You are on page 1of 12

Entire spectrum of fully many-body localized systems using tensor networks

Thorsten B. Wahl,1 Arijeet Pal,1 and Steven H. Simon1


1
Rudolf Peierls Centre for Theoretical Physics, Oxford, 1 Keble Road, OX1 3NP, United Kingdom.
(Dated: September 7, 2016)
We propose a tensor network approximating the set of all eigenstates of a fully many-body localized
system in one dimension. Our construction, conceptually based on the ansatz introduced in Phys.
Rev. B 94, 041116(R) (2016), is built from two layers of unitary matrices which act on blocks of `
contiguous sites. We optimize the unitaries by minimizing the magnitude of the commutator of the
approximate integrals of motion and the Hamiltonian, which can be done in a local fashion. We argue
that our approach yields an exponential reduction in computational time and memory requirement
as compared to all previous approaches for obtaining the complete eigenspectrum of large many-
arXiv:1609.01552v1 [cond-mat.dis-nn] 6 Sep 2016

body localized systems. We test the accuracy of our method by comparing the approximate energy
spectrum to exact diagonalization results for the random field Heisenberg model on 12 sites. We find
that the technique is highly accurate deep in the localized regime and maintains a surprising degree
of accuracy in predicting certain local quantities even in the vicinity of the predicted dynamical
phase transition. To demonstrate the power of our technique, we study a system of 60 sites and we
are able to see clear signatures of the phase transition.

I. INTRODUCTION operator up to corrections that decay as e−x/ξ with ξ


some localization length. (We use the term quasi-local as
Many-body localization (MBL), a phenomenon con- compared to strictly-local which would mean that outside
jectured by Anderson in 1958 for disordered, interact- of some finite sized region we obtain something exactly
ing quantum particles1 , occurs in an isolated quantum proportional to the identity.)
system when it fails to reach thermal equilibrium. It A proof of the existence of qLIOMs for a strongly dis-
was shown to exist within perturbation theory for short- ordered, one-dimensional spin-1/2 model shows that this
ranged interacting models with sufficiently strong disor- characteristic of the non-ergodic phase is true at least
der for states even at a finite energy density2,3 . Strik- deep in the MBL phase29,30 . This emergent integrabil-
ingly, in one dimensional models the entire many-body ity is successful in capturing much of the phenomenology
spectrum can be localized4,5 . As opposed to a thermaliz- in 1D, developed based on exact diagonalization of small
ing system where the eigenstates exhibit volume law en- systems. But in the absence of numerics on sufficiently
tanglement and satisfy the eigenstate-thermalization hy- large systems or a mathematical proof, its generalization
pothesis (ETH)6,7 , for a one-dimensional system exhibit- to weaker disorder or higher dimensions remains under
ing full many-body localization (FMBL) in the strongly intense investigation31,32 .
disordered regime, the eigenstates of the entire spectrum In the FMBL phase, the entire spectrum of the
are expected to obey an area law8,9 . many-body Hamiltonian can be described in terms of
The breakdown of thermalization lends itself to several the quantum numbers of the qLIOMs. A consequence
interesting phenomena which are absent in a thermaliz- of this is that all of the eigenstates obey an area-
ing system10 . Topological and symmetry breaking orders law of entanglement8,9 , which then allows the use of
which are destroyed by thermal fluctuations at equilib- highly efficient approximations such as tensor network
rium can be extended to highly excited states at a finite techniques33–38 . The essence of these techniques is the
energy density due to MBL11–15 . Logarithmic growth realization that states with area-law entanglement can
of entanglement in the FMBL phase may allow the con- be described numerically using exponentially fewer pa-
struction of logical qubits in an interacting system which rameters than are required to describe an arbitrary state
can serve as robust quantum memories16,17 . Even the in Hilbert space. Such techniques are impressively (and
quantum phase transition between the thermal and MBL provably) efficient for describing gapped ground states
phases is not described by any of the conventional the- in one dimension34,38 . Increasingly, similar techniques
ories of phase transition18–21 . Recent developments in have been computationally effective in two dimensions as
cold atoms and various forms of synthetic quantum mat- well37,39 . What is special about FMBL systems is that
ter have allowed the experimental study of the phenom- the area law holds not only for the ground state, but for
ena of thermalization and its breakdown in a controlled the entire spectrum.
manner22–24 . Exploiting this area-law entanglement it was shown
Many of the features of FMBL can be understood in that excited eigenstates of FMBL systems can be ap-
terms of an extensive set of emergent quasi-local, exact proximated efficiently as matrix product states40 . Fur-
integrals of motion (qLIOM)25–28 . The phrase quasi-local thermore, the unitary operator diagonalizing the entire
here indicates that if we trace over the operator within Hamiltonian can be represented as a tensor network,
a region of size x around where the operator is localized, known as a spectral tensor network41 . The algorithm
we should obtain something proportional to the identity to construct such a spectral tensor network proposed by
2

