Skip to content
Publicly Available Published by De Gruyter April 28, 2020

Predicting Imperfect Echo Dynamics in Many-Body Quantum Systems

  • Lennart Dabelow ORCID logo and Peter Reimann EMAIL logo

Abstract

Echo protocols provide a means to investigate the arrow of time in macroscopic processes. Starting from a nonequilibrium state, the many-body quantum system under study is evolved for a certain period of time τ. Thereafter, an (effective) time reversal is performed that would – if implemented perfectly – take the system back to the initial state after another time period τ. Typical examples are nuclear magnetic resonance imaging and polarisation echo experiments. The presence of small, uncontrolled inaccuracies during the backward propagation results in deviations of the “echo signal” from the original evolution and can be exploited to quantify the instability of nonequilibrium states and the irreversibility of the dynamics. We derive an analytic prediction for the typical dependence of this echo signal for macroscopic observables on the magnitude of the inaccuracies and on the duration τ of the process, and verify it in numerical examples.

1 Introduction

Explaining the irreversibility of processes in macroscopic systems based on the time-reversible laws governing their microscopic constituents is a major task of statistical mechanics, dating all the way back to Boltzmann’s H-theorem [1] and Loschmidt’s paradox [2]. Besides its ontological dimension, this question is also intimately related to the special role of nonequilibrium states and their apparent instability in many-body systems, which generically tend toward equilibrium as time progresses. Characterising this instability within the realm of quantum mechanics is one goal of the present work. Whereas a direct investigation of the pertinent states (i.e. Hilbert space vectors or density operators) can, in principle, have academic value, similarly to an analysis of phase-space points in classical systems, we focus here on observable quantities that can be extracted from macroscopic measurements.

Recently, so-called out-of-time-order correlators have gained considerable attention as a suggestion to generalise concepts from classical chaos theory, notably Lyapunov exponents, to quantum systems [3], [4], but the analogy is far from complete [5], [6]. With this in mind, we follow an even simpler route here and consider so-called echo dynamics [7], where a quantum many-body system with Hamiltonian H starts from a (pure or mixed) “target state” ρT and is evolved forward in time for a certain period τ to reach the “return state” ρR. At this point, one switches to the inverted Hamiltonian H, so that the direction of time is effectively reversed, and the system evolves back toward the target state after another time period τ. By introducing inaccuracies (experimentally unavoidable imprecisions) during this backward evolution via a Hamiltonian of the form H:=H+ϵV, the system will not reach the original initial state ρT again, but instead approach a perturbed state ρT that deviates from it to some extent. In summary,

(1)ρTHτρRH+ϵVτρT.

The deviations between the initial and final states and their dependence on the time span τ and on the amplitude ϵ of the considered imperfections then quantify the instability of the nonequilibrium state and the irreversibility of the dynamics.

In case of a pure state, a seemingly direct probe is the Loschmidt echo, that is, the overlap between the original state ρT and the distorted echo state ρT [8], [9]. (In case of a mixed state, the corresponding probe is the quantum fidelity.) However, this quantity is practically inaccessible in a many-body system, and as emphasized above, we will instead concentrate on echoes of macroscopic observables in the following and compare their expectation values at the beginning and at the end of the proposed protocol [9], [10], [11], [12]. As macroscopic observables cannot distinguish between equivalent microstates, one inevitably has to operate in the nonequilibrium regime to be able to characterise the sensitivity of a many-body system toward imperfections.

Taken literally, the suggested protocol (1) can only have the status of a gedankenexperiment as we cannot practically revert the direction of time in a concrete dynamical setup, and also an exchange of the Hamiltonian H for H is still unphysical in many situations. For example, for a gas of particles in a box, it would require negative particle masses. Nevertheless, it is well known that an effective sign change of the Hamiltonian can be achieved in spin systems, which provides the mechanism underlying spin echo and nuclear magnetic resonance (NMR) experiments [7]. In fact, the presence of imperfections, i.e. “nonreversed components” of the Hamiltonian, forms the basis of magnetic resonance imaging (MRI) by exploiting that different imperfections in different tissues lead to distinct imperfect echo signals [13]. By means of so-called magic- or polarisation-echo techniques, these ideas have also been extended to interacting spin systems [14], [15], [16], [17], [18], [19], [20], [21], employing suitably adapted radiofrequency external fields during the “backward” phase of the evolution. More generally, pulse sequences and time-dependent forces have been used to scan quite notable parameter ranges of (effective) spin Hamiltonians in a variety of experimental setups [22], [23], [24], [25]. Yet another experimental approach toward an effective time reversal consists in tuning a cold atomic gas across a Feshbach resonance [26], [27], [28]. Given the neglected unreversed corrections as well as the sophisticated experimental setups necessary, some sort of imperfections of the form considered here are clearly unavoidable and sometimes, like in MRI, even desired.

