Paper The following article is Open access

Stability of exponentially damped oscillations under perturbations of the Mori-Chain

, , and

Published 16 August 2022 © 2022 The Author(s). Published by IOP Publishing Ltd
, , Citation Robin Heveling et al 2022 J. Phys. Commun. 6 085009 DOI 10.1088/2399-6528/ac863b

2399-6528/6/8/085009

Abstract

There is an abundance of evidence that some relaxation dynamics, e.g., exponential decays, are much more common in nature than others. Recently, there have been attempts to trace this dominance back to a certain stability of the prevalent dynamics versus generic Hamiltonian perturbations. In the paper at hand, we tackle this stability issue from yet another angle, namely in the framework of the recursion method. We investigate the behavior of various relaxation dynamics with respect to alterations of the so-called Lanczos coefficients. All considered scenarios are set up in order to comply with the 'universal operator growth hypothesis'. Our numerical experiments suggest the existence of stability in a larger class of relaxation dynamics consisting of exponentially damped oscillations. Further, we propose a criterion to identify 'pathological' perturbations that lead to uncommon dynamics.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

The apparent emergence of irreversibility from the underlying reversible theory of quantum mechanics is a long-standing puzzle that lacks an entirely satisfying answer to this day [1]. Over the course of the last decades, fundamental concepts like the 'eigenstate thermalization hypothesis' [24] and 'quantum typicality' [57] have crystallized, which give conditions under which isolated quantum systems eventually reach an equilibrium state. However, while these mechanisms ensure eventual equilibration, they make no statement in which manner the equilibrium state is actually reached, i.e., they do not narrow down the eligible routes to equilibrium. In contrast, it is evidently true that some relaxation dynamics, e.g., exponential decays, are much more common in nature than others, e.g., recurrence dynamics.

There are recent attempts to explain this prevalence with the idea that some dynamics are stable versus perturbations of the Hamiltonian. These attempts include, for example, investigations based on Hamiltonian perturbations on the level of random matrices [8]. Further, there exist advances that consider an entire ensemble of permissible perturbations [911]. It is then analytically shown that weak 'typical' perturbations lead to an exponential damping of the original dynamics [9]. This renders exponential decays stable since only the decay constant changes, whereas recurrence dynamics are exponentially suppressed.

With the paper at hand, we address the issue of stability of certain [classes of] relaxation dynamics in the framework of the recursion method [12, 13]. Central quantities that appear within this framework are the so-called Lanczos coefficients, real numbers that characterize the complexity growth of operators over the course of time. As will be presented in section 2, the Lanczos coefficients can be interpreted as hopping amplitudes in a tight-binding model. In this manner, many physical problems, like calculating correlation functions, can practically be reduced to a one-dimensional [finite or semi-infinite] chain, which we refer to as the 'Mori-chain' in the title of this paper. In the following, we consider perturbations on the level of Lanczos coefficients and examine the resulting effect on various kinds of relaxation dynamics.

The particular advantage of our approach within the recursion method framework is that all considered scenarios can be set up to directly comply with the universal operator growth hypothesis [14], which previous approaches have been lacking [8]. Said hypothesis concerns the asymptotic growth of the Lanczos coefficients and it basically states that the coefficients should eventually attain linear growth [with a logarithmic correction in one dimension]. The hypothesis is backed up by analytical as well as numerical evidence [1417].

The paper at hand is structured as follows: section 2 constitutes a preliminary section on the recursion method and the operator growth hypothesis. Afterwards, in section 3, the concrete strategy to study the stability of certain classes of dynamics is explained in detail. In section 4 we present and discuss our numerical results. We conclude in section 5.

2. Preliminaries: recursion method and operator growth hypothesis

In this section, we briefly recall the basics of the recursion method [12, 13] as well as the universal operator growth hypothesis [14]. The preliminary considerations of this section follow [14] rather closely. We consider a system described by some Hamiltonian ${ \mathcal H }$. An observable of interest represented by a Hermitian operator ${ \mathcal O }$ gives rise to a corresponding autocorrelation function

Equation (1)

where ${ \mathcal O }(t)={e}^{{\rm{i}}{ \mathcal H }t}{ \mathcal O }{e}^{-{\rm{i}}{ \mathcal H }t}$ is the time-dependent operator in the Heisenberg picture ( = 1). In the following, it is convenient to work directly in Liouville space, i.e., the Hilbert space of operators, and denote its elements ${ \mathcal O }$ as states $| { \mathcal O })$. Elements of the Liouville space evolve under the Liovillian ${ \mathcal L }=[{ \mathcal H },\cdot \,]$, i.e., $| { \mathcal O }(t))={e}^{{\rm{i}}{ \mathcal L }t}| { \mathcal O })$, similar to wave functions that evolve under the Hamiltonian ${ \mathcal H }$. Using the Liovillian superoperator, the autocorrelation function may be written as $C(t)=({ \mathcal O }| {e}^{{\rm{i}}{ \mathcal L }t}| { \mathcal O })$. The Liouville space is equipped with an infinite-temperature inner product $({{ \mathcal O }}_{1}| {{ \mathcal O }}_{2})=\mathrm{Tr}[{{ \mathcal O }}_{1}^{\dagger }{{ \mathcal O }}_{2}]$, which induces a norm via $| | { \mathcal O }| | =\sqrt{({ \mathcal O }| { \mathcal O })}$.

