Entanglement scaling in fermion chains with a localization-delocalization transition and inhomogeneous modulations
Abstract
We study the scaling of logarithmic negativity between adjacent subsystems in critical fermion chains with various inhomogeneous modulations through numerically calculating its recently established lower and upper bounds. For random couplings, as well as for a relevant aperiodic modulation of the couplings, which induces an aperiodic singlet state, the bounds are found to increase logarithmically with the subsystem size, and both prefactors agree with the predicted values characterizing the corresponding asymptotic singlet state. For the marginal Fibonacci modulation, the prefactors in front of the logarithm are different for the lower and the upper bound, and vary smoothly with the strength of the modulation. In the delocalized phase of the quasi-periodic Harper model, the scaling of the bounds of the logarithmic negativity as well as that of the entanglement entropy are compatible with the logarithmic scaling of the homogeneous chain. At the localization transition, the scaling of the above entanglement characteristics holds to be logarithmic, but the prefactors are significantly reduced compared to those of the translationally invariant case, roughly by the same factor.
I Introduction
In the last decade, there has been an increasing interest in the the entanglement properties of quantum systems eisert2008 ; entanglement_review ; laflorencie . The studies on this subject allowed for a deeper understanding of many-body models, in particular with respect to criticality vidal ; Calabrese_Cardy04 ; hsu2009 , simulability VC2006 ; shi2006 ; SWVC2008 , and thermalization EFG2015 ; AC2017 ; serbyn2019 . In many of the early works, the case when the entire quantum system is in a pure state was considered, and one investigated the entanglement between a subsystem and the rest of the system. In this case, the so-called entanglement entropy is a proper measure of the quantum entanglement ryszard2009 . Consider a system in a pure state , and divide the system into two complementary parts and (i.e., in a way that ); calculating the reduced density matrices , , the entanglement entropy is defined as the von Neumann entropy of either of them:
(1) |
However, if one wishes to characterize the entanglement between two subsystems whose union does not cover the whole system , (), the subsystem is no longer in a pure state, and one has to use another measure to quantify the entanglement between and . One possible candidate is the logarithmic entanglement negativity, which is proven to be an entanglement monotone eisert2001 ; vidal2002 and is defined as
(2) |
where is the reduced density matrix of the subsystem , denotes the partial transpose on subsystem , and is the trace norm. The partial transpose is defined as , with and being bases for and , respectively. It corresponds to the local time reversal, and if a state is separable (non-entangled) partial transposed density matrix is still a valid density matrix, while if it is entangled, the partial transpose of the density matrix may have negative eigenvalues peres1996 . For one-dimensional, homogeneous critical models, there exist conformal field theory (CFT) results for the entanglement entropy and the entanglement negativity calabrese2012 . The entanglement entropy depends on the length of the subsystem as
(3) |
where is the so-called central charge, while the entanglement negativity of adjacent subsystems of length scales as
(4) |
In the recent years, a series of numerical bayat2010 ; calabrese2013 ; ruggiero2016 ; alba2013 ; chung2014 ; sherman2016 ; gray2018 and analytical calabrese2012 ; tonni2013 ; eisler2014 ; calabrese2015 ; coser2014 ; hoogeven2015 ; wen2015 ; blondeau2016 ; santos2011 ; chang2016 ; grover2019 ; Lu2019 ; shapourian2019 works have been devoted to the entanglement negativity. For non-interacting bosonic systems, the entanglement negativity is calculable efficiently (in the number of nodes) both for the ground state and for thermal states vidal2002 ; audenaert2002 ; eisert2003 ; adesso2007 . The reason is, that the partial transpose maps the bosonic Gaussian states to bosonic Gaussian operators peres1996 ; simon2000 . However, for non-interacting fermionic systems, no similar calculation method is known. The ground and thermal states are fermionic Gaussian states, but the partial transpose is not a fermionic Gaussian operator eisler2015 . The lack of a simple formula for the entanglement negativity triggered a series of works on upper and lower bounds shapourian2017 ; shapourian2019 ; shapurian2019RevA ; eisler2018 ; negativity-bounds .
