the thermo-
Nevertheless, we find that thermal fluctuations induce dynamic limit (akin to localization lengths or Lyapunov
a rich phenomenology within the insulating phase. In exponents in non-interacting systems). In contrast, the
general it hosts three different regimes, separated by two putative transition between an ergodic and possibly non-
sharp transitions (at the level of our approximate de- ergodic delocalized phase, that was suggested to map
scription). All three regimes are characterized by an ex- the unfreezing of an associated effective polymer prob-
citation decay rate that decreases exponentially in the lem, takes place within the delocalized metallic phase
system size. At strongest disorder, the system is a non- where the participation ratio and fractal properties of
resonating insulator, where even exponentially rare, opti- non-localized wavefunctions evolve in a non-trivial man-
mal fluctuations of the neighbor spin configurations give ner.
rise to exponentially weak decay. In other words, the sys- Note that we have carried out our analysis at high
tem is insulating even in those rare moments in which the temperatures, so as to single out the consequences of
environment is particularly favorable and which therefore the presence or absence of fluctuations of neighboring
dominate the decay. In contrast, in the intermediate in- spins, while excluding thermal and interaction effects on
termittent metall (or rarely-resonating, but path-frozen the global distribution of local fields. This lead to a ro-
insulator) the annealed time averaged decay rate is dom- bust scenario of three insulating regimes. It remains an
inated by rare fluctuations of the neighboring spins that interesting issue to investigate the dependence of their
induce metallic-like behavior; nevertheless, their expo- phase boundaries on the temperature, which governs the
nential rareness guarantees that the system is still in- strength of the fluctuations of the environmental spins,
sulating in the sense that the time averaged decay rate as well as on the strength of the interactions. We re-
is exponentially small in the system size. As in the non- call that the critical point separating the localized from
resonating phase, the dominant decay path remains fixed the delocalized phase is not shifted by temperature in
and does not change, even though the environment fluc- our model where the global distribution of local fields
tuates. This changes closer to the delocalization tran- P (h) in Eq. (5) has no explicit temperature dependence.
sition, where the analogue of a melting transition oc- It is thus unaffected by the presence of thermal fluctu-
16

ations. This follows from the coincidence of the critical has exactly the form (A1) with β = 1 and Es =
point in frozen and liquid environments, respectively, and − log |J⊥ /s |2 , where the s are the random local poten-
from the fact that transition in a frozen environment it tials with distribution f (). We thus set ws = |J⊥ /s |2 .
is a unique functional of P (h), as is seen from Eq. (40). The decay constant γ can be identified with the free en-
In contrast, the metallic intermittency and path melting ergy density of the polymer,
transitions that occur within the insulating phase in a
liquid environment are expected to display a dependence hlog ΓR i
γ = − lim . (A4)
on T , given that the latter affects the width of the dis- R→∞ R
tribution P (h|E) and thus changes the effective locators
The exact results in Ref. 24 show that γ is fully deter-
ωξ relevant in the fluctuating environment.
mined by the function:
We recall that in the model discussed here the local
fields in (8) are sums of uncorrelated contributions from 1
 Z 
environmental spins. More complex models in which cor- ψ(η) = − log k d f ()[ws ()]η , (A5)
η
relations may establish among the environmental spins at
low temperature, or where long range interactions play where k is the branching number of the Cayley tree. This
a crucial role and may lead to a significant temperature can be recognized as the replicated free energy (per unit
dependence of the global distribution of fields [62–66], length) in a replica approach.
might display a more complex phenomenology. Let us call η ∗ the argument that maximizes ψ, satisfy-
ing

