We review recent results on adiabatic theory for ground states of extended gapped fermionic lattice systems under several different assumptions. More precisely, we present generalized super-adiabatic theorems for extended but finite and infinite systems, assuming either a uniform gap or a gap in the bulk above the unperturbed ground state. The goal of this Review is to provide an overview of these adiabatic theorems and briefly outline the main ideas and techniques required in their proofs.

We review four recent results on adiabatic theory for ground states of extended finite and infinite fermionic lattice systems at zero temperature.1–3 These results are generalized super-adiabatic theorems (see Sec. I B) and concern Hamiltonians of the form

where the unperturbed Hamiltonian H0 is a sum-of-local-terms (SLT) operator describing short-range interacting fermions and is assumed to have a spectral gap above its ground state. This gap might be closed by the (small) perturbation ɛV, which is given by a short-range Hamiltonian, a Lipschitz potential, or a sum of both. Consequently, the results presented in this Review are adiabatic theorems for resonances of Hɛ (cf. Refs. 4 and 5).

The most important corollary and main motivation for proving such theorems in the context of extended fermionic lattice systems is the rigorous justification of linear response theory1,6 and the Kubo formula7 for (topological) insulators,8 such as quantum Hall systems,9 where the prototypical relevant perturbation is a linear external potential modeling a constant electric field closing the gap of H0 for every ɛ ≠ 0 (see Fig. 1).

FIG. 1.

Let H0 be a Hamiltonian with a gapped sector and a gap g. Perturbing with a Lipschitz potential v(x) = ɛ x, the gap gets closed (for large enough lattices). However, as indicated, a local gap persists and an electron at location x0 would either need to overcome the gap (vertical arrow) or tunnel along the distance g/ɛ (horizontal arrow) in order to make a transition from the gapped sector.1,6

FIG. 1.

Let H0 be a Hamiltonian with a gapped sector and a gap g. Perturbing with a Lipschitz potential v(x) = ɛ x, the gap gets closed (for large enough lattices). However, as indicated, a local gap persists and an electron at location x0 would either need to overcome the gap (vertical arrow) or tunnel along the distance g/ɛ (horizontal arrow) in order to make a transition from the gapped sector.1,6

Close modal

In the remainder of this Introduction, we first briefly discuss the connection between linear response and adiabatic theory in Sec. I A (see also Refs. 1 and 6). Furthermore, we point out the key ingredients and developments which allowed to prove1–3 the four adiabatic theorems presented in this Review. Afterward, in Sec. I B, we explain the notion of generalized super-adiabatic theorems and thereby introduce (super-adiabatic) non-equilibrium almost-stationary states (NEASSs)1 as the above mentioned resonances of Hɛ. A first brief but somewhat precise statement and overview of the results is given in Sec. I C.

The formalism of linear response theory7 has been widely used in physics to calculate the response of a system in thermal equilibrium to external perturbations. Put briefly, linear response theory provides an answer to the following question: What is the response of a system described by a Hamiltonian H0, which is initially in an equilibrium state ρ0, to a small static perturbation ɛV? Or, in somewhat more mathematical terms, what is the change10 

of the expectation value of an observable A induced by the perturbation ɛV to leading order in its strength 0 < ɛ ≪ 1? Here, ρɛ denotes the state of the system after the perturbation has been (adiabatically) turned on and σA denotes the linear response coefficient.

The answer to this fundamental question of linear response clearly hinges on the problem of determining ρɛ. Although in few particular situations one expects ρɛ to remain an equilibrium state for the perturbed Hamiltonian Hɛ = H0 + ɛV, the original linear response theory7 was developed for situations where the system is driven out of equilibrium, i.e., ρɛ being a resonance state. As prominently formulated by Simon11 in his “Fifteen problems in mathematical physics” from 1984, the latter non-equilibrium situation causes the main challenges in a rigorous mathematical treatment. However, in either case, the linear response coefficient σA is customarily expected to be given by the celebrated Kubo formula,7 and rigorously justifying it was formulated as one of the problems by Simon.11 For a more detailed recent review on the (mathematical) problem of proving Kubo’s formula and its relevance in the context of quantum Hall systems, we refer to Ref. 6.

In a nutshell, the problem of justifying linear response theory and proving Kubo’s formula is thus to verify that a system, initially in an equilibrium state ρ0, is adiabatically driven by a small perturbation ɛV into a non-equilibrium state ρɛρ0. Since the perturbation acts over a very long (macroscopic) time, this problem clearly supersedes standard perturbation theory: The change of the state being small is not a trivial consequence of the smallness of the perturbation ɛV. Instead, verifying that the two states, ρɛ and ρ0, are close to each other requires an adiabatic-type theorem.

However, even in our rather simple setting (zero temperature, assuming that ρ0 is the gapped ground state of H0 describing an extended fermionic lattice system, the perturbation ɛV might close the gap), the problem of justifying the linear response formalism also goes beyond standard adiabatic theory. In fact, the applicability of the standard adiabatic theorem of quantum mechanics is rather restrictive for the following three reasons:

  • The standard adiabatic theorem requires the perturbation ɛV to not close the spectral gap. In that scenario, it asserts that ρɛ is (close to) the gapped ground state of Hɛ = H0 + ɛV and, as such, a (nearly) equilibrium state.

  • Even if we neglect the first issue, the usual adiabatic theorem estimates the difference between ρɛ and the ground state of the perturbed Hamiltonian Hɛ in operator norm, leaving the translation to local differences in expectation values as an additional and potentially non-trivial step.

  • In general, extended systems are plagued by the orthogonality catastrophe: Whenever for single-particle states ψ,ψ̃ we have ψψ̃ε, the non-interacting many-particle states ⊗x∈Λψx and xΛψ̃x satisfy xΛψxxΛψ̃xε|Λ|, i.e., the norm-estimate deteriorates when |Λ| → ∞. This means that the approximation error in the standard adiabatic theorem grows with the systems size, and it is thus not applicable for macroscopic systems.

A major breakthrough in overcoming these obstacles has recently been achieved by Bachmann, De Roeck, and Fraas12 (see also their introductory lecture notes13). They proved the first adiabatic theorem for extended (but finite) lattice systems with short-range interactions, thereby solving the second and third problem in the list above. More precisely, their result concerns differences in expectation values and provides error estimates, which are uniform in the system size.

For these lattice systems with short-range interactions, well-known Lieb–Robinson bounds14–16 ensure a finite speed of correlation and prevent the build-up of long-range entanglement. Having Lieb–Robinson bounds at hand allowed Bachmann et al.18 to prove that the generator of the spectral flow, introduced by Hastings and Wen,17 is an SLT operator and thus preserves good locality properties. The general spectral flow technique can then be used to prove automorphic equivalence of two gapped ground states ρ0 and ρ1 of Hamiltonians H(0) and H(1), respectively: Given a smooth path sH(s) of (uniformly) gapped SLT Hamiltonians, their ground states are automorphically equivalent (equal up to a conjugation by unitaries) with the generator of the automorphism being an SLT operator.18 This automorphic equivalence allowed Bachmann et al.12 to prove a super-adiabatic theorem (see Sec. I B for an explanation of this notion) for such systems, however, still requiring the spectral gap not only for H0 but also for Hɛ, i.e., the gap must remain open.19 

The four theorems presented in this Review also solved the last remaining problem given under item (i) in the above list, i.e., they allow the perturbation ɛV to close the spectral gap of H0. The main idea for establishing this generalization is that a spatially local gap should suffice for an adiabatic theorem to hold. This underlies the space-time adiabatic perturbation theory originally developed for non-interacting fermions by Panati, Spohn, and Teufel,20,21 where one utilizes a gap that exists locally in space (and time) but does not exist globally. It also underlies the recent results by De Roeck, Elgart, and Fraas,22 where an adiabatic theorem holds even if the “spectral gap” is filled with eigenvalues, whose corresponding eigenvectors are spatially localized, leaving a gap (with smaller size) locally open. Finally, this is also the idea behind Theorems III and, especially, IV, where one still has an adiabatic-type theorem although the gap closes at the boundary of the lattices.

Combining the ideas from the space-time adiabatic perturbation theory with the methods invented in Ref. 12, the first of the four theorems presented in this Review was proven by Teufel.1 It concerns extended but finite systems and requires a spectral gap for H0, uniformly in the system size [see Assumption (GAPunif)]. The precise statement is formulated in Theorem I. In order to extend this result from finite lattices to an infinite system, Henheik and Teufel2 adapted ideas from Nachtergaele, Sims, and Young16 on controlling the thermodynamic limit of automorphisms with SLT generators. This result is formulated in Theorem II.