In the following, the central object of interest is the Liouvillian ${ \mathcal L }$ represented in a particular basis $\{| {{ \mathcal O }}_{n})\}$, the so-called Krylov basis. The Krylov basis is an orthonormal basis of the Liouville space and is routinely constructed as part of the Lanczos algorithm. In this basis, which is determined by some 'seed' observable ${ \mathcal O }$, the representation of the Liouvillian is tridiagonal. To initialize the algorithm, we take the normalized state $| {{ \mathcal O }}_{0})=| { \mathcal O })$, i.e., $({ \mathcal O }| { \mathcal O })=1$, and set ${b}_{1}=\parallel { \mathcal L }{{ \mathcal O }}_{0}\parallel $ as well as $| {{ \mathcal O }}_{1})={ \mathcal L }| {{ \mathcal O }}_{0})/{b}_{1}$. Then we iteratively compute

Equation (2)

The tridiagonal representation of the Liouvillian in the Krylov basis $\{| {{ \mathcal O }}_{n})\}$ is then given by

Equation (3)

where the Lanczos coefficients bn are real, positive numbers output by the algorithm. For our purposes, it is sufficient to assume a finite-dimensional space. Then, the algorithm halts at step n = d + 1, where d is the dimension of the Liouville space, and L is a d × d-matrix. Rewriting the Heisenberg equation of motion in the Krylov basis yields

Equation (4)

where we defined ${\varphi }_{n}:= {{\rm{i}}}^{-n}({{ \mathcal O }}_{n}| { \mathcal O }(t))$. The initial condition is given by φn (0) = δn0 and both φn and bn are set to zero by convention when n is 'out of bounds'. The above equation (4) takes the form of a discrete Schrödinger equation and can be numerically solved by familiar means of, e.g., exact diagonalization or iterative schemes like Runge-Kutta and Chebyshev polynomials.

As mentioned, the Lanczos coefficients bn can be interpreted as hopping amplitudes in a tight-binding model. Then, the correlation function coincides with the amplitude of the first site, i.e., C(t) = φ0(t).

For later reference, we introduce the spectral function Φ(ω) as the Fourier transform of the correlation function, i.e.,

Equation (5)

It can be shown that the Lanczos coefficients bn appear in the continued fraction expansion of Φ(ω).

Equation (6)

Consequently, there exists a (non-linear) one-to-one map between the Lanczos coefficients bn and the autocorrelation function C(t). Thus, a set of bn 's uniquely determines C(t) and vice versa.

Lastly, we present the universal operator growth hypothesis as brought forth in [14]. The hypothesis concerns the asymptotic behavior of the Lanczos coefficients bn and basically states that in generic, non-integrable systems the Lanczos coefficients of local, few-body observables grow asymptotically linear, i.e., above some n the growth is given by

Equation (7)

where α > 0 and γ are real constants and o(gn ) denotes some real sequence fn with ${\mathrm{lim}}_{n\to \infty }| {f}_{n}/{g}_{n}| =0$. In the special case of a one-dimensional system, the asymptotic growth is sub-linear due to an additional logarithmic correction [14].

3. Probing for stability

In this section, we devise our strategy to probe the stability of classes of dynamics with respect to certain types of perturbations. This section consists of the following parts: We present ways to find Lanczos coefficients corresponding to a chosen [class of] correlation functions [section 3.1]. Next, we present four physically motivated requirements that the Lanczos coefficients should fulfill [section 3.2]. Then, we discuss the particular type of perturbations we consider [section 3.3]. Finally, we introduce quantifiers that allow to evaluate the stability of the respective dynamics [section 3.4].

3.1. Dynamics and Lanczos coefficients

The preliminary section already touched on the relation between a correlation function C(t) and the corresponding Lanczos coefficients bn . As mentioned, this correspondence is one-to-one, thus, a set of bn 's uniquely determines C(t) and vice versa. However, the convergence can be quite subtle, i.e., there may be fairly similar dynamics with vastly different Lanczos coefficients. On the other side, similar Lanczos coefficients may lead to quite different dynamics.

In the numerical experiments below, we proceed as follows. First, we choose a correlation function $C^{\prime} (t)$ whose stability we want to probe. We may specify $C^{\prime} (t)$ as a concrete analytical function, or as a member of a class of functions, e.g., exponential decays. The exact Lanczos coefficients corresponding to $C^{\prime} (t)$ are denoted by ${b}_{n}^{{\prime} }$. These coefficients are, in general, unknown and obtaining them is quite a difficult task. Our objective is to find Lanczos coefficients bn corresponding to a correlation function C(t) that should be practically indistinguishable from $C^{\prime} (t)$ for all intents and purposes $[C(t)\simeq C^{\prime} (t)$]. These 'approximate' coefficients bn may be obtained in three different ways:

  • (i)  
    The exact coefficients ${b}_{n}^{{\prime} }$ may be analytically known. In this case ${b}_{n}={b}_{n}^{{\prime} }$ as well as $C(t)=C^{\prime} (t)$.
  • (ii)  
    The coefficients may be obtained via an educated guess. In this case, the bn may be quite different from ${b}_{n}^{{\prime} }$ while still $C(t)\simeq C^{\prime} (t)$.
  • (iii)  
    The coefficients may be 'reverse-engineered' from the correlation function $C^{\prime} (t)$. For details see appendix A.

