Lieb-Robinson bounds show that the speed of propagation of information under the Heisenberg dynamics in a wide class of nonrelativistic quantum lattice systems is essentially bounded. We review works of the past dozen years that has turned this fundamental result into a powerful tool for analyzing quantum lattice systems. We introduce a unified framework for a wide range of applications by studying quasilocality properties of general classes of maps defined on the algebra of local observables of quantum lattice systems. We also consider a number of generalizations that include systems with an infinite-dimensional Hilbert space at each lattice site and Hamiltonians that may involve unbounded on-site contributions. These generalizations require replacing the operator norm topology with the strong operator topology in a number of basic results for the dynamics of quantum lattice systems. The main results in this paper form the basis for a detailed proof of the stability of gapped ground state phases of frustrationfree models satisfying a local topological quantum order condition, which we present in a sequel to this paper.
I. INTRODUCTION
Quantum many-body theory comes in two flavors. The first is the relativistic version generically referred to as Quantum Field Theory (QFT), used for particle physics, and the second is non-relativistic many-body theory, which serves as the basic framework for most of condensed matter physics. The close physical and mathematical similarities between the two have long been recognized and exploited with great success. Bogoliubov’s theory of superfluidity and the BCS theory of superconductivity serve as definitive proof that quantum fields are a useful and even fundamental concept for understanding nonrelativistic many-body systems. Condensed matter theorists have developed field theory techniques that are now omnipresent in the subject.1,3,42,62,79,86,125 See Refs. 51 and 117 for reviews on recent progress in the mathematics of Bogoliubov’s theory and the BSC theory of superfluidity.
The absence of Lorentz-invariance (and the associated constant speed of light c leading to the all-important property of locality in the sense of Haag48) in nonrelativistic many-body theories is the most obvious difference between the two perspectives. Given the importance of this invariance in QFT, which plays an essential role in deriving many of the fundamental properties, and the strong constraints it imposes on its mathematical structure, one would expect that its absence would prevent any close analogy between the relativistic and the nonrelativistic setting to hold true. Contrary to this expectation, successful applications of QFT to problems of condensed matter physics have been numerous. Quantum field theories have provided accurate descriptions as effective theories describing important aspects such as excitation spectra and derived quantities. This typically involves a scaling limit of some type. Conformal field theories have been spectacularly effective in describing and classifying second order phase transitions. Also here a scaling limit is often implied.
The quasilocality properties that are the subject of study in this paper partly explain the closer-than-expected similarities between QFT and the nonrelativistic many-body theory of condensed matter systems. More importantly, they make it possible to prove that much of the mathematical structure of QFT can be found in nonrelativistic many-body systems in an approximate sense. Instead of asymptotic statements and qualitative comparisons, we can prove quantitative estimates: the quasilocality of the dynamics is characterized by an approximate light-cone with errors that can be bounded explicitly. These are the Lieb-Robinson bounds, which have been an essential ingredient in a large number of breakthrough results in the past dozen years.
Although the result of Lieb and Robinson dates back to the early 1970s,75,112 the impetus for the recent flurry of activity and major applications came from the work of Hastings on the Lieb-Schultz-Mattis theorem in arbitrary dimension.54 The possibility of adapting some of the major results of QFT to (nonrelativistic) quantum lattice systems was anticipated by others. For example, Wreszinski studied the connection between the Goldstone theorem,73 charges, and spontaneous symmetry breaking.127 A rigorous proof of a nonrelativistic exponential clustering theorem, long known in QFT,43,114 did not appear until the works.57,92 The time-evolution of quantum spin systems turns local observables into quasilocal ones. Lieb-Robinson bounds were applied to approximate such quasilocal observables by strictly local ones with an error bound in Refs. 23, 88, and 91. These (sequences of) strictly local approximations are those that are used in many concrete applications and also have a conceptual appeal. Further extensions of Lieb-Robinson bounds and a sampling of interesting applications are discussed in Sec. III.
Apart from offering a review of the state of the art of quasilocality estimates, in this paper, we also extend existing results in the literature in several directions. First, for most of the results, we allow the quantum system at each lattice site to be described by an arbitrary infinite-dimensional Hilbert space. For many results, the single-site Hamiltonians may be arbitrary densely defined self-adjoint operators. Another generalization in comparison to the existing literature, made necessary by the consideration of unbounded Hamiltonians, is that time-dependent perturbations are assumed to be continuous with respect to the strong operator topology instead of the operator norm topology. In order to handle this more general situation, a number of technical issues need to be addressed related to the continuity properties of operator-valued functions and of the dynamics generated by strongly continuous time-dependent interactions. These technical issues cascade through the better part of this paper. We will understand if the reader is surprised by the length of the paper, since we were taken aback ourselves as we were completing the manuscript. Many proofs can be shortened if one is only interested in particular cases. Indeed, in many cases, results for more restricted cases exist in the literature. There are also places, however, where the published results in the literature provide only weaker estimates or have incomplete proofs.
In Sec. II, we review the construction and basic properties of quantum dynamics in the context of this work. This includes a careful presentation of analysis with operator-valued functions using the strong operator topology. Section III is devoted to Lieb-Robinson bounds and their application to proving the existence of the thermodynamic limit of the dynamics. We also derive an estimate on the dependence of the dynamics on the interactions and introduce a notion of convergence of interactions that implies the convergence of the infinite-volume dynamics. Section IV is devoted to the approximation of quasilocal observables by strictly local ones by means of suitable maps called conditional expectations. Because they are needed for our applications, the continuity properties of a class of such maps are studied in detail. A general notion of quasilocal maps is introduced in Sec. V, and we study the properties of several operations involving such maps that are used extensively in applications. In Sec. VI, we construct an auxiliary dynamics called the spectral flow (also called the quasiadiabatic evolution), which is the main tool in recent proofs of the stability of spectral gaps and gapped ground state phases. A first application of the spectral flow is the notion of automorphic equivalence, discussed in Sec. VII, which allows us to give a precise definition of a gapped ground state phase as an equivalence class for a certain equivalence relation on families of quantum lattice models. In the Appendix, we collect a number of arguments that are used throughout this paper.
Our original motivation for this work was to supply all the tools needed for the results of Paper II.96 However, this work can now be read as a stand-alone review article about quasilocality estimates for quantum lattice systems. Since the sequel of Paper II96 will be devoted to applying the quasilocality bounds and the spectral flow results from this work to prove the stability of gapped ground state phases, the examples and applications here will be chosen in support of the presentation of the general results.
Throughout this paper, we focus on so-called bosonic lattice systems, for which observables with disjoint support commute. Virtually, all results carry over to lattice fermion systems with only minor changes. This is discussed in some detail in Ref. 97. Another extension of quasilocality techniques not covered in this paper is the case of so-called extended operators. An important example is the half-infinite string operators that create the elementary excitations in models with topological order such as the Toric Code model.71 Lieb-Robinson bounds for such non-quasi-local operators are used in Ref. 29.
II. SOME BASIC PROPERTIES OF QUANTUM DYNAMICS
In this paper, the primary object of study is the Heisenberg dynamics acting on a suitable algebra of observables for a finite or infinite lattice system. For finite systems, this dynamics is expressed with a unitary propagator U(t, s), , on a separable Hilbert space . However, in some cases, for example, when one is interested in the excitation spectrum and dynamics of perturbations with respect to a thermal equilibrium state of the system, the generator of the dynamics is not semibounded and the Hilbert space may be nonseparable. Therefore, in general, we will not assume that is separable or that the Hamiltonian is bounded below.
As described in the Introduction, we consider finite and infinite lattice systems with interactions that are sufficiently local. We allow for an infinite-dimensional Hilbert space at each site of the lattice. However, we impose conditions on the interactions that permit us to prove quasilocality bounds of Lieb-Robinson type (in terms of the operator norm) for bounded local observables. This means that we will allow for the possibility of unbounded “spins” and unbounded single-site Hamiltonians, but require that the interaction be given by bounded self-adjoint operators that satisfy a suitable decay condition at large distances (see Sec. III for more details). We do not consider lattice oscillator systems with harmonic interactions in this paper, since one should not expect bounds in terms of the operator norm for this class of systems (see Refs. 4 and 89). An interesting model that does fit in the framework presented here is the so-called quantum rotor model, which has an unbounded Hamiltonian for the quantum rotor at each site, but the interactions between rotors are described by a bounded potential.72,77,115
We will use the so-called interaction picture to describe the dynamics of Hamiltonians with unbounded on-site terms. This requires that we also consider time-dependent interactions. Time-dependent Hamiltonians are, of course, of interest in their own right, for instance, in applications of quantum information theory. Therefore, we begin with a discussion of the Schrödinger equation for the class of time-dependent Hamiltonians considered in this work.
Let be a complex Hilbert space and denote the bounded linear operators on . Let be a finite or infinite interval. In this section, we review some basic properties of the dynamics of a quantum system with a time-dependent Hamiltonian of the form
where H0 is a time-independent self-adjoint operator with dense domain , and for t ∈ I, and t ↦ Φ(t) is continuous in the strong operator topology. This means that for all , the function t ↦ Φ(t)ψ is continuous in the Hilbert space norm. From these assumptions, it follows that for all t ∈ I, H(t) is self-adjoint with time-independent dense domain [see Ref. 126 (Theorem 5.28)].
We will often consider operator-valued and vector-valued functions of one or more real (or complex) variables and impose various continuity assumptions, which we now briefly review. An operator-valued function is said to be norm continuous (norm differentiable) if it is continuous (differentiable) in the operator norm and strongly continuous (strongly differentiable) if it is continuous (differentiable) in the strong operator topology. With a slight abuse of terminology, we will refer to Hilbert space-valued functions as strongly continuous (strongly differentiable) if they are continuous (differentiable) in the Hilbert space norm. For transparency, when we consider maps defined on a linear space of operators, we will indicate the relevant topology and continuity assumptions explicitly.
The dynamics of a system described in (2.1) is determined by the following Schrödinger equation:
For bounded H(t), through a standard construction, we will see that there exists a family of unitaries , s, t ∈ I, that is jointly strongly continuous with ψ(t) = U(t, t0)ψ0 being the unique solution of (2.2) for all . It follows that the family U(t, s) has the cocycle property: for r ≤ s ≤ t ∈ I, U(t, r) = U(t, s)U(s, r) and . In the case that H(t) = H0 + Φ(t), where H0 is an arbitrary unbounded self-adjoint operator and Φ(t) is bounded, we will make use of the well-known interaction picture dynamics to construct an analogous unitary cocycle. This cocycle will, in particular, generate the unique weak solution of the Schrödinger equation. To this end, we first discuss some other aspects of strongly continuous operator-valued functions that we will need.
A. Properties of continuity, measurability, and integration in
In this section, we review some terminology and discuss a number of properties of operator-valued functions that will be used extensively in the rest of this paper.
Let be a finite or infinite interval and be strongly continuous, i.e., for all , t ↦ A(t)ψ is continuous with respect to the Hilbert space norm. By the uniform boundedness principle, if A is strongly continuous, then A is locally bounded, meaning if J ⊂ I is compact, then
The strong continuity of t ↦ A(t) implies that t ↦ ∥A(t)ψ∥ is continuous for all , and from the above, the map t ↦ ∥A(t)∥ is locally bounded. However, strong continuity does not imply that t ↦ ∥A(t)∥ is continuous [see Ref. 95 (Sec. 2) for a counterexample].
We note that in this paper we use the notations ∥A(t)∥ and ∥A∥(t) interchangeably. For ease of later reference, we now state a simple proposition.
Let be a finite or infinite interval and and be Hilbert spaces.
If are strongly continuous (strongly differentiable), then (t, s) ↦ A(t)B(s) is jointly strongly continuous (separately strongly differentiable).
If and are strongly continuous (strongly differentiable), then (t, s) ↦ A(t) ⊗ B(s) is jointly strongly continuous (separately strongly differentiable).
If is strongly continuous, then the function t ↦ ∥A(t)∥ is lower semicontinuous, measurable, locally bounded and, hence, locally integrable.
It is clear that an analog of Proposition 2.1 holds when strongly is replaced with norm in the statements above. Moreover, an argument similar to the one found in Proposition 2.1(i) shows that if and are both strongly continuous (strongly differentiable), then (t, s) ↦ A(t)ψ(s) is jointly strongly continuous (separately strongly differentiable). As will be clear from the proof, we note that the conclusions of part (iii) of this proposition continue to hold even for weakly continuous A(t).
We prove the statements above in the case of strong continuity; the strong differentiability claims follow similarly.
Since we have that f−1((s, ∞)) is open, this set is also Borel measurable, for all . By a standard lemma in measure theory,41 this implies that f is measurable.
We already noted above that ∥A(t)∥ is bounded on compact intervals by the uniform boundedness principle. This concludes the proof.
B. Dynamical equations and the Dyson series
In this section, we review some well-known facts about Dyson series and from them obtain the Schrödinger dynamics generated by a bounded, time-dependent Hamiltonian. A standard result in this direction can be summarized as follows: Let be a Hilbert space, be a finite or infinite interval, and be strongly continuous and pointwise self-adjoint, i.e., H(t)* = H(t), for all t ∈ I. Under these assumptions (see, e.g., Theorem X.69 of Ref. 109) for each t0 ∈ I and every initial condition , the time-dependent Schrödinger equation
has a unique solution in the sense that there is a unique, strongly differentiable function which satisfies (2.10). This solution can be characterized in terms of a two-parameter family of unitaries such that
These unitaries are often referred to as propagators, and an explicit construction of them is given by the Dyson series. Specifically, for any s, t ∈ I and each , the Hilbert space-valued series
is easily seen to be absolutely convergent in norm. One checks that U(t, s), as defined in (2.12), satisfies the differential equation
which is to be understood in the sense of strong derivatives. Of course, under the stronger assumption that is norm continuous, then (2.13) also holds in norm.
The additional observation we want to make here is that U(t, s) is not only the unique strong solution of (2.13); it is also the case that any bounded weak solution of (2.3) necessarily coincides with U(t, s). By weak solution, we mean that for all and any s, t ∈ I, U(t, s) satisfies
A proof of this fact is contained in the following proposition:
There is a unique strong solution of (2.15), and V is norm continuous.
Any locally norm-bounded, weak solution of (2.15) coincides with the strong solution.
Let be dense. Suppose is strongly continuous and satisfies
- (iv)
If V0 is invertible, the strong solution V of (2.15) is invertible for all t ∈ I. Moreover, in this case, the inverse of V is the unique strong solution of
- (ii)
The uniqueness statement in (ii) is proven similarly. In fact, let V1 and V2 be two locally norm-bounded weak solutions of (2.15). In this case, for any and each t ∈ I,
- (iii)
We will show that any V satisfying the assumptions of (iii) is actually the locally bounded weak solution. To this end, note that for any , (2.16) implies
- (iv)
Arguing as in the proof of (i), the function defined by setting
We conclude this section with an estimate on the solution of certain dynamical equations that will be useful in the proof of the Lieb-Robinson bound in Sec. III.