ACKNOWLEDGMENTS d
ψ(η) = 0. (A6)

dη η=η ∗
V. Ros acknowledges funding by the LabEx ENS-
It was shown in Ref. 24 that :
ICFP: ANR-10-LABX-0010/ANR-10-IDEX-0001-02
PSL* and thanks the Paul Scherrer Institute in Villigen
(
ψ(η ∗ ) if η ∗ ≤ 1
and the Galileo Galilei Institute (GGI) in Florence for γ= (A7)
hospitality and support during the completion of this ψ(1) if η ∗ > 1,
work. M. Müller acknowledges funding from the Swiss
National Science Foundation under Grant 200021 166271 in agreement with the heuristic recipe that the replicated
and thanks GGI Florence for hospitality. free energy should be maximized over the domain 0 ≤
η ≤ 1. The first regime corresponds to the polymer being
in its frozen phase (with broken replica symmetry), in
Appendix A: Directed polymer in random medium: which the partition function (A1) is dominated by a sub-
a recap exponential number of paths.
As discussed in the main text, the spatial decay con-
stant in a frozen environment can be obtained as a
The free energy density of a directed polymer on the
straightforward application of these identities, by replac-
Bethe lattice has been computed explicitly in Refs. 24
ing the distribution f () with that of local fields, P (h).
and 67. On a finite Cayley tree of depth R, the partition
In contrast, in the presence of a fluctuating environment
function of the polymer is given by a sum over all paths
these identities cannot be generalized straightforwardly,
P connecting the root 0 to the boundary BR at distance
as one needs to enforce the physical constraint that the
R, and it takes the generic form:
single path weights are bounded by 1. In that case, it
X Y proves convenient to exploit a variational argument to de-
Zβ,R = ws , (A1) termine the distribution of the quenched energy variables
P:0→BR s∈P
along the paths that dominate the decay rate. Here we
recall this argument in the simpler setting of a directed
where ws = e−βEs is the contribution of the site s, with
polymer with no constraints, showing how it allows to
β the inverse temperature and Es a local energy. For
recover (A7).
a single-particle Anderson problem on the Bethe lattice
Let us assume that (A3) is dominated by a single path
with Hamiltonian
among the k R paths contributing to the sum. The lo-
X †  cal fields s along the dominating, optimal path have an
ci cj + c†j ci ,
X
H= i n i − J ⊥ (A2)
atypical statistics: they are not a typical sample from the
i hi,ji
distribution f (), but rather look like a sample from a bi-
ased distribution (to be determined) that we denote by
the decay rate ΓR computed in the forward approxima-
f opt (). The decay rate along this path is simply given
tion,
by
X Y J⊥ 2 Z
ΓR = (A3)
s
γ ≡ γ{f ()} = − d f opt () log[w()].
opt
(A8)
P:0→BR s∈P
17

However, to introduce some formal steps and functions ψ(η ∗ ), where η ∗ satisfies (A6), under the assumption that
whose more complex equivalents are used in the main one path dominates the sum ΓR .
text, we briefly derive it here by a formal detour. De- However, as η ∗ → 1, this assumption breaks down [67]
noting by ws the weights along the optimal path, we can and the system leaves the frozen phase. At that point
write the decay rate as: the quenched average (A4) becomes equivalent to the
! annealed one, meaning that we can replace hlog ΓR i by
R
loghΓR i. This yields the decay rate:
Z
−Rx 1 X
ΓR ≈ dxe δ x+ log ws
R s=1  Z J 2 
(A9) loghΓR i ⊥
Z Z PR γ = − lim = − log k d f () (A16)
=R dxe−Rx dξeiξ Rx+iξ s=1 log ws
. R→∞ R 

which equals ψ(η = 1).


We introduce the function
R
1 X Appendix B: Details of the evaluation of γL
Φ(z) = log wsz , (A10)
R s=1
In this appendix we provide details on the explicit eval-
and its average with respect to the biased distribution uation of the decay constant γL in a liquid environment,
Z the result of which is shown in Fig. 3 for a particular
Φ(z) = d f opt () log[w()]z . (A11) choice of parameters, see below. The main subtlety in
the calculation consists in regularizing the integral:
Taking the saddle point over x in (A9) we obtain that  2