In the following, we will omit the technical distinction between C(t) and $C^{\prime} (t)$, since we assume that Lanczos coefficients can be found for which the two correlation functions are practically indistinguishable.

3.2. Design of the Lanczos coefficients

In the following, we compare dynamics of correlation functions CA (t) and CB (t) with respect to their stability under a certain class of perturbations. Corresponding sets of Lanczos coefficients bA n and bB n may be obtained by one of the three methods mentioned in the last section. We demand that these Lanczos coefficients satisfy the following [partly physically motivated] requirements sufficiently well:

  • (i)  
    The Lanczos coefficients bX n should reproduce the respective correlation function CX (t) to an acceptable accuracy, where X = A, B (this is sort of obvious and has already been addressed above).
  • (ii)  
    The Lanczos coefficients bX n should comply with the universal operator growth hypothesis, i.e., they should eventually attain linear growth.
  • (iii)  
    The resulting correlation functions CA (t) and CB (t) should decay on more or less the same time scale. This is to ensure a fair comparison between the two, since faster dynamics are typically less affected by perturbations than slower ones.
  • (iv)  
    The Lanczos coefficients bA n and bB n should be similar in magnitude. In particular, we demand that the quantity ${\sum }_{n}{\left({b}_{n}^{X}\right)}^{2}$ is comparable for X = A, B. We show in appendix B that this sum is related to the spectral variance of the Hamiltonian. Hence, this condition fosters the notion that the two correlation functions CA (t) and CB (t) originate from different observables while the underlying Hamiltonians are quite similar. In practice, a value of
    Equation (8)
    close to unity is desirable.

3.3. Design of the perturbations

The considered perturbations are designed on the level of Lanczos coefficients. The coefficients bn corresponding to some correlation function C(t) will be slightly altered according to

Equation (9)

where the perturbed coefficients are denoted by ${\tilde{b}}_{n}$. Here, λ is the perturbation strength and vn is specified below. We show in appendix C that this particular form of the perturbation yields a sensible scaling of the perturbed Hamiltonian with λ.

In general, it is an intricate problem to determine how a perturbation in the form of vn corresponds to a perturbation ${ \mathcal V }$ on the level of Hermitian matrices. This is an issue that we do not attempt to tackle in this paper.

Here, we try to make as few assumptions as possible and model the perturbation vn as random numbers [with the only restriction of a minimal correlation length, see below]. Concretely, we set

Equation (10)

where the xk , yk are real, random numbers from a Gaussian distribution with zero mean and unit variance. They are normalized as ${\sum }_{k}{x}_{k}^{2}+{y}_{k}^{2}=1$.

The sum in equation (10) is capped at a number Nf. This corresponds to excluding shorter wavelength or higher frequencies, which induces a minimal correlation length in the coefficients ${\tilde{b}}_{n}$ [reddish noise]. No truncation at all, i.e., Nf = d, would be equivalent to adding uncorrelated random numbers [white noise] to the coefficients bn . This choice can be justified by invoking the tight-binding model interpretation in which the Lanczos coefficients represent hopping amplitudes between neighboring sites. Simply altering the hopping amplitudes in a random, uncorrelated manner leads to localization effects similar to Anderson localization [18]. Through unsystematic varying of Nf, we find that the transition from unlocalized to to localized behavior is quite sharp. Since the aim is to study the stability of certain relaxation dynamics, we choose Nf as large as possible, but small enough to avoid said localization effects. In practice, this amounts to Nfd/3. In section 4.3, we ease this restriction to identify 'pathological' perturbations.

3.4. Assessing stability

Once the perturbation has acted on the coefficients bn and yielded the perturbed coefficients ${\tilde{b}}_{n}$, we can calculate the corresponding perturbed dynamics $\tilde{C}(t)$ as laid out in section 2. Now, some tools are needed in order to judge how strongly the dynamics were altered by the perturbation. In particular, it is important to assess to what extent the perturbed dynamics still falls into the original class of functions to which the unperturbed dynamics belonged. For example, we may ask if $\tilde{C}(t)$ is still exponential, given that the original dynamics C(t) was exponential [the decay constants may differ]. To this end, we attempt to fit the perturbed dynamics with a function that also described the unperturbed dynamics, e.g., we may try to fit $\tilde{C}(t)$ with f(t) = Aeμ t . In practice, the fit is obtained via a standard fitting routine.