C. The dynamics for a class of unbounded Hamiltonians
1. On the interaction picture dynamics
Proposition 2.4 is an important application of Proposition 2.2. As explained in the remarks of Sec. X.12 of Ref. 109, applying the interaction picture representation to Hamiltonians with the form H = H0 + Φ, even if Φ is time-independent, leads one to study a dynamics with time-dependent Hamiltonians. In this situation, one often produces Hamiltonians that are strongly continuous, but not norm continuous. This leads us to consider Hamiltonians of the form H(t) = H0 + Φ(t), where H0 is a self-adjoint operator with dense domain and Φ(t) is a bounded, pointwise self-adjoint operator that is strongly continuous in t.
Let be a Hilbert space and H0 be a self-adjoint operator with dense domain . Let be a finite or infinite interval and be strongly continuous and pointwise self-adjoint. Then, there is a two parameter family of unitaries {U(t, s)}s,t∈I associated with the self-adjoint operator H(t) = H0 + Φ(t) for which
(t, s) ↦ U(t, s) is jointly strongly continuous,
U(t, s) satisfies the cocycle property (2.19),
- U(t, s) generates the unique, locally bounded weak solutions of the Schrödinger equation associated with H(t), i.e., for any t0 ∈ I and , given by ψ(t) = U(t, t0)ψ0 satisfiesfor all and t ∈ I.(2.38)
We need to justify only the uniqueness of the locally bounded weak solutions. Let t0 ∈ I, let , and suppose ψ1 and ψ2 are two locally bounded solutions of the initial value problem (2.38). Consider the functions and . It is easy to check that these functions are locally bounded weak solutions of the Schrödinger equation associated with the bounded Hamiltonian in (2.39). As such, they are unique, which may be argued as in the proof of Proposition 2.2, and therefore, so too are ψ1 and ψ2.
In this work, we define the Heisenberg dynamics on a suitable algebra of observables in terms of the strongly continuous propagator U(t, s) whose existence is guaranteed by Proposition 2.4. We work under assumptions that guarantee the uniqueness of bounded weak solutions. Strictly speaking, the uniqueness of the weak solution and the possible absence of a strong solution to the Schrödinger equation in the Hilbert space will play no role in our analysis. More information about the solutions and their uniqueness could, however, be important for the unambiguous interpretation of our results. Additional results exist in the literature if one is willing to make additional assumptions on H0 and Φ(t). For example, the following theorem establishes the existence of an invariant domain for the generator and, consequently, the existence of a unique strong solution for the situation where H0 is semibounded and Φ(t) is Lipschitz continuous, which is a common physical situation. As explained in the Introduction, there are important applications of the methods in this paper to situations where these additional assumptions are not satisfied.
In Ref. 119 (Theorem II.21), Simon credits a version of this theorem to Yosida, who proved it in a more general Banach space context (Ref. 129, Sec. XIV.4), but with the Lipschitz condition replaced by a boundedness condition on the derivative of Φ(t). Yosida gives credit to Kato65,66 and Kisyński.70
2. A Duhamel formula for bounded perturbations depending on a parameter
In this section, we consider families of Hamiltonians Hλ(t) which depend on a time-parameter t ∈ I and an auxillary parameter λ ∈ J. For such families, we will prove a version of the well-known Duhamel formula (Proposition 2.6) and use it to derive various continuity properties of the corresponding dynamics (Proposition 2.7).
Let H0 be a densely defined, self-adjoint operator on a Hilbert space and denote by the corresponding dense domain. Let be intervals and consider the family of Hamiltonians Hλ(t), t ∈ I and λ ∈ J, acting on given by
where for each t ∈ I and λ ∈ J, . The self-adjointness of Hλ(t) on the common domain, , is clear. We will assume that (t, λ) ↦ Φλ(t) is jointly strongly continuous. We will also assume that for each fixed t ∈ I, the mapping λ ↦ Φλ(t) is strongly differentiable and that the corresponding derivative, which we denote by Φ′λ(t), satisfies that the map (t, λ) ↦ Φ′λ(t) is jointly strongly continuous.
Under these assumptions, Proposition 2.4 guarantees that for each λ ∈ J, there exists a two parameter family of unitaries which generates the weak solutions of the Schrödinger equation associated with Hλ(t) [see (2.38)]. Our goal here is to show that for fixed s, t ∈ I, the map λ ↦ Uλ(t, s) is strongly differentiable, and moreover,
We will obtain this bound as a corollary of Proposition 2.6, which gives a Duhamel formula for the derivative in this setting. Although the Duhamel formula is well-known, we give an explicit proof here that allows us to clarify the continuity properties implied by our assumptions. In the proof we avoid taking derivatives with respect to t or s which, in general, are unbounded operators.
With stronger assumptions, one can prove (2.48) holds in norm. In fact, arguing as below, if
the map (t, λ) ↦ Φλ(t) is jointly norm continuous,
for each t ∈ I, the map λ ↦ Φλ(t) is norm differentiable, with the derivative denoted by Φ′λ(t), and
the map is jointly norm continuous,
The proof of (2.51) is now completed by demonstrating that upon inserting the Dyson series for Ũλ(t, r) and Ũλ(r, s) into the integral on the right-hand-side of (2.51), the result simplifies to the expression on the right-hand-side of (2.55).
Let Hλ(t) be a family of Hamiltonians as described in (2.46). The corresponding dynamics, as in (2.58), has the following properties:
For each λ ∈ J and , the map is jointly strongly continuous.
For each s, t ∈ I and , the map is strongly differentiable (and hence strongly continuous). Moreover, one has the estimate
- (iii)
For fixed s, t ∈ I and λ ∈ J, the map is continuous on bounded sets when both its domain and codomain are equipped with the strong operator topology. This continuity is uniform for λ in compact subsets of J.
The statement in (i) follows from Proposition 2.4 as [see (2.58)] is the product of jointly strongly continuous mappings.
III. LIEB-ROBINSON BOUNDS AND INFINITE VOLUME DYNAMICS OF LATTICE SYSTEMS
The scope of this paper is lattice models with possibly unbounded single-site Hamiltonians and bounded interactions that, in general, may be time-dependent. This is the setting in which one expects to obtain Lieb-Robinson bounds with estimates in terms of the operator norm of the observables. A well-known example of this situation is the quantum rotor model. We will not consider lattice models with unbounded interactions in this work. The only systems with unbounded interactions that have been studied so far are oscillator lattice systems for which the interactions are quadratic34 or bounded perturbations of quadratic interactions.4,89
In this paper, the “lattice” in lattice systems is understood to be a countable metric space (Γ, d) (not necessarily a lattice in the sense of the linear combinations with integer coefficients of a set of basis vectors in Euclidean space). Typically, Γ is infinite (or more specifically, has infinite diameter), and models are given in terms of Hamiltonians for a family of finite subsets of Γ. After an initial analysis of the finite systems, we study the thermodynamic limit through sequences of increasing and absorbing finite volumes {Λn}, i.e., Λn↑ Γ. Often, the goal is to obtain estimates for the finite systems defined on Λn that are uniform in n. The definitions below prepare for this goal. It is possible to consider a finite set and apply the results derived in this paper to finite systems. We note that some of the conditions imposed are trivially satisfied for finite systems.
The points of Γ, also called sites of the lattice, label a family of “small” systems, which are often, but not necessarily, identical copies of a given system such as a spin, a particle in a confining potential such as a harmonic oscillator, or a quantum rotor. The quantum many-body lattice systems of condensed matter physics are of this type. A wide range of interesting behaviors arises due to interactions between the component systems. It is a central feature of extended physical systems that interactions have a local structure, meaning that the strength of the interactions decreases with the distance between the systems. Often, each system only interacts directly with its nearest neighbors in the lattice. The mean-field approximation ignores the geometry of the ambient space and it is often a good first approximation. In more realistic models, however, the interactions between different components depend on the distance between them. In this section, we derive a fundamental property of the dynamics of quantum lattice systems that is intimately related to the local structure of the interactions. This property is referred to as quasilocality and its basic feature is a bound on the speed of propagation of disturbances in the system, which is known as a Lieb-Robinson bound.
Lieb and Robinson were the first to derive bounds of this type.75 In the years following the original article, a number of further important results appeared, e.g., by Radin107 and, in particular, by Robinson112 who gave a new proof of the theorem of Lieb and Robinson (which is included in Ref. 20). Robinson also showed that Lieb-Robinson bounds can be used to prove the existence of the thermodynamic limit of the dynamics and used the bounds to derive fundamental locality properties of quantum lattice systems. It was only much later, however, that Hastings who pointed out how the Lieb-Robinson bounds could be used to prove exponential clustering in gapped ground states in a paper where he provided the first generalization of the Lieb-Schultz-Mattis theorem to higher dimensions.54 Mathematical proofs then followed by Nachtergaele and Sims,92 Hastings and Koma,57 and Nachtergaele, Ogata, and Sims.88 The new approach to proving Lieb-Robinson bounds developed in these works leading to Ref. 94 yields a better prefactor with a more accurate dependence on the support of the observables. This was important for certain applications such as the proof of the split property for gapped ground states in one dimension by Matsui.81,82
Further extensions of the Lieb-Robinson bounds in several directions quickly followed: Lieb-Robinson bounds for lattice fermions,24,57,97 Lieb-Robinson bounds for irreversible quantum dynamics,53,98,105 a bound for certain long-range interactions,45,111,124 anomalous or zero-velocity bounds for disordered and quasiperiodic systems,27,35,36,52 propagation estimate for lattice oscillator systems,4,26,34,89 and other systems with unbounded interactions,106 including classical lattice systems.28,61,108
The list of applications of Lieb-Robinson bounds includes a broad range of topics: Lieb-Schultz-Mattis theorems,54,93 the entanglement area law in one dimension,55 the quantum Hall effect,6,44,58 quasiadiabatic evolution (spectral flow and automorphic equivalence) including stability and classification of gapped ground state phases,12,21,22,59,83,96 the stability of dissipative systems,19,76 the quasiparticle structure of the excitation spectrum of gapped systems,10,50 a stability property of the area law of entanglement,78 the efficiency of quantum thermodynamic engines,118 the adiabatic theorem and linear response theory for extended systems,7,8 the design and analysis of quantum algorithms,49 and the list continues to grow.5–7,25,30,38,47,64,84,123
In order to express the locality properties of the interactions and the resulting dynamics, we introduce some additional structure on the discrete metric space (Γ, d) in Sec. III A.
A. Lieb-Robinson estimates for bounded time-dependent interactions
1. General setup
As described above, we will study quantum lattice models with possibly unbounded single-site Hamiltonians but bounded, in general, time-dependent interactions. In this section, we give the framework for quantum lattice systems and describe the bounded interactions of interest. We will consider the addition of unbounded on-site Hamiltonians in Sec. III B.
The lattice models we consider are defined over a countable metric space (Γ, d). To each site x ∈ Γ, we associate a complex Hilbert space and denote the algebra of all bounded linear operators on by . Let be the collection of all finite subsets of Γ. For any , the Hilbert space of states and algebra of local observables over Λ are denoted by
where we have chosen to define the tensor product of the algebras so that the last equality holds (i.e., the spatial tensor product, corresponding to the minimal C*-norm116). For any two finite sets Λ0 ⊂ Λ ⊂ Γ, each can be naturally identified with . With respect to this identification, the algebra of local observables is then defined as the inductive limit
and the C*-algebra of quasilocal observables, which we denote by , is the completion of with respect to the operator norm. We will use the phrase quantum lattice system to mean the countable metric space (Γ, d) and quasilocal algebra .
A model on a quantum lattice system is given in terms of an interaction Φ. In the time-independent case, this is a map such that for all The quantum lattice model associated with Φ is the collection of all local Hamiltonians of the form
We will also consider time-dependent interactions. Let be an interval. A map is said to be a strongly continuous interaction if
To each t ∈ I, the map is an interaction.
For each , is strongly continuous.
Given such a strongly continuous interaction Φ, we will often denote by Φ(t) the interaction Φ(·, t) as in (i) above and define the corresponding local Hamiltonians
Similarly, a corresponding time-dependent quantum lattice model may be defined. By our assumptions on the interaction, it is clear that for each t ∈ I, HΛ(t) is a bounded, self-adjoint operator on . Moreover, by Proposition 2.1, is strongly continuous. In this case, Proposition 2.2 demonstrates that there exists a two-parameter family of unitaries , defined as the unique strong solution of the initial value problem
In terms of these unitary propagators, we define a Heisenberg dynamics by setting
In some applications, including Theorem 3.1, we will also consider the inverse dynamics,
where the final equality follows from Proposition 2.2(iv).
As discussed above, Lieb-Robinson bounds approximate the speed of propagation of dynamically evolved observables through a quantum lattice system, and this estimate is closely tied to the locality of the interaction in question. To quantify the locality of an interaction, we introduce the notion of an F-function. An F-function on (Γ, d) is a nonincreasing function F:[0, ∞) → (0, ∞), satisfying the following two properties:
F is uniformly integrable over Γ, i.e.,
- (ii)
F satisfies the convolution condition
An equivalent formulation of (ii) is that there exists a constant C < ∞ such that
Let F be an F-function on (Γ, d) and g : [0, ∞) → [0, ∞) be any nondecreasing, subadditive function, i.e., g(r + s) ≤ g(r) + g(s) for all r, s ∈ [0, ∞). Then, the function
also satisfies (i) and (ii) with ∥Fg∥ ≤ ∥F∥ and . We call any F-function of this form a weighted F-function.
It is easy to produce examples of these F-functions when for some ν ≥ 1 and d(x, y) = |x − y| is the ℓ1-distance. In fact, for any ϵ > 0, the function
is an F-function on . It is clear that this function is uniformly integrable, i.e., (3.8) holds. Moreover, one may verify that
In the special case of g(r) = ar, for some a ≥ 0, we obtain a very useful family of weighted F-functions, which we denote by Fa, given by Fa(r) = e−ar/(1 + r)ν+ϵ. See Subsections 1–3 of the Appendix for other examples and properties of F-functions.
We use these F-functions to describe the decay of a given interaction. Let F be an F-function on (Γ, d) and be an interaction. The F-norm of Φ is defined by
It is clear from the above equation that for all x, y ∈ Γ,
Note that for any , there exist x, y ∈ Z for which d(x, y) = diam(Z), the latter being the diameter of Z. In this case, a simple consequence of (3.15) is
We will be mainly interested in situations where the quantity in (3.14) is finite. In this case, the bound (3.16) demonstrates that the F-function governs the decay of an individual interaction term, and moreover, the estimate (3.15) generalizes this notion of decay by including all interaction terms containing a fixed pair of points x and y.
When Γ is finite, then ∥Φ∥F is finite for any interaction Φ and any function F. For infinite Γ, the set of interactions Φ for which ∥Φ∥F < ∞ depends on F. It is easy to check that ∥·∥F is a norm on the set of interactions for which it is finite. In terms of this norm, we define the Banach space
Of course, depends on Γ and on the single-site Hilbert spaces , but that information will always be clear from the context.