Furthermore, an alternative view on the suggested protocol may be as follows: For a many-body system with Hamiltonian H~ (=H in the language from above), suppose that we are given an initial state ρR (previously the return state), for which it is known that the state obtained after time τ, ρT=eiH~τρReiH~τ(=1), is out of equilibrium. Comparing the time evolution from ρR under the Hamiltonian H~ with the dynamics obtained from a perturbed Hamiltonian H~=H~+ϵV, we achieve the same effect as in the above echo gedankenexperiment (1). However, there is no need for a “backward Hamiltonian” or some other sort of (experimentally difficult) reversal procedure. From a physical point of view, it is thus irrelevant whether the return state ρR was obtained by an explicit unitary time evolution or some other preparation method. Its only relevant property is that it reaches a manifestly out-of-equilibrium state within an accessible time scale τ in order for the effects of imperfections to become macroscopically visible. This setting is silently included in the following, even though we will employ the language of the echo protocol (1) in the remainder of this work.

In Section 2, we set the stage and introduce the suggested echo protocol (1) and the considered imperfections V in detail. Our main result, an analytical prediction characterising the decay of echo signals under generic time-reversal inaccuracies, is derived in Section 3. Thereafter, we verify this result numerically in an explicit spin chain model in Section 4. In Section 5, we conclude by summarising the ideas and relating them to numerical and experimental results from the literature.

2 Setup

To begin with, we represent the Hamiltonian H appearing in (1) in terms of its eigenvalues and eigenvectors as

(2)H=nEn|nn|.

Given some initial state ρ(0) at time t = 0, the state at any later time t > 0 follows as ρ(t):=eiHtρ(0)eiHt. Being interested in some macroscopic observable (see Section 1) in the form of a self-adjoined operator A, and denoting its expectation value in an arbitrary state ρ by Aρ:=Tr[ρA], the actual time evolution can thus be written as

(3)Aρ(t)=m,nei(EnEm)tm|ρ(0)|nn|A|m.

Despite the quasiperiodic nature of the right-hand side, a many-body system will usually equilibrate [29], [30], [31] and spend most of the time close to the time-averaged state ρeq with

(4)m|ρeq|n:=δmnn|ρ(0)|n,

which we refer to as the equilibrium state in the following.

In order to study the effect of imperfections in terms of expectation values of the observable A, the considerations from Section 1 imply that the system must spend some time away from equilibrium; i.e. there must be a reasonable time interval during which Aρ(t) differs distinctly from Aρeq. Therefore, we focus on these deviations from equilibrium, denoted by the symbol

(5)𝒜(t):=Aρ(t)Aρeq.

Provided that the system is out of equilibrium at time t = 0, we then ask how special this situation is by investigating how hard it is to return to this state by an effective, but possibly imperfect reversal of time after the system has relaxed for a certain period τ as detailed in the protocol (1).

In the absence of any imperfections (ϵ = 0), the system traces out the same trajectory in the forward and backward stages, such that

(6)𝒜(τ+t)=𝒜(τt)

for t[0,τ], which constitutes our reference dynamics. It is reasonable to expect that uncontrolled inaccuracies in the time-reversed dynamics will generically push the system closer to equilibrium because they spoil the fine-tuned correlations between state and observable needed for nonequilibrium conditions. Hence, the backward dynamics will usually lie closer to the equilibrium state than the forward one,

(7)|𝒜(τ+t)||𝒜(τt)|  (0tτ).

The sensitivity of the deviations between the perfect and perturbed dynamics with respect to the magnitude ϵ of the inaccuracies is thus an indicator for the chaoticity and irreversibility of the many-body dynamics. The faster 𝒜(τ+t) decays with ϵ compared to 𝒜(τt), the harder it is to design a reversible process and the more extraordinary or special are the nonequilibrium states. Consequently, the relative echo signal 𝒜(τ+t)/𝒜(τt) for times t[0,τ] will be our principal object of study in this work, most importantly in the region around the revival or echo peak at tτ, where deviations from equilibrium will be most pronounced.

We denote the time-dependent state of the system in the forward and backward phases by

(8a)ρf(t):=eiHtρTeiHt,
(8b)ρb(t):=ei(HϵV)tρRei(HϵV)t,

respectively. We also write ρ(t) to refer to the state during the entire process, that is, ρ(t):=ρf(t) for t[0,τ] and ρ(t):=ρb(tτ) for t[τ,2τ].

An implicit assumption in all that follows is that the considered many-body system is finite and exhibits a well-defined macroscopic energy E. Consequently, the state ρ(t) at any time can only significantly populate energy levels within a macroscopically small energy window

(9)IE:=[EΔE,E]

and the imperfections are assumed to be sufficiently small so that they do not modify this window. In addition, it is taken for granted that the density of states (DOS) of H is approximately constant throughout this energy window, D0const and that the same holds for the (negative) imperfect backward Hamiltonian HϵV with eigenvalues Eν and eigenstates |ν.

Focusing on the dynamics during the backward phase from (8b), we can use the transformation matrix Uμk:=μ|k between the eigenbases of the forward and backward Hamiltonians to write the time-dependent expectation values of the observable A – similarly as in (3) – as

(10)Aρb(t)=μ,νμ|ρb(t)|νν|A|μ
(11)=μ,νk,l,m,nei(EνEμ)tei(ElEk)τ×k|ρT|lm|A|nUμkUνlUνmUμn.