Before we introduce quantifiers that measure the effect of the perturbation, some remarks on the fitting ansatz are in order. Correlation functions are symmetric in time, thus, all odd moments necessarily vanish. In particular, correlation functions have zero slope at t = 0. Further, we normalize the correlation function to unity at t = 0. However, for fitting we choose functions that lack these features, i.e., the above exponential ansatz may yield an A that slightly differs from unity. Further, the slope of an exponential at t = 0 is never zero. This 'negligence' is due to the fact that we are more interested in an overall description of relaxation dynamics, rather than in the details of the short-time behavior.

We assess the quality of a fit f(t) by calculating 'how far off' it is from the given perturbed dynamics. Concretely, we define a measure of the 'error' or 'deviation' epsilon by the expression

Equation (11)

Here, the respective functions are evaluated at available points in time tn = n δ t, where δ t is the time step used to solve the equation of motion. The upper bound Neq corresponds to a time at which the dynamics in question has seemingly equilibrated [19].

To get a feeling of which numerical value of epsilon constitutes a good or bad fit, we refer to the exemplary numerical data displayed in figures 3, 4, 8 and 9.

We introduce a second quantifier σ that measures how strongly the unperturbed dynamics are altered due to the perturbation in the first place.

Equation (12)

The construction of this quantity is similar to the construction of the quantity epsilon that measures the fit quality.

4. Numerical analysis

In this section, we apply the presented strategy to certain relaxation dynamics. In particular, we investigate and compare the stability of two classes of dynamics. The first class in question is the class of damped oscillations, whose damping is due to an ordinary exponential factor. This class also includes simple exponential decays, which can be viewed as damped oscillations with zero frequency. These dynamics are ubiquitous in nature. For example, slow exponential dynamics may commonly arise whenever a system interacts weakly with an environment or whenever long-wavelength Fourier components of spatial densities of conserved quantities are considered. Further, exponentially damped oscillations are routinely observed as short-wavelength components.

The possible choices for the second class of dynamics are of course manifold. Since an exhaustive analysis is impossible, we decide on a 'Gaussianized' version of the first class. This is supposed to mean that the class includes Gaussian decays as well as oscillations damped by Gaussian factors.

The restriction of the second class of course stifles any aspiration of generality. Thus, the following results should be viewed as mere numerical experiments that corroborate the existence of stability in the first class.

Concretely, in section 4.1, we compare exponential and Gaussian decay. In section 4.2, the comparison is between two damped oscillations, one with an exponential damping factor and the other with a Gaussian damping factor. Lastly, in section 4.3, we identify 'pathological' perturbations that destroy the considered dynamics.

Throughout the next sections, we set the dimension of the Liouville space to d = 10000 and the frequency cut-off to Nf = 3333 ≈ d/3. We repeat the strategy N = 1000 times by drawing random numbers xk , yk and present statistics for the quantifiers epsilon and σ.

4.1. Exponential versus Gaussian decay

In the first round of our stability investigation, an exponential decay competes against a Gaussian decay. Following the strategy laid out in the previous sections, the first step is to find suitable Lanczos coefficients that comply with the four conditions presented in section 3.2.

We begin with the Gaussian decay. The Lanczos coefficients that exactly correspond to Gaussian decay of the form $C(t)={e}^{-{t}^{2}/2}$ are analytically known [method one in section 3.1)], i.e., ${b}_{n}=\sqrt{n}$ [20]. Nevertheless, the second condition, the compliance with the operator growth hypothesis, needs to be satisfied. To this end, we choose a cutoff-point n at which the coefficients are continued in a linear fashion. This yields the following Lanczos coefficients ['g' = 'Gaussian']

Equation (13)

The parameters α and γ are determined by the requirement of a smooth transition from square-root growth to linear growth, i.e., $\alpha =1/(2\sqrt{{n}^{\star }})$ and $\gamma =\sqrt{{n}^{\star }}/2$. The change from square-root to linear growth for n > n does not strongly affect the 'Gaussianity' of the correlation dynamics [compare figure 2] such that condition one remains fulfilled.

Next, coefficients for an (approximate) exponential decay need to be found. As mentioned, a correlation function can never be truly exponential, since it necessarily features zero slope at t = 0. Thus, we only require the exponential decay to be present after a short Zeno-time, which is usually exceedingly short compared to the relaxation time [21, 22]. We achieve an approximate exponential decay via an educated guess [method two in section 3.1]. Slow dynamics are characterized by a relatively small first coefficient b1, followed by a jump to a larger remainder of coefficients bn≥2. Thus, we make the following ansatz for the coefficients ['e' = 'exponential']

Equation (14)

The parameter a is set to 1.2, a similar magnitude as the first Gaussian coefficient ${b}_{1}^{{\rm{g}}}$. The other parameters α and γ are the same as in the Gaussian case. Thus, condition two, i.e., compliance with the universal operator growth hypothesis, is fulfilled. The Lanczos coefficients as defined in equation (13) and equation (14) are depicted in figure 1. Since both sets of coefficients coincide for n > n, condition four is satisfied as the quantity $q({b}_{n}^{{\rm{g}}},{b}_{n}^{{\rm{e}}})=0.999993$ is sufficiently close to unity. The corresponding dynamics are depicted in figure 2. The ansatz for the coefficients ${b}_{n}^{{\rm{e}}}$ in equation (14) indeed yields a nice exponential decay. A fit of the form Aeλ t with A = 1.02 and λ = 0.24 captures the dynamics quite well. Thus, condition one is sufficiently fulfilled for the exponential decay as well.

