This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\newref

conname = Conjecture  \newrefpropname = Proposition  \newrefdefname = Definition  \newrefsecname = Section  \newrefsubname = Section  \newrefthmname = Theorem  \newreflemname = Lemma  \newrefcorname = Corollary  \newreffigname = Figure  \newrefremname = Remark 

Boundary effects and the stability of the low energy spectrum of the AKLT model

Simone Del Vecchio111Dipartimento di Matematica, Università degli Studi di Bari, Italy / email: [email protected] , Jürg Fröhlich222Institut für Theoretiche Physik, ETH-Zürich , Switzerland / email: [email protected] , Alessandro Pizzo333Dipartimento di Matematica, Università di Roma “Tor Vergata”, Italy / email: [email protected] , and Alessio Ranallo 444Section de mathématiques, Université de Genève, Switzerland / email: [email protected]
(29/10/2024)
Abstract

In this paper we study the low-lying spectrum of the AKLT model perturbed by small, finite-range potentials and with open boundary conditions imposed at the edges of the chain. Our analysis is based on the local, iterative Lie Schwinger block-diagonalization method which allows us to control small interaction terms localized near the boundary of the chain that are responsible for the possible splitting of the ground-state energy of the AKLT Hamiltonian into energy levels separated by small gaps. This improves earlier results concerning the persistence of the so called bulk gap in these models, besides illustrating the power of our general methods in a non-trivial application.

1 Introduction

In this paper we study finite-range perturbations of the quantum chain known as the AKLT model, which was introduced and studied in [AKLT]. Our results concern the low-energy spectrum of the perturbed models with so called open boundary conditions imposed at the edges of the chain, as studied in [MN].

The main purpose of our work is to devise a general method allowing us to control effects of small interaction terms localized near the boundary of the chain that entail the splitting of the ground-state energy of the AKLT Hamiltonian into distinct energy levels separated by small gaps. Besides offering a new approach to the study of the low-energy spectrum of Hamiltonians of perturbed AKLT chains, our results improve earlier ones concerning the persistence of the so called bulk gap, (i.e., the gap between the cluster of energy levels corresponding to the four ground-states and the rest of the spectrum of the Hamiltonian separated from these energy levels by a uniformly positive gap).

One of the main purposes of our analysis is to show that the iterative, local Lie-Schwinger block-diagonalization method introduced in [FP] can be applied to small perturbations of Hamiltonians, such as the one of the AKLT model, with a multi-dimensional ground-state subspace. It turns out that, in spite of this complication, such models can be analyzed with the help of a strictly local block-diagonalization method very similar to the one we developed to study quantum chains with a one-dimensional ground-state subspace spanned by a product vector.

The key properties of the models studied in this paper enabling us to apply the methods developed in [FP] are the following ones.

i) The expectations of bulk observables in the four ground-state vectors of the AKLT chain essentially coincide; see Property (1.8), proven in [AKLT], and generalized under the name of LTQO condition in [BHM].

ii) A mechanism, involving so-called Lieb-Robinson bounds, allowing us to treat unperturbed Hamiltonians that are not just sums of on-site terms and yet to use strictly local conjugations as in [FP]. (In the AKLT model the unperturbed Hamiltonian consists of nearest-neighbor interaction terms.)

In this paper, detailed information on the low-lying spectrum is obtained from local control of effective interaction potentials created in the course of our block-diagonalization procedure, including potentials localized near the boundary of the chain.

Our analysis is motivated by recent studies of spectral properties of Hamiltonians appearing in the characterization of “topological phases”; see, e.g., [BN, MZ, BH, BHM, NSY, O1, O2, O3]. Various refinements and extensions of the local Lie-Schwinger block-diagonalization method can be found in [DFPR1, DFPR2, DFPR3, DFP, DFPRa]. Concerning earlier results on small perturbations of the AKLT model it should be mentioned that the first proof of stability of the spectral gap for Hamiltonians with periodic boundary conditions can be traced back to work by Yarotsky [Y], who uses a cluster expansion. This result has also been established in [MZ] by using the spectral flow method. In a paper by Moon and Nachtergaele [MN], the persistence of the bulk gap is established for open boundary conditions by adapting the spectral flow method originally devised for periodic boundary conditions. In more recent papers (see [NSY1], [NSY2]), similar results have been proven for a fairly large class of models of infinite spin chains.

Concerning the AKLT model in higher dimensions, we mention that, in [LSW], the Hamiltonian on the hexagonal lattice has been proven to be gapped. This result has been extended to so called “decorated" lattices (see [AYLLN], [PW1], [PW2]). Stability of the spectral gap against small perturbations has been proven in [LMY] for a class of decorated AKLT models on the hexagonal lattice. We expect that the techniques developed in [DFPR3] can be adapted to treat such models.

1.1 Definition of the Model

To introduce some notation used throughout our paper we recapitulate the definition of the AKLT model and recall its main features.

1.1.1 Definition and properties of the AKLT model

Consider a one-dimensional lattice Λ\Lambda\subset\mathbb{Z} consisting of NN sites. By 𝐒i=(Si1,Si2,Si3){\bf{S}}_{i}=\left(S_{i}^{1},S_{i}^{2},S_{i}^{3}\right) we denote the components of the spin-1 spin operators at the site i{1,,N}i\in\left\{1,\dots,N\right\}. The Hilbert space of the AKLT chain is given by

Λ(N):=j=1Nj,\mathcal{H}_{\Lambda}\equiv\mathcal{H}^{(N)}:=\bigotimes_{j=1}^{N}\mathcal{H}_{j}\,, (1.1)

where, for each jΛj\in\Lambda, j3\mathcal{H}_{j}\simeq\mathbb{C}^{3} is the Hilbert space of the spin-1 (three-dimensional) representation of SU(2)\mathrm{SU}(2). A “local observable” AA is a self-adjoint operator on (N)\mathcal{H}^{(N)} localized in an interval {1,,N}\mathcal{I}\subset\left\{1,\dots,N\right\} (an interval is a subset of Λ\Lambda consisting of successive sites), meaning that

A acts as the identity on jj.A\,\text{ acts as the identity on }\,\,\bigotimes_{j\notin\mathcal{I}}\,\,\mathcal{H}_{j}\,. (1.2)

The interval \mathcal{I} appearing in (1.2) is denoted by supp(A)\text{supp}(A) and called the “support” of AA. The Hamiltonian of the AKLT model is defined by

HΛ0:=12i=1N1[𝐒i𝐒i+1+13(𝐒i𝐒i+1)2+23].H^{0}_{\Lambda}:=\frac{1}{2}\sum\limits_{i=1}^{N-1}[{\bf{S}}_{i}\cdot{\bf{S}}_{i+1}+\frac{1}{3}({\bf{S}}_{i}\cdot{\bf{S}}_{i+1})^{2}+\frac{2}{3}]. (1.3)

This Hamiltonian can be written as  HΛ0=i=1N1Hi,i+1H^{0}_{\Lambda}=\sum\limits_{i=1}^{N-1}H_{i,i+1}, where Hi,i+1:=𝒫i,i+1(2)H_{i,i+1}:=\mathcal{P}^{(2)}_{i,i+1} and

𝒫i,i+1(2):=𝐒i𝐒i+12+(𝐒i𝐒i+1)26+13.\mathcal{P}^{(2)}_{i,i+1}:=\frac{{\bf{S}}_{i}\cdot{\bf{S}}_{i+1}}{2}+\frac{({\bf{S}}_{i}\cdot{\bf{S}}_{i+1})^{2}}{6}+\frac{1}{3}\,. (1.4)

The operator 𝒫i,i+1(2)\mathcal{P}^{(2)}_{i,i+1} is the orthogonal projection onto the subspace of ii+1\mathcal{H}_{i}\otimes\mathcal{H}_{i+1} carrying the spin-22 representation of SU(2)\mathrm{SU}(2) contained in the tensor product of the spin-1 representations with generarors 𝐒i{\bf{S}}_{i} and 𝐒i+1{\bf{S}}_{i+1}.

Next, we recall various important properties of the AKLT model that will be used in the sequel; (see [AKLT] for details and proofs).

  • i)

    For the model with open boundary conditions, the ground-state subspace has dimension 44, independently of the length of the chain. An explicit basis for the ground-state subspace is constructed in [AKLT, Eq. 2.7].

  • ii)

    HΛ0H^{0}_{\Lambda} is frustration free, i.e., {0}Ker(HΛ0)Ker(Hi,i+1),i{1,,N1}\{0\}\neq\text{Ker}(H^{0}_{\Lambda})\subseteq\text{Ker}(H_{i,i+1}),\ \forall i\in\left\{1,\dots,N-1\right\}.

  • iii)

    Let Λ={1,,N}\mathcal{I}\subset\Lambda=\left\{1,\dots,N\right\} be an interval. We define a Hamiltonian H0H^{0}_{\mathcal{I}} associated with \mathcal{I} by

    H0:=i{1,,N1}:i,i+1Hi,i+1,H^{0}_{\mathcal{I}}:=\sum\limits_{i\in\left\{1,\dots,N-1\right\}\,:\,i,i+1\in\mathcal{I}}H_{i,i+1}\,, (1.5)

    and denote by P()P^{(-)}_{\mathcal{I}} the projection onto Ker(H0)\text{Ker}(H^{0}_{\mathcal{I}}), which is a 44-dimensional subspace of :=ii\mathcal{H}_{\mathcal{I}}:=\bigotimes_{i\in\mathcal{I}}\mathcal{H}_{i}; see point i), above. We set P(+):=1P()P^{(+)}_{\mathcal{I}}:=1-P^{(-)}_{\mathcal{I}}, and we denote by tr()\text{tr}_{\mathcal{I}}(\cdot) the normalized trace with respect to the ground-state subspace of H0H^{0}_{\mathcal{I}}; if 𝒥\mathcal{J}\supseteq\mathcal{I} and AA is localized in \mathcal{I} then

    tr𝒥(P𝒥()A)=tr(P()A).\text{tr}_{\mathcal{J}}(P^{(-)}_{\mathcal{J}}A)=\text{tr}_{\mathcal{I}}(P^{(-)}_{\mathcal{I}}A). (1.6)

    Consequently, (1.6) allows us to define the state ω()\omega(\cdot) by

    ω(A):=tr(P()A),\omega(A):=\text{tr}_{\mathcal{I}}(P_{\mathcal{I}}^{(-)}A), (1.7)

    for all operators AA with supp(A)\text{supp}(A)\subset\mathcal{I}.

  • iv)

    For all pairs of intervals 𝒥,\mathcal{J},\mathcal{I}, with 𝒥\mathcal{J}\supseteq\mathcal{I}, the following estimate holds

    P𝒥()(Aω(A))P𝒥()3d(𝒥c,)+1A,Asupported in,𝒥,\|P^{(-)}_{\mathcal{J}}(A-\omega(A))P^{(-)}_{\mathcal{J}}\|\leq 3^{-d(\mathcal{J}^{c},\mathcal{I})+1}\|A\|,\ \forall A\,\text{supported in}\,\mathcal{I},\,\mathcal{J}\supseteq\mathcal{I}, (1.8)

    where d(𝒥c,)d(\mathcal{J}^{c},\mathcal{I}) is the distance of the complement of the set 𝒥\mathcal{J} in Λ\Lambda from the set \mathcal{I}.

Remark 1.1.

A property analogous to iv) is considered in [MN] for a general class of models and referred to as “local topological quantum order” (LTQO) condition (see [BHM]).

One of the results established in [AKLT] on the model described above is that the spectral gap above the ground-state energy, inf spec(HΛ0)=0\text{inf spec}(H^{0}_{\Lambda})=0, is strictly positive, uniformly in the length of the chain.

Theorem 1.2.

[see Theorem 2.1 [AKLT]] There exists an ε>0\varepsilon>0, independent of the length, NN, of the chain such that

(ψ,HΛ0ψ)εψ2,(\psi,H^{0}_{\Lambda}\psi)\geq\varepsilon\|\psi\|^{2}\,, (1.9)

for all ψ\psi belonging to Ker(HΛ0)\text{Ker}(H^{0}_{\Lambda})^{\perp}.

An important ingredient of the mechanism used to analyze this model (alluded to in point 2, at the beginning of Section 1) is a Lieb-Robison bound, which we recall next.

1.1.2 Lieb-Robinson bounds

For the AKLT model, a Lieb-Robinson bound (see [LR]) on the propagation speed of “observables” in the Heisenberg picture has been proven in [NS]. Using the same notation as in [NS], we consider a one-parameter family of functions, FaF_{a}, defined by

Fa:[0,)(0,),Fa(r)=earer1(1+r)3,a0,F_{a}:[0,\infty)\rightarrow(0,\infty)\quad,\quad F_{a}(r)=e^{-ar}e^{-\sqrt{r}}\frac{1}{(1+r)^{3}}\,,\quad a\geq 0\,,

which belong to the class of so-called \mathcal{F}-functions defined in [NS]; namely they have the properties

  • Fa:=i+Fa(i)<\|F_{a}\|:=\sum_{i\in\mathbb{Z}^{+}}F_{a}(i)<\infty;

  • there exists a finite constant Ca>0C_{a}>0 such that, for all i,ji,j\in\mathbb{Z},

    zFa(|iz|)Fa(|zj|)CaFa(|ij|);\sum_{z\in\mathbb{Z}}F_{a}(|i-z|)F_{a}(|z-j|)\leq C_{a}\cdot F_{a}(|i-j|)\,;

see Section 6.1 of [MN]. Let {exp(isH𝒥0)|s}\big{\{}\exp(isH^{0}_{\mathcal{J}})\,\big{|}\,s\in\mathbb{R}\big{\}} be the one-parameter group generated by the Hamiltonian H𝒥0H^{0}_{\mathcal{J}}, 𝒥Λ\mathcal{J}\subseteq\Lambda; then Eq. (16)(16) of [NS] implies that, for two arbitrary observables AA and BB localized in intervals 1,2\mathcal{I}_{1},\mathcal{I}_{2}, respectively,

[exp(isH𝒥0)Aexp(isH𝒥0),B]4ABF0Caea[d(1,2)2ΦaCa|s|a],\|[\exp(isH^{0}_{\mathcal{J}})\,A\,\exp(-isH^{0}_{\mathcal{J}}),B]\|\leq\frac{4\,\|A\|\cdot\|B\|\cdot\|F_{0}\|}{C_{a}}\cdot e^{-a\cdot[d(\mathcal{I}_{1},\mathcal{I}_{2})-\frac{2\,\|\Phi\|_{a}\cdot C_{a}\cdot|s|}{a}]}, (1.10)

where d(1,2)d(\mathcal{I}_{1},\mathcal{I}_{2}) is the distance between the sets 1,2\mathcal{I}_{1},\mathcal{I}_{2} and

Φa=Hi,i+10Fa(1),\|\Phi\|_{a}=\frac{\|H^{0}_{i,i+1}\|}{F_{a}(1)}\,, (1.11)

which, in the context of this paper, is obviously uniformly bounded in aa. In the sequel (see Section 3.2.2) we will set a=1a=1.

1.1.3 Perturbations of the AKLT Hamiltonian

We consider short-range perturbations of HΛ0H^{0}_{\Lambda} given by hermitian matrices acting nontrivially on Hilbert spaces :=jj\mathcal{H}_{\mathcal{I}}:=\bigotimes_{j\in\mathcal{I}}\,\,\mathcal{H}_{j}, where Λ\mathcal{I}\subset\Lambda. In order to keep our exposition as simple as possible, we consider nearest-neighbour interactions denoted by Vi,i+1V_{i,i+1}, which we assume to be uniformly bounded; i.e., without loss of generality,

Vi,i+11.\|V_{i,i+1}\|\leq 1\,. (1.12)

We define a perturbed Hamiltonian, KΛ(t)K_{\Lambda}(t), as the sum of the AKLT Hamiltonian and a perturbation proportional to a real coupling constant tt, namely

KΛ(t):=HΛ0+ti=1N1Vi,i+1.K_{\Lambda}(t):=H_{\Lambda}^{0}+t\sum_{i=1}^{N-1}\,V_{i,i+1}\,. (1.13)

In our proofs we may and will choose tt to be non-negative.

1.2 Main Result

Our main result is the following theorem proven in Section 3 (see Theorem 3.4).
Theorem. There exists some constant t¯>0\bar{t}>0 independent of the number NN of sites in Λ\Lambda such that, for any real coupling constant tt with |t|<t¯|t|<\bar{t} and for all 1<N<1<N<\infty,

  1. (i)

    the spectrum of KΛ(t)K_{\Lambda}(t) is contained in two disjoint, tt-dependent regions σ+\sigma^{+} and σ\sigma^{-} separated by a gap ΔΛ(t)ε4\Delta_{\Lambda}(t)\geq\frac{\varepsilon}{4}, with ε\varepsilon independent of NN, as specified in Theorem 1.2; i.e., EE′′>ΔΛ(t)E^{\prime}-E^{\prime\prime}>\Delta_{\Lambda}(t), for all Eσ+E^{\prime}\in\sigma^{+} and all E′′σE^{\prime\prime}\in\sigma^{-};

  2. (ii)

    for any d[1,N2)d\in\mathbb{N}\cap[1\,,\,\frac{N}{2}), the eigenspace corresponding to the eigenvalues contained in σ\sigma^{-} is four-dimensional; the gaps between the eigenvalues in σ\sigma^{-} coincide with the gaps between the eigenvalues of the symmetric matrix

    PΛ()(ti=1dVi,i+1+ti=NdN1Vi,i+1)PΛ(),P^{(-)}_{\Lambda}\,\Big{(}\,t\sum_{i=1}^{d}\,V_{i,i+1}+\,t\,\sum_{i=N-d}^{N-1}V_{i,i+1}\Big{)}\,P^{(-)}_{\Lambda}\,, (1.14)

    up to corrections bounded by

    |t|3(d1)+o(|t|).|t|\cdot 3^{-(d-1)}\,+\,o(|t|)\,.
Remark 1.3.

One can see by considering a simple example that the small gaps due to interactions localized near the boundaries are typically of order 𝒪(|t|)\mathcal{O}(|t|). As an interaction we take one component (e.g., the zz-component) of the spin operator at the first site, multiplied by tt. The eigenvalues of the matrix PΛ()tS1zPΛ()P^{(-)}_{\Lambda}\,t\,S^{z}_{1}\,P^{(-)}_{\Lambda} splits into two groups of two eigenvalues each, separated by a gap given by 4|t|3\approx\frac{4\cdot|t|}{3}, up to corrections exponentially small in the length of the chain.

Remark 1.4.