In this work, we will use the upper and lower bounds given in eisler2018 to investigate numerically the scaling of the entanglement negativity in various fermionic chains with random or aperiodic inhomogeneities. In the case of sublattice-symmetry, we will present a simplification in the calculation of the lower bound. The scaling of entanglement entropy will also be studied numerically by the correlation matrix method peschel-2003 .
The rest of the paper is organized as follows. In Sec. II, the models are defined. In III, we recapitulate the steps of calculating the negativity upper and lower bounds introduced in Ref. negativity-bounds , and a simplification of the form of the lower bound due to sublattice symmetry is presented in Sec. III.3. In Sec. IV, we present our numerical results. These are then discussed in Sec. V. .
II Models
In this work, we will consider different variants of the fermionic hopping model having the Hamiltonian of general form
(5) |
where and are site-dependent hopping amplitudes and on-site energies, respectively, while and are fermionic creation and annihilation operators obeying the anticommutation rules and , for . Periodic boundary condition is considered, so that site is identified with site . The chemical potential is zero, so the states with are occupied. We mention that this class of models can be mapped to XX spin- chains by the well-known Jordan-Wigner transformation.
II.1 Off-diagonal inhomogeneity
For this class of models, the on-site energies are all zero , , while the hopping amplitudes are position-dependent, either random or follow an aperiodic modulation. In the first case, which we refer to as random model, we assume that the amplitudes are independent, identically distributed quenched random variables, drawn from a uniform distribution in the interval .
In the latter case, we will use two different aperiodic modulations, both defined by an inflation rule. One of them is the Fibonacci modulation, where hopping amplitudes are modulated according to the Fibonacci sequence. It is defined using a two-letter alphabet ( and ) by the substitution rule:
The first few realizations of the Fibonacci sequence are , , , .
Another two-letter sequence, which we refer to as relevant aperiodic modulation (RAM), is obtained by the following inflation rule jz :
Generating a sufficiently long aperiodic sequence by the repeated application of the inflation rule, a modulation pattern can be associated with it, in which letter () corresponds to amplitude (). The strength of the modulation can be characterized by the ratio of two types of amplitudes, .
The non-zero energy eigenstates of the randomly disordered model are exponentially localized, and the localization length diverges for zero energy as it was shown with rigorous tools in connection with the off-diagonal Anderson model in deylon1986 ; igloi2017 . According to the strong-disorder renormalization group (SDRG) method mdh ; im , the ground state of the random model is a random-singlet state fisherxx , which is a product of one-particle states on pairs of sites which can be arbitrarily far away from each other. The method is approximative but is asymptotically exact, giving the low-energy (large-scale) properties of the system correctly. Analogous to this, the ground state of aperiodic models is an aperiodic-singlet state hida , for any in the case of the RAM jz , but, for the Fibonacci modulation, which is a so-called marginal perturbation, only in the limit .
In such a singlet state, the entanglement entropy in units of is simply given by the number of pairs with precisely one constituent in subsystem . The average entanglement entropy of a subsystem of size , which is part of an infinite system increases asymptotically (apart from log-periodic oscillations for aperiodic models) as
(6) |
where the effective central charge depends on the type of modulation. For the random modelrefael , , for the RAM, , where jz , while for the Fibonacci modulation it varies continuously with , and its limiting value is , where is the golden mean jz ; ijz .
The logarithmic negativity in a singlet state is given by the number of singlets connecting and , in units of . As it was shown in Ref. ruggiero2016 for the random singlet state, the average logarithmic negativity of adjacent intervals of size scales as
(7) |
with the prefactor . Recently, a more detailed study has appeared about the negativity spectrum of this model turkeshi2019 . The prefactor is the half of the prefactor of entanglement entropy, which can be intuitively understood since subsystem borders to only on one side, while, in the case of entanglement entropy, borders to the rest of the system on both sides. According to this, the result in Eq. (7) is expected to hold also for aperiodic singlet states with a prefactor .