iξSP = 1, and x drops out of the saddle point value. One
Z ∞ exp − [h−E]
2Vβ

J

2ξ log h⊥
then finds: ωξ (E) = dh p e , (B1)
−∞ 2πVβ
log ΓR
γ = − lim = −Φ(1), (A12) which has a divergence for ξ > 1/2. Following the origi-
R→∞ R
nal argument in [1], this is regularized by putting a cutoff
which is equivalent to Eq. (A8). around h = 0 on a scale ∆ representing the typical value
In order to determine the unknown biased distribution of self-energy corrections.
p In our setting, the relevant
2
f opt (), it suffices to maximize the functional Eq. (A8) scale is ∆ ∼ J⊥ / Vβ + W 2 . We therefore restrict the
subject to the normalization constraint and the require- average in (B1) to the domain R∆ = R/[−∆, ∆], and
ment that the Kullback-Leibler divergence equal log(k). for simplicity, we neglect the correction to the normal-
That is, we should maximize ization factor due to the cutoff ∆. We now provide the
Z explicit expression of the function ΨL obtained with this
F[f opt , µ1 , µ2 ] = − d f opt () log[w()]+ regularization.
The effective locator becomes:
f opt ()
Z 
d f opt () log
 
µ1 − log k + (A13) [h−E]2
f () J⊥2ξ Z exp − 2Vβ 1
 2 ξ
J⊥
ωξ (E)= dh = √ ×
h2ξ
p
π 2Vβ
Z 
2πVβ R∆
µ2 d f opt () − 1 , " ! !#
E ∆ E ∆
× Iξ p ,p +Iξ − p ,p
where the Lagrange multiplier µ1 enforces that the opti- 2Vβ 2Vβ 2Vβ 2Vβ
mal path occurs with a probability scaling as ∼ k −R , see (B2)
(38) in the main text, and µ2 ensures the normalization
of f opt (). The solution to where we introduced
∞ 2
e−(h+a)
Z
δF[f opt , µ1 , µ2 ] ∂F[f opt , µ1 , µ2 ]
=0= (A14) Iξ (a,u)= dh .
δf opt ∂µ2 u h2ξ

reads: A simple power expansion of the term e−2ha leads to the


1 following representation:
opt f ()[w()] µ1
f () = R 1 . (A15) − E
2
 2 ξ
d f () [w()] µ1 2fξ (∆) e 2Vβ J⊥
ωξ (E) = √ ×
π 2Vβ
It remains to optimize with respect to µ1 . Upon injecting   (B3)
2
 
the form of Eq. (A15) into (A13) and changing notation ∞ Γ 2k−2ξ+1 , ∆
X 2 2Vβ
to η ≡ 1/µ1 , we find F(µ1 = 1/η) = ψ(η), where ψ(η) × 1 + 1−k V k (2k)!
E 2k  ,
was given in (A5). We thus recover the exact result γ = k=1
f ξ 2 β
18

where the coefficients of the expansion contain the in- with coefficients:
complete Gamma function:
2k−1 2k − 2z + 1 ∆2
 

α̃k = Γ , .
Z ∞
f1 (∆) Vβk (2k)! ∂z 2 2Vβ

z=1
Γ(s, x) = ts−1 e−t dt. (B4)
x (B12)
It can be checked that upon taking the limit ∆ → 0 of
The term the coefficients of the series:
∆2
 
1 1 ∞
fξ (∆) = Γ − ξ, (B5)
X
2 2 2Vβ Σ2 (E) ≡ α̃k E 2k , (B13)
k=1
is the source of the divergence as ξ → 1/2 if no regularizer
the latter can be re-summed explicitly, too. Plugging the
is present, ∆ = 0. For k ≥ 1 and ξ ≤ 1 the coefficients
resulting expressions into (B10) and integrating over E
of the expansion (B3) remain regular for ∆ → 0, and
allows one to evaluate explicitly xSP as well.
we therefore approximate them with their ∆ → 0 limit.
The evaluation of the decay constant in a frozen environ-
This allows us to re-sum the series into a function
ment, γF , is straightforward using the formula (40): to
1 X∞
2k Γ k − ξ + 12 2k
 be consistent with the treatment in the liquid environ-
Σ1 (E, ξ) ≡ E = ment, we regularize the integral in (40) by restricting the
2fξ (∆)
k=1
Vβk (2k)! integration to the same domain R∆ .
(B6)
 Fig. 3 was obtained by fixing Jz = 0.6 and p varying
 