We introduce an analog of (3.14) for time-dependent interactions as follows: Consider a quantum lattice system composed of (Γ, d) and . Let be an interval and be a strongly continuous interaction. Given an F-function on (Γ, d), we will denote by the collection of all strongly continuous interactions Φ for which the mapping
is locally bounded. As with the operator norm, we will sometimes use the alternate notation ∥Φ∥F(t) for the quantity defined in (3.18). The function t ↦ ∥Φ∥F(t) is measurable since it is the supremum of a countable family of measurable functions. As such, ∥Φ∥F is locally integrable. As in the time-independent case, (3.18) implies that for all t ∈ I and x, y ∈ Γ,
2. Lieb-Robinson estimates for bounded interactions
In Theorem 3.1, we demonstrate that the finite volume Heisenberg dynamics , as defined in (3.6), associated with any satisfies a Lieb-Robinson bound. Such bounds provide an estimate for the speed of propagation of dynamically evolved observables in a quantum lattice system. One can use these bounds to show that for small times the dynamically evolved observable is well approximated by a local operator. For this reason, Lieb-Robinson bounds and other similar results are often referred to as quasilocality estimates.
Before we state the result, two more pieces of notation will be useful. First, to each , we denote by the Φ-boundary of X,
In some estimates, it may be useful to restrict the time interval used to define the Φ-boundary. For instance, given one could find that for some X, while is strictly smaller for a subinterval . From now on, we will drop the time-interval I from the notation and simply write ∂ΦX. We note also that in many situations, not much is lost by using X instead of ∂ΦX in the following estimates.
Second, for , and s, t ∈ I, the quantity It,s(Φ) defined by
will appear in many results we provide, including Theorem 3.1. Clearly, if CF∥Φ(r)∥F ≤ M, for all r ∈ [min(t, s), max(t, s)], we have It,s(Φ) ≤ |t − s|M. For example, we see that
with
It is easy to see that with the definition , F1 is a new F-function in terms of which the bound (3.23) slightly simplifies in the sense that . This is a general feature of our estimates involving F-functions and the associated norms on the interactions. In what follows, a variety of different F-functions will be used. Often, new F-functions are obtained by elementary transformations of old ones, see, e.g., Subsection 3 of the Appendix. Instead of figuring out the normalization constants that make CF = 1 for each of the F-functions, we note that the final result can be expressed with a renormalized F-function such that CF = 1.
We also note that the map equals the restriction of to . It is useful, however, to consider them as separate maps for each X ⊂ Λ because the estimates for their norms depend crucially on X through SΛ(X). Also note that each only depends on interaction terms Φ(Z, r) such that Z ⊂ Λ and r ∈ [min(t, s), max(t, s)].
B. A class of unbounded Hamiltonians
As we now discuss, the methods in Subsection III A extend to models with unbounded on-site terms. Consider a quantum lattice system composed of (Γ, d) and . Let be an interval, F be an F-function on (Γ, d), and be a time-dependent interaction. To each z ∈ Γ, fix a self-adjoint operator Hz with dense domain . For any and t ∈ I, consider the finite-volume Hamiltonian
The noninteracting Hamiltonian
is essentially self-adjoint with domain
[see Ref. 110 (Theorem VIII.33 and Corollary)]. Since the time-dependent terms are bounded, it follows from Ref. 126 (Theorem 5.28) that for each t ∈ I, HΛ(t) is essentially self-adjoint on with domain . We proceed by using the notations and HΛ(t) for the corresponding self-adjoint closures.
As , it is a strongly continuous interaction, and so for any , Proposition 2.4 guarantees the existence of a finite volume unitary propagator corresponding to HΛ(t). Let us briefly review this in order to motivate our definition of the finite volume dynamics. By Stone’s theorem, the noninteracting self-adjoint Hamiltonian generates a free-dynamics
in terms of a group of strongly continuous unitaries . In this case,
is pointwise self-adjoint with strongly continuous. By Proposition 2.2(v), there is a unique strong solution of the initial value problem
for each s ∈ I. In terms of these solutions, we introduce
for any s, t ∈ I. As is demonstrated in the proof of Proposition 2.4, the operators form a two-parameter family of unitaries. They satisfy the cocycle property (2.19), and generate the unique locally norm bounded weak solutions of the time-dependent Schrödinger equation corresponding to HΛ(t). We use these unitaries to define a dynamics associated with HΛ(t), namely, for any s, t ∈ I, we take as
One readily checks that the family of automorphisms on satisfies the cocycle property and that the following analog of Theorem 3.1 holds for this dynamics.
C. The infinite-volume dynamics
In this section, we will prove several convergence and continuity results for the Heisenberg dynamics associated with interactions that make use of Lieb-Robinson bounds. As is well-known (see, e.g., Ref. 20), Lieb-Robinson bounds can be used to prove the existence of a dynamics in the thermodynamic limit for sufficiently short-range interactions. In Theorem 3.5, we show that given an interaction the dynamics corresponding to finite-volume restrictions of Φ converge in the thermodynamic limit. To prove the existence of the thermodynamic limit, we will apply Theorem 3.4, which establishes that the Heisenberg dynamics is continuous in the interaction space. For example, in the case of time-independent interactions, Theorem 3.4 implies that the difference between the dynamical evolution of a local observable A with respect to two different interactions is small if ∥Φ − Ψ∥F is small. The statement of this result for finite-volume Heisenberg dynamics is the content of Theorem 3.4, with the analogous thermodynamic limit statement given in Corollary 3.6. Finally, given a sequence of interactions which converge locally in F-norm [see Definition 3.7], we show that the corresponding dynamics (which necessarily exist by Theorem 3.5) converge as well; this is the content of Theorem 3.8. In particular, this can be used to prove that the thermodynamic limit of the Heisenberg dynamics is unchanged by the addition of (sufficiently local) boundary conditions. If the interactions are norm continuous, the cocycle of automorphisms describing the infinite-volume dynamics is differentiable with a strongly continuous generator. This is shown in Theorem 3.9.
We now begin with the continuity statement. For this result, we will once again make use of the quantity It,s(Φ), which is defined in (3.21).
- For any with X ⊂ Λ, the boundholds for all and s, t ∈ I.(3.68)
- For any with X ⊂ Λ0 ⊂ Λ, the boundholds for all and s, t ∈ I.(3.69)
Using (3.8), the estimates in (3.68) and (3.69) can be interpreted as bounds on the norm of the difference of two dynamics, thought of as maps from to , that are uniform in Λ but grow linearly in |X|. Since the dynamics are, of course, automorphisms, the bounds of Theorem 3.4 are only nontrivial if the RHS of (3.68) and (3.69) are smaller than 2∥A∥. As is well known, this will be true for both cases if |t − s| is sufficiently small. Additionally, the bound in (3.68) will be nontrivial if ∥Φ − Ψ∥F is small, and the bound in (3.69) will be nontrivial if d(X, ΛΛ0) is small. Note that even if a map is bounded on all of , as is the case in Theorem 3.4, norm bounds for their local restriction can be very useful.
A first application of Theorem 3.4 is a proof that given any collection of self-adjoint on-sites and an interaction , there is a corresponding infinite-volume dynamics on . We obtain this infinite-volume dynamics as a limit of finite-volume dynamics. With this in mind, we will say that a sequence is increasing and exhaustive if Λn ⊂ Λn+1 for all n ≥ 1 and given any , there is an N ≥ 1 for which X ⊂ ΛN.
For unbounded Hz, the continuity of τt,s is limited by the continuity of the on-site dynamics . In a suitable representation of on a Hilbert space, one can retrieve weak continuity of the dynamics. See Ref. 90 for an example. Continuity properties of τt,s and other families of maps will be further discussed in Sec. IV B 1.
The proof of the remaining facts in the statement of this theorem is standard and proceeds in the same way as is done, e.g., in Ref. 120 for quantum spin models with time-independent interactions.
Combining Theorems 3.4 and 3.5, we obtain the following useful estimates for the infinite volume dynamics:
Under the assumptions of Theorem 3.4,
- For any such that X ∩ Y = ∅, the boundholds for all , , and t, s ∈ I.
- For any , the boundholds for all and s, t ∈ I.(3.79)
- For any with X ⊂ Λ, the boundholds for all and s, t ∈ I.(3.80)
We now prove a convergence result for the dynamics associated with interactions in . First, we introduce some notation and terminology associated with extensions and restrictions of interactions, and then state the result.
We now introduce the notion of local convergence in F-norm.
Let (Γ, d) and be a quantum lattice system, and let be an interval. We say that a sequence of interactions converges locally in F-norm to Φ if there exists an F-function, F, such that
for all n ≥ 1,
,
for any and each [a, b] ⊂ I,
Let be a sequence of time-dependent interactions on Γ with Φn converging locally in the F-norm to Φ with respect to F.
- If for every [a, b] ⊂ I,then, for any , , and s, t ∈ I, s ≤ t, we have convergence of the dynamics(3.85)Moreover, the convergence is uniform for s, t in compact intervals, and the dynamics is continuous,(3.86)(3.87)
- If, in addition, for all , and r ∈ I, we also have pointwise local convergence,and uniform boundedness of the interactions on compact intervals I0 ⊂ I,(3.88)then the generators converge uniformly for t in compacts: for all compact I0 ⊂ I and , we have(3.89)where(3.90)
By assumption (3.85), it is clear that supnIt,s(Φn) is finite for all s, t ∈ [a, b]. In this case, for any ϵ > 0, the estimates in (3.92) and (3.93) can be made arbitrarily small for all n sufficiently large (for example, less than ϵ/3) with a sufficiently large, but finite choice of Λ ⊂ Γ. For any such choice of Λ, the bound (3.94) can be made equally small with large n by using local convergence in F-norm.
- (ii)
The interactions Φn(t) and Φ(t) have finite F-norm. Hence, the corresponding derivations, and δt, are well-defined on . For any , we then have
The dynamics considered in the proof of Theorem 3.8 corresponds to the one whose existence is established in the proof of Theorem 3.5 in the special case that the on-sites Hz = 0 for all z ∈ Γ. By going to the interaction picture, as is done in the proof of Theorem 3.4, it is clear that the convergence results (3.87) and (3.90) hold in the case of arbitrary self-adjoint on-site terms.
Theorem 3.8 establishes sufficient conditions for the convergence of the sequence of cocycles to the cocycle , as well as the convergence of the generators to densely defined derivations δt. These conditions are by no means necessary, but will serve our purposes well.
We may now ask whether the dynamics satisfies additional properties and, in particular, whether it is differentiable with the derivative given by the derivation δt. Theorem 5.9 addresses this question.
We conclude this section with a few comments.
It is clear how to modify Definition 3.7 in such a way to describe sequences that are locally Cauchy in F-norm. Given any such Cauchy sequence which also satisfies (3.85), an ϵ/3-argument almost identical to the one in the proof of Theorem 3.8 shows that the corresponding dynamics converge to a cocycle of automorphisms of . With this understanding, one sees that Theorem 3.8 implies Theorem 3.5. In fact, let and take to be an increasing, exhaustive sequence of finite subsets of Γ. Define the sequence of interactions and extend Φn to all of Γ as indicated above. In this case, it is clear that is locally Cauchy in the F-norm defined by F, and moreover, (3.85) holds. Thus, the corresponding dynamics converge. Since the sequence also converges locally to Φ in the F-norm defined by F, we know what the generator of the limiting dynamics is by Theorem 3.8. In this manner, we recover the fact that the limiting dynamics is independent of the increasing, exhaustive sequence of finite-volumes. Moreover, we also see that this limiting dynamics is invariant under a class of finite-volume boundary conditions.
As a final comment we note that one can easily find conditions under which the Duhamel formula (3.66) is also inherited by the infinite-volume dynamics. As this will not be needed in this work, we do not discuss this further.
IV. LOCAL APPROXIMATIONS
In Sec. III A 2, we proved a Lieb-Robinson bound for the finite volume dynamics generated by an interaction . Such bounds provide an estimate for the speed of propagation in a quantum lattice system. More specifically, such bounds can be used to show that while the support of a local observable evolved under the Heisenberg dynamics is nonlocal, at any fixed time t, the observable essentially acts as the identity outside of a finite region of space. It is often useful to approximate these dynamically evolved observables by strictly local observables. It is further desirable that the operation of taking these local approximations has good continuity properties. This is the topic of this section.
A. Local approximations of observables
We first review how the support of a local observable can be identified using commutators. For any Hilbert space , the algebra has a trivial center; in the case of a finite-dimensional Hilbert space, this is known as Schur’s Lemma. A first generalization of this fact is that for any two Hilbert spaces and , the commutant of in is given by (see, e.g., Ref. 63, Chap. 11). Given this, and the structure of the quantum models introduced in Sec. III A 1, one concludes the following: given and X ⊂ Λ, if satisfies [A, B] = 0 for all , then . In other words, vanishing commutators can be used to identify the support of local observables. If the commutator [A, B] is small but not necessarily vanishing (in norm), Lemma 4.1, which is proved in Ref. 91, shows that A can be well-approximated (up to error ϵ) by an observable .
Let and be complex Hilbert spaces. There is a completely positive linear map with the following properties:
For all , .
- Whenever satisfies the commutator boundsatisfies the estimate
- (iii)
For all and , we have
A completely positive linear map with the properties (i) and (iii) is called a conditional expectation [see, e.g., Ref. 103 (Sec. 9.2)]. If is finite-dimensional, one can take , where tr denotes the normalized trace over . In this case, it is straightforward to verify the properties listed in the lemma (see, e.g., Refs. 23 and 88). For general , a normalized trace does not exist, but using Lemma 4.1 it is easy to show that, at the cost of a factor 2 in the RHS of (4.1), we can replace with id ⊗ ρ for an arbitrary state ρ on . This is the content of Lemma 4.2 from Ref. 91.
A number of further comments are in order. First, although the mapping Πρ depends on the state ρ, the “error” estimate in (4.3) is independent of ρ. Next, if is finite dimensional, then ρ can be taken to be the normalized trace and we already know that the factor of two in (4.3) is not needed. The bound for Πρ therefore appears to be nonoptimal. The map from Lemma 4.1 is only known to be bounded (specifically, ) and hence continuous with respect to the operator norm topology. As such, it is not guaranteed that is continuous when both the domain and codomain are endowed with the strong operator topology. However, by choosing a normal state ρ, we get a map Πρ that is continuous on bounded subsets of the domain when both and are endowed with their strong (or weak) operator topologies. The case of the strong operator topology is the content of Proposition 4.3. The case of the weak topology follows by a similar argument.
Recall that is continuous on bounded subsets with both its domain and codomain considered with the strong operator topology if given any bounded net that converges strongly to , the net converges strongly to .
Let and be two complex Hilbert spaces, and ρ be any normal state on . The following maps, when restricted to arbitrary bounded subsets of their domain, are continuous when both the domain and codomain are equipped with the strong operator topology:
,
.