The recapitulated SDRG method and its modifications has a wide range of applications from the description of entanglement of star-graph like systems juhasz_ober_2018 and disordered surfaces juhasz2017 ; juhasz2017_jstatmech including the dynamics of disordered systems pekker2014 ; vosk2013 ; vosk2014 ; bardason2012 ; roosz2017 and description of their highly excited states pouravarni2013 ; pouravarni2015 ; huang2014 ; vasseur2015 ; storms2014 ; hsin2015 or even a construction of quasi-periodic tensor network model enable us to investigate AdS/CFT correspondence jahn2019 .
II.2 The Harper model
Another model, which we will consider is the Harper model. Here, the couplings in the general Hamiltonian in Eq. (5) are constant, , while the local potential is a quasiperiodic function of , namely,
(8) |
where is the golden mean. The irrationality of makes the model quasiperiodic. This model, when formulated in terms of spin variables is also known as Aubry-André model aubry-andre . It is well-known that this model shows a delocalization transition, which is exactly at due to its self-dual property wilkinson . In the region , all one-particle eigenstates are extended, whereas, for , they are all localised on a length aubry-andre
(9) |
At the critical point, , the one-particle states show an interesting multi-fractal behaviour evangeleou2000 ; siebesma1987 : they are essentially localised on sites, where the dimension of the effective support of the state varies from state to state. Its maximal value is found numericallyevangeleou2000 to be .
To our knowledge the logarithmic negativity in the ground state of the Harper model was not studied so far in the literature, the entanglement entropy has been studied in harper_ent , there the authors focused on the effect of the chemical potential. Here we investigate the size dependence of the entanglement entropy.
III Negativity of free fermions: upper and lower bounds
In this section, we briefly summarize the calculation of upper and lower bounds introduced in Ref. eisler2018 and used in the present paper. Instead of the general definitions, here, it is sufficient to restrict ourselves to the formulation valid in the special case of particle number conserving states. The latter means that the density matrix commutes with the particle number operator (), . The ground states of the quadratic fermion Hamiltonian in Eq. (5) are such states.
III.1 Upper bound
First, we consider the upper bound . It is formulated in terms of the covariance matrix
(10) |
where is the correlation matrix, with all the other entries of being zero. For every Hamiltonian in the form (5), the covariance matrix can be obtained following a standard canonical transformation lieb61 . For translational invariant systems, one can easily obtain a closed form of the covariance matrix while, for inhomogeneous systems, it is computable in polynomial time in the number of fermionic modes . The covariance matrix characterizes all correlations in the system, so it also determines the entanglement negativity of any subsystems. However, no simple formula is known for the entanglement negativity in terms of the covariance matrix, making the bounds defined in eisler2018 really valuable. The upper bound we use is given by
(11) |
where denotes the Rényi entropy of state :
(12) |
The state is defined by its covariance matrix as
(13) |
where , and .
III.2 Lower bound
For constructing the lower bound, the matrix is divided in the following way:
(14) |
Using the singular value decomposition of , , where is a diagonal matrix with non-negative elements, whereas and are orthogonal matrices, one can transform by to the following form
(15) |
Denoting the diagonal elements of , , and by , , and , respectively, the lower bound is given by
(16) |
where
(17) | |||||
III.3 Simplification of the lower bound by sublattice symmetry
Here, we show that the expression in Eq. (16) can be further simplified making use of the sublattice symmetry, which holds in the absence of an on-site potential for an even . We also assume, that the lattice is half filled. Due to this, the elements of matrix with indices of the same parity are all zero. By replacing rows and columns of , let us arrange, separately in block and , the odd (even) indices to the first (second) places. Then will have the form
(18) |
We are looking for the singular value decomposition , where the diagonal matrix contains the non-negative singular values of , while the columns of and are eigenvectors of and , respectively. These matrices are then block diagonal,
(19) |
and, as a consequence, and are block diagonal, as well:
(20) |
Transforming by will bring the diagonal blocks
(21) |
to the form
(25) | |||||
(29) |
The elements and in Eq. (16) are therefore zero. Using this, the lower bound can be written as
(30) |
where the prime means that the summation goes over the singular values fulfilling only.