We wish to highlight the effectiveness for explicit computations of the mathematically rigorous formula in result (ii) above, which reduces the problem of estimating the eigenvalue splitting to a leading-order calculation in (formal) Rayleigh-Schrödinger perturbation theory, i.e., to calculating matrix elements of bare potentials in the four-dimensional ground state subspace of the AKLT Hamiltonian. Using the “indistinguishability of the ground-state vectors" (see (1.8)) we can neglect all the bare interaction terms located sufficiently far from the end-points of the chain. This implies that, by keeping only the dd interactions closest to the left- and right end-points, respectively, an error in the values of (small) gaps in the energy spectrum of the perturbed AKLT Hamiltonian results that is bounded above by |t|3(d1)|t|\cdot 3^{-(d-1)}. Notice that, already for d=10d=10, the factor 3(d1) 51053^{-(d-1)}\,\approx\,5\cdot 10^{-5} is tiny, and only 2020 potentials have to be kept in the sums shown above.

Notation

1) The symbol “\subset" denotes a strict inclusion of sets; otherwise the symbol “\subseteq" is used.

2) The symbol {i}\mathcal{I}\cup\{i\} indicates a union of sets of sites of the microscopic lattice.

Acknowledgements. A.P. acknowledges support through the MIUR Excellence Department Project awarded to the Department of Mathematics, University of Rome Tor Vergata, CUP E83C18000100006, and also support through GNFM - INDAM.

2 The Block-Diagonalization Algorithm

In this section we describe some important elements of our method of proving the main result, namely an iterative local block-diagonalization of the Hamiltonians KΛ(t)K_{\Lambda}(t). This method has been developed in several previous papers referred to below, starting with [FP]. The original method devised in that paper cannot be applied directly to the Hamiltonians studied in the present paper, for the following reasons:

  • i)

    The unperturbed Hamiltonian HΛ0H^{0}_{\Lambda} (see (1.3)) is not ultralocal; rather, it is a collection of nearest-neighbour interaction terms 𝒫i,i+1(2)\mathcal{P}^{(2)}_{i,i+1} (see (1.4)).

  • ii)

    The ground-state subspace of a Hamiltonian H0H^{0}_{\mathcal{I}} associated with an arbitrary interval Λ\mathcal{I}\subset\Lambda is not one-dimensional; indeed, it is of dimension four, with a basis of matrix product states described in [AKLT] and enjoying the properties listed in items i)-iv) of Section 1.1.1, above.

2.1 Coarse graining

In order to construct the key ingredients of our analysis, namely local Lie-Schwinger conjugations serving to block-diagonalize the Hamiltonians KΛ(t)K_{\Lambda}(t), the perturbation must be split into terms localized in NN-independent intervals. Without loss of generality, we assume that t>0t>0 and that

(N1)t,t1.(N-1)\,\sqrt{t}\quad,\quad\sqrt{t^{-1}}\quad\in\mathbb{N}\,. (2.1)

The perturbation will be split into terms V1,JV_{\mathcal{I}_{1,J}} supported in intervals 1,J\mathcal{I}_{1,J} containing a number of sites approximatively equal to t1\sqrt{t^{-1}} and belonging to a family \mathfrak{I} of intervals introduced in Definition 2.2, below.

Remark 2.1.

We stress that, for all chains of length smaller than t1\sqrt{t^{-1}}, i.e., (N1)t<1(N-1)\cdot\sqrt{t}<1, one is able to block-diagonalize the Hamiltonian in one shot, using standard perturbation theory (in the form of Lie-Schwinger conjugations), provided t<t¯t<\overline{t}, with t¯>0\overline{t}>0 – iterations are not needed. Actually, for a fixed value of t<t¯t<\bar{t}, the smaller the size of the chain the faster is the convergence of the perturbative series. (Thus, the length of the chain does not imply any lower bound on the size of the coupling constant tt, as one might have guessed mistakenly.)

It will be convenient to think of a macroscopic (finite) lattice with left endpoint X=1X=1, right endpoint X=NX=N, and lattice spacing t1\sqrt{t^{-1}}. The MthM^{th} site of this lattice is the point

1+(M1)t1,with1M(N1)t+1,1+(M-1)\sqrt{t^{-1}}\,,\quad\text{with}\quad 1\leq M\leq(N-1)\sqrt{t}+1\,, (2.2)

of the microscopic lattice Λ\Lambda. The set K,J\mathcal{I}_{K,J} is the interval (i.e., a subset of Λ\Lambda consisting of successive sites) whose endpoints coincide555In general, if the range, κ\kappa, of the interaction terms is larger than 11, the intervals K,J\mathcal{I}_{K,J} are defined in such a way that they overlap, i.e., the endpoints coincide with the sites M=JM=J and M=J+KM=J+K only up to corrections depending on κ\kappa and, consequently, Kt1K\cdot\sqrt{t^{-1}} is its length up to corrections of the order the step length, 11, of the microscopic lattice where the model is defined. with the sites M=JM=J and M=J+KM=J+K of the macroscopic lattice. Notice that it can be helpful to think of the sets K,J\mathcal{I}_{K,J} (and of some enlargements of these sets defined later on) as intervals contained in the real line; for examples, see Figures 1, 2, and 3, which are intended to display the overlap between such sets. As an interval of the real line, K,J\mathcal{I}_{K,J} has length KK in units of t1\sqrt{t^{-1}}.

Definition 2.2.

The elements of the set \mathfrak{I} are the intervals K,J\mathcal{I}_{K,J} (see Fig. 1), where

K,J:={i:i[1,N][ 1+(J1)t1, 1+(J1+K)t1]},\mathcal{I}_{K,J}:=\{i\in\mathbb{N}:i\in[1,N]\cap[\,1+(J-1)\sqrt{t^{-1}}\,,\,1+(J-1+K)\sqrt{t^{-1}}\,]\}\,, (2.3)

with K,JK,J\in\mathbb{N} such that 1+(J1+K)t1N1+(J-1+K)\sqrt{t^{-1}}\leq N. Thus the length, |K,J||\mathcal{I}_{K,J}|, of K,J\mathcal{I}_{K,J} is Kt1K\cdot\sqrt{t^{-1}}.

Remark 2.3.

It follows from the above definitions that the set \mathfrak{I} is closed under taking the union of two overlapping elements.

In the following it will be useful to introduce an ordering relation amongst the intervals labeled by the pairs (K,J)(K,J) with the property that shorter intervals precede longer ones. This relation is specified as follows.

Definition 2.4.

The following defines an ordering relation among the pairs (K,Q)(K,Q) labelling the elements of the set \mathfrak{I} (which will be used in this paper):

(K,Q)(K,Q)ifK>K,or,in caseK=K,ifQ>Q.(K,Q)\succ(K^{\prime},Q^{\prime})\quad\text{if}\quad K>K^{\prime},\quad\text{or},\quad\text{in case}\quad K=K^{\prime},\quad\text{if}\quad Q>Q^{\prime}\,. (2.4)

The symbol (K,Q)1(K,Q)_{\mp 1} labels the pair preceding/succeeding (K,Q)(K,Q), respectively, in the ordering relation of Definition 2.4. For convenience we shall denote the pair preceding (1,1)(1,1) by (0,N)(0,N). The last pair is ((N1)t,1)((N-1)\cdot\sqrt{t},1).

The interval 1,J\mathcal{I}_{1,J} is the support of the operator

i:i,i+11,JVi,i+1.\sum_{i\,:\,i,i+1\,\in\mathcal{I}_{1,J}}\,V_{i,i+1}\,. (2.5)

Thanks to (1.12) and to our definition of the size, |1,J||\mathcal{I}_{1,J}|, of the interval 1,J\mathcal{I}_{1,J}, namely |1,J|=t1|\mathcal{I}_{1,J}|=\sqrt{t^{-1}}, the following operator norm estimate holds,

i:i,i+11,JVi,i+1t1.\|\,\sum_{i\,:\,i,i+1\,\in\mathcal{I}_{1,J}}\,V_{i,i+1}\,\|\leq\sqrt{t^{-1}}\,. (2.6)

After dividing them by t1\sqrt{t^{-1}}, we denote such a collection of potentials by V1,JV_{\mathcal{I}_{1,J}}; i.e.,

V1,J:=1t1i:i,i+11,JVi,i+1,V_{\mathcal{I}_{1,J}}:=\frac{1}{\sqrt{t^{-1}}}\,\sum_{i\,:\,i,i+1\,\in\mathcal{I}_{1,J}}\,V_{i,i+1}\,, (2.7)

the support of V1,JV_{\mathcal{I}_{1,J}} being 1,J\mathcal{I}_{1,J}. The observation in (2.6) implies that V1,J1\|\,V_{\mathcal{I}_{1,J}}\,\|\leq 1. In order to implement the block-diagonalization procedure, it is convenient to re-write the Hamiltonian KΛ(t)K_{\Lambda}(t) using these definitions; i.e.,

KΛ(t)=HΛ0+t1,JΛV1,J.K_{\Lambda}(t)=H_{\Lambda}^{0}+\sqrt{t}\,\sum_{\mathcal{I}_{1,J}\subset\Lambda}\,V_{\mathcal{I}_{1,J}}\,. (2.8)

The block-diagonalization is based on spectral projections, PK,Q(±)P_{\mathcal{I}_{K,Q}}^{(\pm)}, associated with intervals K,Q\mathcal{I}_{K,Q}, which we define next.

Definition 2.5.

By P()P^{(-)}_{\mathcal{I}} we denote the orthogonal projection onto the ground-state subspace of H0H^{0}_{\mathcal{I}}, and we define

P(+):=1P().P_{\mathcal{I}}^{(+)}:=1-P^{(-)}_{\mathcal{I}}\,. (2.9)

We will require analogous definitions of projections associated with general subsets of the lattice Λ\Lambda.

2.2 Recap of the method for ultralocal unperturbed Hamiltonians

In order to explain how the method in [FP] has to be modified because of specific features of the AKLT model (as compared to the models with ultralocal Hamiltonians treated in [FP]), we first observe that the procedure presented in that reference is based on an iterative block-diagonalization of the perturbing potentials in the Hamiltonian of those models involving Lie-Schwinger conjugations, assuming that the coupling constant, t(=|t|)t(=|t|), of the perturbation is sufficiently small. In this paper, too, the block-diagonalization is implemented with the help of local Lie-Schwinger conjugations; and “local” means that each conjugation involves only operators supported in an interval 1,J\mathcal{I}_{1,J} (of length 1 and with left endpoint in JJ in the macroscopic lattice). More precisely, in the notations of Section 1.1, the local Hamiltonian supported in the interval 1,J\mathcal{I}_{1,J} is conjugated by a suitable local unitary operator. Specifically,

H1,J0+tV1,J,H^{0}_{\mathcal{I}_{1,J}}+\sqrt{t}\,V_{\mathcal{I}_{1,J}}\,, (2.10)

(where H1,J0H^{0}_{\mathcal{I}_{1,J}} is defined in (1.5)) is conjugated by a suitably defined unitary operator eZ1,Je^{Z_{\mathcal{I}_{1,J}}},

eZ1,J(H1,J0+tV1,J)eZ1,J=H1,J0+tV1,J,e^{Z_{\mathcal{I}_{1,J}}}(H^{0}_{\mathcal{I}_{1,J}}+\sqrt{t}\,V_{\mathcal{I}_{1,J}})e^{-Z_{\mathcal{I}_{1,J}}}=H^{0}_{\mathcal{I}_{1,J}}+\sqrt{t}\,V^{\prime}_{\mathcal{I}_{1,J}}\,, (2.11)

with the purpose to render the new potentials V1,JV1,J(t)V^{\prime}_{\mathcal{I}_{1,J}}\equiv V^{\prime}_{\mathcal{I}_{1,J}}(t) block-diagonal with respect to the projections P1,J(),P1,J(+)P^{(-)}_{\mathcal{I}_{1,J}}\,,\,P^{(+)}_{\mathcal{I}_{1,J}}; see Definition 2.5.

Obviously new effective interaction potentials are created as a byproduct of the block diagonalization of the potentials V1,JV_{\mathcal{I}_{1,J}}. Such new potentials are supported in intervals given by connected unions of intervals 1,Q\mathcal{I}_{1,Q}. Hence, in general, a sequence of further conjugations of the Hamiltonian KΛ(t)K_{\Lambda}(t) must be introduced in order to block-diagonalize the effective interactions created in previous steps, which are supported in ever larger intervals K,Q\mathcal{I}_{K,Q}; (see Definition 2.2).

In the algorithm designed in [FP], the steps of the block-diagonalization are indexed by pairs (K,Q)(K,Q) labelling the intervals K,Q\mathcal{I}_{K,Q} (for which we have introduced an ordering relation in Definition 2.4); that is, in step (K,Q)(K,Q), the potential, VK,Q(K,Q)1V_{\mathcal{I}_{K,Q}}^{\,(K,Q)_{-1}} – which is the potential obtained in the previous step (i.e., in step (K,Q)1(K,Q)_{-1}) and is supported in K,Q\mathcal{I}_{K,Q} – gets block-diagonalized. In the process new terms, given by

n=11n!adnZK,Q(VK,Q(K,Q)1),\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}_{K,Q}}(V_{\mathcal{I}_{K^{\prime},Q^{\prime}}}^{\,(K,Q)_{-1}})\,, (2.12)

(where adad stands for adjoint action; check (2.40) for its definition) are created that contribute to new interaction potentials, VK,QK,Q(K,Q)V^{(K,Q)}_{\mathcal{I}_{K^{\prime},Q^{\prime}}\cup\mathcal{I}_{K,Q}}, supported in larger intervals K,QK,Q\mathcal{I}_{K^{\prime},Q^{\prime}}\cup\mathcal{I}_{K,Q}. All the terms created by the block-diagonalization procedure with support in the interval K′′,Q′′:=K,QK,Q\mathcal{I}_{K^{\prime\prime},Q^{\prime\prime}}:=\mathcal{I}_{K^{\prime},Q^{\prime}}\cup\mathcal{I}_{K,Q} are lumped together. To control the size of VK′′,Q′′V_{\mathcal{I}_{K^{\prime\prime},Q^{\prime\prime}}} one has to count certain growth processes (of intervals) yielding a given interval K′′,Q′′:=K,QK,Q\mathcal{I}_{K^{\prime\prime},Q^{\prime\prime}}:=\mathcal{I}_{K^{\prime},Q^{\prime}}\cup\mathcal{I}_{K,Q}. The number of such growth processes can easily be estimated to be at most exponential, i.e., to be bounded above by CK′′C^{K^{\prime\prime}}, for some universal constant C>1C>1. Since

(2.12)𝒪(tVK,Q(K,Q)1VK,Q(K,Q)1),\|(\ref{new-term})\|\leq\mathcal{O}(\sqrt{t}\cdot\|V_{\mathcal{I}_{K,Q}}^{\,(K,Q)_{-1}}\|\cdot\|V_{\mathcal{I}_{K^{\prime},Q^{\prime}}}^{\,(K,Q)_{-1}}\|\,)\,, (2.13)

it is then quite easy to inductively prove a bound of the type

VK′′,Q′′(K,Q)|t|ρ(K′′1),\|V_{\mathcal{I}_{K^{\prime\prime},Q^{\prime\prime}}}^{\,(K,Q)}\|\leq|t|^{\,\rho\cdot(K^{\prime\prime}-1)}\,, (2.14)

for some constant ρ\rho, with 0<ρ<120<\rho<\frac{1}{2}, provided that |t||t| is small enough, uniformly in the number NN of sites of the chain. In the following, we will always assume (w.l.o.g.) that t0t\geq 0, and our results will hold under the assumption that t<t¯t<\bar{t}, for some constant t¯\bar{t} independent of N(=|Λ|)N(=|\Lambda|).

Remark 2.6.

In this paper, the term “step” can have two different meanings; namely

  1. 1.

    it can be a label of Hamiltonians and potentials defined in the course of the block-diagonalization procedure: KΛ(K,Q)(t)K_{\Lambda}^{(K,Q)}(t) is the Hamiltonian created in step (K,Q)(K,Q) of the block- diagonalization procedure;

  2. 2.

    it can mean the iteration step from (K,Q)1(K,Q)_{-1} to (K,Q)(K,Q), (i.e., from a certain level (K,Q)1(K,Q)_{-1} to the next one, (K,Q)(K,Q)) in the block-diagonalization procedure.

2.3 Modifications of the procedure for AKLT-type models

Before describing the structure of the Hamiltonian obtained in each step of the block-diagonali-zation procedure (see the definitions contained in Section 2.4), we discuss some new ingredients incorporated into the algorithm (see Section 2.5) yielding the new potentials in each step of the block-diagonalization. The need for new ingredients becomes apparent already in the steps performed to block-diagonalize the bare potentials V1,JV_{\mathcal{I}_{1,J}} . The conjugation of the Hamiltonian KΛK_{\Lambda} by the unitary operator eZ1,Je^{Z_{\mathcal{I}_{1,J}}}, 1<J<Nt+11<J<N\cdot\sqrt{t}+1, has the effect to not only “hook up” to bare interaction potentials, for example tV1,J1\sqrt{t}\,V_{\mathcal{I}_{1,J-1}} and tV1,J+1\sqrt{t}\,V_{\mathcal{I}_{1,J+1}}, but to also “hook up” to terms of the unperturbed Hamiltonians H1,J10,H1,J+10H^{0}_{\mathcal{I}_{1,J-1}}\,,\,H^{0}_{\mathcal{I}_{1,J+1}}, namely to the two projections

𝒫i1,i(2),𝒫i+,i++1(2),\mathcal{P}^{(2)}_{i_{-}-1,i_{-}}\quad,\quad\mathcal{P}^{(2)}_{i_{+},i_{+}+1}\,, (2.15)

where ii_{-} and i+i_{+} are the sites of the microscopic lattice corresponding to the endpoints of the interval 1,J\mathcal{I}_{1,J}; hence, in the conjugation, 𝒫i1,i(2)\mathcal{P}^{(2)}_{i_{-}-1,i_{-}} and 𝒫i+,i++1(2)\mathcal{P}^{(2)}_{i_{+},i_{+}+1} (that do not belong to the local Hamiltonian H1,J0H^{0}_{\mathcal{I}_{1,J}}) get “hooked up" to other terms. Indeed, following the strategy of [FP], we should define an anti-symmetric matrix Z1,JZ_{\mathcal{I}_{1,J}} in order to block-diagonalize the interaction potential V1,JV_{\mathcal{I}_{1,J}} and observe that, in the course of the conjugation generated by Z1,JZ_{\mathcal{I}_{1,J}}, new terms of the type

n=11n!adnZ1,J(𝒫i,i+1(2))\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}_{1,J}}(\mathcal{P}^{(2)}_{i,i+1})\, (2.16)

are created whenever

{i,i+1}1,Jand{i,i+1}1,J.\{i\,,\,i+1\}\nsubset\mathcal{I}_{1,J}\,\quad\text{and}\quad\{i\,,\,i+1\}\cap\mathcal{I}_{1,J}\neq\emptyset\,. (2.17)

We refer to this process in the conjugation as a “hooking” of 𝒫i,i+1(2)\mathcal{P}^{(2)}_{i,i+1} terms.