Employing the assumed constant DOS for H and HϵV, we can approximately identify energy differences EνEμEνEμ of the two Hamiltonians within in the relevant time scales [32], [33], so that

(12)Aρb(t)=μ,νk,l,m,nei(EνEμ)tei(ElEk)τ×k|ρT|lm|A|nUμkUνlUνmUμn.

We recall that the Hamiltonian H corresponds to a given many-body quantum system, whereas the perturbation V describes uncontrolled and/or unknown inaccuracies in the time-reversal procedure. In this spirit, we thus model our ignorance about these imperfections by an ensemble of random operators V, such that the matrix elements m|V|n of V in the eigenbasis of H become random variables. The actually considered V ensembles are inspired by the structure of typical perturbations, featuring possible sparsity as well as an interaction strength depending on the energy difference between the coupled states (“bandedness”) [34], [35], [36], [37], [38]. Requiring hermiticity, m|V|n=n|V|m, and assuming independence of the m|V|n for mn, this suggests the general form

(13)dPmn(v):=dμ|EmEn|(v)

for the probability measures of the m|V|n’s with m<n. Here, {dμΔ}Δ>0 denotes a family of probability measures on ℝ or ℂ with mean zero and variance σv2(Δ), so that the smooth function σv2(Δ) captures the announced bandedness of the interaction matrix. Likewise, for m=n, the probability measure dPnn(v):=dμ0(v) of the (real-valued) diagonal elements n|V|n is assumed to have vanishing mean (otherwise, the perturbation would induce an energy shift) and finite variance.

To obtain a useful prediction regarding the behaviour of an actual system, we first compute the average effect of such a perturbation. In a second step, we establish that the resulting prediction satisfies a concentration of measure property, meaning that in a sufficiently high-dimensional Hilbert space a particular realisation of the ensemble becomes practically indistinguishable from the average behaviour. More precisely, deviations from the average will turn out to be suppressed in the number Nv of eigenstates of H that get mixed up by the perturbation ϵV, to be defined explicitly in (19) below. Because of the extremely high-level density in generic many-body systems, this number Nv is typically exponentially large in the system’s degrees of freedom f if the perturbation has any appreciable effect at all [39],

(14)Nv=10𝒪(f)1.

3 Results

According to (12), averaging the echo signal over all possible realisations of the V ensemble requires an average over four transformation matrices Uμk, the overlap of the eigenvectors |k of H and |μ of HϵV. These overlaps inherit their distribution from the distribution (13) of the V matrix elements. Writing 𝔼[] for the average over all V’s, one finds

(15)𝔼[Uμ1k1Uμ2k2Uμ1l1Uμ2l2]=δk1l1δk2l2dk1k2μ1μ2+δk1l2δk2l1(δμ1μ2dk1k2μ1μ2+fk1k2μ1μ2)

in the limit of sufficiently weak ϵ [39]. Here

(16)dklμν:=u(αϵ2,EμEk)u(αϵ2,EνEl),
(17)fklμν:=(αϵ2/2πD0)u(αϵ2,EμEk)u(αϵ2,EνEl)×4α2ϵ4+(EkEl)2+(EμEν)2(Ek+ElEμEν)2[(EμEl)2+α2ϵ4][(EνEk)2+α2ϵ4],

and α:=πσ¯v2D0, where σ¯v2 denotes the mean value of σv2(Δ), introduced below (13), for small argument. Furthermore, the function u(αϵ2,EμEk):=𝔼[|Uμk|2] represents the second moment of the transformation matrices Uμk and is given by the Breit–Wigner distribution

(18)u(γ,E)=γπD0(γ2+E2).

Hence, we can identify

(19)Nv:=2αD0ϵ2=2πσ¯v2D02ϵ2

as the full width at half maximum of the average overlap 𝔼[|Uμk|2] between eigenvectors of H and HϵV, quantifying the number of energy eigenstates mixed by the perturbation as introduced above (14).

Exploiting (15) in the average of (12), we obtain

(20)𝔼[Aρb(t)]=μdklμμk|ρT|kl|A|l+μ,νei(EνEμ)t××k,l[ei(ElEk)τdklμνk|ρT|ll|A|k+fklμνk|ρT|kl|A|l]

If we make use of the constant DOS once again (see below (9)), which allows us to shift summation indices, we find that

(21)μdklμμ=u(2αϵ2,ωkl),
(22)μ,νei(EνEμ)tdklμν=eiωklte2α|t|ϵ2,
(23)μ,νei(EνEμ)tfklμν=u(2αϵ2,ωkl)e2α|t|ϵ2×[cos(ωklt)+2αϵ2ωklsin(ωkl|t|)]

with ωkl:=EkEl. Substituting into (20) yields

(24)𝔼[Aρb(t)]Aρ~=e2α|t|ϵ2[Aρf(τt)Aρ~]+R(|t|)

with the locally averaged equilibrium state ρ~ given by m|ρ~|n:=δmnku(2αϵ2,EnEk)k|ρT|k and

(25)R(t):=e2αtϵ2k,lk|ρT|kl|A|lu(2αϵ2,ωkl)×{1cos(ωklt)2αϵ2ωklsin(ωklt)}.