Figure 1.

Figure 1. Lanczos coefficients corresponding to Gaussian decay (red) and exponential decay (blue). The square-root growth transitions to linear growth at n = n.

Standard image High-resolution image
Figure 2.

Figure 2. Unperturbed correlation functions C(t) calculated from the respective Lanczos coefficients. Dashed curves indicate an exponential fit (cyan) and an exact Gaussian ${e}^{-{t}^{2}/2}$ (black).

Standard image High-resolution image

Further, we can extract from figure 2 that both dynamics decay on more or less the same time scale. Therefore, we also view condition three as satisfied. If any, the Gaussian dynamics is faster and thus less prone to perturbations. Now that we have checked all four conditions presented in section 3.2, we are ready to apply the perturbation as laid out in section 3.3. We set λ = 0.5. Three exemplary perturbed dynamics with respective fits for the exponential case are depicted in figure 3.

Figure 3.

Figure 3. Three exemplary correlation functions originating from the perturbed exponential Lanczos coefficients ${\tilde{b}}_{n}^{{\rm{e}}}$. Dashed, black lines indicate exponential fits. The quantifier epsilon of all three dynamics is relatively close to the mean value, see figure 5. Inset: corresponding perturbed Lanczos coefficients.

Standard image High-resolution image

It is evident that the perturbed dynamics are still nicely described by an exponential, only the decay constant changes. For the Gaussian decay, three exemplary perturbed dynamics with respective fits are depicted in figure 4. Two displayed perturbed Gaussian curves feature oscillations, which can not possibly be captured by a Gaussian fit ansatz. The three depicted fits are much worse than in the exponential case. Insets show corresponding Lanczos coefficients.

Figure 4.

Figure 4. Three exemplary correlation functions originating from the perturbed Gaussian Lanczos coefficients ${\tilde{b}}_{n}^{{\rm{g}}}$. Dashed, black lines indicate Gaussian fits. The respective values of the quantifier epsilon for these three particular dynamics are relatively close to the mean value depicted in figure 5 by a dashed, red line. Inset: corresponding perturbed Lanczos coefficients.

Standard image High-resolution image

The impression from these six exemplary curves is confirmed in figure 5, which shows histograms of deviations epsilon of the fits from the respective perturbed dynamics. The symbol Ω denotes the number of values epsilon within a bin of size 5 · 10−4. There is a clear division between the exponential cases (blue) and the Gaussian cases (red).

Figure 5.

Figure 5. Histogram of the fit quality measure epsilon with a bin size of 5 · 10−4. Dashed lines indicate respective mean values. The stability of the exponential decay is evident. In contrast, the Gaussian decay does not seem to be stable, as the deviations epsilon become quite large. Inset: scatter plot of all points (σi , epsiloni ). Both dynamics are equally affected by the perturbation.

Standard image High-resolution image

The deviations epsilon of the exponential decays are much smaller than those of the Gaussian decays. In particular, the mean value ${\bar{\epsilon }}_{{\rm{e}}}=0.002$ (indicated by a blue, dashed line) in the exponential case is about twenty times smaller than in the Gaussian case ${\bar{\epsilon }}_{{\rm{g}}}=0.042$. For a visualization of this comparison see figures 3 and 4, whose exemplary curves feature values of epsilon close to the respective mean values.

We have to be aware that this apparent stability of the exponential may be due to the possibility that exponential decays are generally less affected by our constructed perturbation than the Gaussian decays. To check this possibility and to show that this is not the case, we consider the quantifier σ, which measures the difference between the perturbed and the unperturbed dynamics. The inset of figure 5 depicts a scatter plot of all 1000 pairs (σi , epsiloni ). If one decay would be consistently less affected than the other, a cluster of points to the left [small values of σ] would emerge. This is evidently not the case, as the distribution along the horizontal axis is relatively similar for both decays. The existence of the diagonal edge is expected, since there can be no fit that is 'farther away' from the perturbed dynamics than the unperturbed dynamics itself, which is, of course, also an eligible candidate for fitting.

Thus, we conclude that exponential decays is indeed stable with respect to perturbations [as in equation (10)]. In contrast, Gaussian decays seem to be quite unstable. This is the first main result of the paper at hand.

4.2. Damped oscillations: exponential versus Gaussian

In the second round, we investigate and compare two types of damped oscillations, one with an exponential damping factor, the other with a Gaussian damping factor. As in the last section, first, two sets of Lanczos coefficients need to be found that satisfy all four conditions imposed in section 3.2.