So far, all the mentioned results were obtained under the assumption of a (uniform) spectral gap for the finite systems (which also implies a gap for the infinite system). However, the recent result on automorphic equivalence with a gap only in the bulk (via the GNS construction) by Moon and Ogata23 opened the door for a new class of adiabatic theorems, where the unperturbed Hamiltonian H0 is no longer required to have a uniform spectral gap. Instead, Theorem III, originally proven by Henheik and Teufel,3 is a result for the infinite volume states and requires a gap in the bulk. This technically means a gap for the infinite system [cf. Assumption (GAPbulk)] but can be understood as requiring a local gap in the interior of the finite lattices (cf. Remark 4).

Moreover, by employing strong locality estimates from Refs. 16 and 23, and assuming fast convergence of ground states, Theorem III can be traced back to extended but finite systems that only have a gap in the bulk. This was also proven in Ref. 3 and is formulated in Theorem IV.

For the results presented in this Review, we consider time-dependent families

(1)

of many-body Hamiltonians for lattice fermions in ΓZd with short-range interactions. Here, Γ will either be a finite box Λ or the whole of Zd. For each tI, we denote by ρ0(t) the instantaneous ground state of H0(t) on the (quasi-local) algebra of observables AΓ. For simplicity of the presentation, we shall assume that the ground state is unique.24 Moreover, we assume that the ground state is separated by a gap from the rest of the spectrum [see Assumptions (GAPunif) and (GAPbulk) in Sec. III for the precise formulation]. The perturbation V(t) can be a Hamiltonian with short-range interactions or a possibly unbounded external Lipschitz potential or a sum of both [see Sec. II and Assumptions (INT1)–(INT4) in Sec. III].

As mentioned above, the main results presented in this Review are so-called generalized super-adiabatic theorems for ρ0(t), which we briefly explain in the following. For ɛ = 0, the results are “standardsuper-adiabatic theorems and establish the existence of super-adiabatic states ρ0η(t) on AΓ close to ρ0(t), i.e.,

such that the adiabatic time-evolution Ut,t0η on AΓ generated by 1ηH0() intertwines the super-adiabatic states to all orders in the adiabatic parameter η > 0, i.e.,

(2)

for all A in a dense subspace DAΓ. Throughout this entire Review, we shall study our system in the Heisenberg picture, meaning that the observable A evolves in time, not the state ρ0η(t0) (see also Proposition 3). Note that the comparison state ρ0η(t) does not involve any time evolution but simply depends on the Hamiltonian at time t (see Definition 1 for details). Here and in the following, we write the arguments of (densely defined) linear operators on AΓ inside the brackets [[⋅]] for better readability.

For ɛ > 0, the scope of the adiabatic theorem (2) extends considerably since the perturbation ɛV(t) might close the spectral gap and turn the ground state ρ0(t) of H0(t) into an instantaneous resonance state Πɛ(t) for Hɛ(t). These states have a lifetime of order O(ε) for the dynamics seisLHε(t), with LHε(t)[[]][Hε(t),] (formally) denoting the derivation associated with Hɛ(t). That is, for all nN and fixed t, it holds that

which is why they were called non-equilibrium almost-stationary states (NEASSs) in this context by Teufel.1 The generalized super-adiabatic theorems then establish the existence of a super-adiabatic NEASS Πɛ,η(t) on AΓ close to Πɛ(t) such that the adiabatic time-evolution Ut,t0ε,η generated by 1ηHε() approximately intertwines the super-adiabatic NEASSs in the following sense: for any n > d and for all ADAΓ, we have

(3)

uniformly for t in compact sets, which we call a generalized super-adiabatic theorem.

In our setting of gapped Hamiltonians H0 describing insulating materials, there is indeed a clear and simple physical picture suggesting the existence of NEASSs for Hɛ, as observed in Refs. 1 and 6 (see Fig. 1). For simplicity, assume that H0 is a periodic one-body operator in one spatial dimension and that the Fermi energy μ (chemical potential) lies in a gap of size g. For the perturbation, we consider the potential of a small constant electric field ɛ. In the initial state ρ0, before the perturbation is turned on, all one-body states with energy smaller than μ are occupied. After the voltage has been applied, the energy of an electron located at position x0 gets substantially shifted by ɛx0, but is only subject to small force of order ɛ. As indicated in Fig. 1, in order to make a transition, such an electron must either overcome the gap of size g or tunnel a macroscopic distance of order g/ɛ. Thus, although ρ0 is neither close to the ground state nor any other equilibrium state of the perturbed Hamiltonian Hɛ = H0 + ɛV, it is still almost stationary for Hɛ. This heuristic picture remains valid if short-range interactions between the electrons are taken into account.

While for ɛ = 0 the generalized super-adiabatic theorem (3) reduces to the standard one (2), for 0 < ɛ ≪ 1, the right-hand side of (3) is small if and only if also η is small, but not too small compared to ɛ, i.e., εn+1d+1η1 for some nN. Physically, this simply means that the adiabatic approximation breaks down when the adiabatic switching occurs at times that exceed the lifetime of the NEASS, an effect that has been observed in adiabatic theory for resonances before; see, e.g., Refs. 4 and 5. It can also be heuristically understood from the tunneling picture given in Fig. 1.

Moreover, in view of the linear response problem discussed in Sec. I A, let us only mention here that a statement such as (3), in fact, yields a solution to this problem after expanding the state Πɛ,η(t) in powers of ɛ, where the linear term [eventually stemming from the first order operator A1 given in (18)] does, in fact, constitute Kubo’s formula. See Refs. 1, 2, and 6 for details.

We shall establish the existence of super-adiabatic NEASSs in four generally quite different situations; the main differences are also summarized in Table I:

  • On finite systems ΛkZd with suitable boundary conditions, assuming that the unperturbed Hamiltonians H0Λk(t) have a gapped ground state uniformly in Λk, there exist NEASSs on AΛk such that the constants in (3) are independent of Λk. See Theorem I and Ref. 1.

  • Additionally assuming convergence of the Hamiltonians (they have a thermodynamic limit; cf. Definition 2) and ground states, there also exists a super-adiabatic NEASS on AZd after taking the thermodynamic limit ΛkZd. See Theorem II and Ref. 2.

  • For the infinite system Zd, assuming that the unperturbed Hamiltonian H0 has a unique gapped ground state (via the GNS construction), there exists a NEASS on AZd, while a (uniform) spectral gap for finite sub-systems is not required. See Theorem III and Ref. 3.

  • Additionally assuming a quantitative control on the convergence of the finite volume Hamiltonians HΛk(t) (they have a rapid thermodynamic limit; cf. Definition 5) and the unperturbed ground states in the thermodynamic limit, there also exist NEASSs on AΛk (again with a uniform constant) up to an error vanishing faster than any inverse polynomial in the distance to the boundary. See Theorem IV and Ref. 3.

TABLE I.

Overview of the adiabatic theorems and the original papers.

Finite volumeInfinite volume
Uniform gap Theorem I; see Ref. 1  Theorem II; see Ref. 2  
Gap in the bulk Theorem IV; see Ref. 3  Theorem III; see Ref. 3  
Finite volumeInfinite volume
Uniform gap Theorem I; see Ref. 1  Theorem II; see Ref. 2  
Gap in the bulk Theorem IV; see Ref. 3  Theorem III; see Ref. 3  

A typical example of a physically relevant class of Hamiltonians,1,6,25 to which the above generalized super-adiabatic theorems apply, is given by

(4)

modeling Chern or topological insulators. In agreement with the precise locality assumptions (INT1)–(INT4) in Sec. III, we suppose that the kinetic term T:ZdL(Cr) is an exponentially decaying function with T(−x) = T(x)*, the potential term ϕ:ZdL(Cr) is a bounded function taking values in the self-adjoint matrices, and the two-body interaction W:[0,)L(Cr) is exponentially decaying and also takes values in the self-adjoint matrices. Note that xy in the kinetic term refers to the difference modulo the imposed boundary condition on Λk. In the first two terms of (4), ax is the column vector of the annihilation operators ax,i (i labels internal degrees of freedom, such as spin) and ax* is the row vector of the creation operators ax,i* (see Sec. II). In addition, with a slight abuse of notation in the third term of (4), we wrote ax*ax for the row vector with entries ax,i*ax,i and ay*ay for the column vector with entries ay,i*ay,i.

It is well known that non-interacting Hamiltonians H0, i.e., with W ≡ 0, of type (4) on a torus (periodic boundary condition) have a uniform spectral gap [see Assumption(GAPunif)] whenever the chemical potential μ multiplying the number operator lies in a gap of the spectrum of the corresponding one-body operator on the infinite domain. It was recently shown26,27 that the spectral gap remains open when perturbing by sufficiently small short-range interactions W ≠ 0. On the other hand, the Hamiltonian H0 on a cube with open boundary condition has, in general, no longer a spectral gap because of the appearance of edge states. However, away from the boundary, a gap in the bulk [see Assumption (GAPbulk)] is still present. While also the uniqueness of the ground state is expected to hold for such models, to our knowledge, it has been shown only for certain types of quantum spin systems; cf. Refs. 28–32. For further details, we refer to the original papers.1–3 Finally, it is an interesting program to extend Table I by further rows representing different notions of a spectral gap for H0, e.g., a local gap as in Ref. 32 or even only a mobility gap (see Ref. 22 for a first result in this direction).

