Irrelevant corrections at the quantum Hall transition
Abstract
The quantum Hall effect is one of the most extensively studied topological effects in solid state physics. The transitions between different quantum Hall states exhibit critical phenomena described by universal critical exponents. Numerous numerical finite size scaling studies have focused on the critical exponent for the correlation length. In such studies it is important to take proper account of irrelevant corrections to scaling. Recently, Dresselhaus et al. [Phys. Rev. Lett. 129, 026801(2022)] proposed a new scaling ansatz, which they applied to the two terminal conductance of the Chalker-Coddington model. In this paper their proposal is applied to our previously reported data for the Lyapunov exponents of that model using both polynomial fitting and Gaussian process fitting.
keywords:
quantum Hall effect, critical exponent, finite size scalingKeith Slevin Tomi Ohtsuki
Keith Slevin
Department of Physics, Graduate School of Science, Osaka University, Toyonaka, Osaka 560-0043, Japan
Email Address:[email protected]
Tomi Ohtsuki
Division of Physics, Sophia University, Chiyoda, Tokyo 102-8554, Japan
Email Address:[email protected]
1 Introduction
Four decades after its discovery[1], the quantum Hall effect continues to attract considerable attention, both experimentally and theoretically. Its most striking feature is the plateau behaviour of the Hall conductance , which is quantised in integer multiple of
(1) |
Here is Planck’s constant, is the elementary charge, and is an integer topological number [2]. As a control parameter such as two dimensional electron density (i.e., gate voltage) or perpendicular magnetic field is varied, the Hall conductance changes from one quantised value to the other. The transition is accompanied by a power law divergence of a correlation length ,
(2) |
with a constant and a critical exponent that is expected to have a universal value.
In early finite size scaling analyses, the value of was estimated to be [3, 4]. The problem was later revisited and a significantly higher estimate [5, 6, 7, 8, 9, 10, 11, 12] was obtained (although a slightly smaller value, , was reported for an analysis of the Chern number [13]). The recent proposal of a conformal field theory [14, 15] with marginal scaling, i.e., with , has stimulated further numerical study of this problem[16, 17].
A major difficulty in finite size scaling analyses of the quantum Hall transition is the presence of slow irrelevant corrections that complicate the estimation of the critical exponent. In Ref. [17], the energy and system size dependence of two terminal conductance of the Chalker-Coddington model[18, 19] was calculated and a finite size scaling analysis performed. In that paper, Dresselhaus et al. proposed a method for taking into account the irrelevant correction based on a factorisation ansatz for the scaling function. In this paper, we re-analyse our previously published data for the Lyapunov exponents of the Chalker-Coddington model following their proposal.
2 Method
The data we re-analyse here is for the smallest positive Lyapunov exponent of the Chalker-Coddington model reported in Refs. [5, 10]. The simulations used the transfer matrix method[20, 21, 22]. A detailed explanation of this method is given in Ref. [23].
Very long quasi-one-dimensional strips were simulated. The parameters in the simulation are the transverse width of the strip and the energy relative to the centre of the Landau level measured in units of the Landau bandwidth. Here, is equal to the number of saddle points in the transverse direction of the strip. It is known that for this model the critical point is at .
Simulations were performed for and and . The output of each simulation was the estimate of and the standard deviation of that estimate for the given parameters and . The precision of the data for the Lyapunov exponents is 0.03% except for and where the precision is 0.05%. Further details can be found in Refs. [5, 10].
The analysis of the data involves fitting the system size and energy dependence of the dimensionless quantity to the finite size scaling form including corrections to scaling due to an irrelevant scaling variable [3, 24].
(3) |
where and are, respectively, the relevant and irrelevant scaling variables. These are distinguished by the sign of the corresponding exponent and . The scaling variables and are analytic functions of the energy . In addition, the relevant variable is zero at the critical point.
The method proposed by Dresselhause et al. is based on two assumptions. First is that the scaling function factorises as follows
(4) |
The second is that the dependence of the irrelevant variable can be neglected. With these assumptions
(5) |
Dresselhause et al. then note that the following ratio,
(6) |
obeys a single parameter scaling law, i.e., that the correction due to the irrelevant variable drops out of the analysis.
If we set in Eq. (4) we find
(7) |
Thus, one way that the presence of a correction due to an irrelevant variable should be manifest is a system size dependence of at the critical point. Such a dependence is clearly visible in the data from our simulations cf. Fig 2 of Ref. [10]. This system size dependence is removed from since
(8) |
by definition. However, this also means that the quantity
(9) |
which is related to the multi-fractal properties of the critical wavefunctions [18, 25, 26, 4], cannot be estimated by fitting .