Pekker and Clark does not scale efficiently with system the total number of spins and its performance close to
size42 . A proposal with an efficient scaling was given the MBL-thermal transition is also discussed in this sec-
by Pollmann et al. using stacked layers of unitaries by tion. In Sec. VI we present a summary of the results and
minimizing the fluctuations in the total energy43 . It was future directions for the method.
suggested that the accuracy of the approximation for a
given chain length can be increased by increasing the
number of layers. Compared to the methods targeting II. MODEL AND ITS PHENOMENOLOGY
eigenstates within an energy window44–46 , this procedure
is constructed to efficiently approximate all the energies We consider the canonical random-field Heisenberg
and eigenstates with sufficient accuracy, providing access model defined on a spin-1/2 chain4 of N sites with open
to dynamical properties of local observables. boundary conditions,
In this work we improve upon the ansatz from Ref. [43] N −1
by increasing the size of the block of spins acted upon by
X
H= (JSi · Si+1 + hi Siz ) + SN
z z
hN (1)
the unitaries, while keeping the number of layers fixed i=1
at two. We show numerically that this gives rise to an
exponential improvement of the computational time and with Si = 12 σi and each of the hzi is chosen from the
memory requirements. We use a figure of merit which uniform distribution bounded between [−W, W ], where
is directly related to the qLIOMs and motivated by a W is called the disorder strength. The model is known
procedure introduced by Kim et al. to identify slow op- to have a dynamical phase transition into the MBL phase
erators in disorder-free non-integrable models47 . For con- where all states are localized for disorder strength greater
creteness, we consider the one dimensional random field than Wc ≈ 3.54,49 .
Heisenberg model. We compare our scheme to that orig- In the FMBL regime the bare physical spins in the
inally proposed by Ref. [43] even extending their scheme model (also known as ‘p-bits’) can be be unitarily trans-
to four layers (albeit with a different figure of merit to im- formed into an extensive set of mutually-commuting
prove computational efficiency). We find that our strat- quasi-local effective spins τiz (also known as ‘l-bits’)
egy to be both more accurate and computationally effi- which are expected to commute exactly with the Hamil-
cient than that of Ref. [43]. tonian.
We quantify the performance of our scheme by mini- [H, τiz ] = [τiz , τjz ] = 0, (2)
mizing the commutator of the Hamiltonian with the ap-
proximate, local integrals of motion defined through our where τiz = U σiz U † . U is the unitary operator which
tensor network ansatz. As we show, this figure of merit exactly diagonalizes the Hamiltonian. In the localized
decomposes into strictly local parts, which allows us to phase, the unitary transformation U can be decomposed
evaluate it with linear cost in the system size, thus en- into a sequence of local unitaries so that the l-bits develop
abling us to reliably assess the performance of our ansatz exponentially decaying tails away from site i29 . More
in the regime where exact diagonalization is unavailable. mathematically,
We corroborate this by comparing the tensors obtained
by optimizing our figure of merit with exact diagonaliza- kTri−r,i−r+1,...,i+r (τiz − σiz )k1 ≤ a e−r/ξL (3)
tion results for 12 sites, where we observe that the numer- with positive constants a, ξL for N  r. We use the
ical value of the figure of merit indeed reflects how well
P
1-norm kAk1 = jk |Ajk | and consider the matrix rep-
the real MBL energy spectrum is approximated. We find resentation of τiz − σiz in a fixed basis. ξL can be defined
a very high accuracy of our ansatz for unitaries acting to be the localization length of the MBL system where
on six contiguous sites, and thus use the same procedure the trace is taken over the collection of spins within a
to tackle a chain with 60 sites as a function of the dis- distance r of site i. It is important to note that the defi-
order strength. Remarkably, the ansatz fares extremely nition of the localization length is not unique. There can
well for local observables at weak disorder and close to even be multiple localization lengths and some of them
the MBL-to-thermal phase transition in this model. We may not diverge at the MBL-thermal transition25 .
use the fluctuations in the half-cut entanglement entropy According to Eq. (2), the Hamiltonian and the set of l-
calculated with this ansatz to estimate the location of bits {τjz } can be simultaneously diagonalized where every
the transition48 which is in agreement with the exact di- eigenstate is a product state in the l-bit basis. Each
agonalization studies. eigenstate can be uniquely labelled by the eigenvalues
In Sec. II we define the model used to perform our ij = ±1 of the set of l-bit operators {τjz } (j = 1, . . . , N ),
calculations and also highlight the phenomenological fea- |ψi1 i2 ...iN i. In the l-bit basis the Hamiltonian can be
tures of the FMBL phase in one dimension. In Sec. III expressed in the following form,
and IV, we give a detailed description of the tensor net- N N N
work ansatz and the figure of merit used to diagonalize H=
X
Ji τiz +
X
Jij τiz τjz +
X
Jijk τiz τjz τkz
the full Hamiltonian efficiently. The numerical results i=1 i,j=1 i,j,k=1
and their comparison to exact digonalization are pre-
sented in Section V. The scaling of the procedure with + ..., (4)
3

is expected to decrease as the inverse of a polynomial


function in computational cost.
In Ref. 43, the best tensor network approximation is
found by minimizing the sum of the energy variances of
all approximate eigenstates |ψ̃i1 ...iN i. The computational
cost for the calculation of this quantity scales as 25n ,
where n is the number of layers. However, it appears
that n needs to grow exponentially with the localization
length ξL in order to keep the accuracy of the approxima-
tion fixed on average: while there is a notation of locality
of Ũ on the scale of the localization length, there may
FIG. 1. Tensor network Ũ as proposed in Ref. 43 with n layers not be any way to parameterize ux,n with the number of
of 4 × 4 unitaries {ux,y }. Since the unitary Ũ is supposed to parameters scaling subexponentially with ξL , especially
approximately diagonalize the Hamiltonian, the correspond- close to the transition into the thermal phase. Hence,
ing approximate eigenstates |ψ̃i1 ,...,iN i are the states obtained the number of real parameters required to describe the
by fixing the lower open indices in the figure to be i1 , . . . , iN . unitary within that range is expected to typically scale as
22ξL . Therefore, in order to reproduce local observables
with a given accuracy, of the order of 22ξL N parameters
where the coefficients Jijk... typically decay exponentially are required. Since the number of parameters of the ten-
with the largest distance between two spins |i − j| occur- sor network in Fig. 1 is only 42 N n2 = 8N n, n is required
ring in a particular cluster in the expansion. The proba- to grow as 22ξL . It follows that the computational re-
bility of a coefficient being substantially larger than the sources need to grow superexponentially with ξL to keep
typical value expected from this exponential decay, is also the accuracy fixed. This makes it hard to approach the
exponentially small29 . transition to the delocalized phase using the multi-layer
ansatz.
We propose to overcome this problem by increasing the
III. TENSOR NETWORK ANSATZ range of sites acted on by the building block unitaries,
instead of varying the number of layers. Thus, we stick
Tensor network states (TNS) are believed to provide to two layers of unitaries with ` lower and upper “legs”
an efficient representation of the ground states of local (` is even) contracted as shown in Fig. 2. Each unitary
gapped Hamiltonians. That is, as the system size is in- has 22` parameters, i.e., the total number of parameters
creased, the number of parameters required to approxi- is 22`+1 N/`. Hence, ` needs to grow only linearly with
mate the ground state wave function with a certain fi- ξL in order to keep the accuracy fixed. The contraction
delity (e.g., with at least 99 % overlap) increases only cost of the tensor network arising in the variational op-
polynomially with the system size. For FMBL systems timization of its unitaries scales only exponentially in `
with sufficiently strong disorder (i.e., in the absence of (as discussed in the following section), which is an ex-
a mobility edge), the whole spectrum of eigenstates ful- ponential improvement over the the multi-layer ansatz.
fills the area law and thus, can be efficiently represented Moreover, for a fixed localization length ξL , we anticipate
by MPS40 . However, since the number of eigenstates the error of our approximation to decrease as exp(−`/ξL )
is exponential in the system size, for large N one can due to Eq. (3), which decays much faster than the multi-
only tackle the eigenstates in a certain energy window layer ansatz does with n, allowing us to describe eigen-
using MPS (see e.g. Refs. [44–46]). On the other hand, states more accurately closer to the MBL transition. As
spectral tensor networks are meant to encode an approx- the computational cost is exponential in `, this corre-
imation to all eigenstates at once, which is a desirable sponds to a decrease in the error of local observables as
property if one aims to calculate dynamical properties the inverse of a polynomial function in computational
of local observables in MBL systems. We build on the resources, which is typical for ground states using TNS.
tensor network ansatz proposed in Ref. [43]. It defines a Finally, note that in both the ansatz of Ref. [43] and in
unitary matrix Ũ , which approximately diagonalizes the our approach, the number of parameters required to cal-
Hamiltonian, in terms of many 4 × 4 unitaries, which culate local observables with a given accuracy, increases
are stacked in several layers and contracted as shown in linearly with N in the localized phase.
Fig. 1.
In the following, we argue that for this tensor network
the approximation of local observables with a given accu- IV. FIGURE OF MERIT
racy requires the computational resources to grow super-
exponentially with the localization length. In contrast, In order to find the unitary Ũ as described by our
for the tensor network we will suggest they scale only ex- tensor network which is as close as possible to the unitary
ponentially with the localization length. In addition, for U exactly diagonalizing the MBL Hamiltonian, we define
a fixed localization length, the error of local observables a figure of merit which reflects the deviation between the
4