After a brief introduction to the relevant mathematical framework in Sec. II, we formulate the four main theorems in Sec. III. Ideas of their proofs are provided in Sec. IV.

In this section, we briefly introduce the (standard) mathematical framework used in the formulation of the adiabatic theorems. More explanations and details are provided in Refs. 13.

We consider fermions with r spin or other internal degrees of freedom on the lattice Zd. Let {XZd}{XZd:|X|<} denote the set of finite subsets of Zd, where |X| is the number of elements in X. For each XZd, let FX be the fermionic Fock space built up from the one-body space 2(X,Cr). The C*-algebra of bounded operators AXL(FX) is generated by the identity element 1AX and the creation and annihilation operators ax,i*, ax,i for xX and 1 ≤ ir, which satisfy the canonical anti-commutation relations (CARs). Whenever XX′, then AX is naturally embedded as a subalgebra of AX. For infinite systems, the algebra of local observables is defined as the inductive limit

with respect to the operator norm ‖ ⋅‖ is a C*-algebra, called the quasi-local algebra. The even elements AZd+AZd form a C*-subalgebra. In addition, note that for any XZd, the set of elements AXN commuting with the number operator NXxXax*axxXi=1rax,i*ax,i forms a subalgebra of the even subalgebra, i.e., AXNAX+AX. As only even observables will be relevant to our considerations, we will drop the superscript + from now on and redefine AZdAZd+.

Since a very similar construction is common for quantum spin systems (see, e.g., Ref. 16), all the results immediately translate to this setting.

We shall consider sequences of Hamiltonians defined on centered boxes Λk ≔ {−k, …,+k}d of size 2k with metric dΛk(,). This metric may differ from the standard 1-distance d(⋅, ⋅) on Zd restricted to Λk if one considers discrete tube or torus geometries, but satisfies the bulk-compatibility condition,

An interaction on a domain Λk is a map

with values in the self-adjoint operators. Note that the maps ΦΛk can be extended to maps on the whole {XZd} or restricted to a smaller Λl, trivially. In order to describe fermionic systems on the lattice Zd in the thermodynamic limit, one considers sequences Φ=ΦΛkkN of interactions on domains Λk and calls the whole sequence an interaction.

An infinite volume interaction is a map,

again with values in the self-adjoint operators. Such an infinite volume interaction defines a general interaction Ψ=ΨΛkkN by restriction, i.e., by setting ΨΛkΨ|{XΛk}.33 With any interaction Φ, one associates an operator family, which is a sequence A=(AΛk)kN of self-adjoint operators,

For any a > 0 and nN0, we define the norm

(5)

on the space of interactions.34 Note that these norms depend on the sequence of metrics dΛk on the cubes Λk, i.e., on the boundary conditions.

Similar constructions for interactions and interaction norms are long known. More commonly, the norms are independent of the particular lattice Λk and the interaction ΦΛkkN is given by restrictions of a single infinite volume interaction. Moreover, in earlier works,35,36 the authors did not require additional decay properties, which were only added later (see, e.g., Refs. 16, 37, and 38). The use of interactions and corresponding norms, which are not simply restrictions of an infinite volume interaction, originates in Ref. 25 to incorporate non-trivial boundary conditions. In order to control commutators with Lipschitz potentials (see Sec. II C), the dependence on the diameter dΛk-diam(X) was added in Ref. 1. Finally, to ensure the existence of the thermodynamic limit, it is necessary to require the bulk-compatibility condition.2,3 Yet another variant of defining interaction norms is to replace dist(x, y) with diam(X) in (5) (see, e.g., Refs. 38 and 39).

In order to quantify the difference of interactions in the bulk (see Sec. III B), we also introduce for any interaction ΦΛl on the domain Λl and any ΛM ⊂ Λl the quantity

where d and diam now refer to the 1-distance on Zd.

Let Ba,n be the Banach space of interactions with finite ‖ ⋅‖a,n-norm, and define the space of exponentially localized interactions as the intersection Ba,nN0Ba,n. In the literature, the vector spaces of operator families, which can be written in terms of such interactions, are denoted by La,n and La,. Moreover, we will be a bit sloppy in the following terminology and call the elements AΛk of an operator sequence Asum-of-local-terms (SLT) operators whenever its interaction ΦA has a finite interaction norm similar to (5), but with the exponential replaced by a function growing faster than any polynomial. This will allow us to formulate the results and the ideas of the proofs without too many details. For the precise conditions, see, e.g., see Ref. 2, Sec. 2.2.

Now, let IR be an open interval. We say that a map Φ:IBa,n is smooth and bounded whenever it is (i) term- and point-wise smooth in tI, i.e., tΦΛk(t,X) are C-functions for all kN and X ⊂ Λk, and (ii) suptIdidtiΦ(t)a,n< for all iN0. The corresponding spaces of smooth and bounded time-dependent interactions and operator families are denoted by BI,a,n and LI,a,n and are equipped with the norm ‖Φ‖I,a,n ≔ suptI‖Φ(t)‖a,n. We say that Φ:IBa, is smooth and bounded if Φ:IBa,n is smooth and bounded for all nN0, and we write BI,a, and LI,a, for the corresponding spaces of time-dependent exponentially localized interactions and operator families, respectively.

For (time-dependent) infinite volume interactions Ψ, we add a superscript ° to the norms and to the normed spaces defined above, emphasizing, in particular, the use of open boundary conditions, i.e., dΛkd. Note that the compatibility condition for the metrics dΛk implies that ‖Ψ‖a,n ≤ ‖Ψ‖°a,n.

For the perturbation, we will allow external potentials v=vΛk:ΛkRkN that satisfy the Lipschitz condition,

(6)

and call them for short Lipschitz potentials.40 With a Lipschitz potential v, we associate the corresponding operator-sequence Vv=(VvΛk)kN defined by

and denote the space of Lipschitz potentials by V. We emphasize that, since supkNsupxΛk|vΛk(x)| might be infinite, Vv is, in general, no SLT operator. However, this is still more restrictive than general onsite potentials because it only varies slowly in space. Moreover, we say that the map v:IV is smooth and bounded whenever (i) vΛk(x,) are C-functions for all kN and x ∈ Λk and (ii) it satisfies suptICdidtiv(t)< for all iN0. The space of smooth and bounded time-dependent Lipschitz potentials is denoted by VI.

As above, we also introduce infinite volume Lipschitz potentialsv:ZdR, which, again by restriction and invoking the compatibility condition for the metrics dΛk, can be viewed as a Lipschitz potential with dΛkd in (6). In addition, analogously to Sec. II B, for (time-dependent) infinite volume Lipschitz potentials, we add a superscript ° to the constant from (6) and to the spaces, emphasizing the use of open boundary conditions. Note that the compatibility condition for the metrics dΛk implies that CvC°v.

As mentioned in the Introduction, we shall distinguish two generally quite different settings regarding the presence of a spectral gap of the unperturbed Hamiltonian H0 grouped as Theorem I and Theorem II in Sec. III A and Theorem III and Theorem IV in Sec. III B. First, in Sec. III A, we will work under the assumption that there exists a sequence of subsystems (Λk)kN equipped with an appropriate metric (reflecting, e.g., periodic boundary conditions), ensuring that all (but finitely many) H0Λk have a uniform gap above their ground state, which is made precise in Assumption (GAPunif). Then, in Sec. III B, however, we drop this assumption and solely assume that H0 has a gap in the bulk, meaning that the GNS Hamiltonian, describing the system in the thermodynamic limit, has a spectral gap above its ground state eigenvalue zero [see Assumption (GAPbulk)]. Note that the second group of results is more general than the first group with regard to the gap condition since a uniform gap for finite systems guarantees a spectral gap for the GNS Hamiltonian describing the infinite system (see Proposition 5.4 in Ref. 41). Therefore, the second row in Table I somewhat improves the results in the first row since finding a suitable geometry for which one already has a spectral gap for finite systems is no longer necessary.

In the precise formulation of the adiabatic theorems, we shall frequently use the abbreviating phrase that a state Πɛ,η(t) is a super-adiabatic NEASS (see Sec. I B), which we generally define as follows: reminiscent of Refs. 13.

Definition 1

(Super-adiabatic non-equilibrium almost-stationary states).