For brevity in what follows, when fitting data for we drop the suffix from the relevant variable and write simply
(10) |
where the constant has been absorbed into a re-definition of the scaling function . The correlation length critical exponent is related to by
(11) |
3 Numerical results
In this section we report the results of finite size scaling fits to the data for obtained as described above. The standard deviations for the estimates of are calculated as described in the Appendix. The fractional error is in the range 0.04% to 0.07%. Since, for , by definition, data points for this energy are not used in the fitting. In our simulation, periodic boundary conditions were imposed in the transverse direction. The Lyapunov exponent is then an even function of . We report results for both polynomial fitting and Gaussian process fitting.
3.1 Polynomial fitting
To ensure that the fitting function is an even function of we make an even function and an odd function. Both and are expanded in Taylor series. The series are truncated at orders and respectively. Thus, we arrive at
(12) |
and
(13) |
The fitting parameters are the coefficients in the truncated Taylor series and the critical exponent . We fix to avoid ambiguity in the parameterisation and ensure universality of . The total number of parameters is
(14) |
The best fit is found by minimising the -squared statistic. The quality of the fit is evaluated using the goodness-of-fit probability . Fits with are rejected. Both and the standard deviations for the parameters were determined by generating and fitting 500 synthetic data sets. Details of this method are given in Ref. [23].
The results of polynomial fitting are tabulated in Table 1. We found that it was not possible to fit all our simulation data using the ansatz of Dresselhaus et al. When data for systems sizes were excluded acceptable fits were obtained, i.e., goodness-of-fit . When data with energy were excluded it was possible to fit our data when data for were also included.
In Figure 1 we plot the data and the finite size scaling fit corresponding to the first line of Table 1. Note that the error in the data is much smaller than the symbol size. In Figure 2 we plot the same data as a function of the relevant scaling variable to demonstrate the collapse of the data onto a single curve that characterises single parameter scaling.
In Figures 3 and 4 we do the same for the fit corresponding to the last line of Table 1 where data for are also included but the energy range is restricted. The estimate of the exponent varies by about 1% or so depending on the data included and the order at which the Taylor expansions of and are truncated. It is noticeable that excluding non-linearity of the relevant scaling variable leads to slightly lower estimates of the critical exponent.
156 | 5 | 165 | 0.2 | 3 | 6 | |||
156 | 7 | 157 | 0.3 | 5 | 8 | |||
116 | 3 | 123 | 0.3 | 1 | 4 | |||
116 | 5 | 98 | 0.8 | 3 | 6 | |||
138 | 5 | 160 | 0.1 | 1 | 8 | |||
138 | 5 | 150 | 0.2 | 3 | 6 |




3.2 Gaussian process fitting
For comparison we also report the results of Gaussian process fitting [27] of the data for . In this method a Gaussian process is used to represent the scaling function rather than a polynomial. Also non-linearity of the relevant scaling variable with respect to is neglected, i.e., . In this fit, in contrast to the polynomial fitting [Eq. (12)], the condition is not imposed. There are 4 fitting parameters: the critical exponent , and three parameters that characterise the Gaussian process. These parameters are varied to maximise the likelihood function [27, 28]. The standard deviation for the estimate of the exponent was obtained from 1000 Monte Carlo samples. Unfortunately a goodness-of-probability is not available for this method of fitting.
The results are tabulated in Table 2 and the scaling collapse of the data is shown in Figure 5. When the range of energy is restricted so that a polynomial fit that neglects non-linearity of the scaling variable is acceptable, i.e., for the polynomial fit, the estimates of for both methods agree. However, when the range of energy is widened so that non-linearity must be included to obtain an acceptable goodness-of-fit for the polynomial fit, the polynomial fit gives a larger estimate of than the Gaussian process fit. Comparing the first lines of Table 1 and Table 2 we see that non-linearity, which is included in the polynomial fit but not in the Gaussian process fit, gives a noticeably lower -squared.
156 | 353 | |||
116 | 117 | |||
138 | 160 |

4 Discussion
Dresselhaus et al. proposed a factorisation ansatz for the scaling function used in finite size scaling analysis of the quantum Hall transition. We applied this method to our previously published data for the smallest Lyapunov exponent of the Chalker-Coddington model. Provided data for smaller system sizes is excluded, we obtain acceptable fits of our data. The standard deviations quoted in Table 1 and Table 2 are statistical errors, i.e., they assume there are no systematic deviations between the model and data. Looking at Table 1 we see that the estimates of which include non-linearity of agree within the limits of error, i.e., intervals with overlap. However, the estimates where non-linearity are neglected are of the order of lower. All the estimates we obtained here are lower than the estimates in the range to obtained in our previous work Refs. [5, 10] with the same data where the factorisation ansatz of Dresselhaus et al. was not used. While these discrepancies are small, we think this indicates that the factorisation ansatz while giving acceptable fits for some ranges of data is possibly not exact.