E2
2Γ 23 − ξ 1 − 1 F1 21 − ξ; 12 ; 2V 2
β
, W, J⊥ in such a way that the ratio ∆ = J⊥ / Jz2 + W 2
2fξ (∆)(2ξ − 1) remained constant and equaled ∆ = 2 × 10−2 .

which remains regular when ξ → 1/2, ∆ → 0. Therefore:


2
Appendix C: On the location of the path-freezing
µE
µ µ − 2Vβ µ (B7) transition in a liquid environment
[ωξ ] = [λ∆ (ξ)] e [1 + Σ1 (E, ξ)] ,

where In this appendix we show that the succession of phases


ξ and transitions as seen in Fig. 3 for a system with a liquid
2

2fξ (∆) J⊥ environment is generic. In particular, this means that the
λ∆ (ξ) = √ (B8) path-unfreezing transition (where µ∗ → 1) always takes
π 2Vβ
place within the rarely resonating phase, between the
accounts for the regularized singularity. With the aid of localization transition and the transition toward the non-
these formulas one obtains: resonating insulator. This follows indeed from simple
convexity arguments.
E2
 Z 
1 −µ µ We start by showing that the path-freezing transition
ΨL (µ, ξ) = − log k dEρ(E)e 2Vβ
[1 + Σ1 (E, ξ)]
µ always occurs within the rarely-resonating phase. This
− log[λ∆ (ξ)], is equivalent to stating that the non-resonating insula-
(B9) tor is always path-frozen. We show this by way of con-
tradiction: Deep in the insulator phase, we are in the
which can be evaluated numerically. non-resonating phase, dominated by a single path and
As discussed in the main text, determining the rele- µ∗ < 1. We now assume that for some parameters the
vant value of the parameter ξ requires to compute the solution of (60) reaches µ∗ = 1, while we are still in the
sign of xSP given in (51). Once (58) is used, Eq. (51) is non-resonating insulator, as defined by the condition that
equivalent to: the dominating fluctuations have a positive decay rate,
xSP > 0, which requires to set ξ = 1 in ΨL (µ, ξ). Un-

der these assumptions, using (51) together with (58) one
dE ρ(E)[ω1 (E)]µ −1 ∂z

R
ωz (E)

z=1 obtains the expression:
xSP = − , (B10)
dE ρ(E)[ω1 (E)]µ∗
R
 2 2 
dh P (h|E) Jh⊥ log Jh⊥
R R
dE ρ(E)

where explicitly (C1)
xSP = − R ,
 2  dEρ(E) ω1 (E)
∂ωz J⊥ J 2 − E 2 ∂fz (∆)
= log ω1 + √ ⊥ e 2Vβ
∂z z=1 2Vβ πVβ ∂z with ω1 (E) defined in Eq. (49). For a generic convex

z=1
2 2 X∞ function F, the following inequality holds:
J f1 (∆) − 2Vβ
E
+ ⊥ √ e α̃k E 2k
πVβ Z Z 
k=1
dh P (h|E) F(g(h)) ≥ F dh P (h|E) g(h) , (C2)
(B11)
19

which is a strict inequality for F strictly convex (provided while fixing the parameter ξ to ξ = 1, implies that
g(h) is not a constant or, equivalently, that the density the decay constant xSP as defined in Eq. (C1) becomes
P (h|E) does not collapse to a delta distribution). Ex- negative and thus is unphysical.
ploiting this for F(y) = y log(y) (which is strictly convex
for y > 0) and g(h) = |J⊥ /h|2 , we obtain Let us now turn to the localization transition, and show
that path-unfreezing (as signaled by µ∗ → 1) always pre-
− xSP cedes it within the insulator. Technically, we need to
 2 
J⊥
 2  show that, upon approaching delocalization the maxi-
dh P (h|E) Jh⊥
R R R
dE ρ(E) dh P (h|E) h log mizing µ∗ reaches 1 before γL → 0. This will imply that