We warn the reader that the conjugation used in the block-diagonalization step (1,J)(1,J) is generated by an anti-symmetric matrix Z1,JZ_{\mathcal{I}^{*}_{1,J}} supported in a somewhat larger interval 1,J1,J\mathcal{I}_{1,J}^{*}\supset\mathcal{I}_{1,J}. In this informal description, we attempt to explain the problems that would arise in the block-diagonalization if the matrix Z1,JZ_{\mathcal{I}_{1,J}} supported in the interval 1,J\mathcal{I}_{1,J} were used.

The new interaction terms (2.16) show some important differences as compared to the operators in (2.12):

  1. 1.

    Although the support of the new term displayed in (2.16) coincides with 1,J\mathcal{I}_{1,J}, up to a single site, the control of its norm is quite difficult, since the counterpart of (2.12) is

    n=11n!adnZ1,J(𝒫i,i+1(2)t),\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}_{1,J}}(\frac{\mathcal{P}^{(2)}_{i,i+1}}{\sqrt{t}})\,, (2.18)

    which cannot be estimated, in a manner similar to (2.13), in terms of

    𝒪(tV1,J(1,J)1𝒫i,i+1(2)t);\mathcal{O}(\sqrt{t}\cdot\|V_{\mathcal{I}_{1,J}}^{\,(1,J)_{-1}}\|\cdot\|\frac{\mathcal{P}^{(2)}_{i,i+1}}{\sqrt{t}}\|\,)\,;

    indeed this might appear to make it impossible to prove an inductive estimate as in (2.14).

  2. 2.

    Unless a term supported in K,Q{i,i+1}\mathcal{I}_{K,Q}\cup\{i,i+1\} is already block-diagonal, it should be lumped to the effective potential VK,Q(K,Q)V^{(K,Q)}_{\mathcal{I}_{K^{\prime},Q^{\prime}}}, for some interval K,Q\mathcal{I}_{K^{\prime},Q^{\prime}} with

    K,QK,Q{i,i+1}.\mathcal{I}_{K^{\prime},Q^{\prime}}\supset\mathcal{I}_{K,Q}\cup\{i,i+1\}\,.

The complications described here force us to modify the method proposed in [FP]: in the present paper, the strict locality of the on-site operators studied in that paper is given up and replaced by a locality property expressed in terms of decay properties of the Green functions of the local Hamiltonians H0H^{0}_{\mathcal{I}}, or, equivalently, by Lieb-Robinson bounds associated with the one-parameter groups generated by the operators H0H^{0}_{\mathcal{I}}. Locality is exploited in a careful study of the operator in (2.16), but associated with an enlarged interval 1,J1,J\mathcal{I}^{*}_{1,J}\supset\mathcal{I}_{1,J} introduced in Definition 2.8, below. Thus, in order to block-diagonalize an effective potential supported in K,Q\mathcal{I}_{K,Q}, we shall use a local unperturbed Hamiltonian with support in a larger interval K,QK,Q\mathcal{I}^{*}_{K,Q}\supset\mathcal{I}_{K,Q}. The Lieb-Robinson bound in (1.10) will be used to show that the off-diagonal part of the operator

adZ1,J(𝒫i,i+1(2)t)=[Z1,J,𝒫i,i+1(2)t]ad\,Z_{\mathcal{I}^{*}_{1,J}}(\frac{\mathcal{P}^{(2)}_{i,i+1}}{\sqrt{t}})=[Z_{\mathcal{I}^{*}_{1,J}}\,,\,\frac{\mathcal{P}^{(2)}_{i,i+1}}{\sqrt{t}}] (2.19)

w.r.t. to spectral projections associated with the enlarged interval 1,J¯\overline{\mathcal{I}^{*}_{1,J}} (introduced in Definition 2.10),

P1,J¯(),P1,J¯(+):=1P1,J¯(),P^{(-)}_{\overline{\mathcal{I}^{*}_{1,J}}}\quad,\quad P^{(+)}_{\overline{\mathcal{I}^{*}_{1,J}}}:=1-P^{(-)}_{\overline{\mathcal{I}^{*}_{1,J}}}\,, (2.20)

has a norm that decays in tt at least as fast as

𝒪(tV1,J(1,J)1).\mathcal{O}(\sqrt{t}\cdot\|V_{\mathcal{I}_{1,J}}^{\,(1,J)_{-1}}\|\,)\,.

Here one uses that the distance between K,Q\mathcal{I}_{K,Q} and the endpoints of K,Q\mathcal{I}^{*}_{K,Q} is of order t1\sqrt{t^{-1}}. This enables us to lump this term, as well as the terms corresponding to n2n\geq 2 in the expression

n=11n!adnZ1,J(𝒫i,i+1(2)t),\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{1,J}}(\frac{\mathcal{P}^{(2)}_{i,i+1}}{\sqrt{t}})\,, (2.21)

together with a potential term supported in a larger interval containing 1,J\mathcal{I}^{*}_{1,J}, which will be block-diagonalized in a subsequent step. As for the diagonal part of the operator in (2.19), no extra power of tt is gained from the argument based on the Lieb-Robinson bounds; but this is not a problem, because this term does not need to be block-diagonalized anymore.

Thanks to the property in (1.8), the use of enlarged intervals also solves a problem666In the following we only try to convey the main ideas underlying our modification of the block-diagonalization procedure. related to the degeneracy of the ground-state eigenvalue of Hamiltonians of the type H0H^{0}_{\mathcal{I}}: the new potential, which we will denote by VK,Q¯(K,Q)V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{K,Q}}}, is supported in the enlarged interval K,Q¯\overline{\mathcal{I}^{*}_{K,Q}} after the block-diagonalization of the potential VK,Q(K,Q)1V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}} and is not given by the full-fledged Lie-Schwinger series (see (2.32)) associated with the conjugation eZK,Qe^{Z_{\mathcal{I}^{*}_{K,Q}}}. To be more explicit, the potential VK,Q¯(K,Q)V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{K,Q}}} will contain the following contributions:

  • i)

    The expression

    ω(VK,Q(K,Q)1)+PK,Q¯(+)PK,Q(+)[VK,Q(K,Q)1ω(VK,Q(K,Q)1)]PK,Q(+)PK,Q¯(+),\omega(V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}})+P^{(+)}_{\overline{\mathcal{I}^{*}_{K,Q}}}\,P^{(+)}_{\mathcal{I}^{*}_{K,Q}}\,\,\Big{[}V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}-\omega(V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}})\Big{]}\,P^{(+)}_{\mathcal{I}^{*}_{K,Q}}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{K,Q}}}\,, (2.22)

    which originates in the zero-order term in the Lie Schwinger series, i.e., in

    PK,Q()VK,Q(K,Q)1PK,Q()+PK,Q(+)VK,Q(K,Q)1PK,Q(+),P^{(-)}_{\mathcal{I}^{*}_{K,Q}}\,\,V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\,P^{(-)}_{\mathcal{I}^{*}_{K,Q}}+P^{(+)}_{\mathcal{I}^{*}_{K,Q}}\,\,V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\,P^{(+)}_{\mathcal{I}^{*}_{K,Q}}\,, (2.23)

    from which we extract

    ω(VK,Q(K,Q)1)(1PK,Q(+))+PK,Q(+)VK,Q(K,Q)1PK,Q(+)\omega(V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}})\,\Big{(}1-P^{(+)}_{\mathcal{I}^{*}_{K,Q}}\Big{)}+P^{(+)}_{\mathcal{I}^{*}_{K,Q}}\,\,V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\,P^{(+)}_{\mathcal{I}^{*}_{K,Q}} (2.24)

    and neglect a remainder whose norm decays exponentially in the distance, 𝒪(t1)\mathcal{O}(\sqrt{t^{-1}}), between K,Q\mathcal{I}_{K,Q} and the endpoints of K,Q\mathcal{I}^{*}_{K,Q}; see (1.8). This remainder term and the higher-order terms in the Lie-Schwinger series are treated as perturbations and lumped together with a potential, supported in a larger interval, that will be block-diagonalized in a later step.

  • ii)

    The diagonal part w.r.t. PK,Q¯(),PK,Q¯(+)P^{(-)}_{\overline{\mathcal{I}^{*}_{K,Q}}},P^{(+)}_{\overline{\mathcal{I}^{*}_{K,Q}}} proportional to the projections 𝒫i,i+1(2)\mathcal{P}^{(2)}_{i,i+1}, hence of terms of the type in (2.19).

Refer to caption
Figure 1: The overlapping of the intervals 1,J\mathcal{I}_{1,J} and 1,J+1\mathcal{I}_{1,J+1} for 2J(N1)t22\leq J\leq(N-1)\sqrt{t}-2.

In replacing (2.23) by (2.24) we must exclude from the block-diagonalization steps all intervals touching the endpoints of the lattice Λ\Lambda. Therefore, we must henceforth distinguish bulk- from boundary-potentials, as explained in Section 2.4 below. The boundary potentials will get block-diagonalized only in the last step that corresponds to the interval Λ\Lambda (given by the entire chain).

Remark 2.7.

In all steps of the block-diagonalization except the last one, the local Hamiltonians have a 4-fold degenerate ground-state energy. The control of the so-called “bulk gap” is however similar to the one used when considering a Hamiltonian with a non-degenerate ground-state energy.

2.4 Enlarged intervals and unitary conjugations

We begin this subsection by introducing enlarged intervals that will be needed in our procedure, as explained in Section 2.3; see also Figures 2 and 3.

Definition 2.8.

\mathfrak{I}^{*} is the set of intervals whose elements are the intervals K,Q\mathcal{I}^{*}_{K,Q} defined by

K,J:={i:i[1,N][ 1+(J43)t1, 1+(J23+K)t1]},\mathcal{I}^{*}_{K,J}:=\{i\in\mathbb{N}:i\in[1,N]\cap[\,1+(J-\frac{4}{3})\sqrt{t^{-1}},\,1+(J-\frac{2}{3}+K)\sqrt{t^{-1}}\,]\}\,, (2.25)

with K,JK,J such that K,J\mathcal{I}_{K,J}\in\mathfrak{I}.

Refer to caption
Figure 2: The interval 1,J\mathcal{I}_{1,J} and its enlargement 1,J\mathcal{I}^{\ast}_{1,J}.
Definition 2.9.

With each interval K,Q\mathcal{I}^{*}_{K,Q}\in\mathfrak{I}^{*} we associate an interval ~K,Q\widetilde{\mathcal{I}}^{*}_{K,Q}\in\mathfrak{I} defined as the smallest interval of type K,Q\mathcal{I}_{K^{\prime},Q^{\prime}} containing the interval K,Q\mathcal{I}^{*}_{K,Q}.

Definition 2.10.

With each interval K,Q\mathcal{I}^{*}_{K,Q}\in\mathfrak{I}^{*} we associate an interval K,Q¯\overline{\mathcal{I}^{*}_{K,Q}} defined as the interval obtained from K,Q\mathcal{I}^{*}_{K,Q} by including (if present) the two sites in the microscopic lattice, nearest to K,Q\mathcal{I}^{*}_{K,Q}, one on the right and one on the left.

In order to implement the block-diagonalization steps, we define two subsets of the set \mathfrak{I} of intervals introduced in Definition 2.2 of Section 1.1.3:

bulk:={K,J:1,NK,J}\displaystyle\mathfrak{I}_{\text{bulk}}:=\{\mathcal{I}_{K,J}\in\mathfrak{I}:1,N\notin\mathcal{I}_{K,J}\} (2.26)
b.ry:={K,J:1K,J or NK,J}.\displaystyle\mathfrak{I}_{\text{b.ry}}:=\{\mathcal{I}_{K,J}\in\mathfrak{I}:1\in\mathcal{I}_{K,J}\text{ or }N\in\mathcal{I}_{K,J}\}. (2.27)
Definition 2.11 (Restricted ordering).

The block-diagonalization steps will be associated with intervals K,Qbulk\mathcal{I}_{K,Q}\in\mathfrak{I}_{\text{bulk}}. We will make use of the ordering introduced in Definition 2.4 (Section 1.1.3) restricted to pairs (K,Q)(K,Q) with K,Qbulk\mathcal{I}_{K,Q}\in\mathfrak{I}_{\text{bulk}}. Thus the symbols (K,Q)1(K,Q)_{-1}, (K,Q)+1(K,Q)_{+1} refer to the preceding and the successive element of (K,Q)(K,Q), respectively, with respect to this restricted ordering, i.e., the interval with coordinates (K,Q)1(K,Q)_{-1} or (K,Q)+1(K,Q)_{+1} is required to belong to bulk\mathfrak{I}_{\text{bulk}}.

Using successive unitary conjugations, we shall derive a transformed Hamiltonian that, in step (K,Q)(K,Q), will coincide with the operator

KΛ(K,Q)(t)\displaystyle K_{\Lambda}^{\,(K,Q)}(t) :=\displaystyle:= HΛ0+\displaystyle H^{0}_{\Lambda}+ (2.30)
+tQV1,Q¯(K,Q)++tQ;(K,Q)(K,Q)VK,Q¯(K,Q)+\displaystyle+{\sqrt{t}}\sum_{Q^{\prime}}V^{\,(K,Q)}_{\overline{\mathcal{I}^{*}_{1,Q^{\prime}}}}+\dots+{\sqrt{t}}\sum_{Q^{\prime}\,;\,(K,Q^{\prime})\preceq(K,Q)}V^{\,(K,Q)}_{\overline{\mathcal{I}^{*}_{K,Q^{\prime}}}}+\quad\quad\quad
+tQ;(K,Q)(K,Q)VK,Q(K,Q)+tQVK+1,Q(K,Q)++tV(N1)t2,2(K,Q)+\displaystyle+{\sqrt{t}}\sum_{Q^{\prime}\,;\,(K,Q^{\prime})\succ(K,Q)}V^{\,(K,Q)}_{\mathcal{I}_{K,Q^{\prime}}}+{\sqrt{t}}\sum_{Q^{\prime}}V^{\,(K,Q)}_{\mathcal{I}_{K+1,Q^{\prime}}}+\dots+{\sqrt{t}}V^{\,(K,Q)}_{\mathcal{I}_{(N-1)\cdot\sqrt{t}-2,2}}+
+tQW1,Q(K,Q)++tW(N1)t,1(K,Q),\displaystyle+{\sqrt{t}}\sum_{Q^{\prime}}W^{\,(K,Q)}_{\mathcal{I}_{1,Q^{\prime}}}+\dots+{\sqrt{t}}W^{\,(K,Q)}_{\mathcal{I}_{(N-1)\cdot\sqrt{t},1}}\,,

where the two types of potentials, “V” and “W,” are specified below:

Refer to caption
Figure 3: How an interval 1,J\mathcal{I}_{1,J} relates to 1,J\mathcal{I}^{\ast}_{1,J} and ~1,J\widetilde{\mathcal{I}}^{\ast}_{1,J}.
  • Depending on whether (K,Q)(K,Q)(K^{\prime},Q^{\prime})\preceq(K,Q) or (K,Q)(K,Q)(K^{\prime},Q^{\prime})\succ(K,Q) potentials of type “V” are labeled by intervals K,Q¯\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}} or by intervals K,Q\mathcal{I}_{K^{\prime},Q^{\prime}}, respectively; in both cases K,Qbulk\mathcal{I}_{K^{\prime},Q^{\prime}}\in\mathfrak{I}_{\text{bulk}}. The first type of potentials, i.e., those corresponding to (K,Q)(K,Q)(K^{\prime},Q^{\prime})\preceq(K,Q), are block-diagonalized, and the block-diagonalization is w.r.t. the two projections PK,Q¯()P^{(-)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}, PK,Q¯(+)P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}} (see Definition 2.5); more precisely, they are of the form

    VK,Q¯(K,Q)=PK,Q¯(+)VK,Q¯(K,Q)PK,Q¯(+)+PK,Q¯()VK,Q¯(K,Q)PK,Q¯().V^{\,(K,Q)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}=P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,V^{\,(K^{\prime},Q^{\prime})}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}+P^{(-)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,V^{\,(K^{\prime},Q^{\prime})}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,P^{(-)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,. (2.31)

    It is straightforward to check that they are block-diagonal w.r.t. any pair P(+),P()P^{(+)}_{\mathcal{I}}\,,\,P^{(-)}_{\mathcal{I}} with K,Q\mathcal{I}\supset\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}, due to the frustration free property of HΛ0H^{0}_{\Lambda}.

  • The potentials WK,Q(K,Q)W^{\,(K,Q)}_{\mathcal{I}_{K^{\prime},Q^{\prime}}} are characterized by the property that K,Qb.ry\mathcal{I}_{K^{\prime},Q^{\prime}}\in\mathfrak{I}_{\text{b.ry}}, i.e., they are zero if K,Qb.ry\mathcal{I}_{K^{\prime},Q^{\prime}}\notin\mathfrak{I}_{\text{b.ry}}. They get block-diagonalized only in the very last step.