Note that ρ~ is the state approached for large times and can thus be identified with the equilibrium state ρeq from (5). It usually corresponds to the pertinent thermal state [40], [41], [42], [43]. We also observe that R(t) vanishes at t = 0 and for t → ∞. Furthermore, as shown in the Appendix its magnitude can be bounded from above for arbitrary t ≥ 0 according to

(26)R2(t)δ2A50Nv,

where Nv is the width of the eigenvector overlaps from (19), and

(27)δ2A:=n:EnIE(AnnAρmc)2,

where ρmc denotes the microcanonical density operator corresponding to the energy window from (9). For generic (nonintegrable) Hamiltonians H, essentially all observables A of actual interest are expected to satisfy the so-called eigenstate thermalisation hypothesis (ETH) [44]. Hence, the right-hand side of (27) can be roughly estimated as A2 (with A denoting the operator norm of A, i.e. the largest eigenvalue in modulus), and R(t) in (24) can be neglected according to (14) and (26).

For integrable systems, the relevant observables are still expected to satisfy the so-called weak ETH [45], [46], [47], and thus the right-hand side of (27) can be roughly estimated as A2N/f, where N is the number of energy levels En contained in IE, whereas f counts the degrees of freedom of the considered system and therefore scales as ln(N) (see also (14)). Again, one can conclude from (26) that R(t) in (24) amounts in many cases to a small correction.

Finally, we point out that the bound (26) is still rather loose as the oscillating character of the summands in (25) with respect to both ωkl and t will usually result in very strong “accidental cancellation” effects, which are entirely disregarded in our derivation of the “worst case” bound (26) in the Appendix.

Indeed, we have not been able to identify any specific example of practical interest where the last term in (24) plays a significant role. Accordingly, this term is henceforth considered as negligible. With (5), (8), and (14), we thus arrive at the first key result of this section,

(28)𝔼[𝒜(τ+t)]𝒜(τt)=e2αtϵ2(0tτ),

which quantifies the average effect of imprecisions during the backward evolution within the considered ensembles of V’s and for small enough ϵ. Analogously to [39], one can then proceed to derive a bound for the variance of 𝒜(τ+t). This leads to

(29)𝔼[𝒜(τ+t)2]𝔼[𝒜(τ+t)] 2CvA2Nv,

where Cv is a constant of order 103 or less. In view of (14), the variance (29) of 𝒜(τ+t) with V is thus exponentially small in the number of degrees of freedom, establishing a so-called concentration of measure property of the considered ensembles of V operators. For instance, invoking Chebyshev’s inequality from probability theory, the estimate (29) implies that the probability for 𝒜(τ+t) to differ by more than A/Nv1/3 from the average 𝔼[𝒜(τ+t)] at a certain instance in time t is less than Cv/Nv1/3. This is our second key result of this section, promoting (28) from a mere statement about the ensemble average to a prediction for individual realisations. As deviations from the average are extremely rare for reasonably large many-body systems, we can conclude that

(30)𝒜(τ+t)𝒜(τt)=e2αtϵ2(0tτ)

is an excellent approximation for the vast majority of time-reversal inaccuracies V captured by the considered ensembles. This relation for the echo dynamics (1) under an imperfect backward Hamiltonian constitutes our main result. It asserts that the echo signal is exponentially suppressed in the propagation time t and the intensity ϵ2 of the imperfections.

4 Example

We consider a spin-12 XXX chain model with Hamiltonian

(31)H=i=1L1𝝈i𝝈i+1,

where 𝝈i=(σix,σiy,σiz) is a vector of Pauli matrices acting on site i. For the perturbation V, we choose

(32)V=i<jα,β=13Jijαβσiασjβ,

where the couplings Jijαβ are drawn independently from a normal distribution (unbiased Gaussian with unit variance). As the observable, we take the staggered magnetisation

(33)A=i=1L(1)iσiz.

Turning to the initial (target) state ρT=|ψψ|, let us first consider a Néel state |ψ¯:=|. In order to account for the requirement that the DOS should be approximately constant (see below (9)), we rescale the probability amplitudes n|ψ¯ according to the corresponding energy eigenvalues En with a Gaussian weight of zero mean and standard deviation σψ, resulting in

(34)|ψ:=CneEn2/2σψ2n|ψ¯|n,

where the normalisation constant C is chosen such that ψ|ψ=1. As discussed around (9), the large isolated systems we have in mind (e.g. an MRI sample) are expected to exhibit a macroscopically well-defined energy, so that σψ should lie below the measurement resolution. In particular, while it is possible to prepare a clean Néel state in few-body experiments with cold atoms, such attempts will most likely result in a filtered or coarse-grained variant as the degrees of freedom increase.

Quantitatively, for the example in Figure 1, we chose a chain length of L = 14 and a standard deviation of σψ=1.3, so that the state (34) is focused around energy E = 0 with approximately 15 % of the total 2L=16,384 levels En within ±σψ. This procedure reduces the staggered magnetisation A of |ψ compared to |ψ¯, but still gives an appreciably out-of-equilibrium expectation value (see also Fig. 1).