FIG. 2. Construction of the unitary Ũ in terms of unitaries


ux,1 , ux,2 acting on ` sites (in this example ` = 4). Again, the
approximate eigenstates |ψ̃i1 ,...,iN i are the states obtained by
fixing the lower open indices in the figure to be i1 , . . . , iN .

two. This can be achieved by defining the approximate


l-bits corresponding to Ũ , τ̃iz = Ũ σiz Ũ † . If they were the
exact l-bits, they would commute with the Hamiltonian
and with each other. The latter property is fulfilled by
construction, so we define the error in our approximation
as the sum of the (squared) trace norms of the individual
commutator of τ̃iz with the Hamiltonian,

N FIG. 3. Sum of
1X Ptensor network contractions which yields the
tr [H, τ̃iz ][H, τ̃iz ]†
 N z 2
f ({ux,y }) := second term, i=1 tr (H τ̃i ) , in Eq. (5). The multiplica-
2 i=1 tions from left to right in Eq. (6) correspond to top to bottom
N
X in the figure. The indices of the lower wiggly lines are to be
tr(H 2 ) − tr (H τ̃iz )2 contracted with those of the corresponding upper wiggly lines.

= . (5)
For a given position i of the σ z operator and arbitrary posi-
i=1
tions j, k of the two-body Hamiltonian terms, all unitaries of
the lower layer (i.e., un,1 ) apart from the ones directly con-
In the following, we call f the sum of the commutator
nected to the σ z operators cancel with their adjoints and can
norms (SCN), which will be our figure of merit. In order be replaced by identities (i.e., straight vertical lines). In the
to minimize f , we evaluate the right hand side of Eq. (5), example in the figure this corresponds to all unitaries un,1 for
which may naively appear exponentially hard in the num- n 6= x. Furthermore, all unitaries of the second layer (un,2 )
ber of sites N to evaluate. However, it is possible to break which are not directly connected to the remaining ones of the
it down into a sum of local terms, rendering the compu- first layer cancel and can be substituted by identities. This
tational complexity linear in the system size. To that implies that the x-th summand in Eq. (6) depends only on the
end, we first express the Hamiltonian as a sum of terms unitaries ux+1,1 , ux,2 , ux+1,2 . The contraction corresponding
PN
hi acting on two neighboring sites i, i + 1, H = i=1 hi . to this term is shown in Fig. 4.
PN z 2

Then, the last term of Eq. (5), i=1 tr (H τ̃i ) , can
be easily written as a sum of tensor networks, see Fig. 3.
V. NUMERICAL RESULTS
This term can be further decomposed into local parts as
depicted in Fig. 4 (using τ̃iz = Ũ σiz Ũ † )
A. Optimization method
N N
X X −1  
f ({ux,y }) = N tr(H 2 ) − tr hj Ũ σiz Ũ † hk Ũ σiz Ũ † In the following section, we will approximate the eigen-
i=1 j,k=1 states of the Hamiltonian defined in (1). The model
N/` possesses U (1) symmetry (it conserves the total spin-z
PN
component), [H, i=1 Siz ] = 0. Furthermore, the Hamil-
X
= const. − fx (ux,1 , ux−1,2 , ux,2 ). (6)
x=1 tonian is time-reversal symmetric and thus real. In con-
ventional TNS, symmetries of the model can be imposed
fx (ux,1 , ux−1,2 , ux,2 ) itself is a sum of tensor networks on the individual tensors50,51 : Any TNS that is invariant
which only depend on ux,1 , ux−1,2 , and ux,2 , the Hamil- under a symmetry can be written as a (possibly differ-
tonian terms hj and the σiz operators which are connected ent) TNS, where all its individual tensors form a projec-
to those unitaries. The most expensive contraction is the tive representation of the corresponding symmetry group,
one where both Hamiltonian terms connect the two uni- that is, they are invariant up to a phase under the ac-
taries ux−1,2 , ux,2 , giving rise to a matrix multiplication tion of the symmetry. In doing so, the dimensions of the
of two 22` × 22` matrices, i.e., the overall computational tensor indices might have to be increased by some fac-
cost is of order 26` . tor that is independent of the system size. The cost of
5