> R exactly at the transition, µ∗ can no longer be chosen as
dEρ(E) ω1 (E) a maximizer of ΨL (µ, ξ) with µ∗ < 1; instead, the maxi-
mum of ΨL has to be assumed on the boundary, µ∗ = 1.
R
dE ρ(E)ω1 (E) log [ω1 (E)]
= R , Let us now show that µ∗ → 1 strictly within the insula-
dEρ(E) ω1 (E)
(C3) tor. Within the rarely-resonating phase ξ ∗ is determined
by the condition (cf. Eq. (55):
The derivative in (60) can be written as: Z Z J 2ξ∗ J 2 
∗ ⊥ ⊥
dEρ(E)ωξµ∗ −1 (E) dhP (h|E) log = 0.
dEρ(E) log[ωξ ]ωξµ h h
R !
∂ΨL (µ, ξ) 1
= ΨL (µ, ξ) − . (C7)
dEρ(E)ωξµ
R
∂µ µ The inequality (C2) applied to g(h) = |J⊥ /h|2ξ and

(C4) F(y) = y log y (for y > 0) gives:


Under our assumptions, µ∗ = 1 is a solution of Z J 2ξ∗ J 2ξ∗
∂ΨL /∂µ = 0 with ξ = 1, thus: ⊥ ⊥
dhP (h|E) log > ωξ∗ (E) log[ωξ∗ (E)],
R h h
dEρ(E) ωξ log ωξ (C8)
ΨL (1, ξ) − = 0. (C5)
which upon injection into Eq.(C7) implies:
R
dEρ(E) ωξ

ξ=1
Z
Using that the liquid decay rate satisfies γL = Ψ(µ∗ , ξ ∗ ) ∗
dEρ(E)ωξµ∗ (E) log[ωξ∗ (E)] < 0. (C9)
(with here µ∗ = ξ ∗ = 1), together with Eq. (C3) we
obtain
From the expression (C4) for the derivative ∂ΨL /∂µ
we conclude that the value of µ∗ for which γL =
− xSP > γL > 0, (C6)
ΨL (µ∗ , ξ ∗ ) = 0 can not be a stationary point of ΨL , since
(C9) implies:
since everywhere in the insulating phase γL > 0. This
is a contradiction to our starting assumption that xSP R µ∗
was non-negative. Hence, we conclude that being in the ∂ΨL (µ, ξ) 1 dEρ(E) log[ωξ∗ ]ωξ∗
∗ ∗ =− ∗ ∗ > 0.
dEρ(E)ωξµ∗
R
non-resonating insulator (xSP > 0) is incompatible with ∂µ µ ,ξ µ
reaching the path unfreezing transition (µ∗ → 1), which (C10)
can thus only occur in the rarely-resonating phase. A Therefore, at the delocalization transition µ∗ must lie at
simple extension of these arguments to an arbitrary the right boundary of the admissible interval 0 ≤ µ ≤ 1,
µ∗ 6= 1 satisfying (60) shows that the delocalization i.e., µ∗ = 1. From this it follows that the delocal-
transition (γL → 0) can only occur within the rarely- ized phase is always adjacent to a path-unfrozen, rarely-
resonating phase (ξ < 1). Namely, imposing γL = 0, resonating insulating phase.

[1] Philip W Anderson. Absence of diffusion in certain ran- anderson transition. Phys. Rev. B, 21:2366–2377, 1980.
dom lattices. Phys. Rev., 109:1492, 1958. [6] Dmitry A Abanin and Zlatko Papić. Recent progress
[2] Denis M Basko, Igor L Aleiner, and Boris L Altshuler. in many-body localization. Annalen der Physik,
Metal–insulator transition in a weakly interacting many- 529(7):1700169, 2017.
electron system with localized single-particle states. An- [7] François Huveneers. Classical and quantum systems:
nals of physics, 321(5):1126–1205, 2006. transport due to rare events. Annalen der Physik,
[3] IV Gornyi, AD Mirlin, and DG Polyakov. Interacting 529(7):1600384, 2017.
electrons in disordered wires: Anderson localization and [8] SA Parameswaran, Andrew C Potter, and Romain
low-T transport. Phys. Rev. Lett., 95:206603, 2005. Vasseur. Eigenstate phase transitions and the emergence
[4] John Z. Imbrie. On Many-Body Localization for Quan- of universal dynamics in highly excited states. Annalen
tum Spin Chains. J. Stat. Phys., 163:998–1048, apr 2016. der Physik, 529(7):1600302, 2017.
[5] L. Fleishman and P. W. Anderson. Interactions and the [9] John Z Imbrie, Valentina Ros, and Antonello Scardic-
20

