Introduction

The Kitaev model1 is one of the greatest examples of two-dimensional (2D) solvable models of quantum spin liquids (QSLs)2,3,4, especially in the perspective of spin-orbital-entangled physics5,6. This model has a bond-dependent anisotropic interaction, which brings about exchange frustration and realizes gapped and gapless spin liquid states depending on its parameters. Amazingly, this interaction can be furnished in materials with a strong spin–orbit coupling7. Iridates and α-RuCl3 are prominent examples of candidate materials for the Kitaev model8,9,10, and thus are called Kitaev materials. As a hallmark, the half-integer quantization of the thermal Hall conductance has been observed in α-RuCl3 with a magnetic field11, suggesting the fractionalization of spin degrees of freedom predicted from the Kitaev model. However, it is also known that these honeycomb materials cannot fully be understood by the original (pure) Kitaev model12,13. While other diagonal or offdiagonal interactions might be important in real materials14, the importance of disorder has been ignored in these materials until recently15,16,17. Indeed, experiments in A3LiIr2O6 (A = H, Ag, etc.) show a universal scaling in the field dependence of the heat capacity5,18, which strongly suggests the existence of disorder19,20. The candidate ground states must be disordered QSLs, and the absence of long-range order can be attributed to the critical role of disorder.

In fact, the role of disorder in QSLs itself is a long-standing problem because of the absence of a solvable model, except for limited cases21. We propose a disordered Kitaev model as a "numerically” solvable model for the disordered QSL, where we can treat the magnetic field effect within the perturbation theory. Thus, this study is not only a model investigation for the disordered Kitaev materials like A3LiIr2O6 (A = H, D, Ag, etc.)5,18,22, but also a systematic examination of a numerically solvable disordered QSL, which would be an attempt towards the universal understanding of various disordered QSLs. Especially, since most QSLs are unsolvable, an unbiased study of disordered QSLs was impossible in the previous method.

Moreover, the bond-disordered Kitaev model with an applied magnetic field is not an “analytically” solvable problem. While it seems that the problem is reduced to the Anderson localization of symmetry class D with a short-range correlated disorder23, we discovered that this is not the end of the story. The nontrivial part comes from the third-order perturbation in an applied field because the computation of the “disorder” in the time-reversal-breaking term requires diagonalization1. This nonlocal operation makes the disorder long-range correlated. Such a long-range correlated random mass disorder could be a relevant perturbation24, and a simple renormalization group analysis may fail. Treating the long-range correlated disorder essentially requires a large-scale calculation and the extrapolation to the thermodynamic limit, which were numerically difficult in the finite-size Monte Carlo methods25,26. For these reasons, we invented a powerful numerical method based on the kernel polynomial method (KPM)27 to overcome the difficulty.

From the materials side, the dispersion of low-energy excitations of H3LiIr2O6 is a mystery. Although the behavior of the heat capacity of H3LiIr2O6 without a magnetic field5 is partially explained by the bond-disordered Kitaev model17, the T2-dependence of the heat capacity of H3LiIr2O6 with a magnetic field5, which suggests the linear dispersion of the low-energy density of states (DOS)5, has never been described by an unbiased calculation. Such heat capacity behaviors are phenomenologically explained by the random-singlet theory with a Dzyaloshinskii–Moriya (DM) interaction19. However, it is not clear whether the DM interaction is important in H3LiIr2O6, and symmetric Kitaev–Γ interactions may be a dominant consequence of the spin–orbit coupling. We rather try to explain the observed linear dispersion with an applied field from the bond-disordered Kitaev model. Indeed, an unbiased numerical simulation of the bond disorder is not only theoretically nontrivial but also experimentally important.

The Anderson transition (or quantum Hall transition) is visible in a Kitaev spin liquid (KSL) by applying a small magnetic field. The quantized thermal Hall conductivity serves as an order parameter, which must go to zero in the large-disorder limit. This limit without a linear slope of the thermal Hall conductivity is named Anderson–Kitaev spin liquid (AKSL). The transition between KSL and AKSL is nothing but a transition between different Hall plateaus. Although we defined the transition in 2D, the transition may be observable even in three-dimensional (3D) generalizations28,29 (mobility edge).

In this article, we simulate the bond-disordered Kitaev model to see a crossover between KSL and AKSL, especially from the topological transition in the thermal Hall effect11,25,30. We discovered that the quantized thermal Hall effect is not as stable as expected from the Anderson transition with a short-range correlated disorder.