ing upper and lower legs together into one single index
each. Each of these blocks, say uB , can be parameterized
by an antisymmetric real matrix AB , uB = eAB , making
the unitaries real, that is, orthogonal matrices.
In order to carry out the optimization, we pick initial
values for the antisymmetric matrices AB parameterizing
the unitaries and optimize the unitaries individually by
sweeping from the left end of the chain to the right and
back, until convergence is achieved. Crucially, each such
minimization step requires only the evaluation of a few
terms in the sum of Eq. (6). As it turns out, better
results are achieved by always optimizing two connected
unitaries at once.
We use a quasi-Newtonian routine supplied with the
gradient with respect to the parameters contained in the
matrices AB . This gradient comes almost for free in the
contraction of the tensor network of Fig. 4 if one contracts
its rows in the right order, as explained in more detail in
Appendix A.
As it turns out, the final SCN figure of merit depends
on the choice of the initial unitaries. Hence, its potential
landscape is too “bumpy” for the quasi-Newtonian algo-
rithm to find its global minimum. In some cases, where
the number of parameters is small enough we can over-
FIG. 4. Decomposition of the figure of merit (5) into local come this problem by carrying out optimizations for the
terms resulting in Eq. (6). Again, the indices of the lower wig- same disorder configuration with many random choices
gly lines are to be contracted with those of the corresponding for the initial unitaries. Additionally, at each step of the
upper wiggly lines. The shown tensor network is obtained af- sweeping process, we initialized the unitaries that are cur-
ter replacing mutually cancelling unitaries in Fig. 3 by identi- rently being optimized with at least 15 random choices,
ties (vertical lines). Those forming closed loops yield a factor and optimized over the latest set of unitaries combined
of 2 each, which results in the prefactor 2−N +2`+2 of fx shown with the ones obtained in the previous part of the sweep,
in the figure. Terms j, k where hj or hk are not connected to and proceeded with the best of them. Sampling the uni-
ux−1,2 or ux,2 yield contributions which are independent of taries in this manner allows the procedure to escape local
all unitaries ux,y and can thus be neglected in the local def-
minima and increases the likelihood of finding the global
inition of our figure of merit. Note that if hj or hk is not
connected to both ux−1,2 and ux,2 (central position), the uni-
minimum. If after optimizing over sufficiently many ini-
tary it is not connected to, contracts with its adjoint to yield tial random choices of unitaries, the best value of the
another identity, which gives some additional computational SCN figure of merit was obtained repeatedly with high
advantage. accuracy, we assumed it to be (close to) the global mini-
mum value.
We take advantage of this approach for ` = 2, where
variational optimization of TNS usually reduces tremen- U (1) and time reversal symmetry imply that the uni-
dously by imposing such symmetries on the tensors, as taries ux,y have only one non-trivial block with a sin-
they become sparse and have fewer variational parame- gle variational parameter. For larger `, each individual
ters. We implement a similar procedure for evaluating minimization starting from a given choice of the initial
the entire spectrum; time reversal symmetry (i.e., that unitaries was too expensive to perform a global minimiza-
the whole tensor network is real) can be imposed by tak- tion. In that case, we obtained best results by initializing
ing all tensors to be real. To ensure that the total spin-z the ` = 4 unitaries, such that they correspond to the best
component is conserved, each individual tensor ux,y is ` = 2 unitaries we obtained previously for the same dis-
assumed to leave the total spin-z of the block invariant, order configuration. This is straightforwardly achieved
P`
i.e. [ux,y , m=1 szm ] = 0, where szm is defined in the by defining
same Hilbert space as ux,y . Graphically speaking, this
u`=4 `=2
 `=2
u2x−1,1 ⊗ u`=2

means that the sum of the spin-z components on the x,1 = 1 ⊗ u2x−1,2 ⊗ 1 2x,1 , (7)
lower legs of each tensor has to equal the sum of the u`=4 `=2
x,2 = 1 ⊗ u2x,2 ⊗ 1, (8)
spin-z components of the upper legs (remember that all
indices have dimension two, corresponding to spin-1/2 where 1 = ( 10 01 ). Note that the obtained ` = 4 unitaries
particles). All tensor entries whose indices do not fulfill are also real and invariant under U (1) symmetry. Un-
this requirement are forced to be zero. This leads to a fortunately, for ` = 6 it is not possible to initialize the
block structure of the matrix which is obtained by group- unitaries with the best ` = 4 result (as no such blocking
6

` = 4 to ` = 6 is less than from ` = 2 to ` = 4 for


this particular disorder configuration, as opposed to the
disorder-averaged behavior, see section V C. We also cal-
culated the mean variance of the Hamiltonian (the figure
of merit used in Ref. 43),
1 X
∆H 2 = N hψi1 ...iN |H 2 |ψi1 ...iN i
2 i ...i
1 N

−hψi1 ...iN |H|ψi1 ...iN i2 ,



(9)