We assume to be in the following general setting, which is made precise in concrete situations: For (small) ɛ > 0, define the time-dependent Hamiltonian,
and let ρ0(t) be (close to)42 the ground state of H0(t). Moreover, denote the Heisenberg time-evolution on the algebra of (quasi-local) observables AΓ generated by 1ηHε(t) as Ut,t0ε,η, where t, t0I for some open interval IR and η > 0 is a (small) adiabatic parameter.

Then, we say that a state Πɛ,η(t) on AΓis a super-adiabatic non-equilibrium almost-stationary state for the state ρ0(t) and the time-evolution Ut,t0ε,η on AΓ if it satisfies the following properties:

  1. Πɛ,η almost intertwines the time evolution: For any nN, there exists a constant Cn such that for any t, t0I and for all X ⋐ Γ and AAXAΓ, we have
    (7)
  2. Πɛ,η is local in time: Πɛ,η(t) only depends on H0 and V and their time derivatives at time t.

  3. Πɛ,η is stationary whenever the Hamiltonian is stationary: If for some fixed tI all time-derivatives of H0 and V vanish at time t, then Πɛ,η(t) equals the NEASS43 Πɛ(t) for the instantaneous ground state ρ0(t) and the time-evolution seisLHε(t) generated by the time-independent Hamiltonian Hɛ(t).

  4. Πɛ,η equals the (approximate) ground state ρ0 of H0 whenever the perturbation vanishes and the Hamiltonian is stationary: If for some tI all time-derivatives of H0 and V vanish at time t and V(t) = 0, then Πɛ,η(t) = Πɛ,0(t) = ρ0(t).

We could have written bound (7) in a more general form as indicated by (3). For example, we could allow (1 + |tt0|d+1) to be replaced by a constant CK < ∞, depending only on a compact subset KI of times or, similarly, |X|2 to be replaced by a constant CX < ∞, depending only on the support XZd of the observable A. In addition, the power of η in the denominator could be allowed to be more general, e.g., some constant Cd < ∞ instead of d + 1. However, the concrete form of (7) indeed matches the precise bounds of the results in Sec. III.

Throughout this section, we assume that H0 has a uniformly gapped unique ground state in the following sense.

(GAPunif) Assumptions on the ground state ofH0.

Let ΦH0=ΦH0ΛkkN be an interaction. There exists LN such that for all tI, kL, and corresponding Λk, the operator H0Λk(t) has a simple gapped ground state eigenvalue E0Λk(t)=infσ(H0Λk(t)), i.e., there exists g > 0 such that distE0Λk(t),σ(H0Λk(t))\{E0Λk(t)}g for all tI, kL. We denote the spectral projection of H0Λk(t) corresponding to E0Λk(t) by P0Λk(t) and write ρ0Λk(t)()trP0Λk(t) for the canonically associated state on AΛk.

A physically relevant class of Hamiltonians satisfying this assumption (possibly up to the uniqueness, which we require for simplicity of the presentation) was given in (4) in Sec. I C. In the following, we shall present adiabatic theorems for extended but finite systems (Theorem I) and for infinite systems (Theorem II) under Assumption (GAPunif).

1. Extended but finite systems

The basic assumption on the Hamiltonian says that it is composed of exponentially localized interactions and/or a Lipschitz potential.

(INT1) Assumptions on the interactions.

Let H0, H1 be the Hamiltonians of two time-dependent exponentially localized interactions, i.e., ΦH0,ΦH1BI,a, for some a > 0, and vVI be a time-dependent Lipschitz potential.

The following results due to Teufel1 marks the starting point for generalized super-adiabatic theorems for extended fermionic lattice systems.

Theorem I

[Adiabatic theorem for finite systems with a uniform gap (see Ref. 1, Theorem 5.1)].

Under Assumptions (GAPunif) and (INT1), there exists a sequence of near-identity44automorphismsβε,η,Λk(t)=eiεLSε,η(t)Λkwith SLT generatorsSɛ,ηfor anyɛ, η ∈ (0, 1] andtIsuch that the states
(8)
are super-adiabatic NEASSs for the Heisenberg time-evolutionUt,t0ε,η,ΛkonAΛkgenerated by1ηHε,Λk()with
uniformly inkL. That is, for everynN, there exists a constantCn, such that for anyAAX,ɛ, η ∈ (0, 1], and allt, t0I, it holds that

The proof of this result fundamentally builds on space-time adiabatic perturbation theory20,21 and technical estimates originally derived in Ref. 12. The latter show that the operations necessary for the construction of the generator of the near-identity automorphism in the definition of the NEASS in (8) (almost) preserve exponential localization required for the Hamiltonian (see Sec. IV). As already mentioned in the Introduction, although the adiabatic theorem in Ref. 12 is at first sight quite similar to the one above, it requires the perturbation to not close the spectral gap of the Hamiltonian H0 and is thus not generalized in the sense explained in Sec. I B.

2. Infinite systems

The next result is obtained from Theorem I by taking ΛkZd. This requires the interactions and the Lipschitz potential composing the Hamiltonian (1) to have a thermodynamic limit2 in the following sense.

Definition 2

(Thermodynamic limit of interactions and potentials).

  • An exponentially localized time-dependent interaction ΦBI,a, is said to have a thermodynamic limit (have a TDL) if there exists an infinite volume interaction ΨB°I,a, such that
    and we write ΦtdΨ in this case.

    An operator family is said to have a TDL if and only if the corresponding interaction does.

    For more general (non-exponentially localized) SLT operators, the definition is completely analogous.

  • A Lipschitz potential vVI is said to have a TDL if there exists an infinite volume Lipschitz potential vV°I such that

    Again, we write vtdv in this case.

Note that whenever Φ = Ψ for some infinite-volume interaction Ψ or v = v for some infinite volume Lipschitz potential v, both Φ and v trivially have a TDL.

The following proposition is a standard consequence of Lieb–Robinson bounds and shows that the property of having a TDL for interactions and Lipschitz potentials guarantees the existence of the thermodynamic limit for the corresponding evolution operators.16,45 We remark that it remains true under less restrictive assumptions on the localization quality of the interaction (see, e.g., Proposition 2.2 in Ref. 2).

Proposition 3

(Thermodynamic limit of evolution operators).

LetK0LI,a,andwVIboth have a thermodynamic limit, i.e., ΦK0tdΨK0andwtdwfor someΨK0B°I,a,andwV°I. SetK = K0 + Vw, and letUη,Λk(t,t0)denote the evolution family generated byKΛk(t)in scaled time withη > 0, i.e., the solution to the Schrödinger equation
withUη,Λk(t0,t0)=id. Then, there exists a co-cycle of automorphismsUt,t0η:AZdAZdsuch that for allAAloc,

The co-cycleUt,t0ηonly depends onΨK0andwand is generated by the time-dependent (closed) derivation(LK(t),D(LK(t)))associated withK(t).

As mentioned above, since Theorem II is deduced from Theorem I by taking ΛkZd, we will need to assume the existence of a thermodynamic limit for the building blocks of the Hamiltonian (1).

(INT2) Assumptions on the interactions.

For ΨH0,ΨH1B°I,a, for some a > 0 and vV°I, there exist ΦH0,ΦH1BI,a, and vVI with appropriate boundary conditions (encoded in the definition of the norms defining the spaces L and the Lipschitz condition) all having a TDL with the respective object as the limit, i.e., ΦH0tdΨH0, ΦH1tdΨH1, and vtdv.

We also assume the convergence of ground states by means of the Banach–Alaoglu theorem (the unit sphere in AZd* is weak*-compact), essentially only in order to avoid the extraction of a subsequence.

(Sunif) Assumptions on the convergence of states.

Assume that for every tI, the sequence (ρ0Λk(t))kN of ground states (naturally extended to the whole of AZd) converges in the weak*-topology to a state ρ0(t) on AZd, which we call the gapped limit ground state attI.

We can now formulate the second generalized super-adiabatic theorem concerning infinite systems with a uniform gap.2 

Theorem II

[Adiabatic theorem for infinite systems with a uniform gap (see Ref. 2, Theorems 3.2 and 3.5)].

Under Assumptions (GAPunif), (INT2), and (Sunif), there exists a near-identity automorphismβε,η(t)=eiεLSε,η(t)with SLT generatorsSɛ,ηfor anyɛ, η ∈ (0, 1] andtIsuch that the state
is a super-adiabatic NEASS for the Heisenberg time-evolution onAZdgenerated by1ηΨHε()with

The crucial point in the Proof of Theorem II in Ref. 2 is to show that the property of having a TDL is designed in such a way that it is preserved under all necessary operations for the construction of the NEASS (see Sec. IV). Therefore, also the near-identity automorphism from (8) converges as ΛkZd by means of Proposition 3.