Acknowledgements
K. S. and T. O. were supported by JSPS KAKENHI Grants No. 19H00658, and T. O. was supported by JSPS KAKENHI Grants No. 22H05114. K. S and T. O. would like to thank E. Dresselhaus, I. Gruzberg and Kenji Harada for fruitful discussion.
Appendix A Error calculation
For our Lyapunov simulation the data are independently normally distributed random variables, i.e., the output of a simulation is of the form
(15) |
where is a normally distributed random variable with mean zero and variance unity. The output of the simulation is the estimate of the expectation and an estimate of the standard deviation of that estimate. To perform the scaling analysis proposed by Dresselhaus et al., we need to estimate
(16) |
using data from two independent simulations performed for the same system size with different energies and . The obvious estimate is
(17) |
where we abbreviate as etc. For all our simulations
(18) |
Therefore, it is reasonable to expand the left hand side as follows
(19) |
Taking the expectation value we find
(20) |
We see that the bias in the estimate can be neglected for our data. To fit the data we also need an estimate for the standard deviation of this estimate. For the variance we find
(21) |
to the same order of approximation. Thus the standard deviation we need is
(22) |
References
- [1] K. v. Klitzing, G. Dorda, M. Pepper, Phys. Rev. Lett. 1980, 45 494.
- [2] D. J. Thouless, M. Kohmoto, M. P. Nightingale, M. den Nijs, Phys. Rev. Lett. 1982, 49 405.
- [3] B. Huckestein, Rev. Mod. Phys. 1995, 67 357.
- [4] F. Evers, A. D. Mirlin, Rev. Mod. Phys. 2008, 80 1355.
- [5] K. Slevin, T. Ohtsuki, Phys. Rev. B 2009, 80 041304.
- [6] M. Amado, A. V. Malyshev, A. Sedrakyan, F. Domínguez-Adame, Phys. Rev. Lett. 2011, 107 066402.
- [7] I. C. Fulga, F. Hassler, A. R. Akhmerov, C. W. J. Beenakker, Phys. Rev. B 2011, 84 245447.
- [8] J. P. Dahlhaus, J. M. Edge, J. Tworzydło, C. W. J. Beenakker, Phys. Rev. B 2011, 84 115133.
- [9] H. Obuse, I. A. Gruzberg, F. Evers, Phys. Rev. Lett. 2012, 109 206804.
- [10] K. SLEVIN, T. OHTSUKI, International Journal of Modern Physics: Conference Series 2012, 11 60.
- [11] I. A. Gruzberg, A. Klümper, W. Nuding, A. Sedrakyan, Phys. Rev. B 2017, 95 125414.
- [12] M. Puschmann, P. Cain, M. Schreiber, T. Vojta, Phys. Rev. B 2019, 99 121301.
- [13] Q. Zhu, P. Wu, R. N. Bhatt, X. Wan, Phys. Rev. B 2019, 99 024205.
- [14] R. Bondesan, D. Wieczorek, M. Zirnbauer, Nuclear Physics B 2017, 918 52.
- [15] M. R. Zirnbauer, Nuclear Physics B 2019, 941 458.
- [16] E. J. Dresselhaus, B. Sbierski, I. A. Gruzberg, Annals of Physics 2021, 435 168676, special Issue on Localisation 2020.
- [17] E. J. Dresselhaus, B. Sbierski, I. A. Gruzberg, Phys. Rev. Lett. 2022, 129, 2 026801.
- [18] J. T. Chalker, P. D. Coddington, Journal of Physics C: Solid State Physics 1988, 21, 14 2665.
- [19] B. Kramer, T. Ohtsuki, S. Kettemann, Physics Reports 2005, 417, 5 211 .
- [20] J. L. Pichard, G. Sarma, Journal of Physics C: Solid State Physics 1981, 14, 6 L127.
- [21] A. MacKinnon, B. Kramer, Phys. Rev. Lett. 1981, 47 1546.
- [22] A. MacKinnon, B. Kramer, Zeitschrift für Physik B Condensed Matter 1983, 53, 1 1.
- [23] K. Slevin, T. Ohtsuki, New Journal of Physics 2014, 16, 1 015012.
- [24] K. Slevin, T. Ohtsuki, Phys. Rev. Lett. 1999, 82 382.
- [25] J. L. Cardy, Scaling and renormalization in statistical physics, Cambridge University Press, Cambridge, 1966.
- [26] M. Janssen, Fluctuations and localization in mesoscopic electron systems, World Scientific, Singapore, 2001.
- [27] K. Harada, Phys. Rev. E 2011, 84 056704.
- [28] R. Yoneda, K. Harada, Y. Y. Yamaguchi, Phys. Rev. E 2020, 102 062212.