which was ∆H`=2 2 = 0.0771, ∆H`=4 2 = 5.30 · 10−3 and


∆H`=62 = 9.39 · 10−4 , decaying in a very similar way as
the SCN.
We find that the SCN reflects reliably the accuracy of
our approximate method and thus, captures the extent
to which the Hamiltonian is diagonalized by the optimal
FIG. 5. Comparison of the optimized tensor network Ũ for unitary matrix Ũ . Therefore, for larger systems where
` = 2 (a), ` = 4 (b) and ` = 6 (c) with exact diagonalization exact diagonalization is unavailable, we can use the SCN
for N = 12 and a single disorder realization at W = 6. The in order to assess the quality of the approximation by our
energy differences ∆E were obtained by ordering the diago- tensor network.
nal elements of Ũ † H Ũ and subtracting them from the ordered
As a further corroboration, we computed the overlaps
exact eigenvalues of the Hamiltonian H. The optimized SCN
figure of merit was f`=2 /2N = 0.2560, f`=4 /2N = 0.0240,
between the exact eigenstates and the states defined by
f`=6 /2N = 0.0067. The mean level spacing was 0.0095, indi- the columns of Ũ for ` = 6. The overlaps are in gen-
cated by dashed lines. eral very high (more than 97 % of them have more than
95 % overlap), see Fig. 6(a). If one orders the eigenstates
according to their energies and the states defined by Ũ ac-
is possible from ` = 4 to ` = 6). In this case, we initial- cording to the diagonal elements of Ũ † H Ũ , most of them
ized the unitaries with the optimal ` = 2 unitaries for a correspond to each other. Mismatches typically occur
given disorder configuration. This corresponds simply to because the separation between two exact energies is of
choosing u`=6 `=2 `=2 `=2 the order of the errors of their approximate values, such
x,y = u3x−3+y,y ⊗ u3x−2+y,y ⊗ u3x−1+y,y .
that they get swapped in the ordering of the columns of
Ũ . To show that the local properties of the eigenstates
also match to high degree of accuracy, we compare the
B. Comparison to exact diagonalization
distribution over all sites of the expectation value of σiz
evaluated in all the eigenstates. In Fig. 6(b) the dis-
In order to demonstrate the precision of our method tributions resulting from exact diagonalization and the
for efficiently diagonalizing the Hamiltonian in the spectral tensor network overlap to a remarkable preci-
FMBL regime, we performed the optimization defined sion, showing that the method has indeed converged to
in Sec. V A for a system of size N = 12 with disorder the eigenstates with the appropriate local features.
strength W = 6 and one generic disorder realization us- For comparison, we also optimized the unitaries using
ing unitaries with ` = 2, 4, and 6 legs. We compare our the ansatz in Fig. 1 with four layers (and l = 2). This cor-
results to the energies and the eigenstates of the Hamil- responds to the scheme proposed in Ref. [43], extending
tonian obtained using exact diagonalization. The results their explicit numerical study of a network of two layers
are shown in Figs. 5 and 6. to four layers (and also using our figure of merit rather
In Fig. 5 the distribution of the differences between the than theirs). The results are shown in Fig. 7: The U (1)
(ordered) diagonal elements of the matrix Ũ † H Ũ and the and time reversal symmetries again imply that there is
exact energies (defined to be ∆E) are plotted for the cho- only one variational parameter per unitary. If we opti-
sen values of `. The distribution narrows tremendously mize them using a global optimization as described at
with increasing ` with a sharp peak at ∆E = 0. For ` = 4 the end of section V A, we find f /2N = 0.2372 and thus
and 6, the bulk of the distribution is well within the mean hardly any improvement as compared to the two-layer
level spacing of the system (shown with dashed vertical case.
lines in Fig. 5), showing that the energies of the Hamil- We also carried out such an optimiziation without
tonian evaluated using this method have an error which imposing any restrictions on the unitaries, i.e., they
is exponentially small in system size. The optimized val- are parameterized by a Hermitian 4 × 4 matrix Hx,y ,
ues of the SCN figure of merit were f`=2 /2N = 0.2560, ux,y = eiHx,y with 16 variational parameters per unitary.
f`=4 /2N = 0.0240, f`=6 /2N = 0.0067, showing a simi- In that case, the result f 0 /2N = 0.0645 is a significant
lar decay with `. (For an explanation of the normaliza- improvement over the two-layer case. (We checked that
tion factor 2−N , see subsection V C.) The decrease from the improvement is almost as large if one does not impose
7

FIG. 6. Comparison of the optimized tensor network Ũ for FIG. 7. Comparison of the optimized tensor network Ũ using
` = 6 with exact diagonalization for N = 12 and a single dis- four layers in the ansatz of Fig. 1 with exact diagonalization
order realization at W = 6. The approximate eigenstates are for N = 12 and the same disorder realization as in Fig. 5
given by the columns of Ũ and the energies are given by the (i.e., W = 6). If one imposes U (1) symmetry (a), we ob-
diagonal elements of Ũ † H Ũ . The exact diagonalization and tain an optimized SCN figure of merit of f /2N = 0.2372 and
tensor network results are compared by ordering the energies thus hardly any improvement as compared to the two-layer
and the diagonal elements, respectively. (a) Distribution of ansatz, which gives 0.2560. For a full parameterization of the
the overlap of the exact and the matched eigenstates. (b) Dis- unitaries (b), we obtain f /2N = 0.0645, which is a signifi-
tribution of the expectation value of σiz over the sites and the cant improvement, but still far behind our result using four
eigenstates obtained from exact diagonalization (light brown) legs per unitary as well as being computationally much more
and the TNS (blue). expensive. The mean level spacing was 0.0095, indicated by
dashed lines.

the symmetries on the tensors of the two-layer scenario.)


However, the improvement when the unitaries are chosen
without any reference to the global conservation law is
still significantly less accurate than the multi-leg ansatz
for four layers. The global optimization as described at
the end of Sec. V A takes two orders of magnitude longer
and requires many calculations to be performed in par-
allel for each disorder realization. Despite our efforts,
we did not succeed in making the multi-layer l = 2 net-
work calculation as accurate or computationally efficient
as our multi-leg ansatz. Thus, the numerical results cor-
roborate that the FMBL system cannot be approximated
as efficiently and accurately by increasing the number of
layers compared to increasing the number of legs per uni-
tary. We also note that the relaxation of the U (1) and
time reversal symmetries for four layers results in a sig-
nificant improvement, as the additional parameters can
partially compensate for the lack of parameters that we
conjectured. FIG. 8. Scaling of the SCN figure of merit as a function of
the system size N for ten different disorder realizations with
W = 6 optimized by unitaries of block sizes ` = 2, 4, 6. For
given N , the same ten disorder configurations were taken for
C. Scaling with the system size all values of `. As discussed in the main text, we expect
f /2N ∝ N , which is consistent with the numerical result.
One of the primary objectives of this work is to es-
tablish our ansatz for the description of fully many-body
localized systems. For a given point in the MBL phase, SCN averaged over many disorder configurations should
increasing the system size does not require an increase grow as N 2N . We corroborated this by optimizing our
in ` to approximate the local properties of eigenstates tensor network ansatz for ` = 2, 4, and 6, and system
with a constant accuracy (averaged over disorder real- sizes in the range between N = 12 and 60 for ten disor-
izations), due to the localized nature of the eigenstates. der realizations, as shown in Fig. 8. We gather that on an
Therefore, the SCN should detect a constant mismatch average f /2N indeed increases linearly with system size
per lattice site and thus, increase linearly with the sys- for all choices of `. We also notice that for a given system
tem size. However, recall that it is defined as a trace size, the SCN decreases by almost an order of magnitude
of an operator in the 2N -dimensional Hilbert space, i.e. as ` is increased by 2: From ` = 2 to ` = 4 it decreases
a mismatch that is not affected by the sites far away is by a factor of 5 on average, and from ` = 4 to ` = 6 we
multiplied by the trace over the identity operator corre- again have an average improvement by a factor slightly
sponding to them, which grows as 2N . As a result, the larger than 4.
8

FIG. 10. Comparison of the eigenstates from the optimized


tensor network Ũ for ` = 6 and exact eigenstates from exact
diagonalization for N = 12 and a single disorder realization
FIG. 9. Figure of merit, SCN, for N = 60 as a function at disorder strengths (a) W = 2, (b) W = 3 and (c) W = 4.
of the disorder strength W for ` = 2, 4, 6. The same ten Distribution of the expectation value of σiz over the sites and
disorder configurations were taken for all ` and all choices of eigenstates eigenstates from exact diagonalization (blue) and
W , adjusting the overall prefactor of the random magnetic TNS (light brown).
fields.