In this section, we drop Assumption (GAPunif) of a uniform gap for finite systems, but merely work under the condition of a gap in the bulk, which is formulated via the Gelfand–Naimark–Segal (GNS) construction in Assumption (GAPbulk): Let ΨH0B°a,0 be an infinite volume interaction and LH0 denote the induced derivation on (a dense subset of) AZd. A state ω on AZd is called an LH0-ground state if and only if ω(A*LH0(A))0 for all AD(LH0). Let ω be an LH0-ground state and (Hω,πω,Ωω) be the corresponding GNS triple [i.e., Hω be a Hilbert space, i.e., πω:AL(Hω) be a representation, and, i.e., ΩωHω be a cyclic vector]. Then, there exists a unique densely defined, self-adjoint positive operator H0,ω ≥ 0 on Hω satisfying

(9)

for all AA and tR. We call this H0,ω the bulk Hamiltonian (or GNS Hamiltonian) associated with ΨH0 and ω. See Ref. 46 for the general theory.

We assume that ΨH0 has a unique gapped ground state in the following sense (cf. Refs. 3 and 23):

(GAPbulk) Assumptions on the ground state ofΨH0.

  • Uniqueness. For each tI, there exists a unique LH0(t)-ground state ρ0(t).

  • Gap. There exists g > 0 such that σ(H0,ρ0(t)(t))\{0}[g,) for all tI.

  • Regularity. For any strictly positive fS(R) (Schwarz functions), define Df as the set of observables AAZd for which AfA+supkN(1EΛk)[[A]]/f(k)<, where EΛk[[]] denotes the conditional expectation (see Ref. 2, Appendix C). Then, for any ADf, tρ0(t)(A) is differentiable and there exists a constant Cf such that

The smoothness of expectation values of (almost) exponentially localized observables as under item (iii) is a rather technical condition and a consequence of a uniform gap as in Assumption (GAPunif) (see Remark 4.15 in Ref. 23 and Lemma 6.0.1 in Ref. 47). Although the uniqueness of the ground state in item (i), which we required throughout this Review, is expected to hold for the physically relevant type of Hamiltonian (4), it has been shown, to our present knowledge, only in very specific quantum spin systems. These include (a) weak perturbations of non-interacting gapped frustration-free systems28,32 and (b) short-range interacting frustration-free models fulfilling local topological quantum order (LTQO).31,39

Remark 4.
As mentioned in the beginning of Sec. III, item (ii) holds, in particular, if one has a uniform gap for finite systems as spelled out in Assumption (GAPunif) since it cannot close abruptly in the thermodynamic limit for the GNS Hamiltonian (see Proposition 5.4 in Ref. 41). However, we observe that a considerably weaker sufficient condition for having a gap for the GNS Hamiltonian as in Assumption (GAPbulk) (ii) is to have a gap in the bulk for the finite systems Λk in the following sense: There exists g > 0 such that for all kN, there exists some l=l(k)N with l(k) → ∞ as k → ∞, and we have
(10)
for all AAΛl and all tI, where ρ0Λk(t) denotes a suitable ground state of H0Λk(t). Indeed, assuming that ρ0Λk(t)ρ0(t) for every tI,48 this simply follows after taking the limit k → ∞ on both sides of (10) and realizing that, as k → ∞, the set of admissible observables AAΛl(k) exhausts Aloc, which is dense in AZd by definition. The resulting inequality immediately yields the desired spectral gap for the GNS Hamiltonian (cf. Ref. 46, Proposition 5.3.19, and Ref. 30, Sec. 7).

In the following, we shall present adiabatic theorems for infinite systems (Theorem III) and for extended but finite systems (Theorem IV) under Assumption (GAPbulk).

1. Infinite systems

Analogously to Sec. III A, the basic assumptions on the Hamiltonian say that it is composed of exponentially localized interactions and/or a Lipschitz potential. In addition, the Hamiltonian H0 satisfies a technical regularity assumption in t, for which we recall that IR denotes an open time interval.

(INT3) Assumptions on the interactions.

  • Let ΨH0,ΨH1B°I,a, be time-dependent infinite volume interactions and vV°I be a time-dependent infinite volume Lipschitz potential.

  • Assume that the map IB°a,, tΨH0(t) is continuously differentiable.49 

We can now formulate the third generalized super-adiabatic theorem concerning infinite systems with a gap in the bulk.3 

Theorem III

[Adiabatic theorem for infinite systems with a gap in the bulk (see Ref. 3, Theorem 3.4)].

Under Assumptions (GAPbulk) and (INT3), there exists a near-identity automorphismβε,η(t)=eiεLSε,η(t)onAZdwith SLT generatorsSɛ,ηfor anyɛ, η ∈ (0, 1] andtIsuch that the state
is a super-adiabatic NEASS forρ0(t) and the Heisenberg time-evolution onAZdgenerated by1ηΨHε()with

The key role of the spectral gap condition is that it allows us to construct an inverse of the Liouvillian LH0(t), appearing in the construction of the NEASS, which maps SLT operators to SLT operators with slightly deteriorated locality properties. Hence, the inverse of LH0(t) is called the quasi-local inverse of the Liouvillian.50 Assuming a gap only in the bulk, as done in (GAPbulk), means that the action of the Liouvillian can only be inverted in the bulk (see Sec. IV).

2. Extended but finite systems

Contrary to the results in Sec. III A, the adiabatic theorem describing an infinite system with a gap in the bulk did not require any notion of having a TDL in its formulation. Instead, in order to derive a finite-volume analog from Theorem III (with qualitative additional error terms; see Theorem IV), we need to introduce the stronger notion of having a rapid thermodynamic limit for the exponentially localized interactions and the Lipschitz potential. We refer to Ref. 3 for a detailed discussion of this property.

Definition 5

(Rapid thermodynamic limit of interactions and potentials).

  • An exponentially localized time-dependent interaction ΦBI,a, is said to have a rapid thermodynamic limit with exponentγ ∈ (0, 1) (have a RTDLγ) if there exists an infinite volume interaction ΨB°I,a, such that
    (11)
    and we write ΦrtdΨ in this case.

    A family of operators is said to have a RTDL if and only if the corresponding interaction does.

    For more general (non-exponentially localized) SLT operators, the definition is completely analogous.

  • A Lipschitz potential vVI is said to have aRTDLγ if it is eventually independent of k, i.e., if there exists an infinite volume Lipschitz potential vV°I such that

    Again, we write vrtdv in this case.

In a nutshell, having a RTDLγ means that the interaction (or the Lipschitz potential) essentially agrees with a corresponding infinite volume object, up to terms located on a thin shell with relative size of order kγ−1 right at the boundary of Λk. Note that whenever Φ = Ψ for some infinite-volume interaction Ψ or v = v for some infinite volume Lipschitz potential v, both Φ and v trivially have a RTDLγ [with any exponent γ ∈ (0, 1)].

Theorem IV is deduced from Theorem III by comparing the time evolution Ut,t0ε,η and the near identity automorphism βɛ,η in the definition of the NEASS on the infinite system Zd with the same objects for large (but finite) systems Λk. Therefore, we will need to assume the existence of a rapid thermodynamic limit for the building blocks of Hamiltonian (1).

(INT4) Assumptions on the interactions.

The interactions ΦH0,ΦH1BI,a, and the Lipschitz potential vVI all have a RTDL, i.e., ΦH0rtdΨH0, ΦH1rtdΨH1, and vrtdv. The limiting objects ΨH0, ΨH1, and v satisfy Assumption (INT3).

In Theorem IV, we shall consider finite volume states ρ0Λk(t), which are close to the infinite volume ground state ρ0(t) away from the boundary in following sense.

(Sbulk) Assumption on the convergence of states.

The sequence (ρ0Λk(t))kN of states on AΛk converges rapidly to ρ0(t) in the bulk: there exist CR, mN, and hS such that for any finite XZd, AAX, and ΛkX,

While the sequence ρ0Λk(t)ρ0(t)|AΛk of simple restrictions satisfies Assumption (Sbulk) trivially, the adiabatic theorem ensures the existence of a super-adiabatic NEASS constructed for any such sequence.51 Most interesting for physical application would be a sequence of ground states ρ0Λk(t) of the finite volume Hamiltonians H0Λk(t). While the above assumption is expected to hold for any sequence of finite volume ground states for Hamiltonians modeling Chern or topological insulators such as in (4), the only result we are aware of indeed proving such a statement is again [see the discussion below Assumption (GAPbulk)] for weakly interacting spin systems.28 In spirit, assuming is very similar to supposing that the system satisfies local topological quantum order (LTQO)31,52 or a strong local perturbations perturb locally (LPPL) principle for perturbations acting at the boundary of the system.32,39

We can now formulate the fourth and last generalized super-adiabatic theorem concerning finite systems with a gap in the bulk.3 

Theorem IV

[Adiabatic theorem for finite systems with gap in the bulk (see Ref. 3, Theorem 4.1)].