We again begin with the Gaussian case. The coefficients ${b}_{n}^{\mathrm{gdo}}$ ['gdo' = 'Gaussian damped oscillation'] that correspond to an oscillation damped by a Gaussian factor a neither analytically known, nor can we make an educated guess. This only leaves the third method, in which we 'reverse-engineer' the coefficients from the dynamics itself. To this end, we choose a particular correlation function $C(t)={e}^{-{t}^{2}/8}\cos (2t)$ and proceed as detailed in appendix A. This procedure allows us to obtain about 50 Lanczos coefficients. After that, the coefficients are continued 'by hand' in a manner that first, respects the pattern exhibited by the coefficients so far, and second, eventually becomes linear.

For the exponentially damped oscillation we can again make an educated guess [method two] for the coefficients ${b}_{n}^{\mathrm{edo}}$ ['edo' = 'exponentially damped oscillation'].

We set the first two coefficients b1, b2 to some values (b1 = 2.0, b2 = 1.6), followed by a jump to larger coefficients bn≥3, which then grow in a linear fashion. The slope is determined by and coincides with the slope in the Gaussian case.

The unperturbed Lanczos coefficients and corresponding correlation functions are depicted in figure 6 and figure 7, respectively. The 'reverse-engineered and continued by hand' coefficients ${b}_{n}^{\mathrm{gdo}}$ indeed reproduce the given correlation function $C(t)={e}^{-{t}^{2}/8}\cos (2t)$. Further, the guess for ${b}_{n}^{\mathrm{edo}}$ yields a nice exponentially damped curve, i.e., the fit ansatz $\tilde{C}(t)={{Ae}}^{-\mu t}\cos (\omega t-\phi )$ with A = 1.04, μ = 0.57, ω = 2.19 and ϕ = − 0.32 captures the dynamics quite well. Therefore, condition one is satisfied for both cases. The two sets of coefficients were designed to eventually attain linear growth, thus condition two is fulfilled. It is evident from figure 7 that both dynamics decay on similar time scales, rendering condition three satisfied. Lastly, the quantity $q({b}_{n}^{\mathrm{gdo}},{b}_{n}^{\mathrm{edo}})$ practically equals unity. Thus, all four conditions are met.

Figure 6.

Figure 6. Lanczos coefficients corresponding to an oscillation damped by a Gaussian (red) and an oscillation damped by an exponential (blue).

Standard image High-resolution image
Figure 7.

Figure 7. Unperturbed correlation functions C(t) calculated from the respective Lanczos coefficients. Dashed curves indicate a fit of an exponentially damped oscillation (cyan) and the oscillation damped by a Gaussian ${e}^{-{t}^{2}/8}\cos (2t)$ (black).

Standard image High-resolution image

Next, we switch on the perturbation and set λ = 0.1. Exemplary perturbed dynamics are displayed in figures 8 and 9, respectively. The perturbed dynamics are fitted with the 4-parametric ansatz $\tilde{C}(t)={{Ae}}^{-\mu t}\cos (\omega t-\phi )$ in the exponential and $\tilde{C}(t)={{Ae}}^{-\mu {t}^{2}}\cos (\omega t-\phi )$ in the Gaussian case. While the perturbed coefficients in the exponential case still lead to correlation functions that are within the class of exponentially damped oscillations [with different decay constants μ and frequencies ω], the same can not be said for the Gaussian case, where the perturbed dynamics decay too slowly and can not by captured by the fit ansatz above.

Figure 8.

Figure 8. Three exemplary correlation functions originating from the perturbed Lanczos coefficients ${\tilde{b}}_{n}^{\mathrm{edo}}$. Dashed, black lines indicate 4-parametric fits of exponentially damped oscillations. The quantifier epsilon of all three dynamics is relatively close to the mean value, see figure 10. The perturbation not only changes the decay constant, but also the frequency. Inset: corresponding perturbed Lanczos coefficients.

Standard image High-resolution image
Figure 9.

Figure 9. Three exemplary correlation functions originating from the perturbed Lanczos coefficients ${\tilde{b}}_{n}^{\mathrm{gdo}}$. Dashed, black lines indicate 4-parametric fits of oscillations damped by a Gaussian. The quantifier epsilon of all three dynamics is relatively close to the mean value, see figure 10. Inset: corresponding perturbed Lanczos coefficients.

Standard image High-resolution image

The impression from the six exemplary curves is confirmed in figure 10, which displays histograms of the quantifier epsilon. Again, the division into two distinct peaks is evident. The respective mean values ${\overline{\epsilon }}_{\mathrm{gdo}}=0.026$ and ${\overline{\epsilon }}_{\mathrm{edo}}=0.006$ are indicated as dashed, vertical lines and differ by a factor of about five. For a visualization of what these values mean for the fit quality, see figures 8 and 9, whose curves feature values of epsilon close to the respective mean values. Thus, we conclude that the exponentially damped dynamics are quite stable and, in particular, more stable than the oscillations damped by a Gaussian. The disparity between the two is not as strong as for the decays in the previous section [however, different values of λ may not be directly comparable.]

Figure 10.