chio. Local integrals of motion in many-body localized 2010.


systems. Annalen der Physik, 529(7):1600278, 2017. [28] G. Kucsko, S. Choi, J. Choi, P. C. Maurer, H. Zhou,
[10] Dmitry A. Abanin, Ehud Altman, Immanuel Bloch, and R. Landig, H. Sumiya, S. Onoda, J. Isoya, F. Jelezko,
Maksym Serbyn. Colloquium: Many-body localization, E. Demler, N. Y. Yao, and M. D. Lukin. Critical thermal-
thermalization, and entanglement. Rev. Mod. Phys., ization of a disordered dipolar spin system in diamond.
91:021001, May 2019. Phys. Rev. Lett., 121:023601, Jul 2018.
[11] Rahul Nandkishore and David A. Huse. Many-body lo- [29] IV Gornyi, AD Mirlin, DG Polyakov, and AL Burin.
calization and thermalization in quantum statistical me- Spectral diffusion and scaling of many-body delocaliza-
chanics. Annual Review of Condensed Matter Physics, tion transitions. Annalen der Physik, 529(7):1600360,
6(1):15–38, 2015. 2017.
[12] Sarang Gopalakrishnan, Markus Müller, Vedika Khe- [30] Frank Epperlein, Michael Schreiber, and Thomas Vojta.
mani, Michael Knap, Eugene Demler, and David A Huse. Quantum coulomb glass within a hartree-fock approxi-
Low-frequency conductivity in many-body localized sys- mation. Physical Review B, 56(10):5890, 1997.
tems. Physical Review B, 92(10):104202, 2015. [31] Qiming Li and Philip Phillips. Effect of quantum hopping
[13] Dmitry A Abanin, Wojciech De Roeck, and François Hu- on the coulomb gap of localized electrons in disordered
veneers. Theory of many-body localization in periodically systems. Physical Review B, 48(20):15035, 1993.
driven systems. Annals of Physics, 372:1–11, 2016. [32] G Vignale. Quantum electron glass. Physical Review B,
[14] Vedika Khemani, Rahul Nandkishore, and S. L. Sondhi. 36(15):8192, 1987.
Nonlocal adiabatic response of a localized system to local [33] M Amini, VE Kravtsov, and M Müller. Multifractal-
manipulations. Nature Physics, 11(7):560–565, Jun 2015. ity and quantum-to-classical crossover in the coulomb
[15] Thimothée Thiery, François Huveneers, Markus Müller, anomaly at the mott–anderson metal–insulator transi-
and Wojciech De Roeck. Many-body delocalization tion. New Journal of Physics, 16(1):015022, 2014.
as a quantum avalanche. Physical review letters, [34] Giulio Biroli, Claudio Chamon, and Francesco Zamponi.
121(14):140601, 2018. Theory of the superglass phase. Phys. Rev. B, 78:224306,
[16] Philipp T Dumitrescu, Anna Goremykina, Siddharth A Dec 2008.
Parameswaran, Maksym Serbyn, and Romain Vasseur. [35] Xiaoquan Yu and Markus Müller. Mean field theory of
Kosterlitz-thouless scaling at many-body localization superglasses. Physical Review B, 85(10):104205, 2012.
phase transitions. Physical Review B, 99(9):094205, 2019. [36] Giuseppe Carleo, Marco Tarzia, and Francesco Zamponi.
[17] Alan Morningstar, David A Huse, and John Z Imbrie. Bose-einstein condensation in quantum glasses. Physical
Many-body localization near the critical point. Physical review letters, 103(21):215302, 2009.
Review B, 102(12):125134, 2020. [37] Markus Müller, Philipp Strack, and Subir Sachdev.
[18] Yu. Kagan and L.A. Maksimov. Localization in a system Quantum charge glasses of itinerant fermions with