Under Assumptions (GAPbulk), (INT4), and (Sbulk), there exists a sequence of near-identity automorphismsβε,η,Λk(t)=eiεLSε,η(t)Λkwith SLT generatorsSɛ,ηfor anyɛ, η ∈ (0, 1] andtIsuch that the states
are super-adiabatic NEASSs for the Heisenberg time-evolutionUt,t0ε,η,ΛkonAΛkgenerated by1ηHε,Λk()with
up to an error vanishing faster than any inverse polynomial in the distance to the boundary. That is, for anynN, there exists a constantCn, and for any compactKIandmN, there exists a constantC̃n,m,Ksuch that for allkN, allX ⊂ Λk, allAAX, and allt, t0K,
(12)

The above theorem asserts that by assuming (GAPbulk), one obtains similar adiabatic bounds also for states of finite systems (without a spectral gap), which are close to the infinite volume ground state in the bulk as formulated in Assumption (Sbulk). Since the adiabaticity potentially breaks at the boundaries of the finite systems, non-adiabatic effects arising close to the boundary may propagate into the bulk. Therefore, an additional error term appears, but it decays faster than any polynomial in the size of the finite system for any fixed η. The actual form of the additional error term in the last line of (12) coming out of the proof in (see Ref. 3, Sec. 5) is slightly better but more complicated, which is why we refrain from stating it here.

The main points in the Proof of Theorem IV, which we discuss in Sec. IV, are to show that (i) the property of having a RTDLγ is preserved under all necessary operations for the construction of the NEASS (similarly as for Theorem II) and (ii) having a RTDLγ for an interaction provides an explicit rate of convergence for the associated evolution family as in Proposition 3.

The goal of the present section is to convey the main ideas relevant for proving the individual theorems from Sec. III, where we already glimpsed the key steps required in their proofs. For many technical details, we refer the reader to the original works.1–3 

The fundamental conceptual idea behind the proof for all four variants of the generalized super-adiabatic theorems is a perturbative scheme, which was called space-time adiabatic perturbation theory in Refs. 20 and 21. The basic structure of this computation is most easily presented for finite systems, where no further technical difficulties arise since all appearing operators are in fact matrices and thus bounded. However, it is still necessary to show that all estimates are uniform in the size of the system Λk.

1. Extended but finite systems: Proof of Theorem I

The form in which we presented Theorem I differs slightly from the original result (see Ref. 1, Theorem 5.1). The original statement concerns a sequence Πnε,η,Λk(t)ρ0Λk(t)βnε,η,Λk(t) of states on Λk (indexed by nN), where

From this, Theorem I (and similarly all other three theorems) follows by a simple resummation of εjLAjε,η(t)Λk, which will be discussed in Sec. IV C.

The main idea of the proof is to choose each operator Ajε,η,Λk(t), j = 1, …, n, in such a way that the jth-order term in the perturbative scheme vanishes. For the n-dependent result (i.e., prior to resummation), we apply the fundamental theorem of calculus to get

(13)

and then aim to bound the integrand. Calculating the derivative by using the chain rule and Duhamel’s formula leaves us with

(14)

where Qnε,η,Λk(s) is a shorthand notation for

(15)

and VΛk=VvΛk+H1Λk. Here, IsΛk(Ḣ0Λk(s)) is the SLT generator of the parallel transport within the vector-bundle Ξ0,IΛk over I defined by tρ0Λk(t).53 This parallel transport is also known as the spectral flow, which plays a fundamental role in proving automorphic equivalence of gapped ground state phases (see, e.g., Refs. 12 and 18). Moreover, the operator IsΛk:AΛkAΛk is called the quasi-local inverse of the Liouvillian38 LH0(s)Λk since it satisfies1,12

(16)

and also preserves good localization of its argument (in particular, it maps SLT operators to SLT operators). This combined property of IsΛk heavily relies on the ground state ρ0Λk(s) being gapped1,12,16 and will be of fundamental importance in the following.

In the last line of (15), we expanded in powers of ɛ and η in the sense that Rjε,η,Λk(s), for jn, are polynomials in η/ɛ of order (at most) j with ɛ- and η-independent SLT operators as coefficients. A more detailed step-by-step calculation can be found in the Proof of Proposition 5.1 in Ref. 1. Let us here only report the general structure

(17)

where the first remainder term is given by

and all other R̃jε,η,Λk(s) are composed of iterated commutators of the operators Aiε,η,Λk(s) and Ȧiε,η,Λk(s), for i < jn, with H0Λk(s) and VΛk(s). In contrast to general onsite potentials, the commutator of a Lipschitz potential with an SLT operator is an SLT operator itself (see Ref. 1, Lemma 2.1). For the commutator of SLT operators, this is easy to see.

We now consider individual terms from (15) when plugged into (14). The zero-order term vanishes because ρ0Λk(s) is the ground state of H0Λk(s). By application of (16), we can iteratively choose

(18)

such that (14) vanishes up to

(19)

Moreover, all the operations involved in calculating the Ajε,η,Λk(s), i.e., taking commutators and applying the quasi-local inverse of the Liouvillian, preserve the locality properties of the operators, as shown in the appendices of Refs. 1 and 25, which are heavily based on Ref. 12. Hence, also all Ajε,η,Λk are SLT operators.

It turns out that also Rn+1ε,η,Λk is a polynomial in η/ɛ of order at most n + 1 and its coefficients, as we just explained, are SLT operators (see Ref. 1, Proof of Proposition 6.1). Thus, the absolute value of (13) is bounded by

(20)

where we essentially used a generalized Lieb–Robinson bound (see Ref. 1, Lemma B.5) to estimate the commutator. Note that the (1+|tt0|/ηd)-factor comes from the Lieb–Robinson bound and the adiabatic 1/η-scaling of the time evolution Ut,s. The (1 + (η/ɛ)n+1)-factor comes from bounding the interaction norm of Rn+1ε,η,Λk(s) by separating the polynomial dependence on η/ɛ such that Cn is independent of Λk, ɛ, and η. We have thus shown that the NEASS almost intertwines the time evolution, i.e., item 1 of Definition 1.

We are left with discussing the remaining three characterizing properties of the NEASS given in Definition 1: By construction, all Ajε,η,Λk(t) depend only on H0Λk(t) and VΛk(t) and their jth derivatives at time t. This shows that the NEASS is local in time, i.e., item 2. Moreover, if all time derivatives of H0 and V vanish for some tI, then all non-constant (i.e., in front of some positive power of η/ɛ) coefficients in Rjε,η,Λk vanish and Πnε,η,Λk(t)=Πnε,0,Λk(t). This shows that the NEASS is stationary whenever the Hamiltonian is stationary, i.e., item 3. If, for some tI, Ḣ0Λk(t) and VΛk(t) vanish, then R̃1ε,η,Λk and thus A1ε,η,Λk also vanish. If additionally all derivatives of H0Λk and VΛk at t vanish, also R̃jΛk(t) and thus Ajε,η,Λk(t) vanish. Hence, βnε,η,Λk(t)=1Λk and the NEASS equals the ground state, i.e., item 4 holds.

The above listed general arguments immediately translate to the other three theorems.

2. Infinite systems: Proof of Theorem II

Without any further assumptions, the sequence Hamiltonian Hε,Λk and its constituents H0Λk and VΛk could have nothing in common for different lattice sizes k (they might even describe different physical systems), so taking the limit ΛkZd might not be well-defined. In order to avoid this somewhat meaningless situation, we assumed that the building blocks of the Hamiltonian have a TDL [see Definition 2 and Assumption (INT2)] and also the sequence of ground states (ρ0Λk(t))kN converges [Assumption (Sunif)]. Since the property of having a TDL guarantees the existence of the thermodynamic limit for the corresponding evolution operators (see Proposition 3 and Ref. 16), it remains to show that the operator sequences Ajε,η,Λk(t)kN, j = 1, …, n, constructed in Sec. IV A 1 also have a TDL. More precisely, one needs to show that taking time-derivatives, sums of commutators with the building blocks of Hɛ (and Ḣ0), and the inverse of the Liouvillian [see (17)] leaves the property of having a TDL for SLT operators invariant, which is in fact the main point of the proof in Ref. 2. It is then straightforward to show that compositions of states and automorphisms, all having a thermodynamic limit, converge as ΛkZd. Since the constant Cn from (20) is uniformly bounded in k, the (sketch of a) Proof of Theorem II is complete.