Results

Magnetic field effect

The pure Kitaev model on the honeycomb lattice can be defined from Fig. 1a. The bonds parallel to the red, green, and blue ones are x-, y-, and z-labeled bonds, respectively. The Hamiltonian is

$${H}_{0}=-J\mathop {\sum}\limits_{\langle jk\rangle \in \gamma }{\sigma }_{j}^{\gamma }{\sigma }_{k}^{\gamma },$$
(1)

where 〈jk〉 means a nearest-neighbor (NN) bond, and J > 0. γ = x, y, or z is determined by a bond label. This model is known to be solvable by representing \({\sigma }_{j}^{\gamma }\) by Majorana fermions \(i{b}_{j}^{\gamma }{c}_{j}\). This representation still works even if we introduce bond disorder as follows:

$${H}_{{\rm{bond}}}=-\mathop {\sum}\limits_{\langle jk\rangle \in \gamma }{J}_{jk}{\sigma }_{j}^{\gamma }{\sigma }_{k}^{\gamma },$$
(2)

where Jjk = J ± δJ is a bond-dependent hopping, and δJ > 0 is the strength of bond disorder. This model is still numerically solvable if we can assume that the ground state is 0-flux when δJ is in the perturbative regime. Under this assumption, all the states with a vison, which is defined as a pair of neighboring π-flux vortices excited by flipping the sign of a single bond, are assumed to be the "first” excited states from the ground state flux sector. This is how the perturbation theory works for this Kitaev model. We employ Kitaev’s trick to solve these Hamiltonians with an applied magnetic field1.

Fig. 1: Schematic phase diagram.
figure 1

a Directional dependence of the bond interaction and the NNN hopping arising for Majorana fermions under the magnetic field. b \({\Delta }_{\min }=\min ({\Delta }_{{\rm{vison}}})\) versus disorder strength δJ. After the gap closing around δJ = δJc2, the flux sector becomes vison glass. c κxy versus disorder strength δJ. From δJ > δJc1, the crossover to the AKSL is observed and κxy/T finally reaches 0 around δJ/J = 1.

Let us first review the perturbation for the pure Kitaev model with a magnetic field

$$H={H}_{0}+V.$$
(3)
$$V=-\mathop {\sum}\limits_{j}\left({h}^{x}{\sigma }_{j}^{x}+{h}^{y}{\sigma }_{j}^{y}+{h}^{z}{\sigma }_{j}^{z}\right),$$
(4)

where \({\bf{h}}={({h}^{x},{h}^{y},{h}^{z})}^{t}\) is an applied magnetic field.

It is well known that V can be treated by the third-order perturbation1. The result after introducing itinerant Majorana fermions cj is

$${H}_{{\rm{eff}}}=\frac{iJ}{2}\mathop {\sum}\limits_{\langle jk\rangle }{c}_{j}{c}_{k}+\frac{i\tilde{\kappa }}{2}\mathop {\sum}\limits_{\langle \langle kl\rangle \rangle }{c}_{k}{c}_{l}+({\text{four-fermion}}\, {\text{terms}}).$$
(5)
$$\tilde{\kappa }=\frac{3{h}^{x}{h}^{y}{h}^{z}}{48{\alpha }_{0}^{2}{J}^{2}},$$
(6)

where α0 = 0.262433 in the thermodynamic limit for the 0-flux state, and α0J is a vison gap in the uniform case. The determination of the prefactor follows a mean-field solution31. The direction of the next-nearest-neighbor (NNN) bond 〈〈kl〉〉 is defined clockwise as shown in Fig. 1a around the site j. A site connected by the γ-bond from j is called γ[j] for γ = x, y, and z, as shown in Fig. 1a. We define \(\tilde{h}={h}^{x}{h}^{y}{h}^{z}/48\).

Kitaev’s trick

Next, let us include binary disorder as H = Hbond + V. Following Kitaev1, we can always do perturbation from any random Hbond by a formula

$${H}_{{\rm{eff}}}^{(3)}={\Pi }_{0}V{G}_{0}^{\prime}({E}_{0})V{G}_{0}^{\prime}({E}_{0})V{\Pi }_{0},$$
(7)

where Π0 is a projection onto the ground state flux sector, \({G}_{0}^{\prime}(E)\) is an unperturbed Green function constructed from Hbond with the ground state flux sector excluded from the Hilbert space, and E0 is an initial energy. Since Hbond is solvable by Majorana fermions, it is in principle possible to calculate \({G}_{0}^{\prime}(E)\) numerically to exhaust every term appearing in the third order. For example, a Green function for excited states is efficiently obtained by KPM27 numerically. However, this strategy is surely overkill for our problem.