of interacting particles diffusing in a regular crystal. Sov. cavity-mediated long-range interactions. Phys. Rev. A,
Phys.JETP, 60:1, 1984. 86:023604, Aug 2012.
[19] Mauro Schiulaz, Vipin Varma, and Markus Mueller. [38] Benjamin Sacépé, Thomas Dubouchet, Claude Chape-
Quasi-localized dynamics in clean quantum systems: in- lier, Marc Sanquer, Maoz Ovadia, Dan Shahar, Mikhail
terplay of frustration and temperature. In preparation. Feigel’man, and Lev Ioffe. Localization of preformed
[20] Valentina Ros and Markus Müller. Remanent magnetiza- cooper pairs in disordered superconductors. Nature
tion: signature of many-body localization in quantum an- Physics, 7(3):239, 2011.
tiferromagnets. Physical review letters, 118(23):237202, [39] M. V. Feigel’man, L. B. Ioffe, and M. Mézard.
2017. Superconductor-insulator transition and energy localiza-
[21] Wojciech De Roeck and François Huveneers. Asymptotic tion. Phys. Rev. B, 82:184534, Nov 2010.
quantum many-body localization from thermal disorder. [40] Valentina Ros and Markus Müller. In preparation. 2021.
Communications in Mathematical Physics, 332(3):1017– [41] R Abou-Chacra, DJ Thouless, and PW Anderson. A
1082, 2014. selfconsistent theory of localization. J. Phys. C, 6:1734,
[22] Emilio Cuevas, Mikhail Feigel’Man, Lev Ioffe, and Marc 1973.
Mezard. Level statistics of disordered spin-1/2 systems [42] L. B. Ioffe and Marc Mézard. Disorder-driven quantum
and materials with localized cooper pairs. Nature com- phase transitions in superconductors and magnets. Phys.
munications, 3:1128, 2012. Rev. Lett., 105:037001, 2010.
[23] Sebastian Krinner, Tilman Esslinger, and Jean-Philippe [43] Guilhem Semerjian, Marco Tarzia, and Francesco Zam-
Brantut. Two-terminal transport measurements with poni. Exact solution of the bose-hubbard model on the
cold atoms. Journal of Physics: Condensed Matter, bethe lattice. Physical Review B, 80(1):014524, 2009.
29(34):343003, jul 2017. [44] Markus Müller. Magnetoresistance and localization in
[24] Bernard Derrida and Herbert Spohn. Polymers on disor- bosonic insulators. Europhys. Lett., 102:67008, 2013.
dered trees, spin glasses, and traveling waves. Journal of [45] Xiaoquan Yu and Markus Müller. Localization of disor-
Statistical Physics, 51(5-6):817–840, 1988. dered bosons and magnets in random fields. Annals of
[25] Bernard Derrida. Random-energy model: An exactly Physics, 337:55 – 93, 2013.
solvable model of disordered systems. Phys. Rev. B, [46] Francesca Pietracaprina, Valentina Ros, and Antonello
24:2613, 1981. Scardicchio. Forward approximation as a mean-field ap-
[26] Alan J Bray and Michael A Moore. Chaotic nature of the proximation for the anderson and many-body localization
spin-glass phase. Physical review letters, 58(1):57, 1987. transitions. Phys. Rev. B, 93:054201, Feb 2016.
[27] Giorgio Parisi and Tommaso Rizzo. Chaos in tempera- [47] Michael Aizenman and Simone Warzel. Resonant de-
ture in diluted mean-field spin-glass. Journal of Physics localization for random schrödinger operators on tree
A: Mathematical and Theoretical, 43(23):235003, may graphs. Journal of the European Mathematical Society,
21

15(4):1167–1222, 2013. Physical Review B, 96(21):214204, 2017.