IV Numerical results
We calculated the lower and upper bounds of the entanglement negativity, and , respectively, which were introduced in Ref. eisler2018 and are defined in the previous section, for the models described in Sec. II. It is worth mentioning that the above bounds, apart from an additive constant for the upper bound, are sharp for the special case of singlet states.
The numerical calculations were performed according to two kinds of schemes. Either the total system size was kept fixed (and large) and the size of adjacent subsystems was varied, or was varied keeping the ratio constant. For a given and , and were calculated for all possible positions of the subsystems ( in number) and an averaging was performed in the case of non-random models. For the random model, different positions of the subsystems was considered in each random sample, while the number of samples was for smaller systems, gradually decreasing down to for the largest system with .
In order to make corrections to an expected large- dependence
(31) |
more visible in the second scheme, we also calculated effective, size-dependent prefactors from consecutive data points at and as
(32) |
For the Harper model, we also considered the average entanglement entropy of a subsystem of size , the scaling of which is so far unexplored, and which can be numerically calculated from the eigenvalues of the correlation matrix of fermions restricted to the subsystem vidal ; peschel-2003 .
IV.1 Homogeneous system
First, we investigated the behavior of the lower and upper bounds in the homogeneous chain. Here, we assume that the system is infinite (), which allows us to use the exact expressions of the correlations igloi2020 , while keep the subsystem size finite. The numerical results are shown in Fig. 1.

The upper bound follows the scaling
(33) |
with , which is compatible with the behavior of the entanglement negativity known from CFT, see Eq. (4).
The behavior of the lower limit is more complicated: the overall logarithmic trend is decorated with log-periodic oscillations of small amplitude, which is unusual in a homogeneous system. Here, the oscillations are the consequences of the definition of the lower limit [see Eq. (30)], in which only the singular values greater than contribute to the sum. Putting the singular values in decreasing order, one can observe that the th singular value () slowly increases with increasing . When a singular value crosses the threshold , so that the number of terms in the sum increases by one, a new period of the oscillations starts.
To investigate the overall logarithmic trend, one has to fit to identical parts of the periods, like the breaking points indicated by arrows in Fig. 1. To do this, at least a few period long data set is needed. The period of these oscillations is about on a logarithmic scale, which means that one has to handle quite large subsystems of size . Fortunately, in this size range, only the largest singular values are needed, which can be obtained from the largest eigenvalues of the matrices and . Since we have a closed formula for the matrix elements of and , we can use the power method and multiplicate with () without storing the matrix elements in the memory. The ratios of the consecutive eigenvalues (in decreasing order) of the matrix () are between and , so the power method converges rapidly.
Using the data for the lower bounds at the four breaking points shown in Fig. 1, we calculated effective prefactors from neighboring data points, which are in order, , , and . Assuming corrections of the form , an extrapolation to results in ; if the data point obtained from the smallest system sizes is excluded, one obtains , which gives an estimate on the error of the fit. Thus, in the homogeneous chain, the prefactor of the lower bound, , is close to the half of the prefactor of entanglement negativity .
IV.2 Off-diagonal inhomogeneity
The dependence of the lower and upper bounds on for fixed ratios , as well as the corresponding effective prefactors for the random model are shown in Fig. 2.


As can be seen, both the lower and upper bounds scale logarithmically with for large . The effective prefactor of the upper bound displays a slow crossover from the clean system’s value toward predicted by the SDRG method with increasing , and, at the system sizes available by the numerical method, it is still considerably far from it. We note that, under the same circumstances, the prefactor of the entanglement entropy has similar deviations from the asymptotic value (not shown). In the case of the lower bound, the prefactor in the homogeneous system () is closer to the asymptotic limit , and, accordingly, it shows a more rapid crossover.