For systems having a spectral gap only in the bulk (i.e., for the GNS Hamiltonian), the characteristic (16) of IsΛk, that it essentially inverts the Liouvillian LH0(s)Λk (and still maps SLT operators to SLT operators), is now only fulfilled for certain B1 and B2 in a dense domain DAZd after taking the limit ΛkZd (see Ref. 3, Proposition 3.3). Presuming that the limit actually exists, this point is the main challenge in proving an adiabatic theorem under the less restrictive gap assumption (GAPbulk).

1. Infinite systems: Proof of Theorem III

As just explained, the main difficulty in proving Theorem III is that (16) only holds if H0Λk is gapped. On top of that, we cannot handle the limit ΛkZd of R̃jε,η,Λk directly nor could they be used in the infinite volume version of (16) because it only holds for B1,B2DAZd. However, the rest of the construction from Sec. IV A 1 is still valid, but the lower order terms in (13) have a non-vanishing contribution in finite domains. We thus repeat this construction but take coefficients Ajε,η,Λk,Λl(t), which are built up from H0Λk(t) but restricting the perturbations Ḣ0(t) and V(t) to Λl with l < k. In this way, one can take the limit ΛkZd in (16) with B1=limkR̃jε,η,Λk,ΛlAZd [see (22) and (23) and the comment thereafter for technical obstructions in taking the limit]. Using this notational convention, we introduce the states

where ρ0(t) is the infinite volume ground state, and compare them to the actual objects in infinite volume while estimating

(21)

by means of the triangle inequality. Since all the interactions (and the Lipschitz potential) have a TDL, one can prove [see Ref. 3, Sec. 5.1(b)] that the first and last summand in (21) can be made arbitrarily small for k,lN large enough, and we can thus focus on the second summand. However, since (16) only holds in the limit ΛkZd and also ρ0(t) is not necessarily a ground state of H0Λk(t), the lower order terms in the analog of (14) and (15) do not vanish for finite k and l. Instead, only

(22)

and

(23)

for all lN and uniformly for s and t in compacts. These statements require a careful analysis of deteriorating localization properties along the expansion and convergence estimates in norms measuring the quality of localization [cf. the norm ‖ ⋅‖f introduced in Assumption (GAPbulk) (iii)] such that the limits really converge to the infinite volume version of (16) with B1 and B2 in a dense domain DAZd. For further details, we refer to Proposition 3.2 and the statements in Appendix B of Ref. 3, which are adaptions of technical estimates that were originally established for the proof of automorphic equivalence with a gap only in the bulk.23 Now, combining (22) and (23) with the estimates on the first and third summand in (21), we conclude that all the lower order terms vanish in the limit k → ∞ followed by l → ∞, which finishes our sketch of the Proof of Theorem III.

2. Extended but finite systems: Proof of Theorem IV

Let us briefly explain the strategy to prove Theorem IV. In order to show (12), we first estimate

(24)

and treat the three summands separately. The second summand corresponds to the infinite system and can be estimated by means of Theorem III such that it accounts for the first contribution on the RHS of (12). We are left with bounding the remaining two summands in (24). These contribute the additional error term on the RHS of (12). To estimate them, we need explicit control on the speed of convergence (it must be faster than any inverse polynomial) for the states [see Assumption (Sbulk)] and automorphisms βnε,η,Λk and Ut,t0ε,η,Λk. For the time evolution Ut,t0ε,η,Λk, the rapid convergence to Ut,t0ε,η is ensured by supposing that the building blocks of Hɛ have a RTDL [see Definition 5 and Assumption (INT4)]. This was carried out in see Ref. 3, Appendix B, building on estimates from see Ref. 16, Sec. 3. We remark that the adiabatic 1/η-scaling of the time evolution is responsible for the factor η appearing in the additional error term in (12). In order to show that also βnε,η,Λkβnε,η converges sufficiently fast, we need to show that all Ajε,η have a RTDL, i.e., the operations involved in constructing the generator of βnε,η leave the property of having a RTDL (essentially) invariant (see Ref. 3, Appendix C). This finishes the sketch of the Proof of Theorem IV and we refer to see Ref. 3, Sec. V B for further details.

As mentioned in the beginning of Sec. IV A 1, the statements formulated in Sec. III require a resummation, which we explain in the following. First, note that the generator εSnε,η of βnε,η constructed above can be rewritten as εSnε,η=j=1ni=0jεiηj1Aj,i, where the coefficients Aj,i are time-dependent SLT operators and independent of ɛ and η. Now, it is easy to show (see Ref. 2, Lemma E.1) that there exist a sequence δj → 0 and constants Cn such that the resummed generator

(25)

satisfies

where ‖ ⋅‖SLT denotes an interaction norm similar to (5). Resummations of this type are standard, e.g., in microlocal analysis,54 and the above estimate immediately leads to the bounds (cf. Ref. 2, Lemmata E.3, E.4),

(26)

and

(27)

uniformly in the size of the system Λk. In the context of Theorems II and III, corresponding estimates hold in infinite volume, i.e., without the subscript Λk.

Next, since the sum in (25) is finite for every fixed ɛ > 0, also the resummed generator Sε,η,Λk has a TDL as soon as Snε,η,Λk has a TDL. Therefore, the states Πε,η,Λk constructed using Sε,η,Λk instead of Snε,η,Λk have a well-defined thermodynamic limit Πɛ,η (see Ref. 2, Lemma E.2), and since the bounds (26) and (27) are independent of Λk, they also hold for the respective objects in the thermodynamic limit. Hence, the results formulated in Sec. III can be concluded by combining the n-dependent statements discussed earlier in this section with the bounds (26) and (27) (or their infinite volume correspondents).

It is a pleasure to thank Stefan Teufel for numerous interesting discussions, fruitful collaboration, and many helpful comments on an earlier version of the manuscript. J.H. acknowledges partial financial support from the ERC Advanced Grant No. 101020331 “Random matrices beyond Wigner-Dyson-Mehta.” T.W. acknowledges financial support from the DFG research unit FOR 5413 “Long-range interacting quantum spin systems out of equilibrium: Experiment, Theory and Mathematics.”

The authors have no conflicts to disclose.

Joscha Henheik: Writing – original draft (equal); Writing – review & editing (equal). Tom Wessel: Writing – original draft (equal); Writing – review & editing (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
S.
Teufel
, “
Non-equilibrium almost-stationary states and linear response for gapped quantum systems
,”
Commun. Math. Phys.
373
,
621
653
(
2020
).
2.
J.
Henheik
and
S.
Teufel
, “
Adiabatic theorem in the thermodynamic limit: Systems with a uniform gap
,”
J. Math. Phys.
63
,
011901
(
2022
).
3.
J.
Henheik
and
S.
Teufel
, “
Adiabatic theorem in the thermodynamic limit: Systems with a gap in the bulk
,”
Forum Math. Sigma
10
,
e4
(
2022
).
4.
W. K.
Abou Salem
and
J.
Fröhlich
, “
Adiabatic theorems for quantum resonances
,”
Commun. Math. Phys.
273
,
651
675
(
2007
).
5.
A.
Elgart
and
G. A.
Hagedorn
, “
An adiabatic theorem for resonances
,”
Commun. Pure Appl. Math.
64
,
1029
1058
(
2011
).
6.
J.
Henheik
and
S.
Teufel
, “
Justifying Kubo’s formula for gapped systems at zero temperature: A brief review and some new results
,”
Rev. Math. Phys.
33
,
2060004
(
2021
).
7.
R.
Kubo
, “
Statistical-mechanical theory of irreversible processes. I. General theory and simple applications to magnetic and conduction problems
,”
J. Phys. Soc. Jpn.
12
,
570
586
(
1957
).
8.
G.
Marcelli
,
G.
Panati
, and
S.
Teufel
, “
A new approach to transport coefficients in the quantum spin Hall effect
,”
Ann. Henri Poincaré
22
,
1069
1111
(
2021
).
9.
D. J.
Thouless
,
M.
Kohmoto
,
M. P.
Nightingale
, and
M.
den Nijs
, “
Quantized Hall conductance in a two-dimensional periodic potential
,”
Phys. Rev. Lett.
49
,
405
408
(
1982
).
10.

To be consistent with the rest of this Review, we view states as linear functionals on the algebra of observables (see Sec. II).