Figure 1: Time-dependent expectation values of the staggered magnetisation (33) for the spin-12\(\frac{1}{2}\) XXX chain (31) with L = 14 under the “imperfect reversal” protocol (1) with various perturbation strengths ϵ and reversal times τ. The initial condition is chosen as a filtered Néel target state ρT=|ψ⟩⟨ψ|\({\rho_{\text{T}}}=|\psi\rangle\langle\psi|\) according to (34) with σψ=1.3\({\sigma_{\psi}}=1.3\). The imperfection Hamiltonian V in (1) is of the “spin-glass” form (32). Solid lines correspond to the numerical results using exact diagonalisation. Dashed lines show the prediction for the backward (echo) dynamics according to (30) and (35). Time reversal is initiated after time τ = 5 (red-toned curves), τ = 7.5 (blue-toned), or τ = 10 (green-toned). Inset: Ratio 𝒜(2τ)/𝒜(0)\(\mathcal{A}(2\tau)/\mathcal{A}(0)\) of the echo peak height (at time t=2τ\(t=2\tau\)) and the initial value (at t = 0) as a function of the perturbation strength ϵ for the different reversal times τ. Data points are the numerical solutions; solid lines are the analytical prediction from (30) and (35).
Figure 1:

Time-dependent expectation values of the staggered magnetisation (33) for the spin-12 XXX chain (31) with L = 14 under the “imperfect reversal” protocol (1) with various perturbation strengths ϵ and reversal times τ. The initial condition is chosen as a filtered Néel target state ρT=|ψψ| according to (34) with σψ=1.3. The imperfection Hamiltonian V in (1) is of the “spin-glass” form (32). Solid lines correspond to the numerical results using exact diagonalisation. Dashed lines show the prediction for the backward (echo) dynamics according to (30) and (35). Time reversal is initiated after time τ = 5 (red-toned curves), τ = 7.5 (blue-toned), or τ = 10 (green-toned). Inset: Ratio 𝒜(2τ)/𝒜(0) of the echo peak height (at time t=2τ) and the initial value (at t = 0) as a function of the perturbation strength ϵ for the different reversal times τ. Data points are the numerical solutions; solid lines are the analytical prediction from (30) and (35).

After diagonalising H numerically, we estimated the DOS D0 by averaging over all states with energies En[2σψ,2σψ] (receiving approximately 95 % of the weight), resulting in D0962. Furthermore, we extracted σ¯v2 [see below (17)] from the squared matrix elements |m|V|n|2 within the relevant energy window in (9) (determined again by the 2σψ criterion) by way of averaging around the diagonal within a band of 1000 states, yielding σ¯v20.0729. According to the definition below (17), this yields

(35)α=πσ¯v2D0220.

All parameters entering our analytical prediction (30) are thus explicitly available; that is, there is no free fit parameter.

In Figure 1, we compare the numerical results obtained by exact diagonalisation with our prediction (30) for different propagation times τ and perturbation strengths ϵ, showing good agreement. The largest deviations become apparent for small τ and large ϵ. By generalising the analysis of [39], we expect that the band profile σv2(Δ) of the perturbation V becomes important in this regime. The exponential form (30) for the suppression of the echo signal is then anticipated to show a transition to a Bessel-like decay of the form 4J1(x)2/x2, where x depends linearly on the noise strength ϵ, the reversal time τ, and the square root of the band width. More generally, within the considered ensemble of inaccuracies, the echo-signal attenuation is essentially given by the Fourier transform of u(γ,E) introduced above (18), which can be viewed as the ensemble-averaged fidelity or survival probability [9], [48] of an eigenstate |n of the clean Hamiltonian H under the imprecise backward Hamiltonian H′. When the influence of the random inaccuracies in H′ increases, the echo signal may therefore be expected to approach the known decay of fidelity in pure random matrix models [48]. Unfortunately, we are not aware of an exact analytic solution of the pertinent equations in the intermediate regime between the exponential and Bessel-type behaviours [39].

5 Conclusions

We investigated the stability of observable echo signals in many-body quantum systems under the influence of uncontrolled imperfections in the pertinent effective time reversal. The considered protocol starts from a nonequilibrium initial state and lets the system evolve under the time-independent Hamiltonian H for some time τ, at which an (effective) time reversal is performed that directs the system back toward the initial nonequilibrium state. By introducing small inaccuracies in the time-reversed Hamiltonian, we obtained a measure for the instability of nonequilibrium states and the irreversibility of the dynamics in terms of directly observable quantities.

Our prediction for the relative echo signal under such a distorted backward Hamiltonian, (30), includes an exponential dependence on both the squared perturbation strength ϵ2 and the propagation time t. In particular, the height of the echo peak at tτ is thus expected to decay exponentially in τ. Systems with this property were labelled “irreversible” in [11], as opposed to “reversible” ones where the decay is (at most) algebraic. In this sense, one may interpret the present result as a prediction that many-body quantum systems are typically irreversible. However, it should be pointed out that the functional dependence of the echo peak on τ rather appears to be a property of the inaccuracies V than of the system itself, even though the structure of V is (in a real system) of course influenced by the properties of the system. In any case, this functional dependence could also be confirmed for an exemplary spin-12 XXX model, whose dynamics shows good agreement with our analytical prediction without any remaining fit parameter.