In the case of the relevant aperiodic modulation, the lower and upper bounds, as well as the effective prefactors as a function of are exemplified in Fig. 3 for . The system sizes were , which are the lengths of words obtained by the repeated application of the inflation rule starting with letter , while the subsystem size was , where stands for the integer part. By this choice of , log-periodic oscillations in the data can be avoided jz . As can be seen, the bounds follow a logarithmic scaling and their asymptotic prefactors shown in the inset for different disorder strengths are in agreement with the expectation valid for singlet states.

For the Fibonacci modulation, which is a marginal one, we calculated the size dependence of lower and upper bounds for different modulation strengths . Here, the system sizes and the subsystem sizes were chosen to be and , respectively, where denotes the th term of the Fibonacci sequence. As can be seen in Fig. 4, the bounds plotted against show log-periodic oscillations with a period of three data points. This is in accordance with that the self-similarity of the aperiodic singlet state is achieved by applying the third power of the inflation transformation jz . Therefore, in estimating the prefactors, we used every third data points. For both bounds, we find a logarithmic dependence on as given in Eq. (31). The prefactor of the upper bound shows a variation with qualitatively similar to that of the prefactor of the entanglement entropy jz : For moderate strengths of the modulation, it has a slight deviation from the clean system’s value, then it tends rapidly to the singlet-state limit. In the case of the lower bound, the prefactor in the clean system () and in the singlet-state limit () hardly differ, therefore the variation with is very weak.

IV.3 The Harper model


In the Harper model, we investigated first the scaling of the entanglement entropy in the ground state. In these calculations, system sizes were chosen to be twice of Fibonacci numbers , and the size of the subsystem was .
The results are shown in Fig. 5. We find that, in the extended phase, the entanglement entropy scales logarithmically as , and the measured prefactor () is compatible with that of the homogeneous system up to the precision of the estimation. The reason of this agreement is the extended nature of the eigenstates, which is similar to the eigenstates of a homogeneous chain.
In the critical point, the entanglement entropy is still found to scale logarithmically as
(34) |
however, the effective central charge differs significantly from the central charge of the homogeneous system.
In the localized phase, the entanglement entropy saturates to finite values in the limit . As it is demonstrated in Fig. 5, in large systems , in which the length scale is set by the localization length rather than the system size, the entanglement entropy follows the law
(35) |
with the same prefactor as found in Eq. (34). Approaching the critical point, this law is, however, deteriorates when the diverging localisation length becomes comparable with the system size.
Numerical results on the entanglement negativity are shown in Fig. 6. In the delocalized phase and in the critical point, the upper bound increases logarithmically with the system size. In the delocalized phase, just like for the entranglement entropy, the prefactor is found to agree with the that of the conformally invariant homogeneous system. In the critical point, the prefactor of the upper bound is reduced to . As can be seen in Fig. 6, the size dependence of the lower bound shows oscillations which may originate both in the discreteness in its definition (see section IV.1) and in the quasiperiodic modulation. For this reason, a completely reliable estimate of the prefactor cannot be obtained.
As can seen in Fig. 6b, the upper and lower bounds become nearly equal in the localized phase. Their ratio approaches to one, if , and the ratio is already for . This makes possible to investigate the behavior of the negativity in the localized phase. Although the difference of the upper and lower bounds becomes larger as one approaches the transition point, there is an intermediate regime, where the transition point is not too far, but the difference of the bounds is still moderate. Similarly to the behavior of the entanglement entropy in the localized phase, the upper bound of entanglement negativity increases according to
(36) |
in the vicinity of the transition point, where is the prefactor characteristic of the critical point.


b) Dependence on the localization length in the localized phase. The black line with slope is drawn to guide the eye.
V Conclusion and outlook
In this paper, we investigated numerically the scaling of the entanglement negativity in fermionic chains with different aperiodic and random modulations. In general, we found a logarithmic increase of both bounds with the subsystem size, provided the model is critical. Some of the off-diagonal modulations, like the random, the relevant aperiodic, and the extremal Fibonacci modulation asymptotically induce a state composed of singlets. In these cases, we have confirmed that the prefactors of the lower and upper bounds coincide and agree with the half of the prefactor of the entanglement entropy, .
