• Keine Ergebnisse gefunden

3.6 Typical time evolution

3.6.3 Discussion

perturbation profile is largely determined by the values ofαv andv.

For completeness, we mention that there are some deviations between the second-order continued-fraction solution (3.166) for even larger times t and values of the coupling λ not shown here [248]. However, for such large values of λ, the premises from Sec. 3.2, notably Prerequisite (ii), will presumably break down, too, meaning that this situation was essentially excluded a priori.

Otherwise, this scenario will still be covered well by the strong-perturbation asymptotics (3.165).

Furthermore, we recall that a related universality was observed in Sec. 3.4.2 (see Figs. 3.4 and 3.5 in particular) for the ensemble-averaged resolventG(z) from (3.57) already.

Altogether, we therefore conclude that the analytic expressions (3.164), (3.165), and (3.166), which were originally obtained for special cases of the perturbation profile, in fact cover a very broad (if not comprehensive) regime of physically relevant coupling strengths and perturbation profiles.

Moreover, the necessary information about the perturbation has been reduced even further. Instead of the full perturbation profile, which was the starting point of our modeling in Sec. 3.2, we infer that, in essence, only the intrinsic strengthαvfrom (3.12) and the band widthv from (3.13) are required for a decent description of the response profilegλ(t).

to the number of digits precision (e.g., Cκ ≈20 for κ= 10−20) and Nv is the number of mixed levels from (3.9). Given that the latter number scales exponentially in the degrees of freedom f, mild violations of the ETH as quantified by (3.182) are thus indeed negligible, even though the scaling ofmc(A) has to be taken into account in any concrete example, too.

The derivation of an even stronger bound applying to the originalR(t) and for all timest can be found in Ref. [260], yielding

R(t)22mc(A) 50Nv

, (3.186)

where

2mc(A) := X

ν:Eν∈IE

(Aνν− hAiρmc)2 (3.187) also measures the violation of the ETH, albeit in terms of the squared magnitude of deviations between the diagonal matrix elements and the microcanonical prediction. Even if the strong ETH is violated, the reference system may still often satisfy the weak ETH (cf. Sec. 2.2.2 and particularly Eq. (2.26)), e.g., if the Hamiltonian has translational symmetry. As a consequence of Eq. (2.26), the violation quantified by (3.187) will then be subextensive inN, whereasNv from (3.9) is expected to grow extensively with N. The bound (3.186) thus entails that R(t) is indeed negligible if the system is sufficiently large.

Similar estimates are also expected to apply to the pertinent remnant terms for larger perturba-tions, arising from the approximation (3.128) of the fourth-order overlap moment, as well as due to (subleading) corrections to any such approximation. Moreover, these bounds are quite conser-vative, considering that the individual terms in the sum in (3.151) are oscillating for any fixedt, meaning that cancellations unaccounted for in the above bounds will arise naturally. Notably, we have so far never encountered a particular example whereR(t) in (3.149) entailed visible deviations from (3.152).

Long-time limit and prethermalization. In light ofR(t) being negligible, we will concentrate ex-clusively on Eq. (3.160) for the typical expectation values. Taking into account the property (3.162) of gλ(t), we understand that the perturbed systems are predicted to equilibrate (even if the un-perturbed system does not), and that the corresponding stationary expectation value approached for long times ishAiρ˜λ with the state ˜ρλ from (3.150). This quantity was already at the focus of Deutsch’s early studies on thermalization [114, 253] (see also Ref. [257]). According to its defini-tion (3.150), the state ˜ρλ is based on the diagonal ensemble ρ0 (see Sec. 2.2.1) of the reference systemH0, but the level occupations are additionally averaged over the energy scaleΓv that mea-sures the mixing of unperturbed eigenvectors caused by the perturbation. If the reference system satisfies the (strong) ETH, then alreadyρ0 can be well approximated by the microcanonical den-sity operator ρmc, and the same will hold for ˜ρλ. Moreover, even if the unperturbed system is, for example, integrable and fulfills just the weak ETH, expectation values with respect to ˜ρλ will usually still coincide with those obtained from the microcanonical ensembleρmc[7, 114, 252, 257], unless the perturbations are very weak or still exhibit conservation laws disregarded inρmc. Quite generally, the prediction (3.160) can therefore be simplified further by setting hAiρ˜λ =hAiρmc and thus

hAiρλ(t)=hAiρmc+|gλ(t)|2

hAiρ0(t)− hAiρmc

. (3.188)

Hence the perturbed dynamics will typically resemble the behavior of the unperturbed system initially, notably even if the latter does not thermalize or even equilibrate, but will eventually approach a thermal state as time progresses. At this point we thus highlight that the predic-tion (3.188) includes, as a special case, a descrippredic-tion of the prethermalizapredic-tion process discussed in Sec. 3.5, as announced towards the end of that subsection (see also Sec. 2.2.3 and Table 3.1).