[48] Michael Aizenman and Simone Warzel. Extended states [58] VE Kravtsov, BL Altshuler, and LB Ioffe. Non-ergodic
in a lifshitz tail regime for random schrödinger operators delocalized phase in anderson model on bethe lattice and
on trees. Phys. Rev. Lett., 106:136804, 2011. regular graph. Annals of Physics, 389:148–191, 2018.
[49] Gabriel Lemarié. Glassy properties of anderson localiza- [59] G. Biroli and M. Tarzia. Anomalous dynamics on the
tion: Pinning, avalanches, and chaos. Physical review ergodic side of the many-body localization transition and
letters, 122(3):030401, 2019. the glassy phase of directed polymers in random media.
[50] Giorgio Parisi, Saverio Pascazio, Francesca Pietracap- Phys. Rev. B, 102:064211, Aug 2020.
rina, Valentina Ros, and Antonello Scardicchio. Ander- [60] Felipe Monteiro, Masaki Tezuka, Alexander Altland,
son transition on the bethe lattice: an approach with David A Huse, and Tobias Micklitz. Quantum ergodicity
real energies. Journal of Physics A: Mathematical and in the many-body localization problem. arXiv preprint
Theoretical, 53(1):014003, 2019. arXiv:2012.07884, 2020.
[51] BI Shklovskii and BZ Spivak. Hopping transport in [61] Marco Tarzia. Many-body localization transition in
solids. Amsterdam: Elsevier, page 271, 1991. hilbert space. Physical Review B, 102(1):014208, 2020.
[52] SV Syzranov, Andreas Moor, and KB Efetov. Strong [62] N. Y. Yao, C. R. Laumann, S. Gopalakrishnan, M. Knap,
quantum interference in strongly disordered bosonic in- M. Müller, E. A. Demler, and M. D. Lukin. Many-
sulators. Physical review letters, 108(25):256601, 2012. body localization in dipolar systems. Phys. Rev. Lett.,
[53] G Biroli, AC Ribeiro-Teixeira, and M Tarzia. Differ- 113:243002, Dec 2014.
ence between level statistics, ergodicity and localiza- [63] Andrii O Maksymov and Alexander L Burin. Many-body
tion transitions on the bethe lattice. arXiv preprint localization in spin chains with long-range transverse in-
arXiv:1211.7334, 2012. teractions: Scaling of critical disorder with system size.
[54] A. De Luca, B. L. Altshuler, V. E. Kravtsov, and Physical Review B, 101(2):024201, 2020.
A. Scardicchio. Anderson localization on the Bethe lat- [64] Alexander L Burin. Energy delocalization in strongly
tice: Nonergodicity of extended states. Phys. Rev. Lett., disordered systems induced by the long-range many-body
113:046806, Jul 2014. interaction. arXiv preprint cond-mat/0611387, 2006.
[55] Cécile Monthus and Thomas Garel. Anderson localiza- [65] Jacob Smith, Aaron Lee, Philip Richerme, Brian Neyen-
tion on the cayley tree: multifractal statistics of the huis, Paul W Hess, Philipp Hauke, Markus Heyl, David A
transmission at criticality and off criticality. Journal of Huse, and Christopher Monroe. Many-body localization
Physics A: Mathematical and Theoretical, 44(14):145001, in a quantum simulator with programmable random dis-
2011. order. Nature Physics, 12(10):907–911, 2016.
[56] KS Tikhonov, AD Mirlin, and MA Skvortsov. Ander- [66] Louk Rademaker and Dmitry A. Abanin. Slow nonther-
son localization and ergodicity on random regular graphs. malizing dynamics in a quantum spin glass. Phys. Rev.
Physical Review B, 94(22):220203, 2016. Lett., 125:260405, Dec 2020.
[57] M Sonner, KS Tikhonov, and AD Mirlin. Multifractality [67] B Derrida. Directed polymers in a random medium.
of wave functions on a cayley tree: From root to leaves. Physica A: Statistical Mechanics and its Applications,
163(1):71–84, 1990.

You might also like