In the case of the Fibonacci modulation of finite strength , the ground state is no longer a singlet state, and the prefactor of the entanglement entropy is known to vary continuously and monotonically with . We have found a similar behavior of the prefactors of both bounds, although for the lower bound the difference between the two extremal values is very small. The prefactor of the upper bound, which correctly gives the prefactor of the entanglement negativity in the extremal cases and , is presumably holds to be correct in the intermediate regime , as well. However, is not expected to be simply related to , as their ratio is different in the two extremal cases ( and ).
The Harper model, which contains a quasi-periodic modulation in the diagonal term, has different ground-state phases depending on the modulation strength. In the extended phase, the entanglement entropy and the upper bound of the negativity are found to scale identically with the conformally invariant homogeneous chain. This is in accordance with the expectations since, in this phase, the Harper model has extended quasi-free eigenstates a continuous spectrum sinai1987 . In the critical point separating the extended phase from the delocalized one, we find, however, that the prefactors of both the entanglement entropy and the upper bound of the negativity are reduced compared to the values of the homogeneous system. It is remarkable that the prefactors are reduced by roughly the same factor ( for the entanglement entropy and for the negativity upper bound). This may suggest that the scaling can be described by the prefactors of the homogeneous system, but using an effective length . Indeed, the one-particle states of the model in the critical point are known to be fractals, with an effective number of sites participating in them scaling as . The fractal dimension varies from state to state and its maximal valueevangeleou2000 is close to . The clarification of a possible relationship between the reduction factor and the fractal dimensions of eigenstates is an open question which is deferred to future research.
As a possible future direction, the present investigations could be extended to various generalized Harper models. For instance, one could add paring terms () to the Hamiltonian and investigate the so called quasi-periodic Ising model chandran2017 with the free-fermion methods applied here. We also mention here recent generalizations of the Harper model with additional long-range couplings monthus2019 ; deng2019 .
Acknowledgements.
This work was supported by the by the Deutsche Forschungsgemeinschaft through the Cluster of Excellence on Complexity and Topology in Quantum Matter ct.qmat (EXC 2147) and the National Research, Development and Innovation Office NKFIH under grants No. K128989, K124152, K124176, KH129601 and through the Hungarian Quantum Technology National Excellence Program (Project No. 2017-1.2.1-NKP-2017-00001). ZZ was also partially funded by the Janos Bolyai and the Bolyai+ Scholarships.References
- (1) J. Eisert and M. Cramer, M. B. Plenio, Rev. Mod. Phys. 82, 277 (2010).
- (2) P. Calabrese, J. Cardy, and B. Doyon (Eds.), Entanglement entropy in extended quantum systems (special issue), J. Phys. A 42, 500301 (2009).
- (3) N. Laflorencie, Phys. Rep. 643, 1 (2016).
- (4) G. Vidal, J. I. Latorre, E. Rico, and A. Kitaev, Phys. Rev. Lett. 90 227902, (2003); J. I. Latorre, E. Rico, and G. Vidal, Quant. Inf. Comput. 4, 48 (2004).
- (5) P. Calabrese and J. Cardy, J. Stat. Mech. P06002 (2004).
- (6) B. Hsu, M. Mulligan, E. Fradkin, and E-A. Kim, Phys. Rev. B 79, 115421 (2009).
- (7) F. Verstraete and J. I. Cirac, Phys. Rev. B 73, 094423 (2006).
- (8) Y. Shi, L. Duan, and G. Vidal, Phys. Rev. A 74, 022320 (2006).
- (9) N. Schuch, M. M. Wolf, F. Verstraete, and J. I. Cirac, Phys. Rev. Lett. 100, 030504 (2008).
- (10) J. Eisert, M. Friesdorf, and C. Gogolin, Nature Physics 11, 124 (2015).
- (11) V. Alba and P. Calabrese, PNAS 114, 7947 (2017).
- (12) D. A. Abanin, E. Altman, I. Bloch, and M. Serbyn, Rev. Mod. Phys. 91, 021001 (2019).
- (13) R. Horodecki, P. Horodecki, M. Horodecki, and K. Horodecki Rev. Mod. Phys. 81, 865 (2009).