it amenable to the study of eigenstates of large systems


The results in subsection V B show that the MBL
with high accuracy. As we approach the transition into
eigenstates are well-represented by the optimized tensor
the thermal phase at weaker disorder, the eigenstates be-
network, i.e., by minimizing the SCN we obtain an over-
come more entangled. The ansatz with a larger number
all unitary matrix Ũ that approximately diagonalizes the
of legs is able to capture the regions of high local en-
Hamiltonian to a high accuracy for ` = 6. The linear de-
tanglement, allowing the method to perform appreciably
pendence on N of the optimized SCN (divided by the
well even close to the phase transition. In Fig. 10 the
dimension of the Hilbert space (2N )) suggests that, in
distributions of the local observable σiz evaluated in all
the localized region, expectation values of local observ-
the eigenstates of N = 12 system for a single disorder
ables in any eigenstate can be approximated with an error
realization, at disorder strengths W = 2, 3, and 4 are
that depends only on `. Hence, our method is able to ap-
shown. The optimized SCN figure of merit f`=6 /2N was
proximate local properties of eigenstates for large system
0.0453, 0.0274, and 0.0177, respectively.
sizes, where exact diagonalization is not available.
The distribution evaluated using the approximate
The disorder dependence of the figure of merit (scaled eigenstates from the tensor network ansatz matches re-
by 2N ) is shown in Fig. 9 for the different values of ` at markably well with the exact diagonalization results in
N = 60. Deep in the localized phase the relevant quan- the vicinity of the MBL transition. The comparison with
tity decays almost by an order of magnitude each time ` the data from exact diagonalization is good even at dis-
is increased by two. As the disorder strength decreases, order strength W = 2 which is expected to be on the
the rate of decay with ` of the figure of merit slows down, ‘thermal’ side of the phase transition. However, we can-
where a larger ` is required to maintain the same level not expect this to be the case if N is increased, as opposed
of accuracy. Besides the fact that the approximation be- to the localized phase, where local observables can be re-
comes worse as one approaches the transition, the quan- produced with a constant accuracy for fixed ` (cf. Fig.
tity under consideration does not show any signature of 8). Instead, ` would need to be scaled with N to keep
the phase transition. In the following subsection, we in- the accuracy fixed43 . In this regime the eigenstates from
vestigate in more detail the accuracy of the eigenstates the tensor network ansatz have larger weight in the dis-
in the weakly disordered regime and the effects of the tribution at hσiz i ≈ ±1, which suggests that the ansatz
approaching phase transition into the thermal phase. does not fully capture the local features of the eigen-
states. The finite number of legs in the local unitaries of
our tensor network ansatz enforces the qLIOMs to be al-
D. Approaching the many-body localization ways approximately conserved, but strictly local. Thus,
transition our TN ansatz cannot resolve whether there are exactly
conserved qLIOMs in the vicinity of the phase transition.
Deep in the FMBL phase, the tensor network structure We finally turn to an extremely sensitive test of how
of the eigenstates due to the existence of qLIOMs makes well this approximate method reproduces subtle details of
9

FIG. 11. Standard deviation of the half-cut entanglement


entropy (σS ) averaged over the eigenstates as a function of FIG. 12. Standard deviation of the eigenstate-averaged entan-
disorder strength for the TNS with ` = 4 (blue), ` = 6 (red), glement entropy, followed by an average over entanglement
and exact diagonalization (black). The system size is N = 12. cuts, hσS icuts (defined in the text), as a function of disor-
The inset shows the average entanglement entropy (S̄) for the der strength. The system size is N = 60 and the calculations
three cases. were performed on 10 disorder realizations using the TNS with
` = 2 (green), ` = 4 (blue) and ` = 6 (red). The inset shows
the corresponding eigenstate-averaged entanglement entropy,
the physics of the MBL system. We evaluate the fluctu- followed by an average over entanglement cuts (hS̄icuts ).
ation in the half-cut entanglement entropy σS (the stan-
dard deviation of the distribution of entanglement en-
tropy) of the eigenstates for ` = 4, 6 and N = 12 for In Fig. 12 we show the statistical mean and stan-
the TNS and exact diagonalization over a wide range of dard deviation over 10 disorder configurations of the
disorder strengths. The quantity was evaluated using eigenstate-averaged entanglement entropy as a function
10 disorder realizations. In exact diagonalization stud- of W . The curves of hσS icuts and hS̄icuts shown are ob-
ies, this quantity has a peak at the MBL-ETH transition tained after averaging over different entanglement cuts to
which is expected to diverge with system size. Although improve smoothness. In the 2-layer TN ansatz, the eigen-
our ansatz cannot represent any volume-law entangled states can be entangled over a distance only one tensor
states, it is expected to capture the entanglement struc- block away from the cut. Therefore, the maximum num-
ture at length scales of order `. In Fig. 11 we indeed see ber of p-bits contributing to the entanglement does not
a broad peak close to the value of disorder where exact change with the position of the entanglement cut. The
diagonalization gives a relatively narrow peak. As ex- positions for the respective entanglement cuts have been
pected at strong disorder the exact diagonalization and chosen such that they are at least one tensor block away
the TNS are tending towards the same value. from the boundaries of the system. We observe a maxi-
We also calculated the entanglement entropy of eigen- mum at around W = 2.5 for ` = 2, which moves towards
states using the TN approximation for system size N = W = 3 for ` = 4 and ` = 6. The standard deviation for
60 as a function of disorder strength W . For a given N = 60 decays much faster for large W than the N = 12
disorder realization, the computational cost to calculate result. This must therefore also be true for the entropy
such an entropy is independent of N and only depends of the exact eigenstates, as in that regime our TNS ap-
on `. In order to exploit this feature, one has to use the proximation is expected to be very accurate for ` = 6.
fact that the partial trace of |ψ̃i1 ...iN ihψ̃i1 ...iN | gives rise On averaging over entanglement cuts combined with the
to a reduced density matrix whose non-zero eigenvalues increased decay of σS at larger disorder makes the peaks
are the same as the ones contained in the boundary re- much more pronounced than the N = 12 case.
duced density matrix, which is obtained by tracing out In the insets of Figs. 11 and 12, the average entangle-
the physical spins and defining a state on the “virtual ment entropy (S̄) increases as the disorder strength goes
bonds” between the tensors which have been truncated down but for N = 12 this increase is still slower compared
by the entanglement cut52 . Those eigenvalues depend to the exact eigenstates. In the quantum critical regime,
only on the configuration i1 , . . . , iN within a distance of there are suggestions that the transition is driven by a
maximally 3`/2 from the entanglement cut, which makes subcluster of spins which are weakly entangled53 . With
it possible to average over all eigenstates efficiently, too. increasing system size, on the thermal side of the tran-
10