The paradigmatic examples of macroscopic echo experiments are spin echoes and NMR [7], where nuclear spins precess in a strong magnetic field at different frequencies due to local inhomogeneities, leading to dephasing of the initially aligned magnetic moments. Applying a π pulse at time τ reverses the relative orientation between the spins and the external field and thus effectively changes the sign of the corresponding term in the pertinent model Hamiltonian. However, interactions among the spins and with the environment are not reversed and amount to a “perturbation” that causes deviations of the echo signal at time 2τ, which – in line with our general echo analysis here – typically decays exponentially with τ. It should be noted that the “imperfections” in this context are usually vital in applications such as MRI, precisely because different imperfections lead to slightly different decay rates and therefore allow one to distinguish different materials.

Experimental implementations of echo protocols are also available in a variety of interacting spin systems, see, for example, [15], [16], [17], [20], [21]. In these experiments, an effective sign flip of the dominant part of the Hamiltonian (including dipole–dipole or even quadrupole interactions) is achieved by means of an elaborate application of radiofrequency magnetic fields. The prevailing inaccuracies leading to deviations from the perfectly reversed signal are again due to nonreversible correction terms in the Hamiltonian, as well as possible experimental imprecisions in carrying out the required protocol. Their major contribution is thus expected to be of the type studied here, too.

Specifically, the experimental study [20] indeed reports an exponential decay of the peak height with the reversal time τ in a polarisation echo experiment involving the nuclear spins of a cymantrene polycrystalline sample. In the same study, data obtained from a ferrocene sample suggest an approximately Gaussian-shaped dependence; see also [21]. The authors explain this with the much larger relative strength of the nonreversible component in the Hamiltonian, compatible with our prediction that a crossover from an exponential to a Bessel-like[1] decay is expected as the relative strength of V increases; see the discussion at the end of Section 4.

From a conceptual point of view, the approach of our present work, where the Hamiltonians of the forward and backward phases differ slightly, assesses (via observable quantities) the stability of many-body trajectories with respect to variations of the dynamical laws. Therefore, it should be no surprise that deviations grow with the propagation time τ, and the exponential dependence might have been anticipated from perturbation-theoretic considerations, even though the applicability of standard perturbation theory is rather limited for typical many-body systems with their extremely dense energy spectra. In that sense, the present derivation by nonperturbative methods is reassuring and also indicates how deviations from the exponential behaviour will manifest themselves if the influence of imperfections increases. For future work, it will be interesting to investigate a complementary approach that studies the sensitivity toward variations of the initial conditions in macroscopic quantum system.

Acknowledgement

L.D. thanks Patrick Vorndamme for inspiring discussions. This work was supported by the Deutsche Forschungsgemeinschaft (DFG) within the Research Unit FOR 2692 under grant no. 397303734 (Funder Id: http://dx.doi.org/10.13039/501100001659) and by the Paderborn Center for Parallel Computing (PC2) within the Project HPC-PRF-UBI2.

Appendix A: Derivation of (26)

Exploiting (18), we rewrite R(t) from (25) as

(36)R(t)=R~(2αϵ2t),
(37)R~(t):=m,nm|ρT|mn|A|nh(mn,t),
(38)h(n,t):=etπNv1cos(tn/Nv)sin(tn/Nv)n/Nv1+(n/Nv)2,

where Nv is defined in (19), and where ωkl:=EkEl (see below (21)) has been approximated by (kl)/D0, as justified below (9).

According to (14), we can and will take for granted that

(39)Nv1.

In view of (38), it follows that the sum nh(n,t) can be very well approximated by the integral dxh(x,t). Moreover, a straightforward but somewhat tedious calculation yields dxh(x,t)=0. Therefore, we can subtract a constant value from the observable A without changing the value of R~(t) in (37). By means of the definition

(40)A~nn:=n|A|nAρmc

we thus can rewrite (37) as

(41)R~(t)=mm|ρT|mQ~(m,t),
(42)Q~(m,t):=nA~nnh(mn,t).

Exploiting the Cauchy–Schwarz inequality yields

(43)Q~2(m,t)[n(A~nn)2][nh2(mn,t)].

As the last sum over n is independent of m, it follows that

(44)|Q~(m,t)|Q(t),
(45)Q(t):=q(t)n(A~nn)2,
(46)q(t):=nh2(n,t).

Accordingly, R~(t) from (41) can be upper bounded as

(47)|R~(t)|Q(t)mm|ρT|m=Q(t),

where we exploited that ρT is a positive semidefinite operator of unit trace.

Because of (38) and (39), one can conclude – similarly as below (39) – that the sum on the right-hand side of (46) is very well approximated by the integral dyh2(y,t). After going over from the integration variable y to x:=y/Nv, one thus obtains in very good approximation

(48)q(t)=Nv1f(t),
(49)f(t):=e2tdx[1cos(xt)sin(xt)/xπ[1+x2]]2.