The way these potentials are produced in each step of the block-diagonalization procedure is explained in Sect. 2.5. The procedure has the property that, in step (K,Q)(K,Q), the potential VK,Q(K,Q)1V^{\,(K,Q)_{-1}}_{\mathcal{I}_{K,Q}} is transformed to a potential VK,Q¯(K,Q)V^{\,(K,Q)}_{\overline{\mathcal{I}^{*}_{K,Q}}} related to the Lie-Schwinger series (for details see point b) in Definition 2.15):

j=1tj12(VK,Q(K,Q)1)jdiag.\sum_{j=1}^{\infty}t^{\frac{j-1}{2}}\,(V^{(K,Q)_{-1}}_{\mathcal{I}^{*}_{K,Q}})^{\text{diag}}_{j}\,. (2.32)

The operators (VK,Q(K,Q)1)jdiag(V^{(K,Q)_{-1}}_{\mathcal{I}^{*}_{K,Q}})_{j}^{\text{diag}} will be defined below, and “diag” stands for the diagonal part w.r.t. to the two projections PK,Q(),PK,Q()P^{(-)}_{\mathcal{I}^{*}_{K,Q}}\,,\,P^{(-)}_{\mathcal{I}^{*}_{K,Q}}; they are determined by

eZK,Q(GK,Q+tVK,Q(K,Q)1)eZK,Q=:GK,Q+tj=1tj12(VK,Q(K,Q)1)jdiag,e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,(G_{\mathcal{I}^{*}_{K,Q}}+{\sqrt{t}}V_{\mathcal{I}_{K,Q}}^{\,(K,Q)_{-1}})\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}}=:G_{\mathcal{I}^{*}_{K,Q}}+{\sqrt{t}}\sum_{j=1}^{\infty}t^{\frac{j-1}{2}}\,(V^{(K,Q)_{-1}}_{\mathcal{I}^{*}_{K,Q}})^{\text{diag}}_{j}\,, (2.33)

where

GK,Q\displaystyle G_{\mathcal{I}^{*}_{K,Q}} :=\displaystyle:= HK,Q0+tJ=1K1J,Q¯K,QVJ,Q¯(K,Q)1.\displaystyle H_{\mathcal{I}^{*}_{K,Q}}^{0}+{\sqrt{t}}\sum_{J=1}^{K-1}\,\,\sum_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}\subset\mathcal{I}^{*}_{K,Q}}V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}}\,\,. (2.34)

The reader should notice that the (second) sum on the right side of (2.34) does not include those intervals J,Q\mathcal{I}^{\ast}_{J,Q^{\prime}} that share one of their endpoints with K,Q\mathcal{I}^{\ast}_{K,Q}. As a consequence, GK,QG_{\mathcal{I}^{*}_{K,Q}} is localized in K,Q\mathcal{I}^{*}_{K,Q}. The operator ZK,QZ_{\mathcal{I}^{*}_{K,Q}} is given by

ZK,Q:=j=1tj2(ZK,Q)jZ_{\mathcal{I}^{*}_{K,Q}}:=\sum_{j=1}^{\infty}{t^{\frac{j}{2}}}(Z_{\mathcal{I}^{*}_{K,Q}})_{j}\, (2.35)

where the terms (ZK,Q)j(Z_{\mathcal{I}^{*}_{K,Q}})_{j} are defined recursively as follows:

  • (ZK,Q)j:=1GK,QEK,QPK,Q(+)(VK,Q(K,Q)1)jPK,Q()h.c.,(Z_{\mathcal{I}^{*}_{K,Q}})_{j}:=\frac{1}{G_{\mathcal{I}^{*}_{K,Q}}-E_{\mathcal{I}^{*}_{K,Q}}}P^{(+)}_{\mathcal{I}^{*}_{K,Q}}\,(V^{(K,Q)_{-1}}_{\mathcal{I}^{*}_{K,Q}})_{j}\,P^{(-)}_{\mathcal{I}^{*}_{K,Q}}-h.c.\,, (2.36)

    where

    EK,Q:=tJ=1K1J,Q¯K,Qω(VJ,Q¯(K,Q)1)E_{\mathcal{I}^{*}_{K,Q}}:=\,{\sqrt{t}}\sum_{J=1}^{K-1}\,\,\sum_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}\subset\mathcal{I}^{*}_{K,Q}}\,\omega(V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}}) (2.37)

    and ω\omega is defined in (1.7);

  • (VK,Q(K,Q)1)1:=VK,Q(K,Q)1,(V^{(K,Q)_{-1}}_{\mathcal{I}^{*}_{K,Q}})_{1}:=V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\,, (2.38)

    and, for j2j\geq 2,

    (VK,Q(K,Q)1)j:=\displaystyle(V^{(K,Q)_{-1}}_{\mathcal{I}^{*}_{K,Q}})_{j}:=
    p2,r11,rp1;r1++rp=j1p!ad(ZK,Q)r1(ad(ZK,Q)r2(ad(ZK,Q)rp(GK,Q)))+\displaystyle\sum_{p\geq 2,r_{1}\geq 1\dots,r_{p}\geq 1\,;\,r_{1}+\dots+r_{p}=j}\frac{1}{p!}ad\,(Z_{\mathcal{I}^{*}_{K,Q}})_{r_{1}}\Big{(}ad\,(Z_{\mathcal{I}^{*}_{K,Q}})_{r_{2}}\dots(ad\,(Z_{\mathcal{I}^{*}_{K,Q}})_{r_{p}}(G_{\mathcal{I}^{*}_{K,Q}}))\dots\Big{)}+\quad
    p1,r11,rp1;r1++rp=j11p!ad(ZK,Q)r1(ad(ZK,Q)r2(ad(ZK,Q)rp((VK,Q(K,Q)1)1))),\displaystyle\sum_{p\geq 1,r_{1}\geq 1\dots,r_{p}\geq 1\,;\,r_{1}+\dots+r_{p}=j-1}\frac{1}{p!}ad\,(Z_{\mathcal{I}^{*}_{K,Q}})_{r_{1}}\Big{(}ad\,(Z_{\mathcal{I}^{*}_{K,Q}})_{r_{2}}\dots(ad\,(Z_{\mathcal{I}^{*}_{K,Q}})_{r_{p}}((V^{(K,Q)_{-1}}_{\mathcal{I}^{*}_{K,Q}})_{1}))\dots\Big{)}\,,\quad\quad\quad (2.39)

    where the adjoint action of an operator AA on an operator BB is defined by

    adA(B):=[A,B],ad\,A\,(B):=\big{[}A\,,\,B\big{]}\,, (2.40)

    and, recursively,

    adnA(B):=[A,adn1A(B)], for n2.{ad}^{n}A\,(B):=\big{[}A\,,\,{ad}^{n-1}A\,(B)\big{]}\,,\text{ for }\,n\geq 2\,. (2.41)

We note that the construction of ZK,QZ_{\mathcal{I}^{*}_{K,Q}} requires control of the spectral gap of GK,QG_{\mathcal{I}^{*}_{K,Q}} above the ground-state energy, i.e., an estimate on

infspec[(GK,QEK,Q)PK,Q(+)],\inf\text{spec}\,[(G_{\mathcal{I}^{*}_{K,Q}}-E_{\mathcal{I}^{*}_{K,Q}})P^{(+)}_{\mathcal{I}^{*}_{K,Q}}]\,,

which we will outline in Section 3.1.

Remark 2.12.

The reader is invited to notice that the operators of type “W” are not included in the definition of the Hamiltonian GK,QG_{\mathcal{I}^{*}_{K,Q}}.

Remark 2.13.

The Lie-Schwinger series and, accordingly, the series defining ZK,QZ_{\mathcal{I}^{*}_{K,Q}} could actually be truncated, thanks to the structure of the algorithm specified in Section 2.5 below.

2.5 The Algorithm

The following definitions iteratively specify two families of effective interaction potentials, VR,J(K,Q)V^{(K,Q)}_{\mathcal{I}_{R,J}} and WR,J(K,Q)W^{(K,Q)}_{\mathcal{I}_{R,J}}, and we note that the step of the algorithm labelled by (K,Q)(K,Q) is such that K,Qbulk\mathcal{I}_{K,Q}\in\mathfrak{I}_{\text{bulk}} and R,J\mathcal{I}_{R,J} belongs to bulk\mathfrak{I}_{\text{bulk}} if it is the support of VR,J(K,Q)V^{(K,Q)}_{\mathcal{I}_{R,J}} or to b.ry\mathfrak{I}_{\text{b.ry}} if it is the support of WR,J(K,Q)W^{(K,Q)}_{\mathcal{I}_{R,J}}.

Definition 2.14.
  • For 1,Jbulk\mathcal{I}_{1,J}\in\mathfrak{I}_{\text{bulk}}, we define

    V1,J(0,N):=1t1{i,i+1}1,JVi,i+1.V_{\mathcal{I}_{1,J}}^{(0,N)}:=\frac{1}{\sqrt{t^{-1}}}\sum_{\{i,i+1\}\subset\mathcal{I}_{1,J}}V_{i,i+1}\,. (2.42)
  • For 1,Jb.dry\mathcal{I}_{1,J}\in\mathfrak{I}_{\text{b.dry}}, we define

    W1,J(0,N):=1t1{i,i+1}1,JVi,i+1.W_{\mathcal{I}_{1,J}}^{(0,N)}:=\frac{1}{\sqrt{t^{-1}}}\sum_{\{i,i+1\}\subset\mathcal{I}_{1,J}}V_{i,i+1}. (2.43)

Furthermore,

  • for K,Jbulk\mathcal{I}_{K,J}\in\mathfrak{I}_{\text{bulk}}, with K2K\geq 2, we define

    VK,J(0,N):=0,V_{\mathcal{I}_{K,J}}^{(0,N)}:=0\,, (2.44)
  • for K,Jb.dry\mathcal{I}_{K,J}\in\mathfrak{I}_{\text{b.dry}}, with K2K\geq 2, we define

    WK,J(0,N):=0.W_{\mathcal{I}_{K,J}}^{(0,N)}:=0\,. (2.45)

We view (0,N)(0,N) as the predecessor of (1,2)(1,2), in accordance with the restricted ordering introduced in Definition 2.11.

Notation: In the following, ω\omega is the state introduced in (1.7); moreover, ii^{\ast}_{-} and i+i^{\ast}_{+} are the two boundary sites in the microscopic lattice of the interval K,Q\mathcal{I}_{K,Q}^{\ast}.

Definition 2.15.

Assuming that, for an arbitrary (K,Q)1(K,Q)_{-1} with (K,Q)1(0,N)(K,Q)_{-1}\succ(0,N), the operators VR,J(K,Q)1,VR,J¯(K,Q)1,WR,J(K,Q)1V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}\,,\,V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{R,J}}}\,,\,{W^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}\,} are well defined for any (R,J)(R,J) in bulk\mathfrak{I}_{\text{bulk}} and in b.dry\mathfrak{I}_{\text{b.dry}}, respectively, and that the operators ZK,QZ_{\mathcal{I}^{*}_{K,Q}} (see (2.35)) are well defined, and assuming that if (K,Q)=(1,2)(K,Q)=(1,2) then Z1,2Z_{\mathcal{I}^{*}_{1,2}} is well defined, then definitions a-1), a-2), b), c-1), and c-2) (see below) are meaningful. Such prescriptions are organized into three groups, 𝒜\mathcal{A}, \mathcal{B}, and 𝒞\mathcal{C}; for each of them we give first a description in words.

  1. 𝒜)\mathcal{A})

    Items a-1) and a-2) below deal with identity maps, that is they describe situations where for a given interval \mathcal{I} the corresponding potential (supported in \mathcal{I}) does not change from step (K,Q)1(K,Q)_{-1} to step (K,Q)(K,Q); in terms of the conjugation associated with step (K,Q)(K,Q), the new potential is either the zero order term in the expansion (in ZK,QZ_{\mathcal{I}^{*}_{K,Q}}) of

    eZK,QOeZK,Q,e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,O_{\mathcal{I}}\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}}\,, (2.46)

    where OO_{\mathcal{I}} stands for the potential under consideration in step (K,Q)1(K,Q)_{-1}, or just the operator OO_{\mathcal{I}} whenever [O,eZK,Q]=0[O_{\mathcal{I}}\,,\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}}]=0;

    • a-1)

      if (K,Q)(R,J)(K,Q)\prec(R,J),  R,Jbulk\mathcal{I}_{R,J}\in\mathfrak{I}_{\text{bulk}} and K,QR,J\mathcal{I}^{*}_{K,Q}\nsubseteq\mathcal{I}_{R,J} we set

      VR,J(K,Q):=VR,J(K,Q)1;V^{(K,Q)}_{\mathcal{I}_{R,J}}:=V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}\,; (2.47)

      if (K,Q)(R,J)(K,Q)\prec(R,J),  R,Jb.ry\mathcal{I}_{R,J}\in{\mathfrak{I}_{\text{b.ry}}} and K,QR,J\mathcal{I}^{*}_{K,Q}\nsubseteq\mathcal{I}^{*}_{R,J} we set

      WR,J(K,Q):=WR,J(K,Q)1;{W^{(K,Q)}_{\mathcal{I}_{R,J}}:=W^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}}\,; (2.48)
    • a-2)

      if (K,Q)(R,J)(K,Q)\succ(R,J), for R,Jbulk\mathcal{I}_{R,J}\in\mathfrak{I}_{\text{bulk}}, we set

      VR,J¯(K,Q):=VR,J¯(K,Q)1,V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{R,J}}}:=V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{R,J}}}, (2.49)

      if (K,Q)(R,J)(K,Q)\succ(R,J), for R,Jb.ry\mathcal{I}_{R,J}\in\mathfrak{I}_{\text{b.ry}}, we set

      WR,J(K,Q):=WR,J(K,Q)1;W^{(K,Q)}_{{\mathcal{I}_{R,J}}}:=W^{(K,Q)_{-1}}_{{\mathcal{I}_{R,J}}}; (2.50)
  2. )\mathcal{B})

    Item b) below describes the process that takes place when the label (K,Q)(K,Q) of the step coincides with the label (R,J)(R,J) of the potential under consideration. By construction, only for potentials of type “V" the labels (R,J)(R,J) and (K,Q)(K,Q) can coincide. As anticipated in Section 2.3, the map which defines the new potential in step (K,Q)(K,Q) consists of two operations:

    - extracting the quantity (2.54) from the leading order term of the Lie-Schwinger series defined in (2.32) and associated with the potential supported in K,QR,J\mathcal{I}_{K,Q}\equiv\mathcal{I}_{R,J} in step (K,Q)1(K,Q)_{-1};

    - extracting the diagonal part from the first order of what we refer to as hooking of the (rescaled) projection terms, i.e.,

    eZK,Q𝒫i1,i(2)teZK,Q𝒫i1,i(2)t,e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,\frac{\mathcal{P}^{(2)}_{i-1,i}}{\sqrt{t}}\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}}-\frac{\mathcal{P}^{(2)}_{i-1,i}}{\sqrt{t}}\,, (2.51)

    where the support of 𝒫i1,i(2)\mathcal{P}^{(2)}_{i-1,i} overlaps with K,Q\mathcal{I}^{*}_{K,Q} but it is not contained in it;

    • b)

      if (K,Q)=(R,J)(K,Q)=(R,J) then

      VR,J¯(K,Q)\displaystyle V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{R,J}}} :=\displaystyle:= ω(VK,Q(K,Q)1)+PK,Q¯(+)PK,Q(+)[VK,Q(K,Q)1ω(VK,Q(K,Q)1)]PK,Q(+)PK,Q¯(+)\displaystyle\omega(V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}})+P^{(+)}_{\overline{\mathcal{I}^{*}_{K,Q}}}\,P^{(+)}_{\mathcal{I}^{*}_{K,Q}}\,\,\Big{[}V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}-\omega(V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}})\Big{]}\,P^{(+)}_{\mathcal{I}^{*}_{K,Q}}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{K,Q}}}\quad\quad\quad (2.54)
      +PR,J¯(+)(adZK,QR,J(𝒫i1,i(2)t))PR,J¯(+)\displaystyle+P^{(+)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,\Big{(}adZ_{\mathcal{I}^{*}_{K,Q}\equiv\mathcal{I}^{*}_{R,J}}(\,\frac{\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}}{\sqrt{t}})\Big{)}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{R,J}}}
      +PR,J¯(+)(adZK,QR,J(𝒫i+,i++1(2)t))PR,J¯(+),\displaystyle+P^{(+)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,\Big{(}adZ_{\mathcal{I}^{*}_{K,Q}\equiv\mathcal{I}^{*}_{R,J}}(\,\frac{\mathcal{P}^{(2)}_{i^{*}_{+},i^{*}_{+}+1}}{\sqrt{t}})\Big{)}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,,

      where ii^{*}_{-} and i+i^{*}_{+} are the sites of the microscopic lattice corresponding to the endpoints of the interval K,QR,J\mathcal{I}^{*}_{K,Q}\equiv\mathcal{I}^{*}_{R,J} .

  3. 𝒞)\mathcal{C})

    Items c-1) and c-2) below describe growth processes for the V and the W potentials, respectively. Regarding the new potential of type V (see c-1)) associated with a given fixed interval R,J\mathcal{I}_{R,J} with K,QR,J\mathcal{I}^{*}_{K,Q}\subset\mathcal{I}_{R,J}, it does not involve W operators. It involves VV operators, but also includes possible contributions coming from the hooking (in step (K,Q)(K,Q)) of rescaled projections, namely higher order and off-diagonal first order terms. The growth process prescribed in c-2) for the W potentials includes all the terms which upon the conjugation of the Hamiltonian in step (K,Q)(K,Q) turned out to be supported in the interval R,J\mathcal{I}_{R,J} supposed to be in b.ry\mathfrak{I}_{\text{b.ry}}. Concerning the nontrivial structure designed in c-1) and c-2), the reader is referred to the explanations in Remarks 2.16, 2.17, and 2.18.

    • c-1)

      if K,QR,Jbulk\mathcal{I}^{*}_{K,Q}\subset\mathcal{I}_{R,J}\in\mathfrak{I}_{\text{bulk}} then

      VR,J(K,Q)\displaystyle V^{(K,Q)}_{\mathcal{I}_{R,J}} :=\displaystyle:= eZK,QVR,J(K,Q)1eZK,Q\displaystyle e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}} (2.62)
      +K,Q[𝒢R,J(K,Q)]1n=11n!adnZK,Q(VK,Q(K,Q)1)\displaystyle+\sum_{\mathcal{I}_{K^{\prime},Q^{\prime}}\in[\mathcal{G}^{(K,Q)}_{\mathcal{I}_{R,J}}]_{1}}\,\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(V^{(K,Q)_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}})
      +K,Q[𝒢R,J(K,Q)]2n=11n!adnZK,Q(VK,Q¯(K,Q)1)\displaystyle+\sum_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}\in[\mathcal{G}^{(K,Q)}_{\mathcal{I}_{R,J}}]_{2}}\,\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}})
      +δ~K,Q=R,JK,Q[𝒢R,J(K,Q)]3n=11n!adnZK,Q(VK,Q¯(K,Q)1)\displaystyle+\delta_{\widetilde{\mathcal{I}}^{*}_{K,Q}=\mathcal{I}_{R,J}}\sum_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}\in[\mathcal{G}^{(K,Q)}_{\mathcal{I}_{R,J}}]_{3}}\,\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}})
      +δ~K,Q=R,J[PK,Q()VK,Q(K,Q)1PK,Q()ω(VK,Q(K,Q)1)PK,Q()]\displaystyle{+\delta_{\widetilde{\mathcal{I}}^{*}_{K,Q}=\mathcal{I}_{R,J}}[P^{(-)}_{\mathcal{I}^{*}_{K,Q}}V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}P^{(-)}_{\mathcal{I}^{*}_{K,Q}}-\omega(V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}})P^{(-)}_{\mathcal{I}^{*}_{K,Q}}]}\,\,\quad\quad
      +δ~K,Q=R,J[m=2tm12(VK,Q(K,Q)1)mdiag]\displaystyle+\delta_{\widetilde{\mathcal{I}}^{*}_{K,Q}=\mathcal{I}_{R,J}}[\sum_{m=2}^{\infty}{t^{\frac{m-1}{2}}}(V^{(K,Q)_{-1}}_{\mathcal{I}^{*}_{K,Q}})^{\text{diag}}_{m}]
      +δ~K,Q=R,J(n=21n!adnZK,Q(𝒫i1,i(2)t+𝒫i+,i++1(2)t))\displaystyle+\delta_{\widetilde{\mathcal{I}}^{*}_{K,Q}=\mathcal{I}_{R,J}}\Big{(}\sum_{n=2}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(\,{\frac{\mathcal{P}^{(2)}_{i^{\ast}_{-}-1,i^{\ast}_{-}}}{\sqrt{t}}}+{\frac{\mathcal{P}^{(2)}_{i^{\ast}_{+},i^{\ast}_{+}+1}}{\sqrt{t}}})\Big{)}
      +δ~K,Q=R,J[PK,Q¯()(adZK,Q(𝒫i1,i(2)t+𝒫i+,i++1(2)t))PK,Q¯(+)+h.c.],\displaystyle+\delta_{\widetilde{\mathcal{I}}^{*}_{K,Q}=\mathcal{I}_{R,J}}[P^{(-)}_{\overline{\mathcal{I}^{*}_{K,Q}}}\Big{(}adZ_{\mathcal{I}^{*}_{K,Q}}(\,{\frac{\mathcal{P}^{(2)}_{i^{\ast}_{-}-1,i^{\ast}_{-}}}{\sqrt{t}}}+{\frac{\mathcal{P}^{(2)}_{i^{\ast}_{+},i^{\ast}_{+}+1}}{\sqrt{t}}})\Big{)}P^{(+)}_{\overline{\mathcal{I}^{*}_{K,Q}}}+h.c.]\,,\quad