- (14) J. Eisert, Entanglement in quantum information theory, PhD thesis, University of Potsdam (2001).
- (15) G. Vidal and R. F. Werner, Phys. Rev. A 65, 032314 (2002).
- (16) A. Peres, Phys. Rev. Lett. 77, 1413 (1996).
- (17) P. Calabrese, J. Cardy, and E. Tonni, Phys. Rev. Lett. 109, 130502 (2012).
- (18) A. Bayat, P. Sodano, and S. Bose, Phys. Rev. B 81, 064429 (2010).
- (19) P. Calabrese, L. Tagliacozzo, and E. Tonni, J. Stat. Mech. P05002 (2013).
- (20) P. Ruggiero, V. Alba, and P. Calabrese, Phys. Rev. B 94, 035152 (2016).
- (21) V. Alba, J. Stat. Mech. P05013 (2013).
- (22) C.-M. Chung, V. Alba, L. Bonnes, P. Chen, and A. M. Läuchli, Phys. Rev. B 90, 064401 (2014).
- (23) N. E. Sherman, T. Devakul, M. B. Hastings, and R. R. P. Singh, Phys. Rev. E 93, 022128 (2016).
- (24) J. Gray, S. Bose, and A. Bayat, Phys. Rev. B 97, 201105(R) (2018)
- (25) P. Calabrese, J. Cardy, and E. Tonni, J. Stat. Mech. P02008 (2013).
- (26) V. Eisler and Z. Zimborás, New J. Phys. 16, 123020 (2014).
- (27) P. Calabrese, J. Cardy, and E. Tonni, J. Phys. A 48, 015006 (2015).
- (28) A. Coser, E. Tonni, and P. Calabrese, J. Stat. Mech. P12017 (2014).
- (29) M. Hoogeveen and B. Doyon, Nucl. Phys. B 898, 78 (2015).
- (30) X. Wen, P.-Y. Chang, and S. Ryu, Phys. Rev. B 92, 075109 (2015).
- (31) O. Blondeau-Fournier, O. A. Castro-Alvaredo, and B. Doyon, J. Phys. A 49, 125401 (2016).
- (32) R. A. Santos, V. Korepin, and S. Bose, Phys. Rev. A 84, 062307 (2011).
- (33) P.-Y. Chang and X. Wen, Phys. Rev. B 93, 195140 (2016).
- (34) C Lu, TH Hsieh, T Grover, arXiv: 1912.04293 (2019)
- (35) TC Lu, T Grover, Phys. Rev. B 99, 075157 (2019)
- (36) H Shapourian, S Ryu, J. Stat. Mech. 043106 (2019)
- (37) K. M. R. Audenaert, J. Eisert, M. B. Plenio, and R. F. Werner, Phys. Rev. A 66, 042327 (2002).
- (38) J. Eisert and M. B. Plenio, Int. J. Quant. Inf. 1, 479 (2003).
- (39) G. Adesso and F. Illuminati, J. Phys. A 40, 7821 (2007).
- (40) R. Simon, Phys. Rev. Lett. 84, 2726 (2000).
- (41) V. Eisler and Z. Zimborás, New J. Phys. 17, 053048 (2015).
- (42) H. Shapourian, K. Shiozaki, and S. Ryu, Phys. Rev. B 95, 165101 (2017).
- (43) H Shapourian, S Ryu, Phys. Rev. A 99, 022310 (2019)
- (44) V. Eisler and Z. Zimborás, Phys. Rev. B 97, 165123 (2018).
- (45) J. Eisert, V. Eisler, and Z. Zimborás, Phys. Rev. B 97, 165123 (2018).
- (46) I. Peschel, J. Phys. A: Math. Gen. 36, L205 (2003).
- (47) R. Juhász and Z. Zimborás, J. Stat. Mech. P04004 (2007).
- (48) F. Deylon, H. Kunz, B. Souillard, J. Phys. A: Math. Gen. 16, 25 (1986).