Figure 10. Histogram of the fit quality measure epsilon with a bin size of 5 · 10−4. Dashed lines indicate respective mean values. The stability of the exponentially damped oscillations is evident. In contrast, the Gaussian counterpart does not seem to be stable, as the deviations epsilon become quite large. Inset: scatter plot of all points (σi , epsiloni ). Both dynamics are more or less equally affected by the perturbation.

Standard image High-resolution image

The inset of figure 10 again shows a scatter plot of all points (σi , epsiloni ). The blue cluster of dots is a little more concentrated to small values of σ, indicating that the exponentially damped oscillations are a little less altered by the perturbation than their Gaussian counterpart. However, the marginal distribution over σ is still less partitioned into two peaks than the one over epsilon [which is the data in the histogram itself]. Hence, the stability of exponentially damped oscillations is still due to the nature of the dynamics and not due to a smaller effect of the perturbation on the dynamics.

We conclude that exponentially damped oscillations are stable with respect to perturbations [as in equation (10)]. In contrast, oscillations damped by a Gaussian seem to be quite unstable. This is the second main result of the paper at hand.

4.3. Pathological perturbations

In this section, we lift the restriction imposed on the vn earlier, which was that frequencies were cut at Nfd/3. Instead, we include all frequencies in the construction of the perturbation, which amounts to setting Nf = d. This case simply corresponds to adding uncorrelated random numbers to the unperturbed coefficients bn . Otherwise, the strategy is pursued in the same manner as above.

Three exemplary dynamics for exponential decays are depicted in figure 11 [for conciseness, we refrain from showing more exemplary data for the other cases]. As is evident, the exponential decay is completely absent and replaced by quite irregular dynamics, which do not seem to reach an equilibrium [Neq is set as large as possible [19] in these cases]. These curves hint at the presence of localization effects, which prevent the 'particle' [in the tight-binding picture] to leave the first site.

Figure 11.

Figure 11. Three exemplary dynamics originating from perturbed Lanczos coefficients ${\tilde{b}}_{n}^{{\rm{e}}}$. The perturbation consists of uncorrelated random numbers, i.e., Nf = d in equation (10). The exponential decay is completely destroyed and the correlation function does not seem to equilibrate at all. Inset: corresponding perturbed Lanczos coefficients.

Standard image High-resolution image

A histogram of the Gaussian versus exponential data can be viewed in figure 12. The mean values read ${\bar{\epsilon }}_{{\rm{g}}}=0.15$ in the Gaussian case and ${\bar{\epsilon }}_{{\rm{e}}}=0.12$ in the exponential case, which is some orders of magnitude larger than before when shorter wavelengths were excluded. Further, the accumulation of points along the diagonal in the inset suggests that the curves no longer resemble their original form, which is expected judging from figure 11.

Figure 12.

Figure 12. Histogram of the fit quality measure epsilon with a bin size of 5 · 10−3. Dashed lines indicate respective mean values. For both oscillating cases the deviations epsilon are quite large. Thus, both relaxation dynamics are unstable with respect to the perturbation including all wave lengths. Inset: scatter plot of all points (σi , epsiloni ). Points accumulate along the diagonal edge indicating that the respective dynamics are heavily altered due to the perturbation.

Standard image High-resolution image

The histogram for the oscillating cases is displayed in figure 13. The mean deviations read ${\bar{\epsilon }}_{\mathrm{gdo}}=0.035$ and ${\bar{\epsilon }}_{\mathrm{edo}}=0.025$. These values are comparable to the Gaussian case in figure 10. However, since the quantifier epsilon measures more or less the deviation 'per time step', the comparison of values $\bar{\epsilon }$ between dynamics that do and do not equilibrate can be void of meaning. In practice, judging from all figures displaying exemplary curves, a value epsilon of about one percent corresponds to fits that 'look good to the eye'.

Figure 13.

Figure 13. Histogram of the fit quality measure epsilon with a bin size of 5 · 10−4. Dashed lines indicate respective mean values. For both the exponential and the Gaussian case the deviations epsilon are extremely large. Thus, both relaxation dynamics are unstable with respect to the perturbation including all wave lengths. Inset: scatter plot of all points (σi , epsiloni ). Points accumulate along the diagonal edge indicating that the respective dynamics are heavily altered due to the perturbation.

Standard image High-resolution image

Based on these observations, we specify a property of perturbations that lead to non-generic or pathological dynamics. Namely, a perturbation is 'untypical', if it gives rise to Lanczos coefficients that vary non-smoothly, i.e., there is no minimal correlation length within the coefficients. Again, what this means on the level of Hermitian matrices is difficult to assess and beyond the scope of this work. This is the third main result of the paper at hand.

5. Conclusion

In this paper, we performed numerical experiments to probe the stability of two classes of relaxation dynamics. The first class consisted of exponentially damped oscillations, which also includes exponential decays. The second class was chosen as a Gaussian counterpart to the first class, i.e., including Gaussian decays and oscillations damped by a Gaussian.

The whole strategy was formulated in the framework of the recursion method, in particular, the perturbations were constructed as an alteration of the Lanczos coefficients. Unperturbed coefficients bn and perturbation vn were chosen to satisfy certain physically motivated conditions.