      where

      [𝒢R,J(K,Q)]1\displaystyle[\mathcal{G}^{(K,Q)}_{\mathcal{I}_{R,J}}]_{1} :=\displaystyle:= {K,Qbulk|(K,Q)(K,Q),K,QK,Q,\displaystyle\Big{\{}\,\mathcal{I}_{K^{\prime},Q^{\prime}}\in\mathfrak{I}_{\text{bulk}}\,\,|\,\,(K^{\prime},Q^{\prime})\succ(K,Q)\,,\,\mathcal{I}_{K^{\prime},Q^{\prime}}\cap\mathcal{I}^{*}_{K,Q}\neq\emptyset, (2.63)
      K,QR,J,and~K,QK,Q=R,J}\displaystyle\,\mathcal{I}_{K^{\prime},Q^{\prime}}\neq\mathcal{I}_{R,J}\,,\,\text{and}\,\,\widetilde{\mathcal{I}}^{*}_{K,Q}\cup\mathcal{I}_{K^{\prime},Q^{\prime}}=\mathcal{I}_{R,J}\,\,\Big{\}}
      [𝒢R,J(K,Q)]2\displaystyle[\mathcal{G}^{(K,Q)}_{\mathcal{I}_{R,J}}]_{2} :=\displaystyle:= {K,Qbulk|(K,Q)(K,Q),K,QK,Q,K,QK,Q\displaystyle\Big{\{}\,\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}\in\mathfrak{I}_{\text{bulk}}\,\,|\,\,(K,Q)\succ(K^{\prime},Q^{\prime})\,,\,\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}\cap\mathcal{I}^{*}_{K,Q}\neq\emptyset\,,\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}\nsubset\mathcal{I}^{*}_{K,Q}
      and~K,Q~K,Q=R,J}\displaystyle\text{and}\,\,\widetilde{\mathcal{I}}^{*}_{K,Q}\cup\widetilde{\mathcal{I}}^{*}_{K^{\prime},Q^{\prime}}=\mathcal{I}_{R,J}\,\,\Big{\}}\,
      [𝒢R,J(K,Q)]3\displaystyle[\mathcal{G}^{(K,Q)}_{\mathcal{I}_{R,J}}]_{3} :=\displaystyle:= {K,Qbulk|K,QK,Q,iK,Q or i+K,Q}\displaystyle\Big{\{}\,\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}\in\mathfrak{I}_{\text{bulk}}\,\,|\,\,\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}\subset\mathcal{I}^{*}_{K,Q},\,i_{-}^{*}\in\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}\text{ or }i_{+}^{*}\in\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}\,\Big{\}}\,
      Remark 2.16.

      The terms in (2.62), (2.62), and (2.62) are related to the block-diagonalization in step (K,Q)(K,Q) and are present only if ~K,Q=R,J\widetilde{\mathcal{I}}^{*}_{K,Q}=\mathcal{I}_{R,J}. More precisely, we observe that: (2.62) originates from the definition of (2.34) in the sense that accounts for the higher order terms of the conjugation of those potentials supported in intervals J,Q\mathcal{I}^{\ast}_{J,Q^{\prime}}, with J,QK,Q\mathcal{I}^{*}_{J,Q^{\prime}}\subset\mathcal{I}^{*}_{K,Q}, that share one of their endpoints with K,Q\mathcal{I}^{\ast}_{K,Q}; (2.62) collects what is left of the first order term of the Lie Schwinger series after extracting the quantity in (2.54) which enters the definition in b); (2.62) is the Lie Schwinger series above first order.

      Remark 2.17.

      The companion off-diagonal terms of (2.54) and (2.54) respectively are two terms of the type in (2.62) for an interval R,J,\mathcal{I}_{R^{\prime},J^{\prime},} such that R,J=~K,Q=~R,J\mathcal{I}_{R^{\prime},J^{\prime}}=\widetilde{\mathcal{I}}^{*}_{K,Q}=\widetilde{\mathcal{I}}^{*}_{R,J}. The term in (2.62) accounts for the higher order terms of the operator resulting from the hooking of the projections in step (K,Q)(K,Q), provided ~K,Q=R,J\widetilde{\mathcal{I}}^{*}_{K,Q}=\mathcal{I}_{R,J}.

    • c-2)

      if K,QR,Jb.ry\mathcal{I}^{*}_{K,Q}\subset\mathcal{I}_{R,J}\in\mathfrak{I}_{\text{b.ry}},

      WR,J(K,Q)\displaystyle W^{(K,Q)}_{{\mathcal{I}_{R,J}}} :=\displaystyle:= eZK,QWR,J(K,Q)1eZK,Q\displaystyle e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,W^{(K,Q)_{-1}}_{{\mathcal{I}_{R,J}}}\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}} (2.72)
      +K,Q𝒢R,J(K,Q)n=11n!adnZK,Q(WK,Q(K,Q)1)\displaystyle+\sum_{\mathcal{I}_{K^{\prime},Q^{\prime}}\in\mathcal{G}^{(K,Q)}_{\mathcal{I}_{R,J}}}\,\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(W^{(K,Q)_{-1}}_{{\mathcal{I}_{K^{\prime},Q^{\prime}}}})
      +K,Q[𝒢R,J(K,Q)]1n=11n!adnZK,Q(VK,Q(K,Q)1)\displaystyle+\sum_{\mathcal{I}_{K^{\prime},Q^{\prime}}\in[\mathcal{G}^{(K,Q)}_{\mathcal{I}_{R,J}}]_{1}}\,\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(V^{(K,Q)_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}})
      +K,Q[𝒢R,J(K,Q)]2n=11n!adnZK,Q(VK,Q¯(K,Q)1)\displaystyle+\sum_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}\in[\mathcal{G}^{(K,Q)}_{\mathcal{I}_{R,J}}]_{2}}\,\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}})
      +δ~K,Q=R,JK,Q[𝒢R,J(K,Q)]3n=11n!adnZK,Q(VK,Q¯(K,Q)1)\displaystyle+\delta_{\widetilde{\mathcal{I}}^{*}_{K,Q}=\mathcal{I}_{R,J}}\sum_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}\in[\mathcal{G}^{(K,Q)}_{\mathcal{I}_{R,J}}]_{3}}\,\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}})
      +δ~K,Q=R,J[PK,Q()VK,Q(K,Q)1PK,Q()ω(VK,Q(K,Q)1)PK,Q()]\displaystyle{+\delta_{\widetilde{\mathcal{I}}^{*}_{K,Q}=\mathcal{I}_{R,J}}[P^{(-)}_{\mathcal{I}^{*}_{K,Q}}V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}P^{(-)}_{\mathcal{I}^{*}_{K,Q}}-\omega(V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}})P^{(-)}_{\mathcal{I}^{*}_{K,Q}}]}\,\,\quad\quad
      +δ~K,Q=R,J[m=2t(m1)2(VK,Q(K,Q)1)mdiag]\displaystyle+\delta_{\widetilde{\mathcal{I}}^{*}_{K,Q}=\mathcal{I}_{R,J}}[\sum_{m=2}^{\infty}{t^{\frac{(m-1)}{2}}}(V^{(K,Q)_{-1}}_{{\mathcal{I}^{*}_{K,Q}}})^{\text{diag}}_{m}]
      +δ~K,Q=R,J(n=21n!adnZK,Q(𝒫i1,i(2)t+𝒫i+,i++1(2)t))\displaystyle+\delta_{\widetilde{\mathcal{I}}^{*}_{K,Q}=\mathcal{I}_{R,J}}\Big{(}\sum_{n=2}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(\,{\frac{\mathcal{P}^{(2)}_{i^{\ast}_{-}-1,i^{\ast}_{-}}}{\sqrt{t}}}+{\frac{\mathcal{P}^{(2)}_{i^{\ast}_{+},i^{\ast}_{+}+1}}{\sqrt{t}}})\Big{)}
      +δ~K,Q=R,J[PR,J¯()(adZK,Q(𝒫i1,i(2)t+𝒫i+,i++1(2)t))PR,J¯(+)+h.c.],\displaystyle+\delta_{\widetilde{\mathcal{I}}^{*}_{K,Q}=\mathcal{I}_{R,J}}[P^{(-)}_{\overline{\mathcal{I}^{*}_{R,J}}}\Big{(}adZ_{\mathcal{I}^{*}_{K,Q}}(\,{\frac{\mathcal{P}^{(2)}_{i^{\ast}_{-}-1,i^{\ast}_{-}}}{\sqrt{t}}}+{\frac{\mathcal{P}^{(2)}_{i^{\ast}_{+},i^{\ast}_{+}+1}}{\sqrt{t}}})\Big{)}P^{(+)}_{\overline{\mathcal{I}^{*}_{R,J}}}+h.c.]\,,\,\,\quad\quad

      where

      𝒢R,J(K,Q):={K,Qb.ry|K,QK,Qand~K,QK,Q=R,J}.\mathcal{G}^{(K,Q)}_{\mathcal{I}_{R,J}}:=\Big{\{}\,\mathcal{I}_{K^{\prime},Q^{\prime}}\in\mathfrak{I}_{\text{b.ry}}\,\,|\,\,\mathcal{I}_{K^{\prime},Q^{\prime}}\cap\mathcal{I}_{K,Q}\neq\emptyset\,\,\text{and}\,\,\widetilde{\mathcal{I}}^{*}_{K,Q}\cup\mathcal{I}_{K^{\prime},Q^{\prime}}=\mathcal{I}_{R,J}\,\,\Big{\}}\,.
      Remark 2.18.

      The terms above are all analogous to the ones in c-1) except for the term in (2.72). A counterpart of (2.72) cannot be present in c-1) since it consists of operators where the hooked potentials are supported in intervals K,Qb.ry{\mathcal{I}_{K^{\prime},Q^{\prime}}}\in\mathfrak{I}_{\text{b.ry}}, hence such operator cannot contribute to a VV term.

In the next theorem we show that the algorithm described above is consistent with the unitary conjugation of the Hamiltonian KΛ(K,Q)1(t)K_{\Lambda}^{\,(K,Q)_{-1}}(t) generated by the operator ZK,QZ_{\mathcal{I}^{*}_{K,Q}}.

Theorem 2.19.

For the Hamiltonian KΛ(K,Q)(t)K_{\Lambda}^{\,(K,Q)}(t), defined iteratively by (2.30)-(2.30) and Definition 2.15 above, the following identity holds

KΛ(K,Q)(t)=eZK,QKΛ(K,Q)1(t)eZK,Q.K_{\Lambda}^{\,(K,Q)}(t)=e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,K_{\Lambda}^{\,(K,Q)_{-1}}(t)\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}}. (2.73)

Proof

We prove the identity claimed in the statement of the theorem by studying the conjugation of each term on the right side of the expression given below

eZK,QKΛ(K,Q)1(t)eZK,Q\displaystyle e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,K_{\Lambda}^{\,(K,Q)_{-1}}(t)\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}} (2.74)
=\displaystyle= eZK,Q[HΛ\displaystyle e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,\Big{[}\,H_{\Lambda}
+tQV1,Q¯(K,Q)1++tQ;(K,Q)(K,Q)VK,Q¯(K,Q)1\displaystyle\quad+\sqrt{t}\sum_{Q^{\prime}}V^{\,(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{1,Q^{\prime}}}}+\dots+{\sqrt{t}}\sum_{Q^{\prime}\,;\,(K,Q^{\prime})\preceq(K,Q)}V^{\,(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{K,Q^{\prime}}}}\quad\quad\quad
+tQ;(K,Q)(K,Q)VK,Q(K,Q)1+tQVK+1,Q(K,Q)1++tV(N1)t2,2(K,Q)1\displaystyle\quad+\sqrt{t}\sum_{Q^{\prime}\,;\,(K,Q^{\prime})\succ(K,Q)}V^{\,(K,Q)_{-1}}_{\mathcal{I}_{K,Q^{\prime}}}+\sqrt{t}\sum_{Q^{\prime}}V^{\,(K,Q)_{-1}}_{\mathcal{I}_{K+1,Q^{\prime}}}+\dots+\sqrt{t}V^{\,(K,Q)_{-1}}_{\mathcal{I}_{(N-1)\cdot\sqrt{t}-2,2}}
+tQW1,Q(K,Q)1++tW(N1)t,1(K,Q)1]eZK,Q\displaystyle\quad+\sqrt{t}\sum_{Q^{\prime}}W^{\,(K,Q)_{-1}}_{\mathcal{I}_{1,Q^{\prime}}}+\dots+\sqrt{t}W^{\,(K,Q)_{-1}}_{\mathcal{I}_{(N-1)\cdot\sqrt{t},1}}\Big{]}\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}}

and subsequently re-assembling the terms according to the rules introduced in Definition 2.15.

The following observations are important.

  1. (i)

    For all intervals R,J\mathcal{I}_{R,J} or R,J\mathcal{I}^{*}_{R,J} with the property that R,JK,Q=\mathcal{I}_{R,J}\cap\mathcal{I}^{*}_{K,Q}=\emptyset or R,JK,Q=\mathcal{I}^{*}_{R,J}\cap\mathcal{I}^{*}_{K,Q}=\emptyset, we have that

    eZK,QVR,J(K,Q)1eZK,Q\displaystyle e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}} =\displaystyle= VR,J(K,Q)1=:VR,J(K,Q),\displaystyle V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}=:V^{(K,Q)}_{\mathcal{I}_{R,J}}\,, (2.77)
    eZK,QVR,J¯(K,Q)1eZK,Q\displaystyle e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{R,J}}}\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}} =\displaystyle= VR,J¯(K,Q)1=:VR,J¯(K,Q),\displaystyle V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{R,J}}}=:V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,, (2.78)
    eZK,QWR,J(K,Q)1eZK,Q\displaystyle e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,W^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}} =\displaystyle= WR,J(K,Q)1=:WR,J(K,Q),\displaystyle W^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}=:W^{(K,Q)}_{\mathcal{I}^{*}_{R,J}}\,, (2.79)

    which follows from a-1) and a-2) in Definition 2.15.

  2. (ii)

    Using a Lie-Schwinger block-diagonalization associated with an “unperturbed” Hamiltonian GK,QG_{\mathcal{I}^{*}_{K,Q}} – see (2.34) – and a “perturbation” tVK,Q(K,Q)1\sqrt{t}V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}, we find that

    eZK,Q(HK,Q0+tJ=1K1J,QK,QVJ,Q¯(K,Q)1+tVK,Q(K,Q)1)eZK,Q\displaystyle e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,\,\Big{(}H_{\mathcal{I}^{*}_{K,Q}}^{0}+{\sqrt{t}}\sum_{J=1}^{K-1}\,\,\sum_{\mathcal{I}^{*}_{J,Q^{\prime}}\subset\mathcal{I}^{*}_{K,Q}}V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}}+\sqrt{t}V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\Big{)}\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}}\, (2.80)
    =\displaystyle= HK,Q0+tJ=1K1J,Q¯K,QVJ,Q¯(K,Q)1+tm=1tm12(VK,Q(K,Q)1)mdiag\displaystyle H_{\mathcal{I}^{*}_{K,Q}}^{0}+{\sqrt{t}}\sum_{J=1}^{K-1}\,\,\sum_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}\subset\mathcal{I}^{*}_{K,Q}}V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}}+\sqrt{t}\sum_{m=1}^{\infty}{t^{\frac{m-1}{2}}}(V^{(K,Q)_{-1}}_{\mathcal{I}^{*}_{K,Q}})^{\text{diag}}_{m}\, (2.82)
    +eZK,QtJ,QK,Q;J,Q¯K,QVJ,Q¯(K,Q)1eZK,Q\displaystyle+e^{Z_{\mathcal{I}^{*}_{K,Q}}}\sqrt{t}\sum_{\mathcal{I}^{*}_{J,Q^{\prime}}\subset\mathcal{I}^{*}_{K,Q}\,;\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}\nsubset\mathcal{I}^{*}_{K,Q}}V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}}e^{-Z_{\mathcal{I}^{*}_{K,Q}}}

    where in the expression within parentheses in (2.80) we have used the identity

    J=1K1J,QK,QVJ,Q¯(K,Q)1=J=1K1J,Q¯K,QVJ,Q¯(K,Q)1+J,QK,Q;J,Q¯K,QVJ,Q¯(K,Q)1,\sum_{J=1}^{K-1}\,\,\sum_{\mathcal{I}^{*}_{J,Q^{\prime}}\subset\mathcal{I}^{*}_{K,Q}}V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}}\,=\,\sum_{J=1}^{K-1}\,\,\sum_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}\subset\mathcal{I}^{*}_{K,Q}}V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}}+\sum_{\mathcal{I}^{*}_{J,Q^{\prime}}\subset\mathcal{I}^{*}_{K,Q}\,;\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}\nsubset\mathcal{I}^{*}_{K,Q}}V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}}\,,\quad (2.83)

    and (2.82) is the result of the Lie-Schwinger conjugation. Next, we split the conjugation in (2.82) into the zero order term and the rest, so as to get

    (2.80)\displaystyle(\ref{2.71}) (2.85)
    =\displaystyle= HK,Q0+tJ=1K1J,QK,QVJ,Q¯(K,Q)+t((2.54))+t((2.62) or (2.72))+t((2.62) or (2.72))\displaystyle H_{\mathcal{I}^{*}_{K,Q}}^{0}+{\sqrt{t}\sum_{J=1}^{K-1}\,\,\sum_{\mathcal{I}^{*}_{J,Q^{\prime}}\subset\mathcal{I}^{*}_{K,Q}}V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}}+\sqrt{t}((\ref{b1main}))+\sqrt{t}((\ref{expdecayerror})\text{ or }(\ref{c-off-diag}))+\sqrt{t}((\ref{lshighorder})\text{ or }(\ref{corrW1}))}
    +t((2.62) or (2.72)),\displaystyle+\sqrt{t}((\ref{Valgo67})\text{ or }(\ref{Valgo66}))\,,

    where the alternatives of the type “(2.62) or (2.72)(\ref{expdecayerror})\text{ or }(\ref{c-off-diag})" on the right side of the formula above depend on whether the resulting operator is a bulk- or a boundary-potential; furthermore we have used Definition 2.15, case a-2), which yields the identity

    J=1K1J,Q¯K,QVJ,Q¯(K,Q)1+J,QK,Q;J,Q¯K,QVJ,Q¯(K,Q)1=J=1K1J,QK,QVJ,Q¯(K,Q).\sum_{J=1}^{K-1}\,\,\sum_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}\subset\mathcal{I}^{*}_{K,Q}}V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}}+\sum_{\mathcal{I}^{*}_{J,Q^{\prime}}\subset\mathcal{I}^{*}_{K,Q}\,;\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}\nsubset\mathcal{I}^{*}_{K,Q}}V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}}=\sum_{J=1}^{K-1}\,\,\sum_{\mathcal{I}^{*}_{J,Q^{\prime}}\subset\mathcal{I}^{*}_{K,Q}}V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}}\,.\quad (2.86)
  3. (iii)

    The action of the conjugation on the terms VR,J(K,Q)1V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}, with K,QR,J\mathcal{I}^{*}_{K,Q}\subset\mathcal{I}_{R,J}, is

    eZK,QVR,J(K,Q)1eZK,Q=(2.62).e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}}=(\ref{identity-c}). (2.87)
  4. (iv)

    For the conjugation of the terms VR,J(K,Q)1V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}, with K,QR,J\mathcal{I}^{*}_{K,Q}\cap\mathcal{I}_{R,J}\neq\emptyset and K,QR,J\mathcal{I}^{*}_{K,Q}\nsubset\mathcal{I}_{R,J}, R,JK,Q\mathcal{I}_{R,J}\nsubset\mathcal{I}^{*}_{K,Q} ,

    eZK,QVR,J(K,Q)1eZK,Q=VR,J(K,Q)1+n=11n!adnZK,Q(VR,J(K,Q)1),e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}}=V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}+\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}})\,, (2.88)

    we notice that the first term on the right side of (2.88) is VR,J(K,Q)V^{(K,Q)}_{\mathcal{I}_{R,J}} (see cases a-1) Definition 2.15); as for the second term:

    • if R,JR,J~K,Qbulk\mathcal{I}_{R^{\prime},J^{\prime}}\equiv\mathcal{I}_{R,J}\cup\widetilde{\mathcal{I}}^{*}_{K,Q}\in\mathfrak{I}_{\text{bulk}} it contributes to VR,J(K,Q)V^{(K,Q)}_{\mathcal{I}_{R^{\prime},J^{\prime}}}, according to (2.62);

    • if R,JR,J~K,Qb.ry\mathcal{I}_{R^{\prime},J^{\prime}}\equiv\mathcal{I}_{R,J}\cup\widetilde{\mathcal{I}}^{*}_{K,Q}\in\mathfrak{I}_{\text{b.ry}} it contributes to WR,J(K,Q)W^{(K,Q)}_{\mathcal{I}_{R^{\prime},J^{\prime}}}, according to (2.72).