sition the size and the entanglement of the subcluster location of the peak is at a weaker disorder than the es-
grows with N , while on the localized side of the quan- timates from exact diagonalization, it is drifting in the
tum critical regime, the entanglement remains small. By direction of increasing W with increasing `.
varying ` and N in our TN ansatz, it may be feasible to The accurate construction of all eigenstates in the
access this regime numerically which is a question suit- FMBL phase and in the vicinity of the MBL-ETH transi-
able for future work. tion for large system sizes opens the door to study several
fascinating phenomena associated with the subject. As a
by-product of the procedure, using our optimized unitary
VI. CONCLUSIONS AND OUTLOOK one directly obtains the approximate qLIOM operators
in the localized phase. The ability to vary ` and study
In this work we have made several significant advances eigenstates in the vicinity of the MBL-ETH transition
in efficiently approximating the entire spectrum of fully suggests that our procedure may be able to capture
many-body localized systems. Besides improving upon some of the scaling properties on the localized side of
the tensor network ansatz proposed in Ref. [43], we also the quantum critical regime of the transition. Given the
optimize the network by minimizing a different figure of efficiency of the method, it may be feasible to scale the
merit (the SCN) given by the magnitude of the commu- procedure to numerically address the question of many-
tator of the Hamiltonian and the approximate qLIOMs body localization in two dimensions. Since, MBL of
produced by the tensor network ansatz. This figure of Floquet systems have a structure similar to that of static
merit can be evaluated by decomposing into strictly lo- Hamiltonians, our method can be generalized to study
cal terms and further provides significant computational the spectrum of Floquet systems exhibiting MBL as well.
advantages from the cancellation of unitaries as shown in
Fig. 4.
Further, we have extended the 2-leg, multi-layer ten-
ACKNOWLEDGMENTS
sor network ansatz for FMBL systems43 to unitaries with
several legs while keeping the number of layers fixed at
two. We have shown that compared to increasing the S.H.S. and T.B.W. are both supported by TOPNES,
number of layers, the extension to multiple legs (l-legs) EPSRC grant number EP/I031014/1. S.H.S. is also sup-
is far more computationally efficient — obtaining (expo- ported by EPSRC grant EP/N01930X/1. The work of
nentially) higher accuracies for the same system size and A.P. was performed in part at the Aspen Center for
computational cost. Physics, which is supported by National Science Founda-
By comparing the energies and eigenstates evaluated tion grant PHY-1066293. Statement of compliance with
using TNS to exact diagonalization for a chain of 12 EPSRC policy framework on research data: This publica-
sites, we demonstrated that our figure of merit (SCN) tion is theoretical work that does not require supporting
reflects the accuracy of our method. In the regime where research data.
the figure of merit is small, the energy eigenvalues from
the TNS and exact diagonalization match extremely well.
Furthermore, the distribution of expectation values of lo- Appendix A: Calculation of the gradient
cal observables in the eigenstates also matches very well
with the exact diagonalization calculation. Therefore, As pointed out in subsection V A, due to the presence
this method is able to approximate the entire spectrum of U (1) and time reversal symmetry, we parameterize the
with all the eigenstates to a high degree of accuracy. unitaries in terms of real antisymmetric matrices AB cor-
We observed that the SCN (normalized by 2N ) in- responding to the blocks B of conserved U (1) charge. For
creases linearly with the system size. This shows that ` = 4, 6 the optimization gets tremendously sped up by
our method only incurs a constant error per lattice site, providing the gradient of the function to be minimized.
i.e. on implementing our scheme to larger systems, for Hence, the derivative is, calling {am x,y } the parameters
fixed `, local observables can be calculated with a size- contained in all blocks AB corresponding to a certain
independent accuracy. Hence, our approximation can be unitary ux,y ,
readily used for large system sizes. 
∂fx (ux,1 ,ux−1,2 ,ux,2 )
In the strongly disordered regime, the error as mea- −

∂am , if y = 1
x,y
sured by the SCN decreases exponentially with the num- ∂f ({u}) 
∂fx (ux,1 ,ux−1,2 ,ux,2 ) ∂f (u ,ux,2 ,ux+1,2 )
= − ∂am − x+1 x+1,1 ∂am ,
ber of legs per unitary. At weaker disorder on approach- ∂amx,y 
 x,y x,y

ing the MBL-ETH eigenstate transition, the local prop-


 if y = 2.
erties of the eigenstates are well approximated even close (A1)
to the transition. The fluctuation of the entanglement
entropy evaluated in the eigenstates has a broad peak at In order to evaluate those derivatives, we contract the
a disorder strength which is lower than the critical dis- local tensor network of Fig. 4 as shown in Fig. 13 and
order strength Wc ≈ 3.5 which is predicted to be the cut out the tensor the derivative is taken of at the very
critical point using exact diagonalization. Although the top or the very bottom, respectively, before taking the
11

overall trace indicated by wiggly lines. Since we cut out


a tensor with ` lower and upper legs, the result of the
contraction is also an 2` × 2` matrix, say M . This matrix
can be used to obtain both fx by putting back the missing
tensor, fx (ux,1 , ux−1 , ux,2 ) = tr(M ux,y
 ), and the desired
∂f (u ,u ,u ) ∂u
derivative, x x,1∂amx−1 x,2 = 4 Re tr(M ∂am x,y
) with-
x,y x,y
out the need for any additional contractions.

FIG. 13. The tensor network contraction results in


2−N +2l+2 M , with the 2` × 2` matrix M described in the main
text, if the unitary of which the derivative is being taken, is
cut out on the very top or very bottom (marked in red) before
the trace is taken (by summing over the indices corresponding
to the remaining wiggly lines).