- (ii)
Let {Aα |α ∈ I} be a net in the unit ball of that converges to . By (i), we know that Bα = Πρ(Aα) converges to B = Πρ(A) in the strong operator topology on . By Proposition 2.1(ii) [see also (2.4) and the preceding discussion], it follows that {Bα ⊗ 1l∣α ∈ I} strongly converges to B ⊗ 1l in .
B. Application to quantum lattice models
We now extend the results of Subsection IV A to infinite quantum lattice systems on Γ. In this setting, states cannot be defined in terms of a single density matrix. Moreover, as explained below, we will want to define a consistent family of conditional expectations with values in , for all . To this end, we consider a locally normal product state ρ, i.e., for each site x ∈ Γ, we fix a normal state ρx on and take the unique state ρ on such that for all finite Λ ⊂ Γ. Then, given , we define conditional expectations similar to those in Lemma 4.2 by
Here, as before, we have taken idX as the identity map on . In our applications, the dependence of these maps on ρ is of minor consequence. Moreover, it will be convenient to view these maps as elements of . For these reasons, we suppress the dependence on ρ and define by
For fixed X, these projections are compatible in the sense that if X ∪ Λ′ ⊂ Λ and , then
We summarize this relation and other consistency relations in Proposition 4.5. First, however, we describe how, given a fixed finite volume X, one can extend the maps , , to an operator ΠX on (and consequently, ).
For any Λ′ ⊆ Λ, recall that we can identify as a subalgebra of , and so we can write (4.10) as . In particular, if X ⊆ Λ′ ⊆ Λ, one has that , from which we see that the following map is well-defined:
Since this map is bounded, in fact of norm one, ΠX has a unique extension to which we also denote by ΠX. We note that ΠX(A) = A if . We refer to the family of conditional expectations , respectively, ΠX, , as a localizing family. By construction, the finite volume local approximations all satisfy the conditions of Lemma 4.2. Corollary 4.4 shows that the results of Lemma 4.2 also extend to .
We now state several consistency properties of the finite and infinite volume conditional expectations. To facilitate the statement of the properties, we use to denote .
Fix a locally normal product state ρ on , and let and be such that X, Y, and Λ′ are all subsets of Λ. The following properties hold for the localizing maps defined with respect to ρ:
If , then .
.
.
If X ⊆ Λ′, then .
for all .
The proofs of these properties for are all elementary and follow from the definition of and the fact that ρ ↾Λ is a product state. The statements for Λ = Γ follow from taking finite volume limits Λ′ ↑Γ of for and using the norm bound ∥ΠX∥ ≤ 1 to extend to in the usual manner.
1. Continuity of local approximations
Given finite sets X ⊂ Λ ⊂ Γ, Proposition 4.3(ii) implies that the projection map preserves continuity in the strong operator topology. In particular, if is strongly continuous for all t in an interval , then is also strongly continuous. In applications, we will be interested in a sequence of strongly continuous functions , with Λn↑ Γ, that converges to a bounded map . It will then be desirable that the localizing projections ΠY(A(t)), , also satisfy certain continuity properties.
While we do not have the standard von Neumann algebra setting where the notion of locally normal is more natural, it is convenient to define a similar notion in our setting with C*-algebras without reference to a representation.
A linear map is called locally normal if there exists an increasing, exhaustive sequence of finite subsets of Γ and corresponding bounded linear transformations with the following properties:
For all n, is continuous on bounded subsets when both its domain and codomain are considered with the strong operator topology;
Local uniform convergence of to : For all X ⊂ Γ finite and any ϵ > 0, there exists N such that for all n ≥ N we have
Note that local uniform convergence implies that converges strongly to . However, since N is allowed to depend on X, this convergence is in general not uniform in . If is finite-dimensional, property (i) is automatically satisfied. Let us now consider an example satisfying Definition 4.6.
A simple consequence of Definition 4.6(i) is the following: for each n ≥ 0, is strongly continuous if t ↦ A(t) is strongly continuous. Lemma 4.8 establishes that the same property holds for the composition for any .
Let , be the extension of the map defined in (4.11), and let be a locally normal map. Then, for every strongly continuous map defined on an interval , the function is also strongly continuous.
Two comments are in order. First, by Proposition 4.3(ii), for any such that Y ⊂ Z, considered as a map into is also strongly continuous. Second, if t ↦ A(t) is, in fact, continuous in the norm topology on (in particular, if ), and is bounded, the result of Lemma 4.8 is trivial since the bounded linear map preserves the norm-continuity.
We will also encounter one-parameter families of locally normal maps, , that are strongly continuous in s and uniformly locally normal in the sense of the following definition:
Let be an interval. A family of linear maps , s ∈ I, is called a strongly continuous family of uniformly locally normal maps if there exists an increasing, exhaustive sequence of finite subsets of Γ and families of bounded linear maps strongly continuous in s, with the following properties:
For all n and s, is continuous on bounded subsets when both its domain and codomain are considered with the strong operator topology, and this continuity is uniform for s ∈ I.
Uniform local convergence of to : For all X ⊂ Γ finite and any ϵ > 0, there exists N such that for all n ≥ N, we have
For families , s ∈ I with I being an infinite interval, the uniformity asked for in part (ii) of this definition will typically not hold and one is led to consider subfamilies parametrized by s ∈ I0 ⊂ I for compact intervals I0. Also note that the properties of a strongly continuous family of uniformly locally normal maps imply that is strongly continuous by the usual ϵ/3 argument. We have not assumed, however, that the maps are bounded. In general, is only locally bounded and cannot be extended to all of .
We now discuss two examples. The first is for a model with uniformly bounded on-sites, while the second does not require this assumption.
For each n ≥ 0, one can show that property (i) of Definition 4.9 holds by arguing as in the proof of Proposition 2.7(iii) and using that I0 is compact and ∥Φ∥F is locally bounded.
Finally, we observe that (ii) is a simple consequence of Corollary 3.6(iii).
For each , (s, t) ↦ Φs(Z, t) is jointly strongly continuous on .
For each and , s ↦ Φs(Z, t) is strongly differentiable, and its derivative is jointly strongly continuous on .
For each s ∈ I0, and moreover, for each T > 0,
We first show that for each n ≥ 0, the map is strongly continuous. In fact, the argument below demonstrates that is uniformly continuous in the operator norm on .
Our next result shows that given a strongly continuous family of uniformly locally normal maps and any , the map is jointly strongly continuous whenever t ↦ A(t) is strongly continuous. In particular, we have continuity on the diagonal t = s. The result, of course, also applies to the finite-volume setting where we can take Y = Λ = Γ.
Let I and J be intervals, , be the extension of the map defined in (4.11), and , s ∈ I be a strongly continuous family of uniformly locally normal maps. Then, for every strongly continuous , t ∈ J, the function is jointly strongly continuous in t and s.
By the assumptions, there exists an increasing, exhaustive sequence of finite subsets of Γ and bounded linear transformations , strongly continuous in s, that approximate as in Definition 4.9.
f(s, t0) is continuous in s.
f(s, t) is an equicontinuous family of functions of t parameterized by s ∈ I.
In all the proofs above, we have used results on the finite volume local approximates to obtain results for the infinite volume local approximates. However, there may be instances where one wants to work in a suitable representation of the infinite-volume algebra. We conclude this section with a result regarding the GNS representation of a locally normal state.
Let , be the family Hilbert spaces defined in (3.1), ω be a locally normal state on , and its corresponding GNS representation. The map is continuous on arbitrary bounded subsets of its domain with respect to the strong operator topology on both its domain and codomain.
V. QUASILOCAL MAPS
In Sec. III, we proved a Lieb-Robinson bound for the dynamics associated with a sufficiently local interaction. In addition to estimating the speed of propagation of a dynamically evolved observable, these bounds imply that the dynamics for a fixed time t is quasilocal. As a result, they can be well approximated by local observables as shown in Corollary 4.4. In recent years, other quasilocal maps have played a key role in proving both locality estimates of the spectral flow12,54,59,93 and spectral gap stability of frustrationfree quantum lattice models.21,22,56,83,85,96,97,130 While we will consider both of these topics, the former in Sec. VI and the latter in Paper II,96 the focus of this section is the general study of quasilocal maps [see (5.1)], and the investigation of the key properties that will be useful in the above-mentioned applications. There exists a broad range of other applications that we will not discuss here.5–9,46
We begin by showing how to apply the techniques from Sec. IV to obtain estimable local approximations of quasilocal maps. In Sec. V B, we provide a number of examples that will arise in our applications, including the difference of two dynamics. We discuss compositions of two quasilocal maps in Sec. V C and prove sufficient conditions for which the composition is again quasilocal. In Sec. V D, consider the composition of a quasilocal map with an interaction. We show that under suitable conditions such a composition can be rewritten as a local interaction. Moreover, we quantify the decay of the resulting interaction in terms of the decays of the original interaction and the quasilocal map. If the transformed interaction has sufficient decay, then the theory developed in Sec. III applies and an infinite volume dynamics exists. We conclude in Sec. V E by returning to the example of the difference of two dynamics and proving a continuity result.
A. General quasilocal maps
Let (Γ, d) and be a quantum lattice system defined as in Sec. III A 1. A linear map is said to satisfy a quasilocality bound of order q ≥ 0 if there is C < ∞ and a nonincreasing function G : [0, ∞) → [0, ∞) with limr→∞G(r) = 0 such that for all , and , ,
Any linear mapping satisfying (5.1) will be referred to as quasilocal. When relevant, we will denote by the smallest constant for which (5.1) holds. Since the function G in (5.1) governs the decay of the quasilocal map, we may refer to G as a decay function associated with . In this work, we will always assume that quasilocal maps are linear. However, there may be other contexts in which it might be appropriate to generalize this definition.
The dependence of the bound in (5.1) on the support of the observable A through the factor |X|q is a choice we made based on the applications we have in mind. However, under appropriate assumptions, most of the estimates proved in this section also hold for quasilocal maps with more general functions of |X|.
In most of our applications, the metric space (Γ, d) is equipped with an F-function F. In this case, one can often estimate a quasilocal map as follows: there is C < ∞ such that for all , any , and , we have
As we will see below, in certain estimates, the bound (5.2) has advantages over (5.1). A simple over-counting argument shows that
It follows from the uniform integrability of F [see (3.8)] that G(r) → 0 as r → ∞. When the corresponding F-function is weighted, i.e., F = Fg as defined (3.11), one has that
and so, in this case, an estimate of the form (5.2) reduces to that of (5.1). For more detailed information on F-functions, including weighted F-functions, see Subsections 1–3 of the Appendix.
We now demonstrate an important estimate concerning quasilocal maps. For this result, we use the concepts introduced in Sec. IV B and, in particular, the localizing maps ΠX and ΔX(n), defined with respect to a locally normal product state ρ, as in (4.11) and (4.16), respectively.
Note that the result above, of course, also applies to finite systems. In particular, for any finite Λ, the result holds for any quasilocal map .
B. Examples of quasilocal maps
In this section, we discuss a few of the most common examples of quasilocal maps [defined as in (5.1)]. In applications, quasilocal maps are often constructed as the thermodynamic limit of appropriate finite-volume maps. We first describe this class of quasilocal maps as a general example. Each of the more concrete examples we present later in this section will be of this general form.
(A general example). Let q ≥ 0, C < ∞, and G be a nonincreasing function G : [0, ∞) → [0, ∞) with limr→∞G(r) = 0. Let be a sequence of increasing and exhaustive finite subsets of Γ. In particular, this means that Λn ⊂ Λn+1 for all n ≥ 1, and given any x ∈ Γ, there exists N ≥ 1 for which x ∈ ΛN. Suppose that for each n ≥ 1, there is a linear map for which
- Given any sets X, Y ⊂ Λn, the boundholds for all observables and .(5.13)
For each finite X ⊂ Γ and any ϵ > 0, there is an N ≥ 1 for which
For later applications, we now describe a variant of the estimate in Lemma 5.1 which is particularly relevant to quasilocal maps of the type from Example 5.2.
The bound above is particularly useful as it expresses the decay of the quantity on the left-hand-side above in both large n and m.
C. Compositions of quasilocal maps
In applications, we find it useful to recognize certain mappings as the composition of quasilocal maps. When Γ is finite, these compositions are well-defined and estimates, as indicated below, follow readily. For sets Γ of infinite cardinality, more care must be taken when defining such compositions. This section discusses two classes of examples where these compositions are well-defined and we describe the estimates that follow.
It will be convenient to make an additional assumption on the metric space (Γ, d). We say that (Γ, d) is ν-regular if the cardinality of balls in Γ grows at most polynomially, i.e., there exist non-negative κ and ν for which
More comments about ν-regular metric spaces (Γ, d) can be found in Subsection 1 of the Appendix. Under this assumption, given any and n > 0, the cardinality of X(n), the inflation of X defined in (4.14), satisfies the following rough estimate:
Let (Γ, d) and be a quantum lattice system on a ν-regular metric space. We will say that a linear map is locally bounded if there are non-negative numbers p and B for which
More general growth in X, i.e., the support of A, could be allowed, but the above moment condition covers all of the applications we have in mind. As discussed in Sec. V A, we say that a linear map is quasilocal if there are non-negative numbers q and C as well as a nonincreasing function G, G:[0, ∞) → [0, ∞), with limr→∞G(r) = 0 for which given any ,
We will refer to C, q, and G as the parameters of the quasilocal map.
We first consider compositions of linear maps for the following situation. Suppose that is locally bounded and quasilocal in that it satisfies both (5.37) and (5.38). Furthermore, assume that is linear, bounded, and quasilocal. So, in particular, there are non-negative numbers q2 and C2 and a decay function G2 for which the analog of (5.38) holds for . In many applications, the mapping arises as the unique linear extension of a bounded, quasilocal map . In this situation, we can define the composition in the usual way, i.e.,
Moreover, any such map satisfies the following estimate.
Let (Γ, d) be ν-regular, be a locally bounded, quasilocal map, and be a bounded, quasilocal map. For i = 1 or 2, denote by Bi, Ci, pi, qi, Gi the corresponding parameters from (5.37) and (5.38). Then, the following holds for the composition :
- is locally bounded: for any with ,where one may take and p = p1.(5.40)
- For any and , where ,where the numbers and p may be taken as in (5.40), one may take q = p1 + q2, and(5.41)
We note that if the function G described in (5.42) above is nonincreasing and satisfies limr→∞G(r) = 0, then the above estimates show that is quasilocal.
From Proposition 5.9, we now have conditions under which there is a well-defined composition of two locally bounded, quasilocal maps. Lemma 5.10 provides local bounds and quasilocal estimates for the resulting composition.
Let (Γ, d) be a ν-regular metric space. For i = 1, 2, let be locally bound, quasilocal maps. Suppose that G1, the decay function in (5.38) associated with , has a finite νp2-th moment and let denote the composition from (5.50).