Adopting the results from Ref. [97] (see also Sec. 2.3.2 and Eq. (2.40) in particular), if applicable, one can even devise an additional prediction for the unperturbed dynamics hAiρ0(t). In this case, the entire prethermalization scenario is captured by an analytical theory that merely depends on the initial value hAiρ0(0), the nonthermal stationary value hAiρ0, and the temperature (for the prediction of hAiρ0(t), see Eq. (2.40) and Ref. [97]) as well as the thermal value hAiρmc, the intrinsic perturbation strengthαv, and (possibly) the perturbation band widthv(for the present prediction ofhAiρλ(t)from hAiρ0(t)).

Fermi’s golden rule. Another interesting and important special case entailed in the result (3.160) is a form of Fermi’s golden rule [1]. If we choose the initial state to be an eigenstate of the reference system,ρ(0) =iihνi|, and take the projectorA=|νfihνf|onto another unperturbed eigenstate as our observable, the time-dependent expectation valuehAiρλ(t) is just the probability pνi→νf(t) to observe the transition from|νiito|νfiafter timet. Upon substitution into (3.160), we find that

pνi→νf(t) = ˜u(EνfEνi)

1− |gλ(t)|2

+δνiνf|gλ(t)|2. (3.189) For sufficiently weak perturbations, in particular, the response profile gλ(t) assumes the expo-nential form (3.164), such that the transition probability according to (3.189) approaches its pre-dicted equilibrium value ˜u(EνfEνi) at the rate Γ = 2πλ2σv2 (see Eq. (3.65)). Observing that σ2v ' E[|Vµν|2] in this case, Eq. (3.189) is the ensemble average of Fermi’s golden rule in dis-guise. It is noteworthy that this relation was derived here by nonperturbative methods, contrary to the “traditional” approach. This is particularly reflected by the fact that (3.189) is expected to hold beyond the traditional (exponential) golden-rule regime whengλ(t) is given by the solutions of (3.173) in general. Yet the analysis from Sec. 3.6.2 suggests that the traditional rule with the exponentialgλ(t) will eventually apply at late times.

Insignificance of level fluctuations. As promised in Sec. 3.2 (see below Eq. (3.18)), we finally come back to the question of when the fluctuations of energy levels are negligible for the dynamics so that we can, in particular, approximateEλn−Emλ in (3.7) byEn−Emas required in Prerequisite (v) (see also Eq. (3.142)). As recorded in (3.18), the influence of level fluctuations on the relaxation dynamics is expected to be negligible as long as the relaxation time tR is much larger than the inverse perturbation strength,tR(λσ0)−1.

According to our main result (3.160), the relaxation timetRis set by the characteristic time scale of the response profile gλ(t). Focusing on sufficiently weak perturbations first, such that gλ(t) is given by (3.164), we associate tR with Γ−1 from (3.65). The conditiontR (λσ0)−1 is then equivalent to

σ0

σv

2πλσv

ε ≈p

Nv (3.190)

with Nv from (3.9), and where we identifiedΓv =Γ (see above Eq. (3.9) and Eq. (3.33)) in the last step. In view of (3.9), the fluctuations σ0 of the diagonal matrix elements Vµµ may exceed the fluctuations σv of the off-diagonal Vµν by many orders of magnitude without violating the condition (3.190) and hence (3.18).

For stronger perturbations, such that gλ(t) assumes the form (3.165), the typical relaxation time isγ−1 from (3.69). The condition (3.18) then leads to

σ0 σv

rv

ε . (3.191)

Since the level spacing ε decreases exponentially in the degrees of freedom f, whereas the band width v (i.e., the energy range of the perturbation) should be roughly independent of f, we eventually reach the same conclusion that the fluctuations of the diagonal matrix elements of V remain insignificant with respect to the relaxation dynamics even if they exceed the off-diagonal fluctuations by many orders of magnitude.

Finally, we remark that we would have arrived at the same conclusions, too, if the rigorous bound (3.19) had been employed in lieu of (3.18). Observing that the time scale tR decreases gradually as the response profile crosses over from (3.164) to (3.165), we have thus justified the generic validity of Prerequisite (v)a posteriori.

Applicability to real systems. The original goal formulated at the beginning of this chapter was to describe the dynamics of an actual physical system with Hamiltonian (3.1). So far, we showed that the vast majority of perturbations within any of the admitted ensembles from Sec. 3.3 result in the relaxation behavior (3.160) or (3.188). It thus remains to be argued that the true perturbation of the system of interest is a typical member of one of those considered ensembles. Unfortunately, it is virtually impossible to prove this for any concrete given system, so we can only collect evidence supporting this conjecture and investigate disqualifying properties (see also Sec. 2.3). In the

subsequent considerations, we take the prerequisites collected in Sec. 3.2 for granted. That is to say, we exclude violations of these requirements, which were explicitly exploited in the derivation of the prediction (3.160), from the following discussion since a system may fail to follow this prediction for obvious reasons in that case.

Eventually, the decisive question is whether or not the key features of the true perturbation with regard to the relaxation behavior are shared by the majority of perturbations in a suitable ensemble.