  5. (v)

    Concerning the conjugation of the terms of the type VR,J¯(K,Q)1V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{R,J}}}, we notice that they appear in (2.5) only for (K,Q)1(R,J)(K,Q)_{-1}\succeq(R,J). Thus, for (K,Q)(R,J)(K,Q)\succ(R,J), we study the possible situations:

    • if R,JK,Q=\mathcal{I}^{*}_{R,J}\cap\mathcal{I}^{*}_{K,Q}=\emptyset we refer to (2.78);

    • if R,JK,Q\mathcal{I}^{*}_{R,J}\cap\mathcal{I}^{*}_{K,Q}\neq\emptyset

      eZK,QVR,J¯(K,Q)1eZK,Q=VR,J¯(K,Q)1+n=11n!adnZK,Q(VR,J¯(K,Q)1),e^{Z_{\mathcal{I}^{*}_{K,Q}}}V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{R,J}}}\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}}=V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{R,J}}}+\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{R,J}}})\,, (2.89)

      where the first term is VR,J¯(K,Q)V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{R,J}}}, by a-2) of Definition 2.15; regarding the second term, i.e.,

      n=11n!adnZK,Q(VR,J¯(K,Q)1),\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{R,J}}})\,,
      • if R,J~K,Q~R,Jbulk\mathcal{I}_{R^{\prime},J^{\prime}}\equiv\widetilde{\mathcal{I}}^{*}_{K,Q}\cup\widetilde{\mathcal{I}}^{*}_{R,J}\in\mathfrak{I}_{bulk}, it contributes to VR,J(K,Q)V^{(K,Q)}_{\mathcal{I}_{R^{\prime},J^{\prime}}} according to (2.62) of Definition 2.15.

      • if R,J~K,Q~R,Jb.ry\mathcal{I}_{R^{\prime},J^{\prime}}\equiv\widetilde{\mathcal{I}}^{*}_{K,Q}\cup\widetilde{\mathcal{I}}^{*}_{R,J}\in\mathfrak{I}_{b.ry}, it contributes to WR,J(K,Q)W^{(K,Q)}_{\mathcal{I}_{R^{\prime},J^{\prime}}} according to (2.72) of Definition 2.15.

  6. (vi)

    With regard to the terms WR,J(K,Q)1W^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}, we observe that:

    • the case R,JK,Q=\mathcal{I}_{R,J}\cap\mathcal{I}^{*}_{K,Q}=\emptyset has already been discussed;

    • if K,QR,J\mathcal{I}^{*}_{K,Q}\subset\mathcal{I}_{R,J} the expression

      eZK,QWR,J(K,Q)1eZK,Q=WR,J(K,Q)1+n=11n!adnZK,Q(WR,J(K,Q)1)e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,W^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}}=W^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}+\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(W^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}})\, (2.90)

      contributes to WR,J(K,Q)W^{(K,Q)}_{\mathcal{I}^{*}_{R,J}} according to (2.72);

    • if K,QR,J\mathcal{I}^{*}_{K,Q}\cap\mathcal{I}_{R,J}\neq\emptyset and K,QR,J\mathcal{I}^{*}_{K,Q}\nsubset\mathcal{I}_{R,J}, R,JK,Q\mathcal{I}_{R,J}\nsubset\mathcal{I}^{*}_{K,Q}, in the expression

      eZK,QWR,J(K,Q)1eZK,Q=WR,J(K,Q)1+n=11n!adnZK,Q(WR,J(K,Q)1)e^{Z_{\mathcal{I}^{*}_{K,Q}}}W^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}}=W^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}+\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(W^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}) (2.91)

      the first term defines WR,J(K,Q)W^{(K,Q)}_{\mathcal{I}_{R,J}}, by a-1) and a-2); the other terms, i.e.,

      n=11n!adnZK,Q(WR,J(K,Q)1)\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(W^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}) (2.92)

      contribute to WR,J(K,Q)W^{(K,Q)}_{\mathcal{I}_{R^{\prime},J^{\prime}}}, with R,J~K,QR,J\mathcal{I}_{R^{\prime},J^{\prime}}\equiv\widetilde{\mathcal{I}}^{*}_{K,Q}\cup\mathcal{I}_{R,J}, according to (2.72) in c-2).

  7. (vii)

    We finally consider the terms of the unperturbed Hamiltonian HΛ0H^{0}_{\Lambda} supported in the intervals of type (i,i+1)(i,i+1) which overlap with K,Q\mathcal{I}^{*}_{K,Q} but are not contained in it; for these terms we have:

    eZK,Q(𝒫i+,i++1(2)t+𝒫i1,i(2)t)eZK,Q\displaystyle e^{Z_{\mathcal{I}^{*}_{K,Q}}}\,\,(\frac{\mathcal{P}^{(2)}_{i^{\ast}_{+},i^{\ast}_{+}+1}}{\sqrt{t}}+\frac{\mathcal{P}^{(2)}_{i^{\ast}_{-}-1,i^{\ast}_{-}}}{\sqrt{t}})\,e^{-Z_{\mathcal{I}^{*}_{K,Q}}}
    =\displaystyle= 𝒫i+,i++1(2)t+𝒫i1,i(2)t+n=11n!adnZK,Q(𝒫i+,i++1(2)t+𝒫i1,i(2)t)\displaystyle\frac{\mathcal{P}^{(2)}_{i^{\ast}_{+},i^{\ast}_{+}+1}}{\sqrt{t}}+\frac{\mathcal{P}^{(2)}_{i^{\ast}_{-}-1,i^{\ast}_{-}}}{\sqrt{t}}+\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(\frac{\mathcal{P}^{(2)}_{i^{\ast}_{+},i^{\ast}_{+}+1}}{\sqrt{t}}+\frac{\mathcal{P}^{(2)}_{i^{\ast}_{-}-1,i^{\ast}_{-}}}{\sqrt{t}})
    =\displaystyle= 𝒫i+,i++1(2)t+𝒫i1,i(2)t+(2.54)+(2.54)+[((2.62)+(2.62)) or ((2.72)+(2.72))],\displaystyle\frac{\mathcal{P}^{(2)}_{i^{\ast}_{+},i^{\ast}_{+}+1}}{\sqrt{t}}+\frac{\mathcal{P}^{(2)}_{i^{\ast}_{-}-1,i^{\ast}_{-}}}{\sqrt{t}}+(\ref{b-hop1})+(\ref{b-hop2})+[((\ref{b-21})+(\ref{b-212}))\text{ or }((\ref{b-23})+(\ref{b-232}))]\,,

    where the first two terms contribute (once multiplied by t\sqrt{t}) to HΛ0H^{0}_{\Lambda} and the alternative “[((2.62)+(2.62)) or ((2.72)+(2.72))][((\ref{b-21})+(\ref{b-212}))\text{ or }((\ref{b-23})+(\ref{b-232}))]” depends on whether ~K,Q\widetilde{\mathcal{I}}^{*}_{K,Q} is a bulk- or a boundary-interval.

\Box

3 Operator norms and control of the flow

In this section we shall provide proofs of the following claims.

1) The block-diagonalization flow is well defined.

2) It yields quantitative information on the low energy spectrum of KN(t)K_{N}(t), as stated in the Theorem of Section 1.2; (see Theorem 3.4).

The main tool used in our proofs is induction in the steps (K,Q)(K,Q) of our block-diagonalization procedure. This induction is described in Theorem 3.2. The induction hypothesis used to carry out step (K,Q)(K,Q) consists of certain norm bounds on the effective interaction potentials appearing in step (K,Q)1(K,Q)_{-1} and of a lower bound on the spectral gap of the local Hamiltonian GK,QG_{\mathcal{I}^{*}_{K,Q}}. The induction step consists in showing that the same bounds then hold after step (K,Q)(K,Q).

In order to make the proof of Theorem 3.2 a little less heavy, some ingredients of our induction step are deferred to Lemma 3.1 and Lemma 3.3, where we carry out the induction step for some of the quantities appearing in Theorem 3.2, and to Sections 3.2.1 and 3.2.2, where we estimate norms of so-called “hooked terms”, starting from the norms of the interaction potentials involved in the “hooking”.

3.1 Gap estimate

In order to prove the lower bound on the spectral gap of the local Hamiltonian in step (K,Q)+1(K,Q)_{+1}, it is sufficient to bound the operator

P(K,Q)+1(+)(G(K,Q)+1E(K,Q)+1)P(K,Q)+1(+)P^{(+)}_{\mathcal{I}^{*}_{(K,Q)_{+1}}}\,(G_{{\mathcal{I}^{*}_{(K,Q)_{+1}}}}-E_{{\mathcal{I}^{*}_{(K,Q)_{+1}}}})\,P^{(+)}_{{\mathcal{I}^{*}_{(K,Q)_{+1}}}}\, (3.1)

from below, where the local ground-state energy is defined in (3.19). The argument is essentially the same as in [FP], but with some non-trivial twists caused by having to deal with macroscopic and microscopic quantities at the same time; see 3\mathfrak{I}3) below. We assume that, for all (K,Q)(K,Q)(K^{\prime},Q^{\prime})\,\preceq\,(K,Q),

VK,Q(K,Q)1tK116(K)2,VK,Q¯(K,Q)CεtK116(K)2,Cε:=(3+22A1ε),\|V^{(K^{\prime},Q^{\prime})_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}}\|\,\leq\frac{t^{\frac{K^{\prime}-1}{16}}}{(K^{\prime})^{2}}\,,\quad\|V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\|\,\leq C_{\varepsilon}\cdot\frac{t^{\frac{K^{\prime}-1}{16}}}{(K^{\prime})^{2}}\,,\quad C_{\varepsilon}:=\left(3+2\cdot\frac{2A_{1}}{\varepsilon}\right)\,, (3.2)

where A1A_{1} is a universal constant introduced in Lemma 3.3. The assumptions in (3.2) above are shown to hold within the proof by induction, in Theorem 3.2 and Lemma 3.3. Next, we describe some consequences of these assumptions which will be used later on.

  • 1\mathfrak{I}1)

    For (K,Q)(K,Q)(K^{\prime},Q^{\prime})\,\preceq\,(K,Q), VK,Q¯(K,Q)V^{(K,Q)}_{\overline{{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}} is a block-diagonalized potential by construction and – see (2.54) – corresponds to

    VK,Q¯(K,Q)\displaystyle V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}
    =\displaystyle= ω(VK,Q(K,Q)1)+PK,Q¯(+)PK,Q(+)[VK,Q(K,Q)1ω(VK,Q(K,Q)1)]PK,Q(+)PK,Q¯(+)\displaystyle\omega(V^{(K^{\prime},Q^{\prime})_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}})+P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,P^{(+)}_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}\,\,\Big{[}V^{(K^{\prime},Q^{\prime})_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}}-\omega(V^{(K^{\prime},Q^{\prime})_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}})\Big{]}\,P^{(+)}_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\quad\quad
    +PK,Q¯(+)(adZK,Q(𝒫i1,i(2)t))PK,Q¯(+)+PK,Q¯(+)(adZK,Q(𝒫i+,i++1(2)t))PK,Q¯(+)\displaystyle+P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,\Big{(}ad\,Z_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}(\,\frac{\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}}{\sqrt{t}})\Big{)}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}+P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,\Big{(}ad\,Z_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}(\,\frac{\mathcal{P}^{(2)}_{i^{*}_{+},i^{*}_{+}+1}}{\sqrt{t}})\Big{)}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}

    Furthermore, provided K,Q¯(K,Q)+1¯\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}\subset\overline{\mathcal{I}^{*}_{(K,Q)_{+1}}}, it is block-diagonal w.r.t. to the pair of projections P(K,Q)+1¯()P^{(-)}_{\overline{\mathcal{I}^{*}_{(K,Q)_{+1}}}}, P(K,Q)+1¯(+)P^{(+)}_{\overline{\mathcal{I}^{*}_{(K,Q)_{+1}}}}, thanks to PK,Q¯(+)P(K,Q)+1¯()=0P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,P^{(-)}_{\overline{\mathcal{I}^{*}_{(K,Q)_{+1}}}}=0 which follows easily from Definition 2.5 and the frustration free property of the AKLT model.

  • 2\mathfrak{I}2)

    Note that, except for the unperturbed Hamiltonian, H(K,Q)+10H_{\mathcal{I}^{*}_{(K,Q)_{+1}}}^{0}, the general term in (3.1) is given by

    {PK,Q¯(+)PK,Q(+)[VK,Q(K,Q)1ω(VK,Q(K,Q)1)]PK,Q(+)PK,Q¯(+)\displaystyle\Big{\{}P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,P^{(+)}_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}\,\,\Big{[}V^{(K^{\prime},Q^{\prime})_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}}-\omega(V^{(K^{\prime},Q^{\prime})_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}})\Big{]}\,P^{(+)}_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}} (3.4)
    +PK,Q¯(+)(adZK,Q(𝒫i1,i(2)t))PK,Q¯(+)+PK,Q¯(+)(adZK,Q(𝒫i+,i++1(2)t))PK,Q¯(+)}\displaystyle+P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\Big{(}ad\,Z_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}(\,\frac{\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}}{\sqrt{t}})\Big{)}P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}+P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\Big{(}ad\,Z_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}(\,\frac{\mathcal{P}^{(2)}_{i^{*}_{+},i^{*}_{+}+1}}{\sqrt{t}})\Big{)}P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,\Big{\}}\,\,\quad\quad (3.5)

    since ω(VK,Q¯(K,Q))=ω(VK,Q(K,Q)1)\omega(V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}})=\omega(V^{(K^{\prime},Q^{\prime})_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}}).

    We thus focus on (3.4)+(3.5)(\ref{uno-bis})+(\ref{due-bis}) and we observe that (see (1\mathfrak{I}1)))

    (3.4)+(3.5)\displaystyle(\ref{uno-bis})+(\ref{due-bis}) (3.6)
    =\displaystyle= PK,Q¯(+)(VK,Q¯(K,Q)ω(VK,Q(K,Q)1))PK,Q¯(+)\displaystyle P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,\,(V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,-\,\omega(V^{(K^{\prime},Q^{\prime})_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}}))\,P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\, (3.7)

    Henceforth, making use of the inequality

    PK,Q¯(+)1εHK,Q¯0P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\leq\frac{1}{\varepsilon}\,H^{0}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}} (3.8)

    (recall that ε\varepsilon is a lower bound on the spectral gap of HK,Q¯0H^{0}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,; see Theorem 1.2), which follows from the definitions of PK,Q¯(+)P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}} and HK,Q¯0H^{0}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,, we find that

    ±{(3.4)+(3.5)}\displaystyle\pm\{(\ref{uno-bis})+(\ref{due-bis})\} (3.9)
    =\displaystyle= ±{PK,Q¯(+)(VK,Q¯(K,Q)ω(VK,Q(K,Q)1))PK,Q¯(+)}\displaystyle\pm\{P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,\,(V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,-\,\omega(V^{(K^{\prime},Q^{\prime})_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}}))\,P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\} (3.10)
    \displaystyle\leq 2CεεtK116HK,Q¯0,\displaystyle\frac{2C_{\varepsilon}}{\varepsilon}\cdot t^{\frac{K^{\prime}-1}{16}}\,H^{0}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,, (3.11)

    where we have used that

    (3.10)[VK,Q¯(K,Q)+VK,Q(K,Q)1]1εHK,Q¯0\|(\ref{fin-est-penultimo})\|\leq\Big{[}\|V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\|+\|V^{(K^{\prime},Q^{\prime})_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}}\|\Big{]}\frac{1}{\varepsilon}\,H^{0}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,

    and the bounds in (3.2).