A much simpler solution is to use a trick introduced by Kitaev. Though we still need an O(N4) calculation cost to decide all terms by usual matrix diagonalization, where N is the number of sites, there is no need for matrix exponentiation or integration. Kitaev’s trick is done by replacing \({G}_{0}^{\prime}({E}_{0})\) by − (1 − Π0)/Δvison, assuming that the virtual state energy is constant determined just by a vison gap Δvison. This is a bold approximation to simplify the problem drastically, but as we will see essential features, such as the modulation of the mass term, can be captured even within Kitaev’s approximation.

In this way, a typical third-order term is as follows:

$${H}_{{\rm{eff}}}=\frac{i}{2}\mathop {\sum}\limits_{\langle jk\rangle }{J}_{jk}{c}_{j}{c}_{k}+\frac{i}{2}\mathop {\sum}\limits_{\langle \langle kl\rangle \rangle }{\tilde{\kappa }}_{kl}{c}_{k}{c}_{l}+\cdots$$
(8)

where \({\tilde{\kappa }}_{kl}\) depends on the intermediate site j in the third-order perturbation process. Seen from j, \(\tilde{\kappa }\) can be calculated by replacing \(3/{({\alpha }_{0}J)}^{2}\) by 1/(ΔxΔy) + 1/(ΔyΔz) + 1/(ΔzΔx), where Δγ is the energy of the vison excited by flipping the sign of the bond between j and γ[j].

$${\tilde{\kappa }}_{kl}={\tilde{\kappa }}_{j}=\frac{\tilde{h}}{{\Delta }_{x}{\Delta }_{y}}+\frac{\tilde{h}}{{\Delta }_{y}{\Delta }_{z}}+\frac{\tilde{h}}{{\Delta }_{z}{\Delta }_{x}}.$$
(9)

We note that three bonds have the same value of \({\tilde{\kappa }}_{kl}\) around j. Thus, the disorder simply modulate the mass term of Dirac cones via random NNN hoppings, and the problem is still solvable numerically.

In this case, four-fermion terms are short-ranged and we have just ignored them as we are only interested in the Hall conductivity in the h0 limit. Though we will assume the ground state of Hbond to be 0-flux in the following discussions, the perturbation can be done from any flux configuration. We note that a second-order perturbation in h is ignored because it just renormalizes bond-dependent hoppings Jjk and does not break the time-reversal symmetry. Indeed, there is a priori no way to determine the ratio of the coefficients of the second- and third-order perturbations, although we can always use a mean-field solution of the pure Kitaev model to estimate it31. As stressed in the Introduction, the computation by Kitaev’s trick is an essential part of this study, producing long-range correlation in the mass disorder.

Thermal conductivity

We only consider zero temperature and ignore thermal flux fluctuations above the 0-flux sector. Lieb’s theorem32 no longer applies, but we can expect it to be applicable on average. Anyway, the calculation is relevant only in the regime where the flux gap is not closed by thermal fluctuation or bond disorder (δJ < δJc2 in Fig. 1b, c).

We employed Kitaev’s trick to calculate a Majorana spectrum with an external magnetic field for each quenched bond disorder. From this, we can compute an inplane thermal Hall conductivity κxy(T), especially behavior of κxy/T at T → 0. Here xy does not coincide with the Cartesian axis but means a transverse component of the thermal conductivity.

As described in Methods, the thermal Hall conductivity can be computed directly by the Kubo formula. However, we can alternatively use the so-called noncommutative Chern number (NCCN)33, which is defined by a spectral projector for occupied free fermions. This formula is advantageous because it is proven to become integer after disorder average with some conditions, whereas it only makes sense in the thermodynamic limit.

$${\rm{Ch}}=-\frac{2\pi i}{{L}^{2}}\ {\rm{tr}}\left\{{P}_{F}\left[[{r}_{1},{P}_{F}],[{r}_{2},{P}_{F}]\right]\right\},$$
(10)

where PF is a spectral projector, i.e. the projection operator onto the occupied states, and L is a linear scale of the system. This formula must agree with the Kubo formula in the thermodynamic limit by a well-known relation \({\kappa }^{xy}/T=\pi {k}_{B}^{2}{\rm{Ch}}/(12\hslash )\) for Majorana fermions. We note that there are other ways to detect the topological nontriviality34,35,36.