11.
B.
Simon
, “
Fifteen problems in mathematical physics
,” in
Perspectives in Mathematics: Anniversary of Oberwolfach
, edited by
W.
Jäger
,
J.
Moser
, and
R.
Remmert
(
Birkhäuser
,
Basel
,
1984
), pp.
423
454
.
12.
S.
Bachmann
,
W.
De Roeck
, and
M.
Fraas
, in (
American Mathematical Society
,
2020
), Vol. 741.
13.
S.
Bachmann
,
W.
De Roeck
, and
M.
Fraas
, “
The adiabatic theorem in a quantum many-body setting
,” in (
American Mathematical Society
,
2020
), Vol. 741.
14.
E. H.
Lieb
and
D. W.
Robinson
, “
The finite group velocity of quantum spin systems
,”
Commun. Math. Phys.
28
,
251
257
(
1972
).
15.
B.
Nachtergaele
,
R.
Sims
, and
A.
Young
, “
Lieb-Robinson bounds, the spectral flow, and stability of the spectral gap for lattice fermion systems
,” in
Mathematical Results in Quantum Physics. Contemporary Mathematics
, edited by
F.
Bonetto
,
D.
Borthwick
,
E.
Harrell
, and
M.
Loss
(
American Mathematical Society
,
2018
), Vol. 717, pp.
93
115
.
16.
B.
Nachtergaele
,
R.
Sims
, and
A.
Young
, “
Quasi-locality bounds for quantum lattice systems. I. Lieb-Robinson bounds, quasi-local maps, and spectral flow automorphisms
,”
J. Math. Phys.
60
,
061101
(
2019
).
17.
M. B.
Hastings
and
X.-G.
Wen
, “
Quasiadiabatic continuation of quantum states: The stability of topological ground-state degeneracy and emergent gauge invariance
,”
Phys. Rev. B
72
,
045141
(
2005
).
18.
S.
Bachmann
,
S.
Michalakis
,
B.
Nachtergaele
, and
R.
Sims
, “
Automorphic equivalence within gapped phases of quantum lattice systems
,”
Commun. Math. Phys.
309
,
835
871
(
2012
).
19.

A slight generalization of their result can be found in Ref. 25, where the authors used an alternative gauge with a time-dependent vector potential for a quantum Hall model.

20.
G.
Panati
,
H.
Spohn
, and
S.
Teufel
, “
Space-adiabatic perturbation theory
,”
Adv. Theor. Math. Phys.
7
,
145
204
(
2003
).
21.
G.
Panati
,
H.
Spohn
, and
S.
Teufel
, “
Effective dynamics for Bloch electrons: Peierls substitution and beyond
,”
Commun. Math. Phys.
242
,
547
578
(
2003
).
22.
W.
De Roeck
,
A.
Elgart
, and
M.
Fraas
, “
Locobatic theorem for disordered media and validity of linear response
,” arXiv:2203.03786 (
2022
).
23.
A.
Moon
and
Y.
Ogata
, “
Automorphic equivalence within gapped phases in the bulk
,”
J. Funct. Anal.
278
,
108422
(
2020
).
24.

We refer to the original papers1–3 for the most general assumptions. However, note that the results from Ref. 3, corresponding to our Theorems III and IV, are only formulated for a unique ground state although the underlying result on automorphic equivalence of gapped phases23 can be easily generalized to any gapped pure state (see 23, Remark 1.4). In general, allowing for a degenerate ground state (or even a gapped spectral patch) requires understanding an enhanced modification of the spectral flow.

25.
D.
Monaco
and
S.
Teufel
, “
Adiabatic currents for interacting fermions on a lattice
,”
Rev. Math. Phys.
31
,
1950009
(
2019
).
26.
M. B.
Hastings
, “
The stability of free Fermi Hamiltonians
,”
J. Math. Phys.
60
,
042201
(
2019
).
27.
W.
De Roeck
and
M.
Salmhofer
, “
Persistence of exponential decay and spectral gaps for interacting fermions
,”
Commun. Math. Phys.
365
,
773
796
(
2019
).
28.
D. A.
Yarotsky
, “
Uniqueness of the ground state in weak perturbations of non-interacting gapped quantum lattice systems
,”
J. Stat. Phys.
118
,
119
144
(
2005
).
29.
J.
Fröhlich
and
A.
Pizzo
, “
Lie–Schwinger block-diagonalization and gapped quantum chains
,”
Commun. Math. Phys.
375
,
2039
2069
(
2020
).
30.
B.
Nachtergaele
,
R.
Sims
, and
A.
Young
, “
Quasi-locality bounds for quantum lattice systems. Part II. Perturbations of frustration-free spin models with gapped ground states
,”
Ann. Henri Poincaré
23
,
393
511
(
2022
).
31.
B.
Nachtergaele
,
R.
Sims
, and
A.
Young
, “
Stability of the bulk gap for frustration-free topologically ordered quantum lattice systems
,” arXiv:2102.07209 (
2021
).
32.
J.
Henheik
,
S.
Teufel
, and
T.
Wessel
, “
Local stability of ground states in locally gapped and weakly interacting quantum spin systems
,”
Lett. Math. Phys.
112
,
9
(
2022
).
33.

We will use the convention that Φ denotes general interactions and Ψ denotes infinite volume interactions.

34.

One should be aware that the norm definition (5) is slightly modified compared to the original works1–3 for the sake of simplicity in presentation. For more general and precise statements of the theorems, we refer the reader to the original works.

35.
D. W.
Robinson
, “
Statistical mechanics of quantum spin systems
,”
Commun. Math. Phys.
6
,
151
160
(
1967
).
36.
D.
Ruelle
,
Statistical Mechanics. Rigorous Results
(
W. A. Benjamin Inc.
,
New York
,
1969
).
37.
B.
Simon
,
The Statistical Mechanics of Lattice Gases
(
Princeton University Press
,
Princeton, NJ
,
1993
), Vol. I.
38.
M. B.
Hastings
and
T.
Koma
, “
Spectral gap and exponential decay of correlations
,”
Commun. Math. Phys.
265
,
781
804
(
2006
).
39.
S.
Bachmann
,
W.
De Roeck
,
B.
Donvil
, and
M.
Fraas
, “
Stability of invertible, frustration-free ground states against large perturbations
,”
Quantum
6
,
793
(
2022
).
40.

Teufel1 instead allowed slightly more general slowly varying potentials. And while the phrase captures the idea very well, the technical definition is less transparent and slightly complicates the presentation of the proofs. Hence, we here, as in Refs. 2 and 3, restrict to the subclass of Lipschitz potentials.

41.
S.
Bachmann
,
W.
Dybalski
, and
P.
Naaijkens
, “
Lieb-Robinson bounds, Arveson spectrum and Haag-Ruelle scattering theory for gapped quantum spin systems
,”
Ann. Henri Poincaré
17
,
1737
1791
(
2016
).
42.

See the comment below Assumption (Sbulk) for a precise definition.

43.

It follows from the construction sketched in Sec. IV A 1 that Πɛ(t) = Πɛ,0(t). Moreover, Πɛ(t) is almost stationary with a bound as in (7) where the fraction is replaced by ɛn+1.

44.

Indeed, supkNAβε,η,Λk(t)[[A]](ε+η)CA|X||t| for AAX and t in a bounded interval by Ref. 16, Theorem 3.4(i).

45.
J.-B.
Bru
and
W.
de Siqueira Pedra
,
Lieb-Robinson Bounds for Multi-Commutators and Applications to Response Theory
(
Springer
,
2017
).
46.
O.
Bratteli
and
D. W.
Robinson
,
Operator Algebras and Quantum Statistical Mechanics 2. Equilibrium States. Models in Quantum Statistical Mechanics
(
Springer
,
1997
).
47.
A.
Moon
, “
Stable properties of gapped ground state phases in quantum spin chains
,” Ph.D. thesis (
UC Davis
,
2020
); available at https://www.math.ucdavis.edu/~tdenena/dissertations/202003_Moon_Dissertation.pdf.
48.

Note that the sequence (ρ0Λk(t))kN is compact for every fixed tI (Banach–Alaoglu theorem). Moreover, it is shown in Proposition 5.3.25 in Ref. 46 that every limit point of a sequence of ground states associated with a converging sequence of derivations LH0(t)ΛkLH0(t) is a ground state of the limiting derivation.

49.

Note that this technical assumption does not follow from ΨH0B°I,a, as the spaces of smooth and bounded interactions are defined via term-wise and point-wise time derivatives (cf. Sec. II B).

50.

This particular phrase was used in Ref. 1. Others call it local inverse12,25 or just inverse2,6,16. In Ref. 3, it was called SLT inverse. To avoid confusion, we want to reserve the SLT prefix for operators, i.e., SLT operator or SLT generator, but not for maps between SLT operators.

51.

This is why we wrote “(close to) a ground state” in Definition 1.

52.
S.
Michalakis
and
J. P.
Zwolak
, “
Stability of frustration-free Hamiltonians
,”
Commun. Math. Phys.
322
,
277
302
(
2013
).
53.

Since we assumed the uniqueness of the ground state ρ0Λk(t), the vector-bundle Ξ0,IΛk is one-dimensional. If this were not the case, one had to include further terms generating the internal dynamic in Ξ0,IΛk (see Refs. 1 and 2).

54.
A.
Martinez
,
An Introduction to Semiclassical and Microlocal Analysis
(
Springer
,
2002
).