The first main message of the paper at hand is that the exponential class of dynamics is relatively stable under the considered perturbations. In contrast, the Gaussian counterpart is found to be quite unstable. These findings confirm and extent upon previous results based on random matrices [8], which did not comply with the operator growth hypothesis.

We want to emphasize that within this work the main focus should be put on the stability of the former class of relaxation dynamics, as these are ubiquitous in nature, examples are given at the beginning of section 4.

The investigation of the latter class should just be taken as an exemplary comparison. In fact, the choice of a Gaussian counterpart to the first class is quite arbitrary. Any number of relaxation dynamics could have been investigated instead.

The second main message of the paper at hand concerns the nature of the perturbations themselves. Not only, but also in the context of the works on 'typicality of perturbations' [911], it could be interesting to find properties of perturbations that lead to non-generic, 'pathological' dynamics. Here, we identified such a criterion, namely that the perturbation should yield Lanczos coefficients whose minimal correlation length is still above some threshold value.

As is evident from figure 11, including short correlations seems to lead to unorthodox dynamics. The displayed curves do not seem to reach an equilibrium, at least on the available time scale. Rather, localization-like effects are introduced, which cause some part of the wave function to remain on the first site. On the other hand, sufficiently smooth coefficients alter the relaxation dynamics in a 'controlled yet non-trivial' manner.

Perturbations in various numerical investigations based in random matrices [8, 9] as well as spin lattice models [23, 24] seem to naturally possess the above specified property. Hence, a more systematic investigation on the existence of a minimal correlation length in the coefficients for realistic setups could be a possible prospect for future research.

Acknowledgments

This work was supported by the Deutsche Forschungsgemeinschaft (DFG) within the Research Unit FOR 2692 under Grant No. 397 107 022 (GE 1657/3-2) and No. 397 067 869 (STE 2243/3-2).

Data availability statement

All data that support the findings of this study are included within the article (and any supplementary files).

Appendix A.: Reverse-engineering bn from C(t)

In this section, we present the procedure to reverse-engineer the Lanczos coefficients bn from a given correlation function C(t). The idea is to employ the Lanczos algorithm from section 2, but on the level of spectral functions Φ(ω) rather than on the level of observables $| { \mathcal O })$. The inner product of operators turns into an inner product of functions, i.e.,

Equation (A1)

Further, the application of the Liouvillian ${ \mathcal L }$ to an operator ${ \mathcal O }$ corresponds to a multiplication of the respective function in Fourier space with − ω [since the commutator with ${ \mathcal H }$ corresponds to a time derivative, which is equivalent to a multiplication with iω in Fourier space],

Equation (A2)

In practice, we choose a specific correlation function C(t) and calculate its Fourier transform Φ(ω). The initial 'seed' function is set to the (normalized) $\sqrt{{\rm{\Phi }}(\omega )}$. Then, the Lanczos algorithm operates as laid out in section 2. In this manner, about 50 coefficients can be obtained before numerical instabilities become too pronounced.

Appendix B.: Link to spectral width

The eigenvalue equation for the Hamiltonian reads ${ \mathcal H }| {E}_{i}\rangle ={E}_{i}| {E}_{i}\rangle $, where Ei and ∣Ei 〉 denote eigenvalues and eigenstates, respectively. The corresponding eigenvalue equation for the Liouvillian superoperator is given by ${ \mathcal L }{{ \mathcal M }}_{\beta }={{ \mathcal E }}_{\beta }{{ \mathcal M }}_{\beta }$ with eigenvalues ${{ \mathcal E }}_{\beta (i,j)}={E}_{i}-{E}_{j}$ and eigenoperators ${{ \mathcal M }}_{\beta (i,j)}=| {E}_{i}\rangle \langle {E}_{j}| $. Without loss of generality, we can set $\mathrm{Tr}[{ \mathcal H }]={\sum }_{i}{E}_{i}=0$. Further, we denote the dimension of the Hilbert space by ${d}_{{ \mathcal H }}$, i.e., ${d}_{{ \mathcal H }}^{2}=d$. Then we have

Equation (B1)

Thus, we see that the quantity in question is indeed linked to the spectral width of the Hamiltonian.

Appendix C.: Scaling with perturbation strength

Consider a Hamiltonian ${ \mathcal H }={{ \mathcal H }}_{0}+\lambda { \mathcal V }$ consisting of an unperturbed part ${{ \mathcal H }}_{0}$ and a perturbation $\lambda { \mathcal V }$ with $\mathrm{Tr}[{{ \mathcal H }}_{0}{ \mathcal V }]=0$. For the spectral variance of ${ \mathcal H }$ it holds true that

Equation (C1)

On the other hand, with the particular choice of perturbation in equation (9), we get that

Equation (C2)

The term 2λ bn vn is negligible due to the specific choice of vn . Recalling the relation in equation (B1), we see from equation (C2) that the scaling in equation (9) reproduces the scaling in equation (C1) and is therefore appropriate.

Please wait… references are loading.