An analytical evaluation of f(t) from (49) is possible but quite arduous, whereas a numerical evaluation is straightforward; see Figure 2. In either case, one finds that

(50)0f(t)1/50=:c

for all t ≥ 0. Taking into account (36), (45), (47), and (48), we thus arrive at

(51)R2(t)cNvn(A~nn)2.
Figure 2: Numerical evaluation of the function f(t)\(f(t)\) from (49).
Figure 2:

Numerical evaluation of the function f(t) from (49).

As discussed above (9), the diagonal matrix elements n|ρ(0)|n appearing in (3) vanish whenever EnIE. Accordingly, we can arbitrarily modify the corresponding n|A|n’s without any further consequences in (3). Specifically, we can modify them so that all A~nn in (40) are zero if EnIE. Therefore, the summation on the right-hand side of (51) can be restricted to those n with EnIE. Altogether, we thus recover (26) and (27) from the main text.

References

[1] L. Boltzmann, Sitzungsberichte Akad. Wiss. Wien 66, 275 (1872); reprint: Boltzmann’s Wissenschaftliche Abhandlungen, Vol. I, J. A. Barth, Leipzig 1909, p. 316.Search in Google Scholar

[2] J. Loschmidt, Sitzungsberichte Akad. Wiss. Wien 73, 128 (1876).Search in Google Scholar

[3] J. Maldacena, S. H. Shenker, and D. Stanford, J. High Energy Phys. 2016, 106 (2016).10.1007/JHEP08(2016)106Search in Google Scholar

[4] B. Swingle, Nat. Phys. 14, 988 (2018).10.1038/s41567-018-0295-5Search in Google Scholar

[5] Q. Hummel, B. Geiger, J. D. Urbina, and K. Richter, Phys. Rev. Lett. 123, 160401 (2019).10.1103/PhysRevLett.123.160401Search in Google Scholar PubMed

[6] S. Pilatowsky-Cameo, J. Chávez-Carlos, M. A. Bastarrachea-Magnani, P. Stránský, S. Lerma-Hernández, et al., Phys. Rev. E 101, 010202(R) (2020).10.1103/PhysRevE.101.010202Search in Google Scholar PubMed

[7] E. L. Hahn, Phys. Rev. 80, 580 (1950).10.1103/PhysRev.80.580Search in Google Scholar

[8] A. Peres, Phys. Rev. A 30, 1610 (1984).10.1103/PhysRevA.30.1610Search in Google Scholar

[9] T. Gorin, T. Prosen, T. H. Seligman, and M. Žnidarič, Phys. Rep. 435, 33 (2006).10.1016/j.physrep.2006.09.003Search in Google Scholar

[10] B. V. Fine, T. A. Elsayed, C. M. Kropf, and A. S. de Wijn, Phys. Rev. E 89, 012923 (2014).10.1103/PhysRevE.89.012923Search in Google Scholar PubMed

[11] M. Schmitt and S. Kehrein, EPL 115, 50001 (2016).10.1209/0295-5075/115/50001Search in Google Scholar

[12] M. Schmitt and S. Kehrein, Phys. Rev. B 98, 180301 (2018).10.1103/PhysRevB.98.180301Search in Google Scholar

[13] M. T. Vlaardingerbroek and J. A. den Boer, Magnetic Resonance Imaging – Theory and Practice, Third Edition, Springer, Berlin, Heidelberg 2003.10.1007/978-3-662-05252-5Search in Google Scholar

[14] H. Schneider and H. Schmiedel, Phys. Lett. 30A, 298 (1969).10.1016/0375-9601(69)91005-6Search in Google Scholar

[15] W.-K. Rhim, A. Pines, and J. S. Waugh, Phys. Rev. Lett. 25, 218 (1970).10.1103/PhysRevLett.25.218Search in Google Scholar

[16] W.-K. Rhim, A. Pines, and J. S. Waugh, Phys. Rev. B 3, 684 (1971).10.1103/PhysRevB.3.684Search in Google Scholar

[17] S. Zhang, B. H. Meier, and R. R. Ernst, Phys. Rev. Lett. 69, 2149 (1992).10.1103/PhysRevLett.69.2149Search in Google Scholar

[18] R. Kimmich, J. Niess, and S. Hafner, Chem. Phys. Lett. 190, 503 (1992).10.1016/0009-2614(92)85181-9Search in Google Scholar

[19] S. Hafner, D. E. Demco, and R. Kimmich, Solid State Nucl. Mag. Res. 6, 275 (1996).10.1016/0926-2040(96)01234-9Search in Google Scholar

[20] P. R. Levstein, G. Usaj, and H. M. Pastawski, J. Chem. Phys. 108, 2718 (1998).10.1063/1.475664Search in Google Scholar

[21] G. Usaj, H. M. Pastawski, and P. R. Levstein, Mol. Phys. 95, 1229 (1998).10.1080/00268979809483253Search in Google Scholar