After taking an average of Ch κxy/T over a number of disorder configurations, we plot a physical thermal Hall conductivity as a function of δJ. The error bar is estimated from a statistical deviation. From now on we set \(\hbar\) = kB = J = 1.

Numerical results

We first note that, since we only include the third-order perturbation, the results here are not simply comparable with experiments. However, it was proposed that the contribution from hxhyhz can be picked up by applying an inplane magnetic field37, so we only take an odd component under every sign change (hx hx, hy hy, and hz hz) of the three components of h from total κxy. From now on we denote κxy as an odd component under every sign change and ignore other components.

The approximate correspondence between the Kubo formula and NCCN is confirmed for the pure Kitaev model (see Fig. 2a). We note that Haar-random vectors are used in this calculation, but are not used in the following as described in Supplementary Methods and Supplementary Fig. 1. From here we will prefer NCCN to the Kubo formula because we can use KPM to approximate the spectral projector PF to avoid diagonalization. We note that the application of KPM to the Kubo formula requires efforts38.

Fig. 2: Main results.
figure 2

a Kubo formula versus NCCN. NCCN-Diag means that the Chern number is calculated by diagonalization, while NCCN-KPM means that the Chern number is calculated by KPM with \({M}^{\prime}=512\) and R = 100. In order to put errorbars, random vectors are chosen to be Haar-random. Only \(L\,\mathrm{mod}\,\,6=2,\ 4\) is plotted. The ordinary periodic boundary condition is used. b Mean and minimum values of flux gaps calculated by KPM. The error bar is smaller than the line width and only plotted for L = 10. KPBC is used. c NCCN calculated by diagonalization (solid lines) and KPM (scatter plots). Nsample = 24 is used for the diagonalization. For KPM we used R = 24, and Nsample = 360. KPBC is used. d NCCN calculated by KPM and the value extracted for L. The margin of error at 5% significance level is used for the ribbon for the extrapolation. Error bars are standard errors (SE) if not mentioned.

Next, we would move on to a large-scale calculation by Kitaev’s trick. We only take (Kitaev’s) L × L periodic boundary condition (for spins) from L = 10, where the vison gap gets close to the thermodynamic limit. As long as we are interested in the topological property, the h0 limit does not have to be taken. We set \({h}^{x}={h}^{y}={h}^{z}\equiv h={\Delta }_{\min }\), where \({\Delta }_{\min }\) is the minimum vison gap for each disorder configuration. In order to reduce the finite-size effect, we adopt Kitaev’s torus basis where the finite-size effects cancel out, which is defined from a torus basis (Ln1, Ln2 + n1)1. We call it Kitaev’s periodic boundary condition (KPBC) for simplicity. The NCCN formula for KPBC has to be modified as described in Supplementary Methods. This arbitrary choice of boundary conditions does not matter in the thermodynamic limit. The averaged 〈κxy〉/T for T → 0 is shown as a function of δJ, and drops rapidly to 0 from the quantized value as the disorder strength δJ grows. From here 〈κxy〉/T is plotted in the unit of a quantum π/12. We used R = 24 vectors to approximate the trace39.

The mean and minimum values of vison gaps are plotted for each δJ in Fig. 2b. When δJ > δJc2 ~ 0.3, the vison gap approaches 0 for some plaquette, and the 0-flux ground state is destabilized. We tentatively define δJc2 as the point where \({\Delta }_{\min }=0.05\). From here, the perturbation from the 0-flux sector cannot be justified. Moreover, some flux sectors get almost degenerate and the first-order perturbation in h now becomes relevant. Beyond this point, a quantized thermal Hall current is no longer a well-defined notion. Flux excitations and (itinerant) Majorana fermions are not separable, and the discussion based only on free Majorana fermions breaks down.

When δJ δJc2, the calculation by Kitaev’s trick can be justified. Fig. 2c shows NCCN calculated by diagonalization (line plot) and KPM (scatter plot). These two methods agree well. From the data of KPM we extrapolated the thermodynamic limit. The finite-size data are fit by exponential functions, and extracted the converged value for L. The extrapolation is plotted in Fig. 2d and the thermodynamic limit is shown in a line plot with a ribbon. Around δJ = δJc1 = 0.05, NCCN deviates from unity, which suggests the existence of the topological transition into the gapless phase. For the calculations we took Nsample = 360 quenched disorder samples and used M = 1024, and R = 24, where M is the expansion order of KPM.