- (49) G. Roósz, Y.-C. Lin, and F. Igloi, New J. Phys. 19, 023055 (2017).
- (50) S. K. Ma, C. Dasgupta, and C.-K. Hu, Phys. Rev. Lett. 43, 1434 (1979); C. Dasgupta and S. K. Ma, Phys. Rev. B 22, 1305 (1980).
- (51) F. Iglói and C. Monthus, Phys. Rep. 412, 277 (2005); Eur. Phys. J. B 91, 290 (2018).
- (52) D. S. Fisher, Phys. Rev. Lett. 69, 534 (1992); D. S. Fisher, Phys. Rev. B 50, 3799 (1994).
- (53) K. Hida, Phys. Rev. Lett. 93, 037205 (2004).
- (54) G. Refael and J. E. Moore, Phys. Rev. Lett. 93, 260602 (2004).
- (55) F. Iglói, R. Juhász, and Z. Zimborás, Europhys. Lett. 79, 37001 (2007).
- (56) X. Turkeshi, P. Ruggiero, and P. Calabrese, Phys. Rev. B 101, 064207 (2020)
- (57) R. Juhász, J. M. Oberreuter and Z. Zimborás, J. Stat. Mech. 2018, 123106 (2018)
- (58) R. Juhász, I. A. Kovács, G. Roósz, F. Iglói, J. Phys. A: Math. Theor. 50 324003 (2017)
- (59) R. Juhász, J. Stat. Mech. 2017 083107 (2017)
- (60) D. Pekker, G. Refael, E. Altman, E. Demler, V. Oganesyan, Phys. Rev. X 4 011052 (2014)
- (61) R. Vosk, E. Altman, Phys. Rev. Lett. 110 067204 (2013)
- (62) R. Vosk, E. Altman, Phys. Rev. Lett. 112 217204 (2014)
- (63) J. H. Bardarson, F. Pollmann, J. E. Moore, Phys. Rev. Lett. 109 017202 (2012)
- (64) G. Roósz, Yu-Cheng Lin, F. Iglói, New J. Phys. 19, 023055 (2017)
- (65) M. Pouranvari, K. Yang, Phys. Rev. B 88, 075123 (2013)
- (66) M. Pouranvari, K. Yang, Phys. Rev. B 92, 245134 (2015)
- (67) Y. Huang, J. E. Moore, Phys. Rev. B 90 220202(R) (2014)
- (68) R. Vasseur, A.C. Potter, and S.A. Parameswaran, Phys. Rev. Lett. 114, 217201 (2015)
- (69) M. Storms and R. R. P. Singh, Phys. Rev. E 89, 012125 (2014)
- (70) Hsin-Hua Lai, Kun Yang, Phys. Rev. B 91, 081110 (2015)
- (71) A. Jahn, Z. Zimborás, J. Eisert, Preprint arXiv:1911.03485 (2019
- (72) S. Aubry and G. André, Ann. Israel Phys. Soc. 3, 133 (1980).
- (73) M. Wilkinson and J. Austin, Phys. Rev. B 50, 1420 (1994).
- (74) S. N. Evangeleou, J.-L. Pichard, Phys. Rev. Lett. 84, 1643 (2000).
- (75) A. P. Siebesma, L. Pietronero, Europhys. Lett. 4, (5) 597 (1987).
- (76) N. Roy, A. Sharma, Phys. Rev. B 100, 195143 (2019).
- (77) E. Lieb, T. Schultz, and D. Mattis, Ann. Phys. (N.Y.) 16, 407 (1961).
- (78) G. Roósz, I. A. Kovács, F. Iglói, Eur. Phys. J. B 93, 8 (2020).
- (79) Ya. G. Sinai J. Stat. Phys., Vol. 46, 861 (1987).
- (80) A. Chandran and C. R. Laumann, Phys. Rev. X. 7, 031061 (2017).
- (81) C. Monthus, Fractals 27, 195007 (2019).
- (82) X. Deng, S. Ray, S. Sinha, G. V. Shlyapnikov, L. Santos, Phys. Rev. Lett. 123, 025301 (2019).