[22] M. Gärttner, J. G. Bohnet, A. Safavi-Naini, M. L. Wall, J. J. Bollinger, et al., Nat. Phys. 13, 781 (2017).10.1038/nphys4119Search in Google Scholar

[23] K. X. Wei, C. Ramanathan, and P. Cappellaro, Phys. Rev. Lett. 120, 070501 (2018).10.1103/PhysRevLett.120.070501Search in Google Scholar PubMed

[24] K. X. Wei, P. Peng, O. Shtanko, I. Marvian, S. Lloyd, et al., Rev. Lett. 123, 090605 (2019).10.1103/PhysRevLett.123.090605Search in Google Scholar PubMed

[25] M. Niknam, L. F. Santos, and D. G. Cory, Phys. Rev. Research 2, 013200 (2020).10.1103/PhysRevResearch.2.013200Search in Google Scholar

[26] A. Widera, S. Trotzky, P. Cheinet, S. Fölling, F. Gerbier, et al., Phys. Rev. Lett. 100, 140401 (2008).10.1103/PhysRevLett.100.140401Search in Google Scholar PubMed

[27] F. M. Cucchietti, J. Opt. Soc. Am. B 27, A30 (2010).10.1364/JOSAB.27.000A30Search in Google Scholar

[28] C. Weiss, J. Phys. Conf. Ser. 414, 012032 (2013).10.1088/1742-6596/414/1/012032Search in Google Scholar

[29] P. Reimann, Phys. Rev. Lett. 101, 190403 (2008).10.1103/PhysRevLett.101.190403Search in Google Scholar PubMed

[30] N. Linden, S. Popescu, A. J. Short, and A. Winter, Phys. Rev. E 79, 061103 (2009).10.1103/PhysRevE.79.061103Search in Google Scholar PubMed

[31] A. J. Short and T. C. Farrelly, New J. Phys. 14, 013063 (2012).10.1088/1367-2630/14/1/013063Search in Google Scholar

[32] P. Reimann and L. Dabelow, Phys. Rev. Lett. 122, 080603 (2019).10.1103/PhysRevLett.122.080603Search in Google Scholar PubMed

[33] P. Reimann, New J. Phys. 21, 053014 (2019).10.1088/1367-2630/ab1a63Search in Google Scholar

[34] S. Genway, A. F. Ho, and D. K. K. Lee, Phys. Rev. A 86, 023609 (2012).10.1103/PhysRevA.86.023609Search in Google Scholar

[35] W. Beugeling, R. Moessner, and M. Haque, Phys. Rev. E 91, 012144 (2015).10.1103/PhysRevE.91.012144Search in Google Scholar PubMed

[36] N. P. Konstantinidis, Phys. Rev. E 91, 052111 (2015).10.1103/PhysRevE.91.052111Search in Google Scholar PubMed

[37] F. Borgonovi, F. M. Izrailev, L. F. Santos, and V. G. Zelevinsky, Phys. Rep. 626, 1 (2016).10.1016/j.physrep.2016.02.005Search in Google Scholar

[38] D. Jansen, J. Stolpp, L. Vidmar, and F. Heidrich-Meisner, Phys. Rev. B 99, 155130 (2019).10.1103/PhysRevB.99.155130Search in Google Scholar

[39] L. Dabelow and P. Reimann, Phys. Rev. Lett. 124, 120602 (2020).10.1103/PhysRevLett.124.120602Search in Google Scholar PubMed

[40] J. M. Deutsch, Phys. Rev. A 43, 2046 (1991).10.1103/PhysRevA.43.2046Search in Google Scholar PubMed

[41] P. Reimann, New J. Phys. 17, 055025 (2015).10.1088/1367-2630/17/5/055025Search in Google Scholar

[42] C. Gogolin and J. Eisert, Rep. Prog. Phys. 79, 056001 (2016).10.1088/0034-4885/79/5/056001Search in Google Scholar PubMed

[43] C. Nation and D. Porras, New. J. Phys. 20, 103003 (2018).10.1088/1367-2630/aae28fSearch in Google Scholar

[44] L. D’Alessio, Y. Kafri, A. Polkovnikov, and M. Rigol, Adv. Phys. 65, 239 (2016).10.1080/00018732.2016.1198134Search in Google Scholar

[45] G. Biroli, C. Kollath, and A. M. Läuchli, Phys. Rev. Lett. 105, 250401 (2010).10.1103/PhysRevLett.105.250401Search in Google Scholar PubMed

[46] T. N. Ikeda, Y. Watanabe, and M. Ueda, Phys. Rev. E 87, 012125 (2013).10.1103/PhysRevE.87.012125Search in Google Scholar PubMed

[47] V. Alba, Phys. Rev. B 91, 155123 (2015).10.1103/PhysRevB.91.155123Search in Google Scholar

[48] E. J. Torres-Herrera and L. F. Santos, Phys. Rev. A 89, 043620 (2014).10.1103/PhysRevA.89.043620Search in Google Scholar

Received: 2019-12-31
Accepted: 2020-02-27
Published Online: 2020-04-28
Published in Print: 2020-05-26

© 2020 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/zna-2019-0383/html
Scroll to top button