Local density of states

When δJ δJc2, free Majorana fermions are only relevant low-energy excitations, and we can use many tools of free fermions to discuss properties of the transition, such as DOS and a localization length. DOS around the ground state can be measured from the information of the 0-flux sector. As is often the case, we only calculated local density of states (LDOS), instead. LDOS for a site i is defined as follows27:

$${\rho }_{i}(\varepsilon )=\mathop{\sum }\limits_{k = 1}^{N}{\left|\langle i| k\rangle \right|}^{2}\delta (\varepsilon -{\varepsilon }_{k}),$$
(11)

where \(\left|i\right\rangle\) is a site basis, and \(\left|k\right\rangle\) is an eigenstate of the Hamiltonian with an energy ε = εk.

The nonlocality of Majorana fermions does not matter as averaged LDOS approximates DOS well enough. Both of the quantities are easily computed using KPM, and LDOS is enough for our purpose. The ratio of the arithmetic and geometric means of LDOS also works as the order parameter of an Anderson transition instead of the localization length. From the gapped Dirac spectrum (see Fig. 3a) the LDOS becomes gapless as the disorder strength increases. In the gapless region, DOS behaves linearly around ε = 0 (see Fig. 3b–d). The localization in Fig. 3b–d is clear from the discrepancy between the arithmetic and geometric averages of LDOS. Details are included in Supplementary Methods.

Fig. 3: Arithmetic and geometric means of LDOS.
figure 3

a δJ = 0.0. b δJ = 0.1. c δJ = 0.2. d δJ = 0.3. For every figure L = 100, R = 24, and Nsample = 360.

Discussion

Even in a finite-size calculation, the transition between KSL and AKSL was well-observed and the schematic phase diagram in Fig. 1c was confirmed. From the extrapolation, δJc1 is very small and δJc1/J ~ 0.05. δJc1/J α0 ~ 0.26 suggests the fragility of KSL. The fragility potentially explains the observed sample dependence of the half-integer quantized thermal Hall effect in α-RuCl340. There is a possibility that δJc1/J → 0 as \(h/{\Delta }_{\min }\to 0\), which means that the bond disorder in the Kiteav model is a relevant perturbation24. After the transition, the V-shaped behavior of DOS completely agrees with an observed linear low-energy DOS for H3LiIr2O6 with an applied magnetic field5.

It was proposed that the spin-polarized scanning tunneling microscopy (SP-STM) could be a local probe for edge states of the Kitaev model41. During the crossover to AKSL, edge states should disappear gradually and thus SP-STM should be a good probe for the topological transition.

The \(h/{\Delta }_{\min }\to 0\) limit must be taken carefully. A small \(h/{\Delta }_{\min }\) suffers from the finite-size effect, so we need to increase N according to the \(h/{\Delta }_{\min }\to 0\) limit, as described in Supplementary Methods and Supplementary Fig. 2. Whether or not δJc1/J → 0 in this limit is an important future problem. We note that the vortex disorder is known to be relevant, so the introduction of random vortices may change the universality42.

The fragility of the quantization has many implications to experiments. Disorder always exists in real materials, especially in any 2D layered system, and even in clean samples of α-RuCl3 stacking faults must exist43. Thus, the situation is quite similar to that of the fractional quantum Hall effect (FQHE). The observation of FQHE requires a really clean sample, and the recently observed quantized thermal Hall current of FQHE is more sensitive to disorder44. The sensitivity also resembles unconventional superconductors45. It might be universal in strongly correlated systems. Thus, we need to reconsider the importance of cleanness for the topological order in general. Further discussions can be found in Supplementary Discussion.

Note added

After the early version of the present paper on arXiv, a finite-temperature calculation of the disordered Kitaev models has been reported46.

Methods

We would like to introduce the KPM27,47,48. KPM is used twofold in this work. The first is for the vison gap calculation and the second is for the Chern number calculation. The calculation cost for the former is O(N2) and for the latter O(N).