1 13
P. Anderson, Phys. Rev. 109, 1492 (1958). R. Vosk and E. Altman, Phys. Rev. Lett. 112, 217204
2
D. Basko, I. Aleiner, and B. Altshuler, Annals of physics (2014).
14
321, 1126 (2006). A. Chandran, V. Khemani, C. R. Laumann, and S. L.
3
I. Gornyi, A. Mirlin, and D. Polyakov, Physical review Sondhi, Phys. Rev. B 89, 144201 (2014).
15
letters 95, 206603 (2005). Y. Bahri, R. Vosk, E. Altman, and A. Vishwanath, Nature
4
A. Pal and D. A. Huse, Phys. Rev. B 82, 174411 (2010). communications 6 (2015).
5 16
V. Oganesyan and D. Huse, Physical Review B 75, 155111 M. Žnidarič, T. Prosen, and P. Prelovšek, Physical Review
(2007). B 77, 64426 (2008).
6 17
J. M. Deutsch, Phys. Rev. A 43, 2046 (1991). J. H. Bardarson, F. Pollmann, and J. E. Moore, Phys.
7
M. Srednicki, Phys. Rev. E 50, 888 (1994). Rev. Lett. 109, 017202 (2012).
8 18
B. Bauer and C. Nayak, Journal Of Statistical Mechanics- R. Vosk, D. A. Huse, and E. Altman, Phys. Rev. X 5,
Theory And Experiment 2013, P09005 (2013). 031032 (2015).
9 19
A. Pal, Many-body localization, Ph.D. thesis, Princeton A. C. Potter, R. Vasseur, and S. A. Parameswaran, Phys.
University (2012). Rev. X 5, 031033 (2015).
10 20
R. Nandkishore and D. A. Huse, Annual Re- S. L. Sondhi, S. M. Girvin, J. P. Carini, and D. Shahar,
view of Condensed Matter Physics 6, 15 (2015), Rev. Mod. Phys. 69, 315 (1997).
21
http://dx.doi.org/10.1146/annurev-conmatphys-031214- S. Sachdev, Quantum phase transitions (Wiley Online Li-
014726. brary, 2007).
11 22
D. A. Huse, R. Nandkishore, V. Oganesyan, A. Pal, and M. Schreiber, S. S. Hodgman, P. Bordia, H. P. Lüschen,
S. L. Sondhi, Phys. Rev. B 88, 014206 (2013). M. H. Fischer, R. Vosk, E. Altman, U. Schneider, and
12
D. Pekker, G. Refael, E. Altman, E. Demler, and I. Bloch, Science 349, 842 (2015).
V. Oganesyan, Phys. Rev. X 4, 011052 (2014).
12

23
J.-y. Choi, S. Hild, J. Zeiher, P. Schauß, A. Rubio-Abadal,
T. Yefsah, V. Khemani, D. A. Huse, I. Bloch, and
C. Gross, Science 352, 1547 (2016).
24
J. Smith, A. Lee, P. Richerme, B. Neyenhuis, P. W. Hess,
P. Hauke, M. Heyl, D. A. Huse, and C. Monroe, Nature
Physics , 1745 (2016).
25
D. A. Huse, R. Nandkishore, and V. Oganesyan, Physical
Review B 90, 174202 (2014).
26
M. Serbyn, Z. Papić, and D. A. Abanin, Physical review
letters 111, 127201 (2013).
27
V. Ros, M. Mueller, and A. Scardicchio, Nuclear Physics
B 891, 420 (2015).
28
A. Chandran, I. H. Kim, G. Vidal, and D. A. Abanin,
Physical Review B 91, 085425 (2015).
29
J. Z. Imbrie, Journal of Statistical Physics 163, 998 (2016).
30
J. Z. Imbrie, Phys. Rev. Lett. 117, 027201 (2016).
31
A. Chandran, A. Pal, C. Laumann, and A. Scardicchio,
arXiv preprint arXiv:1605.00655 (2016).
32
W. De Roeck and F. Huveneers, arXiv preprint
arXiv:1608.01815 (2016).
33
M. Fannes, B. Nachtergaele, and R. F. Werner, Comm.
Math. Phys. 144, 443 (1992).
34
F. Verstraete and J. I. Cirac, Phys. Rev. B 73, 094423
(2006).
35
D. Perez-Garcia, F. Verstraete, M. M. Wolf, and J. I.
Cirac, Quantum Info. Comput. 7, 401 (2007).
36
N. Schuch, M. M. Wolf, F. Verstraete, and J. I. Cirac,
Phys. Rev. Lett. 100, 030504 (2008).
37
F. Verstraete, V. Murg, and J. I. Cirac, Advances in
Physics 57, 143 (2008).
38
U. Schollwock, Annals of Physics 326, 96 (2011).
39
R. Orus, Annals of Physics 349, 117 (2014).
40
M. Friesdorf, A. H. Werner, W. Brown, V. B. Scholz, and
J. Eisert, Phys. Rev. Lett. 114, 170505 (2015).
41
A. Chandran, J. Carrasquilla, I. H. Kim, D. A. Abanin,
and G. Vidal, Phys. Rev. B 92, 024201 (2015).
42
D. Pekker and B. K. Clark, arXiv preprint arXiv:1410.2224
(2014).
43
F. Pollmann, V. Khemani, J. I. Cirac, and S. L. Sondhi,
Phys. Rev. B 94, 041116 (2016).
44
V. Khemani, F. Pollmann, and S. L. Sondhi, Phys. Rev.
Lett. 116, 247204 (2016).
45
S. P. Lim and D. N. Sheng, Phys. Rev. B 94, 045111 (2016).
46
X. Yu, D. Pekker, and B. K. Clark, arXiv preprint
arXiv:1509.01244 (2015).
47
H. Kim, M. C. Bañuls, J. I. Cirac, M. B. Hastings, and
D. A. Huse, Phys. Rev. E 92, 012128 (2015).
48
J. A. Kjäll, J. H. Bardarson, and F. Pollmann, Physical
review letters 113, 107204 (2014).
49
D. J. Luitz, N. Laflorencie, and F. Alet, Physical Review
B 91, 081103 (2015).
50
M. Sanz, M. M. Wolf, D. Pérez-Garcı́a, and J. I. Cirac,
Phys. Rev. A 79, 042308 (2009).
51
D. Pérez-Garcı́a, M. Sanz, C. Gonzalez-Guillen, M. M.
Wolf, and J. I. Cirac, New Journal of Physics 12, 025010
(2010).
52
J. I. Cirac, D. Poilblanc, N. Schuch, and F. Verstraete,
Phys. Rev. B 83, 245134 (2011).
53
V. Khemani, S. Lim, D. Sheng, and D. A. Huse, arXiv
preprint arXiv:1607.05756 (2016).

You might also like