  • 3\mathfrak{I}3)

    Next, for 1QminQmax(N1)tK+11\leq Q_{\text{min}}\leq Q_{\text{max}}\leq(N-1)\,\sqrt{t}-K^{\prime}+1, we set :=Q=QminQmaxK,Q¯\mathscr{I}:=\bigcup\limits_{Q^{\prime}=Q_{\text{min}}}^{Q_{\text{max}}}\,\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}} and observe that

    Q=QminQmaxHK,Q¯0\displaystyle\sum_{Q^{\prime}=Q_{\text{min}}}^{Q_{\text{max}}}H^{0}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}} =\displaystyle= Q=QminQmaxi,i+1K,Q¯𝒫i,i+1(2)\displaystyle\sum_{Q^{\prime}=Q_{min}}^{Q_{max}}\,\sum_{i,i+1\,\in\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,\mathcal{P}^{(2)}_{i,i+1} (3.12)
    \displaystyle\leq (K+1)i,i+1𝒫i,i+1(2)\displaystyle(K^{\prime}+1)\sum_{i,i+1\,\in\mathscr{I}}\,\mathcal{P}^{(2)}_{i,i+1}\, (3.13)
    =\displaystyle= (K+1)H0.\displaystyle(K^{\prime}+1)\,H^{0}_{\mathscr{I}}\,. (3.14)
  • 4\mathfrak{I}4)

    Using the bound in (3.9)-(3.11) and inequalities (3.12)-(3.14), we conclude that

    ±{K,Q¯(K,Q)+1PK,Q¯(+)PK,Q(+)[VK,Q(K,Q)1ω(VK,Q(K,Q)1)]PK,Q(+)PK,Q¯(+)\displaystyle\pm\Big{\{}\sum_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}\,\subset\,\mathcal{I}^{*}_{(K,Q)_{+1}}}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,P^{(+)}_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}\,\,\Big{[}V^{(K^{\prime},Q^{\prime})_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}}-\omega(V^{(K^{\prime},Q^{\prime})_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}})\Big{]}\,P^{(+)}_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}
    +K,Q¯(K,Q)+1PK,Q¯(+)(adZK,Q(𝒫i1,i(2)t))PK,Q¯(+)+PK,Q¯(+)(adZK,Q(𝒫i+,i++1(2)t))PK,Q¯(+)}\displaystyle\quad+\sum_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}\,\subset\,\mathcal{I}^{*}_{(K,Q)_{+1}}}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\Big{(}adZ_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}(\,\frac{\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}}{\sqrt{t}})\Big{)}P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}+P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\Big{(}adZ_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}(\,\frac{\mathcal{P}^{(2)}_{i^{*}_{+},i^{*}_{+}+1}}{\sqrt{t}})\Big{)}P^{(+)}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\,\Big{\}}
    \displaystyle\leq 2CεεtK116(K+1)H(K,Q)+10.\displaystyle\frac{2C_{\varepsilon}}{\varepsilon}\cdot t^{\frac{K^{\prime}-1}{16}}\,(K^{\prime}+1)\,H^{0}_{\mathcal{I}^{*}_{(K,Q)_{+1}}}\,. (3.16)

The items discussed above are the ingredients of the proof of the following result (for more details concerning this proof we refer to [FP, Section 2.3]).

Lemma 3.1.

Assuming that the bound in (3.2) holds in step (K,Q)(K,Q) of the block-diagonalization, and choosing t>0t>0 so small that

{14Cεεt122Cεεt12l=3ltl216}>0,\Big{\{}{1}-\frac{4C_{\varepsilon}}{\varepsilon}\cdot t^{\frac{1}{2}}-\frac{2C_{\varepsilon}}{\varepsilon}\cdot t^{\frac{1}{2}}\sum_{l=3}^{\infty}l\cdot t^{\frac{l-2}{16}}\Big{\}}>0\,, (3.17)

the inequality

P(K,Q)+1(+)(G(K,Q)+1E(K,Q)+1)P(K,Q)+1(+)ε{14Cεεt122Cεεt12l=3ltl216}P(K,Q)+1(+)P^{(+)}_{\mathcal{I}^{*}_{(K,Q)_{+1}}}\,(G_{{\mathcal{I}^{*}_{(K,Q)_{+1}}}}-E_{{\mathcal{I}^{*}_{(K,Q)_{+1}}}})\,P^{(+)}_{{\mathcal{I}^{*}_{(K,Q)_{+1}}}}\geq\,\varepsilon\cdot\Big{\{}1-\frac{4C_{\varepsilon}}{\varepsilon}\cdot t^{\frac{1}{2}}-\frac{2C_{\varepsilon}}{\varepsilon}\cdot t^{\frac{1}{2}}\sum_{l=3}^{\infty}l\cdot t^{\frac{l-2}{16}}\Big{\}}\,P^{(+)}_{\mathcal{I}^{*}_{(K,Q)_{+1}}} (3.18)

holds, where

E(K,Q)+1:=tJ=1K1J,Q¯(K,Q)+1ω(VJ,Q¯(K,Q)).E_{\mathcal{I}^{*}_{(K,Q)_{+1}}}:=\,{\sqrt{t}}\sum_{J=1}^{K-1}\,\,\sum_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}\,\subset\,\mathcal{I}^{*}_{(K,Q)_{+1}}}\,\omega(V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{J,Q^{\prime}}}})\,. (3.19)

.

3.2 Preliminary estimates of the operator norms of potentials

3.2.1 Estimate of the “hooked” potentials

Assuming the induction hypothesis (3.2) and the bounds (3.78)-(3.76) proven in Lemma 3.3, we readily conclude that, for sufficiently small t>0t>0,

n=11n!adnZK,Q(VK,Q¯(K,Q)1)\displaystyle\Big{\|}\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}})\Big{\|} (3.20)
\displaystyle\leq CA12εtVK,Q(K,Q)1VK,Q¯(K,Q)1\displaystyle C\cdot A_{1}\cdot\frac{2}{\varepsilon}\cdot\sqrt{t}\cdot\|V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\|\cdot\|V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}\| (3.21)
\displaystyle\leq CεCA12εtVK,Q(K,Q)1VK,Q(K,Q)1,\displaystyle C_{\varepsilon}\cdot C\cdot A_{1}\cdot\frac{2}{\varepsilon}\cdot\sqrt{t}\cdot\|V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\|\cdot\|V^{(K^{\prime},Q^{\prime})_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}}\|\,, (3.22)

where CC is a universal constant and CεC_{\varepsilon} is defined in (3.2). Similarly, we can prove that

n=11n!adnZK,Q(VK,Q(K,Q)1)CA12εtVK,Q(K,Q)1VK,Q(K,Q)1\Big{\|}\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(V^{(K,Q)_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}})\Big{\|}\leq C\cdot A_{1}\cdot\frac{2}{\varepsilon}\cdot\sqrt{t}\cdot\|V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\|\cdot\|V^{(K,Q)_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}}\| (3.23)

and

n=11n!adnZK,Q(WK,Q(K,Q)1)CA12εtVK,Q(K,Q)1WK,Q(K,Q)1.\Big{\|}\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(W^{(K,Q)_{-1}}_{{\mathcal{I}_{K^{\prime},Q^{\prime}}}})\Big{\|}\leq C\cdot A_{1}\cdot\frac{2}{\varepsilon}\cdot\sqrt{t}\cdot\|V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\|\cdot\|W^{(K,Q)_{-1}}_{{\mathcal{I}_{K^{\prime},Q^{\prime}}}}\|\,. (3.24)

3.2.2 Estimate of the off-diagonal part of the hooked projections

In this section we assume (3.2) and (for K,Q=R,J\mathcal{I}_{K,Q}=\mathcal{I}_{R,J}) the gap bound stated in 𝒮2\mathcal{S}2 of Theorem 3.2 through Lemma 3.1, i.e.,

PR,J(+)(GR,JER,J)PR,J(+)ε2PR,J(+),P^{(+)}_{\mathcal{I}^{*}_{R,J}}(G_{\mathcal{I}^{*}_{R,J}}-E_{\mathcal{I}^{*}_{R,J}})P^{(+)}_{\mathcal{I}^{*}_{R,J}}\geq\frac{\varepsilon}{2}P^{(+)}_{\mathcal{I}^{*}_{R,J}}\,, (3.25)

then we prove that for tt small

PR,J¯(+)(adZR,J(𝒫i1,i(2)t))PR,J¯()𝒪(t14ε2VR,J(R,J)1).\Big{\|}P^{(+)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,\Big{(}ad\,Z_{\mathcal{I}^{*}_{R,J}}(\,{\frac{\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}}{\sqrt{t}}})\Big{)}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\Big{\|}\leq\mathcal{O}(\frac{t^{\frac{1}{4}}}{\varepsilon^{2}}\cdot\|V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\|)\,. (3.26)

From the definition of adad, and using 𝒫i1,i(2)PR,J¯()=0\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}=0 in the step from (3.27) to (3.28), we have

PR,J¯(+)(adZR,J(𝒫i1,i(2)t))PR,J¯()\displaystyle P^{(+)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,\Big{(}ad\,Z_{\mathcal{I}^{*}_{R,J}}(\,{\frac{\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}}{\sqrt{t}}})\Big{)}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}} =\displaystyle= PR,J¯(+)[ZR,J,𝒫i1,i(2)t]PR,J¯()\displaystyle P^{(+)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,\Big{[}\,Z_{\mathcal{I}^{*}_{R,J}}\,,\,{\frac{\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}}{\sqrt{t}}}\Big{]}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}} (3.27)
=\displaystyle= PR,J¯(+)𝒫i1,i(2)tZR,JPR,J¯()\displaystyle-P^{(+)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,{\frac{\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}}{\sqrt{t}}}\,Z_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}} (3.28)
=\displaystyle= j=1tj2PR,J¯(+)𝒫i1,i(2)t(ZR,J)jPR,J¯().\displaystyle-\sum_{j=1}^{\infty}{t^{\frac{j}{2}}}P^{(+)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,{\frac{\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}}{\sqrt{t}}}\,(Z_{\mathcal{I}^{*}_{R,J}})_{j}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,. (3.29)

The tail, starting from j=2j=2, of the series above, i.e.,

j=2tj2PR,J¯(+)𝒫i1,i(2)t(ZR,J)jPR,J¯(),-\sum_{j=2}^{\infty}{t^{\frac{j}{2}}}P^{(+)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,{\frac{\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}}{\sqrt{t}}}\,(Z_{\mathcal{I}^{*}_{R,J}})_{j}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,, (3.30)

is norm bounded777This can be actually shown within the proof of Lemma 3.3. by 𝒪(t1/2VR,J(R,J)12ε2)\mathcal{O}(t^{1/2}\cdot\frac{\|V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\|^{2}}{\varepsilon^{2}}). Henceforth, we can neglect it since the bound in (3.26) is fulfilled for the summand in (3.30) due to the assumption in (3.2). As for the leading quantity

t12PR,J¯(+)𝒫i1,i(2)t(ZR,J)1PR,J¯()\displaystyle-t^{\frac{1}{2}}P^{(+)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,\frac{\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}}{\sqrt{t}}\,(Z_{\mathcal{I}^{*}_{R,J}})_{1}\,P^{(-)}_{\overline{\mathcal{I}^{*}_{R,J}}} (3.31)
=\displaystyle= PR,J¯(+)𝒫i1,i(2)1GR,JER,JPR,J(+)VR,J(R,J)1PR,J()PR,J¯(),\displaystyle-P^{(+)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}\,\frac{1}{G_{\mathcal{I}^{*}_{R,J}}-E_{\mathcal{I}^{*}_{R,J}}}P^{(+)}_{\mathcal{I}^{*}_{R,J}}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,, (3.32)

(where (3.32) follows from (3.31) by using the definition in (2.38)) we exploit the resolvent identity

1GR,JER,J=1GR,JER,J+iδt+iδtGR,JER,J1GR,JER,J+iδt;\frac{1}{G_{\mathcal{I}^{*}_{R,J}}-E_{\mathcal{I}^{*}_{R,J}}}=\frac{1}{G_{\mathcal{I}^{*}_{R,J}}-E_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}+\frac{i\delta_{t}}{G_{\mathcal{I}^{*}_{R,J}}-E_{\mathcal{I}^{*}_{R,J}}}\,\frac{1}{G_{\mathcal{I}^{*}_{R,J}}-E_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}\,; (3.33)

here δt\delta_{t} is set equal to t14t^{\frac{1}{4}}. Next, from the estimate in (3.25), for tt sufficiently small, we can write

(3.32)\displaystyle(\ref{tre}) =\displaystyle= PR,J¯(+)𝒫i1,i(2)1GR,JER,J+iδtPR,J(+)VR,J(R,J)1PR,J()PR,J¯()\displaystyle-P^{(+)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}\,\frac{1}{G_{\mathcal{I}^{*}_{R,J}}-E_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}P^{(+)}_{\mathcal{I}^{*}_{R,J}}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}} (3.35)
+R1\displaystyle+R_{1}

with R1𝒪(δtε2VR,J(R,J)1)\|R_{1}\|\leq\mathcal{O}(\frac{\delta_{t}}{\varepsilon^{2}}\cdot\|V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\|). R1R_{1} is a remainder term which does not need further treatment since it fulfills the bound in (3.26). On the contrary, the first term requires some further manipulation: namely we start implementing a Neumann expansion of (GR,JER,J+iδt)1(G_{\mathcal{I}^{*}_{R,J}}-E_{\mathcal{I}^{*}_{R,J}}+i\delta_{t})^{-1} (see (2.34), (2.37), and (1\mathfrak{I}1)), with obvious adaptation of the indexes, in order to follow the computation below)

1GR,JER,J+iδtPR,J(+)\displaystyle\frac{1}{G_{\mathcal{I}^{*}_{R,J}}-E_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}P^{(+)}_{\mathcal{I}^{*}_{R,J}} (3.36)
=\displaystyle= 1PR,J(+)(GR,JER,J+iδt)PR,J(+)PR,J(+)\displaystyle\frac{1}{P^{(+)}_{\mathcal{I}^{*}_{R,J}}(G_{\mathcal{I}^{*}_{R,J}}-E_{\mathcal{I}^{*}_{R,J}}+i\delta_{t})P^{(+)}_{\mathcal{I}^{*}_{R,J}}}P^{(+)}_{\mathcal{I}^{*}_{R,J}} (3.37)
=\displaystyle= 1HR,J0+iδtPR,J(+)\displaystyle\frac{1}{H^{0}_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}P^{(+)}_{\mathcal{I}^{*}_{R,J}}
+1HR,J0+iδtj=1PR,J(+)×\displaystyle+\frac{1}{H^{0}_{\mathcal{I}^{*}_{R,J}}+i{\delta_{t}}}\sum_{j=1}^{\infty}P^{(+)}_{\mathcal{I}^{*}_{R,J}}\times
×{(tJ=1K1K,Q¯R,JPR,J(+)[VK,Q¯(R,J)1ω(VK,Q(K,Q)1)𝟙]PR,J(+))1HR,J0+iδt}jPR,J(+)\displaystyle\quad\times\Big{\{}\,\Big{(}\,-{\sqrt{t}}\sum_{J=1}^{K-1}\,\,\sum_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}\subset\mathcal{I}^{*}_{R,J}}P^{(+)}_{\mathcal{I}^{*}_{R,J}}\,\Big{[}V^{(R,J)_{-1}}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}}-\omega(V^{(K^{\prime},Q^{\prime})_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}})\mathbbm{1}\Big{]}\,P^{(+)}_{\mathcal{I}^{*}_{R,J}}\,\Big{)}\frac{1}{H^{0}_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}\Big{\}}^{j}\,P^{(+)}_{\mathcal{I}^{*}_{R,J}}

that we combine with (3.16) so as to get the bound

(3.2.2)𝒪(tε2).\|\,(\ref{rem})\,\|\leq\mathcal{O}(\frac{\sqrt{t}}{\varepsilon^{2}})\,. (3.40)

Therefore we can split the expression as follows

PR,J¯(+)𝒫i1,i(2)1GR,JER,J+iδtPR,J(+)VR,J(R,J)1PR,J()PR,J¯()\displaystyle-P^{(+)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}\,\frac{1}{G_{\mathcal{I}^{*}_{R,J}}-E_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}P^{(+)}_{\mathcal{I}^{*}_{R,J}}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}} (3.41)
=\displaystyle= PR,J¯(+)𝒫i1,i(2)1HR,J0+iδtPR,J(+)VR,J(R,J)1PR,J()PR,J¯()\displaystyle-P^{(+)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}\,\,\frac{1}{H^{0}_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}\,P^{(+)}_{\mathcal{I}^{*}_{R,J}}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}} (3.43)
+R2\displaystyle+R_{2}

where R2𝒪(tε2VR,J(R,J)1)\|R_{2}\|\leq\mathcal{O}(\frac{\sqrt{t}}{\varepsilon^{2}}\cdot\|V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\|). Next, we discard R2R_{2} and substitute PR,J(+)=𝟙PR,J()P^{(+)}_{\mathcal{I}^{*}_{R,J}}=\mathbbm{1}-P^{(-)}_{\mathcal{I}^{*}_{R,J}} in (3.43); the latter expression reads

PR,J¯(+)𝒫i1,i(2)1HR,J0+iδtPR,J(+)VR,J(R,J)1PR,J()PR,J¯()\displaystyle-P^{(+)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}\,\frac{1}{H^{0}_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}P^{(+)}_{\mathcal{I}^{*}_{R,J}}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}} (3.44)
=\displaystyle= PR,J¯(+)𝒫i1,i(2)1HR,J0+iδtVR,J(R,J)1PR,J()PR,J¯()\displaystyle-P^{(+)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,{\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}}\,\frac{1}{H^{0}_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}} (3.46)
+PR,J¯(+)𝒫i1,i(2)1HR,J0+iδtPR,J()VR,J(R,J)1PR,J()PR,J¯().\displaystyle+P^{(+)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,{\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}}\,\frac{1}{H^{0}_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,.