Vison gap calculation

First, let us consider a Majorana Hamiltonian with the following form.

$$H=\frac{1}{4}\mathop {\sum}\limits_{j,k}{{\mathcal{H}}}_{jk}{c}_{j}{c}_{k},$$
(12)

where \({\mathcal{H}}\) is a Hermitian matrix. For Majorana fermions cj, \({\mathcal{H}}\) has a form \({\mathcal{H}}=iA\), where A is a real skew-symmetric matrix. From now on, we assume \({\mathcal{H}}\) to be the ones considered in the main text, either with or without a magnetic field. The eigenvalues of the N × N matrix \({\mathcal{H}}\) is denoted by Ek with k = 1, …, N.

A Green function can be expanded by a Chebyshev polynomial Tm(x) as follows:

$${G}_{jj}(E+i\varepsilon )=i\frac{{\tilde{\mu }}_{0}+2\mathop{\sum }\nolimits_{m = 1}^{M-1}{\tilde{\mu }}_{m}\exp [-im\arccos (E/s)]}{\sqrt{{s}^{2}-{E}^{2}}}.$$
(13)
$${\tilde{\mu }}_{m}={g}_{m}\langle j| {T}_{m}({\mathcal{H}}/s)| j\rangle .$$
(14)
$${g}_{m}=\frac{\sinh [\lambda (1-m/M)]}{\sinh \lambda },$$
(15)

where λ = 4.0 was used in the Lorentz kernel gm. M is the expansion order and m = 0, …, M − 1. ε is a small parameter which goes to 0 when M. The scaling factor s is necessary so that the spectrum of the Hamiltonian \({\mathcal{H}}/s\) falls within the domain of the Chebyshev polynomials [−1, 1]. We note that this expression is for diagonal components, but almost the same is true for offdiagonal components. We simply set \(s={E}_{\max }+0.1\), where \({E}_{\max }\) is the maximal eigenvalue.

Elements of Chebyshev moments \({T}_{m}({\mathcal{H}}/s)\) can be computed recursively by using Tm(x) = 2xTm−1(x) − Tm−2(x) and T2m+i(x) = 2Tm(x)Tm+i(x) − Ti(x) for i = 0, 1. The total O(N2) cost is required to compute all the necessary elements.

From the expanded Green function, we can compute the energy change by the local modification of the Hamiltonian \({\mathcal{H}}\to {\mathcal{H}}+\delta {\mathcal{H}}\). We define

$${\mathcal{D}}(E)=\det [1+G(E)\delta {\mathcal{H}}].$$
(16)

By extending this function to a complex number, the energy difference, i.e. vison gap Δvison, can be computed as follows:

$${\Delta }_{{\rm{vison}}}=\frac{1}{\pi }{\int\nolimits_{-\infty }^{\infty }}F(E)\mathop {\text{lim}}\limits_{\varepsilon \to 0}{\rm{Im}}\ {\mathrm{log}}\,[{\mathcal{D}}(E+i\varepsilon )]dE,$$
(17)

where F(E) is a Fermi-Majorana function at zero temperature.

$$F(E)=-\mathop {\text{lim}}\limits_{\beta \to \infty }\frac{1}{2}\tanh \frac{\beta E}{2}=\vartheta (-E)-\frac{1}{2}.$$
(18)

The evaluation of the integral in the Green function requires fast Fourier transformation (FFT) or discrete cosine transformation (DCT)27. Using this, the integral is reduced to a discrete weighted sum of the Fermi(-Majorana) function evaluated at some specific points. The discretization size \(\tilde{M}\) for the summation was set \(\tilde{M}=2M\) for simplicity. We compared M = 64, 128, 256, 512, 1024, and 2048. Among these, we found that M = 1024 has the best performance for our purpose, where the error is always about 0.01J.

KPM can reproduce the vison gap with an enough precision and reduce the computational cost to O(N) with a truncation algorithm49,50,51. However, later we found that the truncation causes a problem in our simulation, and thus we used the abovementioned O(N2) algorithm without a truncation27,47,48.