- is locally bounded: for any with ,where one may take p = p2 + max{p1, q1} and(5.59)(5.60)
- For any and where , one has thatwhere one may take , q = max{p1, q1} + max{p2, q2}, and(5.61)(5.62)
Again we stress that the bounds above demonstrate that the composition is quasilocal if the function G in (5.62) is nonincreasing with limr→∞G(r) = 0.
D. Quasilocal transformations of interactions
Important applications of quasilocal maps arise in the classification of gapped ground state phases12,14,16,99,100,101 and recent proofs of stability of the spectral gap.21,22,83,85,96,97 In these proofs, key insights come from analyzing the composition of a quasilocal map with an interaction, . It is important to note that such maps are not necessarily interactions themselves, as the image lies in the quasilocal algebra, , rather than the algebra of local observables, . In our applications, the interaction and quasilocal map often depend on an auxiliary parameter and we allow for this in our construction and results. In what follows, we provide a general framework under which one can construct a bona fide interaction from such a composition and derive estimates that determine conditions under which these transformed interactions have a finite F-norm.
We begin with a general description of transformed interactions in Sec. V D 1. In Sec. V D 2, we prove estimates on these transformed interactions in finite volume. In Sec. V D 3, we give conditions under which the finite-volume results proven in Sec. V D 2 extend to the thermodynamic limit. A concrete application of these results will be given in Sec. VI E, where we show that the spectral flow automorphisms can be realized as the dynamics generated by a time-dependent interaction with good decay properties.
1. Transformed finite-volume Hamiltonians
To investigate the spectral properties of a given Hamiltonian H, it is often convenient to work with a unitarily equivalent Hamiltonian H′ = U*HU. When the original Hamiltonian is a sum of local terms, the strict locality of these terms is typically not preserved under the mapping H ↦ H′. In recent applications, most notably the proof of stability, it is shown that locality based arguments, such as Lieb-Robinson bounds, still apply to H′ if the automorphism implemented by the unitary U is sufficiently quasilocal.
In this section, we discuss this situation more generally. Specifically, we analyze the transformation of a given interaction by a quasilocal map. Briefly, we first argue that the composition of a quasilocal map with an interaction can, using the methods of Sec. IV B, still be realized as an interaction with strictly local terms. Moreover, we show that the spatial decay associated with this new interaction can be estimated in terms of the decays of the original interaction and the quasilocal map. Finally, we discuss quasilocality estimates for the dynamics of this transformed interaction.
To establish some notation, let us first consider a simple, time-independent case in finite volume. As before, fix a quantum lattice system composed of (Γ, d) and . Let Φ be an interaction on and recall that for any finite Λ ⊂ Γ, we denote by
the finite-volume Hamiltonian generated by Φ. Our goal here is to analyze the transformation of this local Hamiltonian by a linear map . In particular, we consider
Generally, the map will not preserve locality, and in such cases, each term in (5.68) will be global in the sense that for each X ⊂ Λ. For this reason, the sum on the right-hand-side of (5.68) does not represent an interaction in the sense defined in Sec. III A.
Using the methods of Sec. IV B, one can rewrite the right-hand-side of (5.68) as a sum of strictly local terms. To see this, fix a locally normal product state ρ on . In this finite-volume context, we only use the restriction of ρ to and again refer to it as ρ. In terms of ρ, we have defined local decompositions with respect to any X ⊂ Λ and each n ≥ 0 as the maps given by (4.15). Recall further that has a range contained in (using again the identification of the former as a subalgebra of ), and moreover, . In this case, each term appearing in (5.68) can be written as a finite telescopic sum as in (4.17) by
For any Z ⊂ Λ, define
with the understanding that empty sums are taken to be zero. By construction, and under the additional assumption that for all , we see that ΨΛ is a well-defined (finite-volume) interaction in the sense of Sec. III A. Moreover,
2. Finite-volume results
In this section, we give a finite-volume analysis of the transformed interactions briefly introduced at the end of Sec. V D 1. In Sec. V D 3, we will discuss appropriate conditions under which these results will extend to the thermodynamic limit. For many of our applications, both the interaction and the quasilocal map will be time-dependent. As a consequence, we state and prove our estimates for families of interactions and quasilocal maps.
We make two useful observations in this section. First, we indicate a set of continuity assumptions under which a finite-volume transformed interaction corresponds to a well-defined dynamics. These assumptions will also guarantee that the interaction which generates this transformed interaction is strongly continuous in the sense of Sec. III A 1. Next, we will show that certain decay assumptions on the initial interaction Φ and quasilocal map lead to explicit estimates on the decay of the interaction ΨΛ; here, we are using the notation introduced at the end of Sec. V D 1. Technically, the continuity and decay assumptions are independent; however, in most applications, the models we consider satisfy both sets of assumptions simultaneously.
Let (Γ, d) and be a quantum lattice system, and be an interval. We once again work with strongly continuous interactions , meaning that, for all ,
for all t ∈ I.
The map t ↦ Φ(X, t) is continuous in the strong operator topology on .
For any , we define a finite-volume, time-dependent Hamiltonian associated with Φ by
From the assumptions, it is clear that is also pointwise self-adjoint and strongly continuous as it is the finite sum of such terms.
For the remainder of this subsection, we fix a finite volume and are interested in studying time-dependent transformed finite-volume Hamiltonians analogous to those considered in Sec. V D 1. Specifically, given any family of linear maps , we consider the set of all operators of the form
and will refer to such collections as a finite-volume family of transformed interactions. Under assumptions which guarantee that is pointwise self-adjoint and strongly continuous, the methods of Sec. II B [see also Sec. III A 1] demonstrate that these transformed interactions correspond to a dynamics. More precisely, Proposition 2.2 shows that for any s, t ∈ I, the strong solution of
defines a two-parameter family of unitaries, and thus, a cocycle of automorphisms of with
We refer to as the dynamics corresponding to the transformed Hamiltonian in (5.73).
A main goal of this subsection is to establish assumptions under which the dynamics in (5.75) satisfies a quasilocality estimate, also known as a Lieb-Robinson bound (see Theorem 3.1). In order to do so, we first rewrite the family of transformed interactions in (5.73) as a sum of strictly local terms. Fix a locally normal product state ρ on . For any Z ⊂ Λ and each t ∈ I, set
where, as in Sec. V D 1, we have made local decompositions of the global terms on the right-hand-side of (5.73) [compare with (5.69) and (5.70)]. We stress that for all t ∈ I, we make local decompositions with respect to the same locally normal product state ρ. As in (5.71), it is clear that for each t ∈ I,
We now introduce a set of assumptions on the family of functions which guarantee that (i) the dynamics in (5.75) is well-defined and (ii) the mapping is a strongly continuous interaction in the sense of Sec. III A 1.
We assume that the collection of finite-volume linear maps is a strongly continuous family of strongly continuous transformations that are compatible with the involution in the sense that
for each t ∈ I, , for all ;
for each , the function is norm continuous;
for each t ∈ I, the map is continuous on bounded subsets when both its domain and codomain are equipped with the strong operator topology, and moreover, this continuity is uniform on compact subsets of I.
Assumption 5.11(i), together with Proposition 4.5(v), is used to ensure that the terms defined in (5.76) are pointwise self-adjoint. This is important in defining the unitary propagator, but it plays no role in establishing various continuity properties. Next, as is discussed in Sec. IV B 1, Assumption 5.11 (ii) and (iii) guarantee that is strongly continuous. As such, the finite-volume dynamics associated with this transformed interaction [see (5.74) and (5.75)] is well-defined. In particular, this dynamics is independent of the choice of ρ. Note further that Assumption 5.11 (ii) and (iii) also ensure that for each X ⊂ Λ, is strongly continuous. Given this, Proposition 4.3 shows that each of the finitely many terms on the right-hand-side of (5.76) is strongly continuous as well, and as a result, ΨΛ is a strongly continuous interaction. The interaction ΨΛ, which does depend on the choice of ρ, will be useful in proving a quasilocality bound on the finite-volume dynamics in (5.75).
The goal for the remainder of this section is to quantify a quasilocality estimate for the finite-volume dynamics in (5.75) in terms of decay properties of the original interaction Φ and the finite-volume transformations . For these results, we assume that (Γ, d) is ν-regular and equipped with an F-function F.
Let us again fix an interval and a finite-volume . We make the following decay assumptions on a family of finite volume transformations:
We assume that the family of finite-volume linear maps is a time-dependent family of locally bounded, quasilocal maps in the sense that
There is some p ≥ 0 and a measurable, locally bounded function B : I → [0, ∞) so that given any X ⊂ Λ,
- (ii)
There is some q ≥ 0, a nonincreasing function G : [0, ∞) → [0, ∞) with G(r) → 0 as r → ∞, and a measurable, locally bounded function C : I → [0, ∞) for which given any sets X, Y ⊂ Λ, one has that
We state our basic estimate on these finite-volume transformed interactions. In the statement, we make use of the quantities p, q, and G from Assumption 5.12.
It is clear from the statement, as well as the proof, that the estimate proven in Theorem 5.13 does not require that the mappings satisfy Assumption 5.11. As indicated previously, Assumption 5.11 is convenient because it guarantees that the mapping ΨΛ satisfies the continuity requirements needed to be a strongly continuous interaction, as defined in the beginning of this subsection.
We end this subsection with the following corollary:
3. Results in infinite volume
In this section, we show how the results of Sec. V D 2 extend to the thermodynamic limit. We begin with an assumption on a collection of quasilocal maps . In essence, this definition combines the notion of uniformly locally normal from Definition 4.9 with Assumptions 5.11 and 5.12. As always, we consider a quantum lattice system composed of (Γ, d) and , and let be an interval.
We assume that the family of linear maps is strongly continuous, uniformly locally normal, and uniformly quasilocal in the following sense: there is an increasing, exhaustive sequence of finite subsets of Γ with a family of linear maps for each n ≥ 1 such that
For each n ≥ 1, the family satisfies Assumption 5.11.
There is some p ≥ 0 and a measurable, locally bounded function B : I → [0, ∞) for which given any and n ≥ 1 large enough so that X ⊂ Λn,
- (iii)
There is some q ≥ 0, a non-negative, nonincreasing function G with G(x) → 0 as x → ∞, and a measurable, locally bounded function C : I → [0, ∞) for which given any sets and n ≥ 1 large enough so that X ∪ Y ⊂ Λn,
- (iv)
There is some r ≥ 0, a non-negative, nonincreasing function H with H(x) → 0 as x → ∞, and a measurable, locally bounded function D : I → [0, ∞) for which given any there exists N ≥ 1 such that for n ≥ N,
Before proving the theorem, we make the following comments: First, if is a family of linear maps which satisfies Assumption 5.15, then for any compact I0 ⊂ I, the family is clearly a strongly continuous family of uniformly locally normal maps in the sense of Definition 4.9. Moreover, conditions (ii) and (iii) of Assumption 5.15 guarantee that the sequence of finite-volume approximates satisfies Assumption 5.12 with estimates that are uniform in n. In Sec. VI, an explicit family of weighted integral operators of the type discussed in Example 4.11 will be shown to satisfy all conditions of Assumption 5.15.
In the remainder of this section, we will show that if the initial interaction Φ decays sufficiently fast, then the transformed interactions converge locally in F-norm to Ψ in the sense of Definition 3.7. Moreover, our assumptions will allow for an application of Theorem 3.8 from which we will conclude that the finite-volume dynamics in (5.97) converge. For ease of later reference, let us declare the relevant decay of Φ as an assumption.
Given a ν-regular metric space (Γ, d), and a family of maps satisfying Assumption 5.15, we assume Φ is a strongly continuous interaction such that for m = max{p, q, r}, where p, q, and r are the numbers in Assumption 5.15.
We can now state the main result of this section, for which it will be useful to review Definition 3.7.
Consider a quantum lattice system composed of a ν-regular metric space (Γ, d) and quasilocal algebra . Let be an interval, and let F be an F-function on (Γ, d). Assume that is a family of linear maps satisfying Assumption 5.15, Φ is an interaction satisfying Assumption 5.16, and ρ is a locally normal product state on .
Suppose the quasilocal decay function G from (5.93) has a finite 2ν + 1 moment and is an F-function on (Γ, d) satisfying (5.89), then .
Suppose there is some 0 < α < 1 for which Gα has a finite 2ν + 1 moment, where G is as in (5.93). Suppose also that is an F-function on (Γ, d) satisfying (5.89) with G replaced by Gα. Then, and Ψn converges locally in F-norm to Ψ with respect to .
Some comments are in order. First, under the assumptions of Theorem 5.17(i), the finite-volume interactions Ψn, as defined in (5.98), satisfy the assumptions of Theorem 5.13 and hence the estimate (5.85). In this case, for any F-function on (Γ, d) satisfying (5.89), the corresponding finite-volume dynamics, i.e., the automorphisms defined in (5.97), satisfy the quasilocality bound proven in Corollary 5.14 [see (5.90)]. A main point of Theorem 5.17(i) is that both of these observations extend to the thermodynamic limit. In fact, the assumptions of Theorem 5.17(i) also guarantee that the arguments in Theorem 5.13, and hence an analog of the bound (5.85), also apply to the infinite-volume interaction Ψ as defined in (5.100). Here, we are using that the uniform local convergence in (5.94) guarantees that both the local bound [see (5.92)] and the quasilocal bound [see (5.93)] extend to the limiting map , and in this case, Lemma 5.1 applies. Given this, for any F-function on (Γ, d) satisfying (5.89), one concludes that . As a result, we can apply Theorem 3.5, where we take the case of trivial on-sites Hz = 0 for all z ∈ Γ. This then shows that there exists an infinite volume dynamics, which we denote by τt,s, associated with Ψ. By construction, this infinite-volume dynamics τt,s also satisfies Corollary 5.14.
Theorem 5.17(ii) implies that, under the slightly stronger decay assumptions, the finite-volume dynamics converge to the infinite-volume dynamics τt,s in the sense given by Theorem 3.8. Since the interactions Ψn are constructed using finite-volume local decompositions [see (5.98)], they are not finite-volume restrictions of Ψ, and so an additional argument is required here. We remark that the decay assumptions in Theorem 5.17(ii) imply the decay assumed in Theorem 5.17(i). As a result, the better quasilocality estimates for the dynamics, which follow as a consequence of the assumptions in Theorem 5.17(i), may be used generally.
Next, a careful look at the proof of Theorem 5.17(i) shows that we actually only require for m = max(p, q). The proof of Theorem 5.17(ii) requires the stronger condition of Assumption 5.16, namely, for m = max(p, q, r).
Finally, we note that if the decay function G in (5.93) is a weighted F-function, the arguments below can be simplified a bit.
The proof of Theorem 5.17(i) is argued in the paragraphs above.
E. Quasilocality for the difference of two dynamics
In this section, we prove a quasilocality estimate for the difference of two dynamics as discussed in Example 5.6 of Sec. V B.