Summarizing the preceding subsections, we established that the relaxation dynamics induced by a large variety of perturbations is essentially determined by the perturbation profile (3.10), i.e., the coarse-grained squared magnitude of the perturbation matrix elementsVµν in the unperturbed basis. In fact, the discussion of the response profile from Sec. 3.6.2 suggests that an even stronger reduction to just the parametersαvfrom (3.12) andvfrom (3.13) is legitimate. At the same time, minor fluctuations of the Vµν around their “true” values will not entail any noticeable deviations since, from a mathematical point of view, the propagator e−iHλtis continuous in these variables, and from a physical point of view, no experiment (real or numerical) would be reproducible otherwise.

Thereby, the large freedom to choose the precise distribution (3.24) of the perturbation operator leaves room to tailor an ensemble in such a way that the true perturbation of interest is realized with reasonably high probability. Moreover, the structure of the considered ensembles was designed to emulate common features of real perturbations such as bandedness, sparsity, etc. (see Sec. 3.3), reinforcing that real perturbations can be faithfully modeled or embedded in such an ensemble. Yet the fact that the true perturbation is sampled with reasonable probability within a single ensemble does not necessarily imply that it leads to the typical behavior of that ensemble when it comes to the relaxation behavior of observable expectation values. Indeed, if this behavior depended on some subtle details of the perturbation (for instance, the value of a single matrix element Vµν as an extreme example), the observed dynamics could still deviate from the typical behavior of the majority in case that this subtle feature happens to assume a rare value. Fortunately, again, the predicted behavior (3.160) is remarkably robust and—as demonstrated in Sec. 3.6.2 in particular—

independent of any fine-tuned perturbation characteristics.

Nevertheless, there are certain possible features of physical perturbations that are not explicitly accounted for. Such features are, for example, special operators commuting with the perturbation (notably the observable A [238, 259]) or quenches from systems with few symmetries (such as nonintegrable ones) to systems with more symmetries (such as integrable ones). Interestingly, it may still be possible to model such setups within the present approach if the considered observable and initial state do not explicitly “probe” these features. In any case, those examples constitute rather special situations.

In contrast, a potentially severe shortcoming is the only rudimentary modeling of the local and few-body character of interactions in common physical systems. While the considered ensembles explicitly allow for a banded and sparse matrix structure as it is often found as a result of local and few-body perturbations (e.g., if the reference system is noninteracting, see also below Eq. (3.10)), there is no geometry or notion of physical entities (“particles” or “bodies”) incorporated, mean-ing that sparsity and bandedness arise randomly in the considered perturbations and not in the way implied by the physical structure. It is known that such issues may render physical systems atypical with respect to certain random matrix ensembles [87, 107, 237], i.e., certain aspects of the true system may therefore indeed behave atypically compared to a selected ensemble, even though all considered characteristics (that is, the perturbation profile (3.10)) formally agree. Moreover, the outcome of our theoretical prediction (3.160) that essentially all experimentally relevant ob-servables behave similarly can certainly not be upheld if the initial state exhibits macroscopic inhomogeneities. For instance, Lieb-Robinson bounds [7, 105] (see also Sec. 2.2.1) limit the speed at which a local perturbation can spread across a locally interacting lattice system, implying that observables probing regions far away from the perturbation will notice the change later than those monitoring the vicinity of the perturbation, and thus the two will relax on different time scales.

On the other hand, these issues are again not expected to matter if the considered setup does not explicitly probe them. In particular, the ad hoc modeling of locality and sparsity should there-fore still be satisfactory if the initial state (or the observable) are sufficiently homogeneous on a macroscopic scale.

This raises an interesting point about admissible initial states and observables in general. The only explicit requirement on the initial state for the derivation was a well-defined macroscopic energy (cf. Prerequisite (i)). Whereas the prediction (3.160) and especially the response profile (3.146) do not exhibit an explicit dependence on any more specific properties of the state, there is an implicit dependence mediated by the density of states (or rather, the mean level spacingε), which sets a basic energy and thus also time scale for the reference system as well as a scale to gauge the perturbation strength. Since the mean level spacing may change with the state’s energy (see also Fig. 3.1), initial states pertaining to different energy windows will generally also lead to different relaxation characteristics in (3.160). On the other hand, besides a finite spectral range and resolution (cf. Eqs. (2.18) and (2.19)), there are no obvious restrictions as far as the observableAis concerned. However, we remark that there will always be special combinations of initial state and observable correlated in such a way that the resulting dynamics is atypical with respect to a given perturbation ensemble. Put differently, every combination of an observable Aand an initial state ρ(0) entails a set of atypical perturbations (which could in principle contain the true perturbation of interest), and furthermore these atypical perturbations will generally differ for differentA and ρ(0). A priori, however, it is unfortunately not immediately obvious whether a given combination ofA,ρ(0), andV is correlated such that it behaves atypically compared to the remaining members of a certain ensemble.

In summary, there is compelling evidence to believe that real perturbations can be modeled in terms of the ensembles considered here, unless there are specific reasons to the contrary. These reasons may be explicitly comprehensible in a given setup (macroscopically inhomogeneous initial state, quench from general to special case, ...) or they may be rooted more subtly in correlations between the Hamiltonian, the observable, and/or the initial state. Consequently, it is imperative to verify the prediction (3.160) in explicit numerical or experimental examples, and this will be the subject of the next subsection.