Chern number calculation

A Kubo formula for κxy at zero temperature is reduced to the generalized Thouless-Kohmoto-Nightingale-den Nijs (TKNN) formula52 for noninteracting Majorana Hamiltonians:

$${\kappa }^{xy}=\frac{\hslash {\pi }^{2}{k}_{B}^{2}T}{6{L}^{2}}\mathop{\sum }\limits_{m,n}\vartheta (-{\varepsilon }_{m})\frac{2{\rm{Im}}[\langle m| {v}_{1}| n\rangle \langle n| {v}_{2}| m\rangle ]}{{\left({\varepsilon }_{m}-{\varepsilon }_{n}\right)}^{2}},$$
(19)

where m and n label eigenvalues of \({\mathcal{H}}\), εm and εn, corresponding to eigenstates \(\left|m\right\rangle\) and \(\left|n\right\rangle\), respectively53,54. ϑ(x) is a Heaviside theta and \({v}_{\alpha }=i[{\mathcal{H}},{r}_{\alpha }]/\hslash\) is a velocity operator along the α-direction. We define a position operator rα for the nα-direction for α = 1, 2. A gravitomagnetic term should be added to derive this formula. We note the generalized formula without a translation symmetry was originally discussed by Kitaev1 using a flow of unitary matrices and extended by Kapustin and Spodyneiko55. This Kubo-TKNN formula is nothing but a real-space formulation of the Chern number calculation.

However, as mentioned in the main text, it is better to use a noncommutative Chern number (NCCN) Eq. (10)33 to compute the same quantity in the thermodynamic limit. Though Eq. (10) makes sense only in the thermodynamic limit, we can make use of a discretized version, which exponentially converges to the thermodyanamic limit by an artificial k-space quantization of a size L × L. This version can be derived by replacing the commutator33 as follows:

$$-i[{r}_{\alpha },{P}_{F}]\mapsto \mathop{\sum }\limits_{q = -Q}^{Q}{c}_{q}{e}^{-iq\Delta {r}_{\alpha }}{P}_{F}{e}^{iq\Delta {r}_{\alpha }},$$
(20)

where Δ = 2π/L, c0 = 0 and cq = −cq are determined to hold \(x-\mathop{\sum }\nolimits_{q = -L/2}^{L/2}{c}_{q}{e}^{iq\Delta x}=O({\Delta }^{L})\), and Q ≤ L/2. Thus, we can expect that these two methods to compute κxy may agree with a large L, while the Hall conductivity and the Chern number are a priori different quantities. We fixed Q = 15 for L > 30 because otherwise the calculation cost becomes O(N3).

The spectral projector PF can be rewritten as:

$${P}_{F}=-\frac{1}{2\pi i}{\oint _{{\mathcal{C}}}}\frac{1}{{\mathcal{H}}-E}dE,$$
(21)

where \({\mathcal{C}}\) is a contour which encloses every negative eigenvalue of \({\mathcal{H}}\). The integrant is nothing but a Green function, and can be expanded by KPM.

It is better to use the following Fermi function instead of approximating a spectral projector directly.

$${P}_{F}^{{\rm{eff}}}=F({\mathcal{H}}/s)=\vartheta (-{\mathcal{H}}/s)-\frac{1}{2}.$$
(22)

Due to the particle-hole symmetry, this deformation also gives a correct Chern number. This expression can again be expanded by Chebyshev polynomials. Especially, the Fermi function has been expanded as

$${f}_{m}={\int\nolimits_{-1}^{1}}\frac{dx}{\pi \sqrt{1-{x}^{2}}}{T}_{m}(x)F(x).$$
(23)
$$F(x)={f}_{0}+2\mathop{\sum }\limits_{m = 1}^{{M}^{\prime}-1}{f}_{m}{T}_{m}(x),$$
(24)

where \({M}^{\prime}\) is a cutoff of the expansion for NCCN39. We used the Jackson kernel for the Chern number:

$${g}_{m}^{\prime}=\frac{({M}^{\prime}-m+1)\cos \frac{\pi m}{{M}^{\prime}+1}+\sin \frac{\pi m}{{M}^{\prime}+1}\cot \frac{\pi }{{M}^{\prime}+1}}{{M}^{\prime}+1},$$
(25)

where \(m=0,\ldots ,{M}^{\prime}-1\). We here used \({M}^{\prime}=512\), instead. Thus,

$${P}_{F}^{{\rm{eff}}} \sim {g}_{0}^{\prime}{f}_{0}{T}_{0}({\mathcal{H}})+2\mathop{\sum }\limits_{m = 1}^{{M}^{\prime}-1}{g}_{m}^{\prime}{f}_{m}{T}_{m}({\mathcal{H}}).$$
(26)

This can indeed be evaluated with a calculation cost of O(N). Detailed information for the O(N) approximation is included in Supplementary Methods.