In the estimates below, we use an argument similar to that of Theorem 3.4 [see, in particular, (3.74)] to show that the claim holds with replacing in the definition of . However, by reordering the dynamics in (5.114) [or equivalently, by considering ], we see that the analog of (5.119) holds with the roles of the dynamics τt,s and αt,s interchanged. Since the argument given below applies equally well in this case, it will be clear that can then be expressed in terms of . We now continue with our estimate of the right-hand-side of (5.119).
Collecting the estimates in (5.121), (5.125), (5.128), and (5.130), we find (5.115) as claimed.
VI. THE SPECTRAL FLOW
In this section, we consider a family of finite volume quantum lattice Hamiltonians HΛ(s) acting on a Hilbert space that depend smoothly on a parameter s ∈ [0, 1]. We assume that the spectrum of HΛ(s) can be decomposed into two nonempty disjoint sets, i.e., spec(HΛ(s)) = Σ1(s) ∪ Σ2(s), where Σ1(s) is bounded, and the distance between Σ1(s) and Σ2(s) is greater than a positive value independent of s. The main goal of this section is to show that if the interaction defining HΛ(s) is smooth and decays sufficiently fasts, then we can use the theory described in Sec. V to construct a quasilocal automorphism , which we call the spectral flow, that maps the spectral projection of HΛ(s) onto Σ1(s) back to the spectral projection of HΛ(0) onto Σ1(0). In Sec. VII, we use the spectral flow to discuss the classification of gapped ground state phases. A second important application concerns models with a spectral gap above their ground states, for which s parameterizes a perturbation of the system; this is the main topic we analyze in Paper II.96 While both these applications are for ground states, the methods we introduce here are more general and work equally well for isolated bounded subsets anywhere in the spectrum.
Denoting by P(s) the spectral projection of HΛ(s) onto Σ1(s), the existence of an automorphism αs satisfying
is well-known. As shown by Kato in Ref. 67, under certain conditions which guarantee the smoothness of P(s), the unique strong solution of
is unitary and satisfies
The automorphism studied by Kato was for a family of Hamiltonians defined on a general Hilbert space , and so his results do not take into account the locality structure of a quantum lattice system. As a result, the automorphism induced by UK(s) is not obviously quasilocal. Hastings and Wen were the first to introduce a technique for constructing an automorphism on a quantum lattice system that both satisfies (6.1) and is quasilocal.59 In that work, they referred to the quasilocal automorphism as the quasiadiabatic evolution (or continuation). It is this approach that we follow in this section. Neither name, spectral flow, or quasiadiabatic continuation, accurately and unambiguously captures the essence of this quasilocal automorphism. It suffices to say that it is a unitary dynamics with useful properties. In other works, Hastings introduced novel ways to combine particular instances of the spectral flow with quasilocality properties of quantum lattice systems, most notably in Ref. 54. This work inspired a string of new results in the theory of quantum lattice systems, and so it seems appropriate to refer to the generator of this spectral flow as the Hastings generator.
A. Setup and main results
We first consider a family of parameter dependent Hamiltonians on a general complex Hilbert space and later return to apply our results to the setting of quantum lattice systems. Specifically, we consider operators that depend on a parameter s ∈ [0, 1], and we note that the choice of interval [0, 1] is a matter of convenience. We begin with the following definition:
Let be a complex Hilbert space. We say that a map is strongly C1 if Φ(s) is strongly differentiable for all s ∈ [0, 1], and the derivative is continuous in the strong operator topology.
In many concrete examples, one can pick the interval I(s) as the smallest interval containing Σ1(s), and in that case, μ = γ′.
We give some context for this result. Suppose that is such that the local Hamiltonians HΛ(s) are strongly C1. Recall that for any γ > 0, the Hastings generator , which is defined in terms of [see (6.14)], is strongly continuous and self-adjoint. The automorphism defined as in (6.10) can then be recognized as the Heisenberg dynamics associated with DΛ(s). If Theorem 6.4 holds, then DΛ(s) is itself a local Hamiltonian associated with a strongly continuous interaction . Applying the Lieb-Robinson bound, i.e., Theorem 3.1, to shows that the spectral flow is quasilocal as claimed. In the proof of Theorem 6.4, we show that the norm is bounded from above by a constant independent of Λ, from which local F-norm convergence will follow. The interaction Ψ then defines an infinite volume spectral flow automorphism that is also quasilocal.
Note that we do not require Assumption 6.2 for Theorem 6.4, and, in particular, the quasilocality result holds where the spectrum of HΛ(s) is or is not gapped. If, however, HΛ(s) satisfies Assumption 6.2 with gap , then the finite-volume automorphisms generated by for any will both be quasilocal and satisfy (6.12). In applications to stability, one is interested in the situation that there is some sequence of finite volumes for which both Theorems 6.3 and 6.4 hold simultaneously and that the gaps as in (6.9) are uniformly bounded from below by a positive constant independent of n.
In what follows, we will typically work with a Hastings generator and spectral flow automorphism that depend on a fixed value of γ. As such, we will often suppress the dependence of γ from our notation.
The remainder of this section is organized as follows: In Sec. VI B, we define the explicit weight function Wγ used in our results and prove some basic decay estimates on this function. The reader can skip these details on first reading. Recall that the definition of the Hastings generator is given in terms of a specific weighted integral operator. In Sec. VI C, we define several general weighted integral operators in terms of appropriate L1 functions and prove some useful properties. We use the results from this section to give the proof of Theorem 6.3 in Sec. VI D. We consider quantum lattice systems in Sec. VI E where we show that, in this context, the weighted integral operators introduced in Sec. VI C are quasilocal when defined using the weight functions from Sec. VI B. We then use these results to prove Theorem 6.4 (which is restated as Theorem 6.14). We end the section by showing that there is a well-defined spectral flow automorphism in the thermodynamic limit when the conditions of Theorem 6.4 hold.
B. An explicit weight
To write down the generator of the spectral flow dynamics [see (6.11)], requires a weight function with certain properties. In Sec. VI C, we will define a class of transformations on the algebra of observables of the form
where . In fact, we will make increasingly detailed assumptions of in order to prove useful properties of the map I. At some point, it becomes more efficient to work with a specific family of functions for which the assumptions hold. Having such a family of functions will make it possible to state explicit decay estimates that are useful for applications. As such, in this section, we introduce this family of functions, for which interesting properties were already investigated in Ref. 60, and prove some basic estimates; these will be particularly relevant in Sec. VI E 2. It will be clear that other functions can be used to derive similar results. The details of this section can be skipped on first reading. Its main importance is to demonstrate the existence of functions with all the desired properties.
Consider the sequence defined by
In terms of this sequence, define a function by setting
where c > 0 is chosen so that
It follows from Lemma 6.5 that and so this constant is well-defined.
It is clear that is non-negative and even. Moreover, if we denote by the unitary Fourier transform of , i.e., for each ,
then it is easy to check (see, e.g., Ref. 12) that supp(ŵ) ⊂ [−1, 1]. Lemma 6.5 provides a useful estimate on .
C. On weighted integrals of dynamics
In this section, we briefly discuss some general facts about weighted integrals of a dynamics. Such operators arise as the generator of the spectral flow, and in this case, a number of their basic properties are relevant.
1. Some generalities
Let H be a densely defined self-adjoint operator on a Hilbert space . Denote by τt the associated Heisenberg dynamics, i.e., the one parameter family of automorphisms of given by
For any , a bounded mapping is defined by setting
In fact, Stone’s theorem guarantees that this integral is well-defined in both the weak and strong sense. We refer to the operator I above as the integral of the dynamics τt weighted by , or more briefly, as a weighted integral operator.
Our applications will mainly concern families of these weighted integral operators. In fact, suppose H(s) = H + Φ(s) is as described in (6.3) and for each 0 ≤ s ≤ 1, consider with
where is the dynamics corresponding to H(s) [see (6.3) and (6.4)] and is real-valued. Lemma 6.7 is a useful observation.
2. Two particular weighted integrals
For the applications that follow, two particular weighted integral operators play a key role. We introduce a notation for them here and discuss some basic properties.
Generally, the setup is as before. Let H be a densely defined self-adjoint operator on a Hilbert space and denote by τt the corresponding dynamics [see, e.g., (6.42)].
For any fixed γ > 0, let be any real-valued functions so that (6.33), (6.38), and (6.39) hold. Define two linear maps by setting
As we will see, the properties of and depend crucially on the choice of γ > 0.
In the remainder of this subsection and Subsection VI D, we do not require the more detailed properties of and Wγ that we have proved for the specific functions constructed in Sec. VI B [see (6.18), (6.32), and (6.34)]. These properties will become important later when we analyze the quasilocality properties of the spectral flow. In particular, in Lemma 6.8 and the proof of Theorem 6.3, the specific functions defined in Sec. VI B are not required.
A useful observation for certain applications (see, e.g., Refs. 9 and 84) is that the map is a (left-) inverse of the Liouvillean [H, ·] on the space of off-diagonal operators.
In applications to quantum spin systems, either finite or infinite, the domain condition on in this proposition is quite generally satisfied due to the quasilocality properties of both and the generator of the Heisenberg dynamics. See, e.g., the discussion of the domain of the generator of the dynamics in the proof of Theorem 7.6.
D. The proof of Theorem 6.3
The goal of this section is to complete the proof of Theorem 6.3. Let us recap our progress so far.
Let H(s) = H + Φ(s) be as defined in (6.3). For any γ > 0, a map is defined by
where is the dynamics associated with H(s) as in (6.4) and Wγ is the particular weight function defined in (6.34). By Lemma 6.7, D(s) is pointwise self-adjoint and continuous in the strong operator topology. In this case, for any 0 ≤ s ≤ 1, an automorphism αs of is defined by setting
where the unitary U(s) is the unique strong solution of
The proof of Theorem 6.3 is completed by showing that if H(s) satisfies Assumption 6.2 for some γ > 0, then the automorphisms αs introduced above satisfy (6.12), i.e.,
E. Quasilocality of the spectral flow
For the remainder of this section, let us assume that (Γ, d) is a ν-regular metric space, in the sense of (5.35), and is the complex Hilbert space of the quantum system at x ∈ Γ. We start by considering a finite system corresponding to . Recall that for any X ⊂ Λ, we denote and .
1. Quasilocality for two weighted integral operators
In Sec. VI C 2, we introduced two particular weighted integral operators that will appear frequently in our applications. We now demonstrate that, under certain additional conditions, each of these weighted integral operators satisfies an explicit quasilocality estimate in the sense of Sec. V.
Let us assume that there is a one-parameter family of automorphisms of , which we denote by τt that satisfies a quasilocality estimate. More precisely, suppose that there are positive numbers C and as well as a non-negative, nondecreasing function g for which given any X, Y ⊂ Λ,
for all , , and . Here, d = d(X, Y) is the distance between the sets X and Y . As discussed in Sec. III, such a bound is known for the dynamics generated by a short range Hamiltonian; it is, e.g., a consequence of the Lieb-Robinson bounds in Theorem 3.1. In order to prove the quasilocal bounds below, we need only to know (6.78) and that g(d) becomes sufficiently large [see (6.90)]. In applications, we typically have (6.78) with
In terms of these automorphisms τt, for each γ > 0, define by
for any [compare with (6.50)]. Here again and Wγ are the specific weight functions introduced in Sec. VI B.
Before we state our first result, recall that (t) = γ(γt) and therefore with c being the L1-normalization of [see (6.18)]. Moreover, Corollary 6.6 [see specifically (6.40)] demonstrates that there is an η ∈ (2/7, 1) for which given any x ≥ γ−1e9 the bound
holds. Here, for any b > 0, we have introduced the subadditive, nondecreasing function
Our quasilocality estimate on the weighted integral operator follows.
It can be verified that the function given in (6.84) is monotone and strictly decreasing when .
Depending on the application one has in mind, more decay of the function governing the locality of the dynamics, specifically e−g(d), may be needed. For example, many applications require certain moments of the function to be finite. Let us make two observations in this regard.
Lemma 6.11 is the analog of Lemma 6.10 applicable to .
The proof of this lemma is almost identical to that of Lemma 6.10, except that one uses the estimate (6.41) from Corollary 6.6, instead of (6.40). We also use that .
It can also be verified here that the function given in (6.96) is monotone and strictly decreasing when .
2. Quasilocality of the spectral flow automorphism
We will now consider the spectral flow in the thermodynamic limit. In order to derive explicit estimates useful for applications, we work with F-functions of the ν-regular metric space (Γ, d) of the form F(r) = e−g(r)F0(r), where g is nondecreasing and subadditive, and
for a suitable ξ > 0. As is shown in the Appendix (Subsection 1 a of the Appendix), any choice of ξ > ν + 1 will define an F-function on a ν-regular (Γ, d). In the case , ξ > ν is sufficient. We will say that F is a weighted F-function on (Γ, d) with base F0.
Let us now introduce the models we consider through an assumption.
There is a collection of densely defined, self-adjoint on-site Hamiltonians. For each 0 ≤ s ≤ 1, there is an interaction Φ(s) on for which
For each , for all 0 ≤ s ≤ 1.
For each , strongly C1 in the sense of Definition 6.1.
- F is a weighted F-function on (Γ, d) with base F0 as in (6.97) and there is a bounded, measurable function ∥Φ∥1,1:[0, 1] → [0, ∞) for which given any x, y ∈ Γ, the estimateholds for all 0 ≤ s ≤ 1.(6.98)
Proposition 6.13 relates the quantities introduced above to the methods discussed in Sec. V D 3.
Consider a quantum lattice system composed of a ν-regular metric space (Γ, d) and . Suppose Assumption 6.12 holds with a weighted F-function F(r) = e−g(r)F0(r) for which limr→∞g(r) = ∞ and F0 is as in (6.97). Then, for any γ > 0, the family of maps [as defined in (6.102)], satisfies the conditions of Assumption 5.15.
As is clear from the statement, the results we prove below hold for any γ > 0. For convenience of presentation, we now fix such a value γ > 0 and then suppress it in our notation.
In the first part of Assumption 5.15, we show that for each n ≥ 1, the finite volume families of maps satisfy Assumption 5.11. Since is real-valued, Assumption 5.11(i) is easily verified. We note that integrability of Wγ is a consequence of the estimate (6.41) in Corollary 6.6. To check the remaining parts of Assumption 5.11, we recall that the properties of maps with the form (6.103) were discussed in Example 4.11. Assumption 6.12 guarantees that the methods of Example 4.11 apply, and the remaining details are readily checked.
We can now introduce the spectral flow for the class of models under consideration. As above, all comments below are valid for any choice of γ > 0, which is suppressed in the notation. The spectral flow automorphism can be defined for any choice of γ > 0, and under the general assumptions above, the spectral flow is quasilocal with γ-dependent estimates. It is only for the special relation with the spectral projection P(s) as in (6.12) that γ needs to be a lower bound for the gap in the spectrum as described in Assumption 6.2.
The main result of this subsection is as follows:
We make some comments and point out two corollaries before proving the theorem.
First, in principle, one can do better than the growth assumption in (6.119); in fact, one needs only that there is some 0 < ϵ < 1 for which the decay function Gϵ [see (6.107)] satisfies the conditions of Theorem 5.17 (ii). As can be seen from our comments in Sec. VI E 1 after Lemma 6.10, any weight function g satisfying (6.119) corresponds to such a decay function Gϵ which satisfies the conditions of Theorem 5.17 (ii). However, no weight function g which grows proportional to a logarithm [see (6.91)] corresponds to a decay function Gϵ satisfying the conditions of Theorem 5.17 (ii).
For ease of later reference, we now state two corollaries providing explicit estimates on the quasilocality of the spectral flow.
The following corollary is a direct application of Theorem 3.8 (i):
It is clear that, under the decay assumptions (6.119), Proposition 6.13 holds. In this case, for each γ > 0, the family of maps satisfies Assumption 5.15, and moreover, Φ′ is a suitable initial interaction in the sense of Assumption 5.16. As before, we will suppress the dependence on γ > 0 in what follows.
Since the form of the decay function Gϵ, as in (6.107), is explicit [see (6.96)], it is clear that has a finite 2ν + 1 moment. Arguing as above, a similar, yet different F-function can be produced which satisfies the assumptions of Theorem 5.17 (ii) with α = 1/2; in fact, any choice of 0 < α < 1 suffices. In this case, Ψn converges locally in F-norm to Ψ with respect to this function . As indicated previously, for the estimates in Corollaries 6.15 and 6.16, one can use the original F-function having the form (6.121).
VII. AUTOMORPHIC EQUIVALENCE OF GAPPED GROUND STATE PHASES
A. Uniformly gapped curves and automorphic equivalence
In this section, we use the spectral flow to study gapped ground state phases of a quantum lattice model (Γ, d) and . As before, we will discuss both finite and infinite volume systems and take the thermodynamic limit along a sequence of increasing and absorbing finite volumes. To this end, we will consider the following setup for this section.
Throughout this section, let (Γ, d) be a fixed ν-regular metric space with a weighted F-function of the form F(r) = e−g(r)F0(r), where F0 is an F-function for (Γ, d) of the form (6.97) and g is a non-negative, nondecreasing, subadditive function bounded below by arθ for some θ ∈ (0, 1]. In addition, we consider a fixed sequence of increasing and absorbing finite volumes Λn↑Γ and with the convention that we always take the thermodynamic limit with respect to a subsequence of this sequence. We will use the notation to denote the space of differentiable curves of interactions , s ∈ [0, 1], satisfying Assumption 6.12. At each x ∈ Γ, we may have a densely defined self-adjoint Hx, but these we regard as fixed. Specifically, we only consider here relations between models with different interactions Φ(s) but with the same {Hx∣x ∈ Γ}.
For simplicity, we will assume that the finite-volume Hamiltonians for the models parametrized by s ∈ [0, 1], are defined by
Within the context described above, we now introduce the notion of a uniformly gapped curve of models or, equivalently, a curve of uniformly gapped interactions for which we use the notation EΛ(s) = inf specHΛ(s) to denote the ground state energy of HΛ(s).
We can leave γ unspecified and call the curve simply uniformly gapped if there exists γ > 0 such that it is uniformly gapped with gap γ.
It is well-known that the spectral gap generally depends on the boundary conditions. Our choice to define the path of Hamiltonians (7.1) with respect to a single interaction leads to boundary conditions that are not necessarily the most general ones of interest; studying all possible cases at once would lead to quite onerous notation, which we want to avoid in this discussion. Suffice it to note that everything in this section could be generalized to the situation where we have boundary conditions expressed by a sequence , by requiring that Φn converges locally (in a uniform version of Definition 3.7) in a suitable norm to some . In this case, Φn is then used to define the Hamiltonian (7.1) on the volume Λn for each n. For example, this can be used to study finite systems in with periodic boundary conditions. Even without considering n-dependent interactions, the present setup allows one to study the effects of certain boundary conditions. For example, by replacing Γ by a subset Γ0 ⊂ Γ and different sequences of finite volumes Λn, models defined with the same interaction Φ may show different behavior. An example of this is discussed in detail for a class of so-called PVBS models in Refs. 11 and 18. There, Γ0 is the half-space in defined by an arbitrary hyperplane. For these models, the spectral gap is shown to depend nontrivially on the orientation of the hyperplane.
We use the notion of a uniform gap to define a relation ∼ on as follows:
For , we say that Φ0 and Φ1 are equivalent, denoted by Φ0 ∼ Φ1, if there exists a uniformly gapped curve such that Φ(0) = Φ0 and Φ(1) = Φ1.
In the physics literature, two models Φ0 and Φ1 are said to be in the same gapped ground state phase if Φ0 ∼ Φ1.32,33 Studying curves of models has proved to be fruitful also in mathematical studies.12–16 In this section, we explore some essential properties of models that belong to the same gapped ground state phase. First, however, we show that the relation ∼ used to define this notion is indeed an equivalence relation.
We found it convenient to state Definition 7.2 in terms of differentiable curves because we rely on the differentiability in the construction of the spectral flow automorphisms. Equivalently, however, one can just assume the existence of a piecewise differentiable continuous curve. As we show in the proof of the Proposition 7.3, a simple reparametrization of the concatenation of two differentiable curves yields a differentiable curve with the desired properties.
The relation ∼ defined in Definition 7.2 is an equivalence relation on .
Note that, without loss of generality, we can assume that the sequence (δn) in Definition 7.1 is nonincreasing. It is also easy to see that for uniformly gapped Φ, the spectral projection Pn(s) of associated with the interval becomes independent of the choice of sequence (δn) for large n in the sense that for any two sequences (δn) and (δn′) for which (7.2) holds, the spectral projections associated with the intervals and coincide for sufficiently large n.
Let us now fix a uniformly gapped curve with gap γ. As an application of Theorem 6.14 and the comments following it, we have strongly continuous spectral flow automorphisms for the curve of finite-volume on Λn and αs for the infinite system on Γ. Here, the uniform gap of the curve plays the role of the parameter γ in the construction of the spectral flow. Moreover, Theorem 6.14 [see specifically (6.122)] establishes that αs is the strong limit of , and the convergence of this limit is uniform for s ∈ [0, 1]. Moreover, we can use the spectral flow αs to construct a cocycle of automorphisms , for all t, s ∈ [0, 1]. We can similarly define a collection of finite volume cocycles, .
This is a direct consequence of (7.4), Theorem 6.14, and Lemma 7.5.
Let be a strongly convergent sequence of automorphisms of a C*-algebra , converging to α, and let be a sequence of states on . Then, the following are equivalent:
ωn converges to ω in the weak-* topology;
ωn ◦ α converges to ω ◦ α in the weak-* topology;
ωn ◦ αn converges to ω ◦ α in the weak-* topology.
We conclude this section with the following result regarding the continuity of the spectral gap above the ground state energy of the GNS Hamiltonian of an infinite volume ground state for the case of quantum spin systems, i.e., the single-site Hilbert spaces are finite-dimensional and the dynamics is generated by an interaction for a suitable F-function F. The restriction to the case of quantum spin systems is because we rely on some well-known properties of the dynamics and, in particular, its generator in that case. The finite-dimensionality of the single-site Hilbert spaces is not essential, but the boundedness of the interactions is, in addition to the general setup described at the beginning of this section, including Assumption 6.12.
Recall that for each s, the infinite volume dynamics is a strongly continuous group of automorphisms of generated by a closed operator δ(s), i.e., , and that is a core for δ(s).
As far as we are aware, for all models that satisfy the conditions of the theorem, the gap appears to be continuous in the parameter, not just semicontinuous. In particular, the gap is continuous when perturbation theory applies. This raises the question whether one indeed has continuity of the spectral gap as long as it is strictly positive, or whether additional assumptions are needed for continuity. Needless to say, the gap is not always stable and so should not be expected to be continuous, in general, on a domain where it vanishes at some points.
B. Automorphic equivalence with symmetry
In Sec. VII A, we introduced the classification of gapped ground state phases through equivalence classes of interactions for which there exists an interpolation by a uniformly gapped curve. We showed that within each equivalence class, the sets of ground states are mapped into each other by an automorphism with good quasilocality properties (the spectral flow derived from the uniformly gapped curve of interactions interpolating between the models). Implicit in this description is the idea that any curve of interactions interpolating between two models in distinct phases (different equivalence classes) must contain at least one point where the gap vanishes. Such points are called quantum critical points and one says that a quantum phase transition occurs in the system.115
Physical systems often have symmetries that play an important role. In the description of certain phenomena, it may be essential that a certain symmetry be present in the model. This led to the concept of symmetry protected gapped phases31,39 due to the observation that if one only allows curves of interactions that all possess a given symmetry, a finer classification of gapped ground state phases may arise. A nice example of this are the protected phases of the spin-1 chain.33,102,104,122 In general, the equivalence classes break up into subclasses if a restricted set of uniformly gapped curves of interactions is used to define the equivalence relation. This prompts us to revisit the notion of automorphic equivalence in the presence of symmetry.
A symmetry is usually specified by the action of a group G (as automorphisms) on the algebra of observables of the system. Although there are interesting symmetries that do not fit the framework of group representations by automorphisms, such as dualities and quantum group symmetries, we limit the discussion here to that setting. In general, we use the label G to specify the presence of a certain symmetry. So, we will consider spaces of interactions and of curves of interactions . To be clear, in this context, G stands for the full specification of the symmetry including its action on the system, not just the abstract group.
Here are four important classes of symmetries:
Local symmetries are described by automorphisms β of with the property that they leave the single-site algebras, , invariant. Specifically, we assume that the restrictions of β to , x ∈ Γ, are inner automorphisms given in terms of a unitary : for . These types of symmetries are sometimes called gauge symmetries because gauge symmetries are of this form. Thus, any local symmetry β is determined by a family of unitaries . We say that β is a symmetry of Φ if β(Φ(X)) = Φ(X), for all . It is easy to see that this implies that β commutes with the dynamics τt generated by Φ: β ◦ τt = τt ◦ β. If Φ depends on a parameter s or on the time t, the symmetry condition is assumed to hold pointwise in s and/or t. The set of all local symmetries form a group under the law of composition of automorphisms. It is often useful to consider the (projective) representations of this group, G, given by the local unitaries Ux(g), g ∈ G.
Lattice symmetries are, in general, described by a bijection R: Γ → Γ. It is usually important that R preserves the local structure of (Γ, d), e.g., one requires that R is isometric: d(R(x), R(y)) = d(x, y), x, y ∈ Γ. Examples include translations of lattices such as and reflection symmetries satisfying R2 = id. If we assume that , R can be lifted to an automorphism of as follows. Denote by the natural isomorphism (or a well-chosen one) and define the automorphism βR of , by putting
The symmetry of the interaction is expressed by the property Φ(R(X)) = βR(Φ(X)). In the case of lattice translations, this yields a representation of on , i.e., for denotes the action of translations on Γ, and X + a = {x + a∣x ∈ X}. Correspondingly, Φ is called translation invariant if βa(Φ(X)) = Φ(X + a), for all .
- (iii)
Time-reversal symmetry is expressed as a local symmetry [discussed in (i)] given by an antiautomorphism, implemented on each site by an antiunitary transformation. The latter are, in general, the composition of a unitary transformation and a complex conjugation. Besides taking into account the antilinearity, time reversal symmetry can be treated in the same way as linear local symmetries.
- (iv)
Chiral symmetry is described by a unitary, say C, that anticommutes with the Hamiltonian. So, at each point in the curve of Hamiltonians, we have C*H(s)C = −H(s). For the dynamics, this implies that for all s ∈ [0, 1], , and ,
It should be noted that the basic types of symmetries can be combined. For example, some models are invariant under a combined lattice reflection and time-reversal transformation, without possessing either of these symmetries separately.
We assume the same setup as described in the beginning of Sec. VII of a fixed ν-regular metric space (Γ, d) with a specified weighted F-function F. Let G denote the symmetries under consideration. The fixed family of on-site Hamiltonians Hx, x ∈ Γ is assumed to have the symmetry G as if it were a zero-range interaction. For example, if describes a local unitary symmetry, we assume that the domain of Hx is invariant under Ux and that Hx and Ux commute. Or, as another example, if and the symmetry is the full translation invariance of the lattice, Hx is assumed to be the same self-adjoint operator at each site x.
For the interactions, let denote the space of interactions with finite F-norm that possess the symmetry G, and
Definitions 7.1 and 7.2 of uniformly gapped curves and the equivalence relation now carry over the situation with a symmetry G in the obvious way, as does the proof of the analog of Proposition 7.3. The resulting equivalence classes are called symmetry protected phases. Since the uniformly gapped curves with symmetry are a special case of the general situation, Theorem 7.4 applies and the spectral flow automorphism establishes a bijection between the sets of states along the curve.
For the study of the stability of gapped ground state phases with symmetry breaking we present in Paper II,96 it will be important that the automorphisms αt,s commute with the automorphisms βg, g ∈ G, representing the symmetry on . Moreover, it will be desirable that the interaction Ψ(s) generating αt,s and its finite-system analogs all have the symmetry. There are a few subtleties that merit further discussion concerning the construction of a spectral flow with the desired symmetry properties.
As mentioned above, it is important that both the spectral flow αt,s and its generating interaction Ψ(s) respect the symmetry of the initial interaction Φ(s). Recall that the conditional expectations ΠX from (4.11) play a crucial role in the quasilocality properties of αt,s and the definition of Ψ(s) [see Corollary 6.15, (6.118) and Sec. IV B]. In the presence of a local symmetry β, it is useful to choose the locally normal product state in the definition of the conditional expectations ΠX that is β-invariant, meaning or, equivalently, β(ρx) = ρx [see (4.8)]. This requirement guarantees that if A is invariant under β, then so is ΠX(A), i.e.,
If , then a β-invariant locally normal product state always exists. For example, setting ρx to be the tracial state will produce a β-invariant state. Given any Φ(s) with a local symmetry and any symmetric ρ, it is easy to see using (7.13) that the Hastings interactions Ψn defined in (6.117) and, consequently, the spectral flow derived from Φ(s) both inherit this symmetry. The same holds true for the corresponding infinite volume objects, Ψ and αs. In particular, αs commutes with any local symmetry automorphism β that leaves Φ′(s) invariant for all s ∈ [0, 1].
For infinite-dimensional , a symmetric normal state on may or may not exist. One may have to relax either the normality or the symmetry requirement. Which of the two is more relevant would depend on the situation at hand but for the type of applications we are considering here, it is important to use normal states. In the case of a gauge symmetry described by a compact Lie group, constructing symmetric normal states is not a problem. However, even when such a state does not exist, the symmetry of the spectral flow is restored in the thermodynamic limit. This follows from the observation that although the infinite-volume interaction Ψ(s) depends on the choice of the locally normal state ρ used in its construction, the infinite-volume flow is the thermodynamic limit of automorphisms generated by self-adjoint operators that commute with the symmetry. This is apparent, e.g., from expression (5.71) in which is to be replaced by (s) and is defined in (6.50).
In the presence of a lattice symmetry such as translation invariance, it makes sense to pick a translation invariant product state ρ to define the conditional expectations ΠX. This is obviously always possible and yields a covariant family of conditional expectations, meaning that βa ◦ ΠX = ΠX+a ◦ βa, where for , X + a denotes the action of the translations on Γ and βa denotes the corresponding action on . Finite subsystems defined on quotient lattices, e.g., /() with n ≥ m, can have the corresponding quotient symmetry , which is equivalent to considering the system with periodic boundary conditions. In general, finite systems will not have an exact translation symmetry but, again, the symmetry is recovered in the thermodynamic limit.
The case of time-reversal symmetry can be treated in the same way as local unitary symmetries. Due to the oddness of the function Wγ in (6.35), the Hastings interaction Ψ(s) changes sign under time-reversal. Since the time-reversal automorphism is antilinear, however, this is exactly the requirement for and αs to commute with it.
The case of a fixed chiral symmetry C along the curve of Hamiltonians H(s) = H + Φ(s) implies that Φ′(s) anticommute with C. Using again the oddness of the function Wγ, and the property (7.11), it is straightforward to check that C then commutes with the generator of the spectral flow, i.e., it is a symmetry of the spectral flow.
The list of types of symmetries we have discussed here is not exhaustive. For example, another type of symmetry relevant for applications is duality symmetries. We postpone the discussion of those to Paper II,96 where we will study the stability of gapped ground state phases.
C. Examples of uniformly gapped curves
The construction of the automorphisms αt,s assumes the existence of a uniform lower bound for the spectral gap above the ground states along the curve of models in the sense of Definition 7.1. Establishing a uniform bound for the gap is generally a very hard problem. Fortunately, there are a good number of interesting examples where the existence of a positive uniform lower bound can be proved.
The largest variety of examples is found as a result of perturbing models for which the ground state and the existence of a spectral gap above it are known. We will review the state of the art of perturbative results of this type in Paper II.96 For this reason, we limit ourselves here to citing a few works that illustrate the broad range of examples that exist in the literature: some exactly solvable models such as the anisotropic XY chain,74 quantum perturbations of classical spin models,69,80 perturbations of the AKLT chain2,128 and similar models,121 perturbations of simple models with topological order in the ground state such as the Toric code model,21 general perturbations of frustrationfree models satisfying a local topological order condition,83 and perturbations of quasifree fermion systems.37
Other interesting examples for which explicit lower bounds for the gap can be obtained and classes of models for which the equivalence classes can be explicitly determined are the frustrationfree spin chains with finitely correlated ground states, also known as matrix product states.13,15,40,87,99,100,101 Allowing for general perturbations of such models typically leads to splitting of the degenerate ground states found in the frustrationfree model. The so-called Kennedy triplet of “excited’”states of the spin-1 Heisenberg antiferromagnetic chain of even length can be regarded as an example of this phenomenon.68 In general, sufficiently small perturbations of one-dimensional frustrationfree models with a gap above the ground state will have a group of eigenvalues near the bottom of the spectrum separated by a gap (uniform in the size of the system) from the rest of the spectrum. The associated eigenstates all converge to ground states in the thermodynamic limit. Both statements are proved in Ref. 85.
We postpone a more comprehensive discussion of examples of models with distinct gapped ground state phases until after the presentation of the stability results of gapped phases in Paper II.96
ACKNOWLEDGMENTS
We would like to thank Valentin Zagrebnov for illuminating discussions about the nonautonomous Cauchy problems that arise in quantum dynamical systems. We also thank Martin Gebert for reading an early version of this paper and asking good questions and Derek Robinson for several useful remarks and informative comments about the history of the subject. All three authors wish to thank the Department of Mathematics of the University of Arizona and the University of California, Davis, for extending their kind hospitality to us and for the stimulating atmosphere they offered during several visits back and forth over the years it took to complete this project. B.N. also acknowledges the support of a CRM-Simons Professorship for a stay at the Centre de Recherches Mathématiques (Montréal) during Fall 2018, which created the perfect circumstances to complete this paper. B.N. was supported by the National Science Foundation under Grant Nos. DMS-1515850 and DMS-1813149.
APPENDIX: TECHNICAL ESTIMATES INVOLVING -functions AND -norms
This section collects a number of facts about the decay bounds used throughout this paper. In general, we will assume that Γ is a countable set equipped with a metric, and we denote this metric by d. A good example to keep in mind is with the ℓ1-metric. When necessary, we will also assume Γ is ν-regular [see (A13)].
When considering the Heisenberg dynamics associated with a Hamiltonian, our quasilocality estimates require a short-range assumption on the corresponding interaction. For general sets Γ, which need not have the structure of a lattice, a sufficient condition for the existence of a dynamics in the thermodynamic limit can be expressed in terms of a norm on the interaction. We have found it convenient to express the decay of interactions with distance by a so-called F-function, which we discuss below. Depending on the application one has in mind, more explicit forms of decay, again expressed in terms of a family of F-functions, is convenient. These are by no means the only ways to express decay assumptions for interaction. If generality is not the concern, one can easily re-express decay into a more suitable form for the case at hand, say, e.g., for systems with pair interactions only. In this appendix, our goal is to briefly summarize various notions of decay which occur frequently in the main text.
1. On F-functions
Let (Γ, d) be a countable metric space. When Γ is finite, most notions introduced below are trivial, and for that reason, we will mainly consider the situation where Γ has infinite cardinality. We will say that Γ is equipped with an F-function if there is a nonincreasing function F:[0, ∞) → (0, ∞) for which
F is uniformly integrable:
- (ii)
F satisfies a convolution condition:
Any function F satisfying (A1) and (A2) will be called an F-function on Γ. We note that an immediate consequence of (A2) is that for any pair x, y ∈ Γ, we have the bound
The constant CF enters into a number of our estimates. We say that an F-function on Γ is normalized if CF = 1. Of course, for any F-function F, the function defines a new F-function on Γ for which .
Note that if Γ is equipped with an F-function F, then
where the left-hand-side above is a uniform estimate on the cardinality of the ball of radius n centered at x ∈ Γ. The above follows immediately from the estimate
This estimate also demonstrates that the existence of an F-function guarantees that Γ is uniformly, locally finite.
Moreover, if Γ is infinite, the existence of an F-function implies that the diameter of Γ is infinite. In this situation, if is an increasing, exhaustive sequence of finite subsets of Γ (i.e., Λn ⊂ Λn+1 for all n ≥ 1 and Λn↑Γ), then for any finite X ⊂ Γ,
This follows by observing that for any m ≥ 1,
is a finite subset of Γ. Since is absorbing, there is an N ≥ 1 for which X(m) ⊂ ΛN. Since Γ has infinite cardinality, the set ΓΛN is nonempty. It immediately follows that d(x, y) ≥ m for all x ∈ X and y ∈ΓΛN, from which (A6) follows.
a. Two common examples of F-functions
First, many well-studied quantum spin models are defined on the hypercubic lattice for some integer ν ≥ 1. For concreteness, consider equipped with the ℓ1-metric
Other translation invariant metrics can be treated similarly. For any ϵ > 0, the function
is an F-function on . Integrability follows from
Moreover, for any metric space (Γ, d): if p ≥ 1, the bound
holds for all x, y, z ∈ Γ, since the function t ↦ tp is (midpoint) convex. In this case, the function defined in (A9) satisfies (A2) with
Next, we note that for many of our results, it is not necessary that Γ has the structure of a lattice. We will say that a metric space (Γ, d) is ν-regular if there exist ν > 0 and κ < ∞ for which
Here, for any x ∈ Γ and n ≥ 0, bx(n) is the ball of radius n centered at x and | · | denotes cardinality. From (A4), we see that if Γ has an F-function for which F(r) ≤ Cr−ν, then Γ is ν-regular.
If (Γ, d) is ν-regular, then for any ϵ > 0, the function
is an F-function on Γ.
To see that this is the case, we need only to check uniform integrability, i.e., (A1), as an argument using (A11) shows that this F satisfies (A2). Fix x ∈ Γ. Set Bx(1) = bx(1) and Bx(n) = bx(n)bx(n − 1) for any n ≥ 2. It is then clear that (A13) implies
We can combine the above discussion with (A5) to prove the following result:
Let (Γ, d) be a countable metric space.
If (Γ, d) is ν-regular then F(r) = (1 + r)−(ν+1+ϵ) is an F function of (Γ, d) for all ϵ > 0.
If F(r) = (1 + r)−(ν+ϵ) is an F-function of (Γ, d) for all ϵ > 0, then Γ is ν-regular.
2. On weighted F-functions
For certain applications, it is convenient to consider families of F-functions of a specific form, which we call weighted F-functions.
Let (Γ, d) be a metric space equipped with an F-function F as described in Subsection 1 of the Appendix. Let g:[0, ∞) → [0, ∞) be a non-negative, nondecreasing, sub-additive function, i.e.,
Corresponding to any such g, the function
is an F-function on Γ. In fact, since g is non-negative, Fg satisfies (A1) with ∥Fg∥ ≤ ∥F∥. Moreover, since g is nondecreasing and sub-additive, one also has that
Thus, (A2) holds with .
We may refer to F as the base F-function associated with Fg; note that F0 = F for g = 0. The function g induces a factor r ↦ e−g(r) which is often referred to as a weight. We may also loosely refer to g as a weight and similarly Fg as a weighted F-function. One readily checks that sums, non-negative scalar multiples, and compositions of weights are also weights; in the sense that if g1 and g2 are both non-negative, nondecreasing, sub-additive functions, then so too are g1 + g2, ag1 (for a ≥ 0), and g1 ◦ g2.
In certain applications, it is useful to introduce a one-parameter family of weighted F-functions by taking a base F-function F on Γ, fixing a weight g, and associating to any a ≥ 0, the function ga(r) = ag(r), for which . When g is understood, we often just write to describe this family of weighted F-functions. For example, if F(r) = (1 + r)−p and g(r) = r, then Fa(r) = e−ar(1 + r)−p is the family of weighted F-functions defined in Sec. III A 1.
As a motivation for introducing these weights, consider again . As discussed in Subsection A 1 a of the Appendix, the polynomially decaying function F in (A9) is an F-function associated with . Such an F-function is appropriate for interactions Φ with terms that decay polynomially with the diameter of their support: ∥Φ(X)∥ ≤ C(1 + diam(X))−ν+ϵ. However, we are typically interested in interactions whose terms decay much faster, in particular, exponentially fast. One readily checks that for any a > 0, the exponential function g(r) = e−ar fails to satisfy (A2) and as such is not an F-function on , but does satisfy the criterion to be a weight. Since exponential functions often govern the decay of our interactions, it is convenient that one can obtain an exponentially decaying F-function on by making an appropriate choice of weight.
Before moving on to discussing several useful weights, we point out one added benefit of these functions. In the situation where we do not assume to have a weighted F-function and given , we will often use the simple bound
when applying a Lieb-Robinson bound or a quasilocality bound [see (3.24) and (5.2)]. For weighted F-functions, however, the following is also frequently used:
Here, one typically is considering a quantum lattice system defined on a (large) finite volume Λ, and in the situation that Y = ΛX(n), then the RHS of (A21) decays as e−g(n).
a. Three common weights
With an eye toward our specific applications, we now introduce three particular classes of weights.
First, let μ ∈ [0, 1]. The function g:[0, ∞) → [0, ∞) given by g(r) = rμ is non-negative, nondecreasing, and sub-additive in the sense of (A17). The constant function g(r) = 1 corresponding to μ = 0 is of minor interest; however, the choice of μ = 1 generates exponentially decaying weights. When 0 < μ < 1, the function is often called a stretched exponential.
Next, we provide an example between exponential and stretched exponential decay. As we will show, for any p > 0, the function
is non-negative, nondecreasing, and sub-additive. In our applications of the spectral flow (see, e.g., Sec. VI), the choice of p = 2 is particularly relevant.
Note that at r = ep, the nonconstant part of g has a zero derivative, and for r > ep, g is strictly increasing. That motivates this particular choice of cutoff. Also, it is easy to see that this function is subadditive by taking cases: Let r, s ≥ 0. Consider (i) r + s ≤ ep and (ii) r + s > ep. Both cases are easy to see. For the second case, use
Note that
the latter fact using that r + s > ep.
Finally, the function g:[0, ∞) → [0, ∞) given by
is clearly non-negative and nondecreasing. Since for any r, s ≥ 0, we have that
3. Simple transformations of F-functions
In certain applications, it is convenient to know that various decaying functions are in fact F-functions. Quantities of interest can be estimated in terms of translations or rescalings of known F-functions. For ν-regular Γ, these modifications preserve the basic properties of an F-function. The following two propositions show that suitably defined truncations, shifts, and dilations of F-functions are again F-functions.
If F(r) = e−g(r)F0(r) is a weighted F-function on (Γ, d), then for any ϵ > 0 the function defined by is an F-function on (Γ, d). Moreover,
- (ii)If F(r) = e−g(r)F0(r) is a weighted F-function on (Γ, d), then for any a > 0 the function defined byis an F-function on (Γ, d). In fact,(A33)
4. Basic interaction bounds
In the main text, we frequently use a number of basic estimates concerning interactions that are expressed using F-functions. Here, we collect a few results for later reference.
We begin by recalling some of the basic notation associated with interactions. Let (Γ, d) be a metric space equipped with an F-function F as discussed in Subsection 1 of the Appendix. Let denote the set of finite subsets of Γ. We say that a mapping Φ is an interaction on Γ if with the property that for every . If Φ is an interaction on Γ, we write that if and only if
A basic consequence of (A38) is that for all x, y ∈ Γ,
a. Estimates based on distance
We first provide a basic F-norm estimate based on summing interaction terms whose distance from a specific set is given. Recall that if X ⊂ Γ and n ≥ 0, the set X(n) ⊂ Γ is defined as
, for all t ∈ I;
The map is continuous in the strong operator topology.
Moreover, we will write if , for all t ∈ I, and ∥Φ(t)∥F, which we sometimes write as ∥Φ∥F(t), is a locally bounded function of t.
We now state a corollary of Proposition A.4.
b. Estimates based on diameter
In some situations, it is more convenient to form decay arguments based on the diameters of sets, rather than the distance between sets. For these cases, we will further assume that (Γ, d) is ν-regular so that (A13) holds.
Before we state our first result, let us introduce some convenient notation. Our estimates will often be in terms of moments of certain decay functions. To this end, let G : [0, ∞) → (0, ∞) be a decay function. For any p ≥ 0 and each m ≥ 0, set
whenever the sum on the right-hand-side above is finite. We will refer to as the p-th moment of G. The notation
will be used for iterated moments. A rough estimate involving exchanging the order of the summations shows
We now state our first result, compare with Proposition A.4.
When considering a weighted F-function F(r) = e−g(r)F0(r), one can often use the weight with (A60) to prove that the pth moment of Φ has a finite F-norm. This allows us to apply the Lieb-Robinson bound theory to Φp.