We notice the identity

1HR,J0+iδtPR,J()VR,J(R,J)1PR,J()PR,J¯()=1iδtPR,J()VR,J(R,J)1PR,J()PR,J¯()\frac{1}{H^{0}_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}\,{P^{(-)}_{\mathcal{I}^{*}_{R,J}}}V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}=\frac{1}{i\delta_{t}}P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}} (3.47)

as a consequence of HR,J0PR,J()=0H^{0}_{\mathcal{I}^{*}_{R,J}}P^{(-)}_{\mathcal{I}^{*}_{R,J}}=0; next, by exploiting (1.8), we can estimate

PR,J()VR,J(R,J)1PR,J()PR,J¯()\displaystyle P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\, =\displaystyle= ω(VR,J(R,J)1)PR,J()PR,J¯()+R3\displaystyle\,\omega(V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}})\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}+R_{3} (3.48)
=\displaystyle= ω(VR,J(R,J)1)PR,J¯()+R3\displaystyle\omega(V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}})\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}+R_{3} (3.49)

where R3𝒪(3t13VR,J(R,J)1)\|R_{3}\|\leq\mathcal{O}(3^{-\frac{\sqrt{t^{-1}}}{3}}\cdot\|V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\|) and ω(VR,J(R,J)1)\omega(V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}) is defined in (1.7). Hence, since 𝒫i1,i(2)PR,J¯()=0\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}{P^{(-)}_{\overline{\mathcal{I}^{*}_{R,J}}}}=0, we deduce that

(3.46)𝒪(1δt3t13VR,J(R,J)1)𝒪(t1/2VR,J(R,J)1).\|(\ref{remainder-2})\|\leq\mathcal{O}(\frac{1}{\delta_{t}}\cdot 3^{-\frac{\sqrt{t^{-1}}}{3}}\cdot\|V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\|)\leq\mathcal{O}(t^{1/2}\cdot\|V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\|)\,. (3.50)

The expression in (3.46), i.e.,

PR,J¯(+)𝒫i1,i(2)1HR,J0+iδtVR,J(R,J)1PR,J()PR,J¯(),-P^{(+)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,{\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}}\,\frac{1}{H^{0}_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{{\overline{\mathcal{I}^{*}_{R,J}}}}\,, (3.51)

is to be controlled now. For this purpose, we make use of

1HR,J0+iδt=i0t13ei(HR,Jo+iδt)s𝑑sit13+ei(HR,Jo+iδt)s𝑑s\frac{1}{H^{0}_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}=-i\,\int_{0}^{t^{-\frac{1}{3}}}\,e^{i(H^{o}_{\mathcal{I}^{*}_{R,J}}+i\delta_{t})s}\,ds-i\,\int_{t^{-\frac{1}{3}}}^{+\infty}\,e^{i(H^{o}_{\mathcal{I}^{*}_{R,J}}+i\delta_{t})s}\,ds (3.52)

and define

R4:=iPR,J¯(+)𝒫i1,i(2)t1/3+ei(HR,J0+iδt)s𝑑sVR,J(R,J)1PR,J()PR,J¯()R_{4}:=i\,P^{(+)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}\,\int_{t^{-1/3}}^{+\infty}\,e^{i(H^{0}_{\mathcal{I}^{*}_{R,J}}+i\delta_{t})s}\,ds\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{\overline{\mathcal{I}^{*}_{R,J}}} (3.53)

with

R4𝒪(eδtt13δtVR,J(R,J)1)𝒪(t14VR,J(R,J)1).\|R_{4}\|\leq\mathcal{O}(\frac{e^{-\delta_{t}\cdot t^{-\frac{1}{3}}}}{\delta_{t}}\|V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\|)\leq\mathcal{O}(t^{\frac{1}{4}}\,\|V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\|)\,.

Then, by using (3.52), we can write

PR,J¯(+)𝒫i1,i(2)1HR,J0+iδtVR,J(R,J)1PR,J()PR,J¯()R4\displaystyle-P^{(+)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}\,\frac{1}{H^{0}_{\mathcal{I}^{*}_{R,J}}+i\delta_{t}}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{\overline{\mathcal{I}^{*}_{R,J}}}-R_{4} (3.54)
=\displaystyle= i0t1/3PR,J¯(+)𝒫i1,i(2)ei(HR,J0+iδt)sVR,J(R,J)1PR,J()PR,J¯()𝑑s\displaystyle i\,\int_{0}^{t^{-1/3}}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}\,e^{i(H^{0}_{\mathcal{I}^{*}_{R,J}}+i\delta_{t})s}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,ds (3.55)
=\displaystyle= i0t1/3PR,J¯(+)𝒫i1,i(2)eδtseiHR,J0sVR,J(R,J)1eiHR,J0sPR,J¯()𝑑s\displaystyle i\,\int_{0}^{t^{-1/3}}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}\,e^{-\delta_{t}\cdot s}\,e^{i\cdot H^{0}_{\mathcal{I}^{*}_{R,J}}\cdot s}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,e^{-i\cdot H^{0}_{\mathcal{I}^{*}_{R,J}}\cdot s}P^{(-)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,ds (3.56)
=\displaystyle= i0t1/3PR,J¯(+)eδts[𝒫i1,i(2),eiHR,J0sVR,J(R,J)1eiHR,J0s]PR,J¯()𝑑s,\displaystyle i\,\int_{0}^{t^{-1/3}}\,P^{(+)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,\,e^{-\delta_{t}\cdot s}\,\Big{[}\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}\,,\,e^{i\cdot H^{0}_{\mathcal{I}^{*}_{R,J}}\cdot s}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,e^{-i\cdot H^{0}_{\mathcal{I}^{*}_{R,J}}\cdot s}\Big{]}\,P^{(-)}_{\overline{\mathcal{I}^{*}_{R,J}}}\,ds\,, (3.57)

where from from (3.55) to (3.56) we have used

PR,J()=eiHR,J0sPR,J()P^{(-)}_{\mathcal{I}^{*}_{R,J}}=e^{-i\cdot H^{0}_{\mathcal{I}^{*}_{R,J}}\cdot s}P^{(-)}_{\mathcal{I}^{*}_{R,J}}

and the frustration-free property of the unperturbed Hamiltonian, which in turn implies that PR,J()PR,J¯()=PR,J¯()P^{(-)}_{\mathcal{I}^{*}_{R,J}}\,P^{(-)}_{\overline{\mathcal{I}^{*}_{R,J}}}=P^{(-)}_{\overline{\mathcal{I}^{*}_{R,J}}}, and from (3.56) to (3.57) we have used

𝒫i1,i(2)PR,J¯()=0.\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}P^{(-)}_{\overline{\mathcal{I}^{*}_{R,J}}}=0\,.

Our last tool is the Lieb-Robinson bound (1.10), by which for tt sufficiently small we can estimate (recall δt=t14\delta_{t}=t^{-\frac{1}{4}}),

(3.54)\displaystyle\|(\ref{leading-2})\| \displaystyle\leq t1/4sup0st1/3[𝒫i1,i(2),eiHR,J0sVR,J(R,J)1eiHR,J0s]\displaystyle t^{-1/4}\cdot\sup_{0\leq s\leq t^{-1/3}}\Big{\|}\,\Big{[}\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}\,,\,e^{i\cdot H^{0}_{\mathcal{I}^{*}_{R,J}}\cdot s}\,V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\,e^{-i\cdot H^{0}_{\mathcal{I}^{*}_{R,J}}\cdot s}\Big{]}\,\Big{\|} (3.58)
\displaystyle\leq t1/4.4𝒫i1,i(2)VR,J(R,J)1F0C1e[d(i,R,J)2Φ1C1t1/3]\displaystyle t^{-1/4}\cdot\,.\frac{4\,\|\mathcal{P}^{(2)}_{i^{*}_{-}-1,i^{*}_{-}}\|\cdot\|V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\|\cdot\|{F_{0}}\|}{C_{1}}\cdot e^{-[d(i^{*}_{-}\,,\,\mathcal{I}_{R,J})-2\,\|\Phi\|_{1}\cdot C_{1}\cdot t^{-1/3}]} (3.59)
\displaystyle\leq et14VR,J(R,J)1\displaystyle e^{-\frac{\sqrt{t^{-1}}}{4}}\cdot\|V^{(R,J)_{-1}}_{\mathcal{I}_{R,J}}\| (3.60)

where d(i,R,J)=t13d(i^{*}_{-}\,,\,\mathcal{I}_{R,J})=\frac{\sqrt{t^{-1}}}{3}, and C1C_{1}, Φ1\|\Phi\|_{1} and F0F_{0} are positive constants defined in Section 1.1.2.

This concludes the proof of the bound in (3.26).

3.3 Main theorem

We recall that the first step of the block-diagonalization is associated with the pair (1,2)(1,2). By definition (1,2)1=(0,N)(1,2)_{-1}=(0,N), moreover the potentials VR,J(0,N)V^{(0,N)}_{\mathcal{I}_{R,J}}, with R,Jbulk\mathcal{I}_{R,J}\in\mathfrak{I}_{\text{bulk}}, and WR,J(0,N)W^{(0,N)}_{\mathcal{I}_{R,J}}, with R,Jb.dry\mathcal{I}_{R,J}\in\mathfrak{I}_{\text{b.dry}}, coincide with the operators VR,JV_{\mathcal{I}_{R,J}} appearing in the bare Hamiltonian KΛ(t)K_{\Lambda}(t) (see (2.8)).

Theorem 3.2.

There exists t¯>0\bar{t}>0 independent of NN, such that for all |t|<t¯|t|<\bar{t}, for any (K^,Q^)((N1)t,1)1(\hat{K},\hat{Q})\preceq((N-1)\cdot\sqrt{t},1)_{-1}, the Hamiltonians GK^,Q^G_{\mathcal{I}^{*}_{\hat{K},\hat{Q}}} are well defined, and

  1. 𝒮1)\mathcal{S}1)

    for any interval R,J\mathcal{I}_{R,J}, with R1R\geq 1, the following operator norms estimates hold

    1. (a)

      VR,J(K^,Q^)tR116R2\|V^{(\hat{K},\hat{Q})}_{\mathcal{I}_{R,J}}\|\leq\frac{t^{\frac{R-1}{16}}}{R^{2}}\, for (R,J)(K^,Q^)(R,J)\,\succ\,(\hat{K},\hat{Q}),

    2. (b)

      WR,J(K^,Q^)tR116R2\|W^{(\hat{K},\hat{Q})}_{\mathcal{I}_{R,J}}\|\leq\frac{t^{\frac{R-1}{16}}}{R^{2}}\,,

  2. 𝒮2)\mathcal{S}2)

    let (K^,Q^)+1(\mathcal{I}^{*}_{\hat{K},\hat{Q}})_{+1} be the interval of type \mathcal{I}^{*} associated with the pair (K^,Q^)+1(\hat{K},\hat{Q})_{+1}, then the Hamiltonian G(K^,Q^)+1G_{(\mathcal{I}^{*}_{\hat{K},\hat{Q}})_{+1}} has a spectral gap Δ(K^,Q^)+1\Delta_{(\mathcal{I}^{*}_{\hat{K},\hat{Q}})_{+1}} above its ground-state energy bounded below by ε2,\frac{\varepsilon}{2}, where GK,QG_{\mathcal{I}^{*}_{K,Q}} is defined in (2.34) for K>1K>1, and G1,Q:=H1,Q0G_{\mathcal{I}^{*}_{1,Q}}:=H^{0}_{\mathcal{I}^{*}_{1,Q}}.

Proof

The inductive proof in the pair index (K,Q)(K,Q) is implemented as follows. We consider a fixed (R,J)(R,J) and we show that 𝒮1)\mathcal{S}1) and 𝒮2)\mathcal{S}2) hold from (K,Q)=(0,N)(K,Q)=(0,N) up to (K,Q)=((N1)t,1)1(K,Q)=((N-1)\cdot\sqrt{t},1)_{-1}. In turn, by assuming that 𝒮1)\mathcal{S}1) holds for all VR,J(K,Q)V^{(K^{\prime},Q^{\prime})}_{\mathcal{I}_{R,J}}, WR,J(K,Q)W^{(K^{\prime},Q^{\prime})}_{\mathcal{I}_{R,J}} with (K,Q)(K,Q)(K^{\prime},Q^{\prime})\prec(K,Q) and 𝒮2)\mathcal{S}2) for all (K,Q)(K,Q)(K^{\prime},Q^{\prime})\prec(K,Q), the same properties are proven to hold for VR,J(K,Q)V^{(K,Q)}_{\mathcal{I}_{R,J}}, WR,J(K,Q)W^{(K,Q)}_{\mathcal{I}_{R,J}}, and for G(K,Q)+1G_{(\mathcal{I}^{*}_{K,Q})_{+1}}. Next we invoke Lemma 3.3 and rigorously define ZK,QZ_{\mathcal{I}^{*}_{K,Q}} and KΛ(K,Q)K_{\Lambda}^{(K,Q)}.

In order to check that 𝒮1)\mathcal{S}1) and 𝒮2)\mathcal{S}2) are verified at the initial step corresponding to (K^,Q^)=(0,N)(\hat{K},\hat{Q})=(0,N), we observe since that 𝒮1)\mathcal{S}1) can be verified by direct computation, because

V1,J(0,N)=V1,J1,W1,J(0,N)=V1,J1\|V^{(0,N)}_{\mathcal{I}_{1,J}}\|=\|V_{\mathcal{I}_{1,J}}\|\leq 1\,\quad,\quad\|W^{(0,N)}_{\mathcal{I}_{1,J}}\|=\|V_{\mathcal{I}_{1,J}}\|\leq 1

and VR,J(0,N)=WR,J(0,N)=VR,J=0\|V_{\mathcal{I}_{R,J}}^{(0,N)}\|=\|W^{(0,N)}_{\mathcal{I}_{R,J}}\|=\|V_{\mathcal{I}_{R,J}}\|=0 otherwise; then 𝒮1)\mathcal{S}1) follows. As far as 𝒮2)\mathcal{S}2) is concerned, the statement is true given that (0,N)+1=(1,2)(0,N)_{+1}=(1,2) and G1,2=H1,2(0)G_{\mathcal{I}^{*}_{1,2}}=H^{(0)}_{\mathcal{I}^{*}_{1,2}}.

Within the single induction step, the proof consists of different parts where the allowed interval of t(0)t(\geq 0) is progressively reduced. One of these parts is provided by Lemma 3.3. The induction ensures that the same tt-interval works for all steps.

Concerning 𝒮1)\mathcal{S}1), we show the proof for the potentials VR,J(K^,Q^)V^{(\hat{K},\hat{Q})}_{\mathcal{I}_{R,J}}; with minor modifications the same result can be proved for the potentials WR,J(K^,Q^)W^{(\hat{K},\hat{Q})}_{\mathcal{I}_{R,J}}.

Induction step in the proof of 𝒮1)\mathcal{S}1)

In order to prove 𝒮1)\mathcal{S}1) in step (K^,Q^)(\hat{K},\hat{Q}), we re-expand down to (1,2)(1,2), step by step, i.e., we relate the norm of VR,J(K,Q)V^{(K,Q)}_{\mathcal{I}_{R,J}} to the ones of the operators in step (K,Q)1(K,Q)_{-1} in terms of which VR,J(K,Q)V^{(K,Q)}_{\mathcal{I}_{R,J}} is expressed according to the algorithm. It is then clear that for most of the steps the norm is preserved, i.e., VR,J(K,Q)=VR,J(K,Q)1\|V^{(K,Q)}_{\mathcal{I}_{R,J}}\|=\|V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}\|, and only for special steps we have nontrivial relations.

We recall that, due to the rules of the algorithm displayed in Definition 2.15, a potential of the type VR,J(K^,Q^)V^{(\hat{K},\hat{Q})}_{{\mathcal{I}_{R,J}}} has been defined only for (R,J)(K^,Q^)(R,J)\succ(\hat{K},\hat{Q}); henceforth the following constraints hold: R>K^R>\hat{K} or R=K^R=\hat{K} and J>Q^J>\hat{Q}. In addition, we observe that in view of the prescribed enlargement in Definition 2.9, the RR cannot be equal to 22.

We observe that if R=1R=1 the proof is straightforwad by taking into account that (1,J)(K^,Q^)(1,J)\succ(\hat{K},\hat{Q}) and by applying a-1) in Definition 2.15 repeatedly, so as to get

V1,J(K^,Q^)=V1,J(0,N)=1.\|V^{(\hat{K},\hat{Q})}_{\mathcal{I}_{1,J}}\|=\|V^{(0,N)}_{\mathcal{I}_{1,J}}\|=1\,. (3.61)

General case (R3R\geq 3)

We study the re-expansion step (K,Q)(K,Q) to (K,Q)1(K,Q)_{-1}, by considering various cases with the help of Definition 2.15 (recall that VR,J(K^,Q^)V^{(\hat{K},\hat{Q})}_{\mathcal{I}_{R,J}} is defined for RK^R\geq\hat{K}):

  • 1)

    in case a-1), and, similarly, in case c-1) along with the constraint i+,i~K,Qi_{+},i_{-}\notin\widetilde{\mathcal{I}}^{*}_{K,Q} where i+,ii_{+},i_{-} are the endpoints of R,J\mathcal{I}_{R,J}, it turns out that

    VR,J(K,Q)=VR,J(K,Q)1\|V^{(K,Q)}_{\mathcal{I}_{R,J}}\|=\|V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}\|\, (3.62)

    for which we notice that: in case c-1) only (2.62) contributes thanks to i+,i~K,Qi^{+},i_{-}\notin\widetilde{\mathcal{I}}^{*}_{K,Q}; in case a-1) the equality is straightforward.

  • 2-i)

    in case c-1) along with the property that ~K,Q\widetilde{\mathcal{I}}^{*}_{K,Q} contains one amongst i+,ii_{+},i_{-} (the endpoints of R,J\mathcal{I}_{R,J}), the contributions to the re-expansion are given in (2.62) and (2.62), from which we have

    VR,J(K,Q)\displaystyle\|V^{(K,Q)}_{\mathcal{I}_{R,J}}\| \displaystyle\leq VR,J(K,Q)1\displaystyle\|V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}\| (3.65)
    +K,Q[𝒢R,J(K,Q)]1n=11n!adnZK,Q(VK,Q(K,Q)1)\displaystyle+\sum_{\mathcal{I}_{K^{\prime},Q^{\prime}}\in[\mathcal{G}^{(K,Q)}_{\mathcal{I}_{R,J}}]_{1}}\,\Big{\|}\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(V^{(K,Q)_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}})\Big{\|}
    +K,Q[𝒢R,J(K,Q)]2n=11n!adnZK,Q(VK,Q¯(K,Q)1)\displaystyle+\sum_{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}\in[\mathcal{G}^{(K,Q)}_{\mathcal{I}_{R,J}}]_{2}}\,\Big{\|}\sum_{n=1}^{\infty}\frac{1}{n!}\,ad^{n}Z_{\mathcal{I}^{*}_{K,Q}}(V^{(K,Q)_{-1}}_{\overline{\mathcal{I}^{*}_{K^{\prime},Q^{\prime}}}})\Big{\|}
  • 2-ii)

    in case c-1) along with the property that ~K,Q\widetilde{\mathcal{I}}^{*}_{K,Q} contains both i+i_{+} and ii_{-} (the endpoints of R,J\mathcal{I}_{R,J}) the re-expansion consists of terms (2.62), (2.62), (2.62), (2.62), (2.62), from which we have

    VR,J(K,Q)VR,J(K,Q)1\displaystyle\|V^{(K,Q)}_{\mathcal{I}_{R,J}}\|\leq\|V^{(K,Q)_{-1}}_{\mathcal{I}_{R,J}}\| (3.66)
    +(2.62)+(2.62)+(2.62)+(2.62)+(2.62).\displaystyle+\|(\ref{Valgo67})\|+\|(\ref{expdecayerror})\|+\|(\ref{lshighorder})\|+\|(\ref{b-21})\|+\|(\ref{b-212})\|\,. (3.67)

The control of (3.65)+(3.65)(\ref{hook-V})+(\ref{hook-V^*}) relies on the computations in Section 3.2.1 together with the assumption of 𝒮1)\mathcal{S}1) in step (K,Q)1(K,Q)_{-1}; hence we can bound as follows

(3.65)+(3.65)\displaystyle(\ref{hook-V})+(\ref{hook-V^*}) \displaystyle\leq CεCA1tK=RK2R1VK,Q(K,Q)1VK,Q(K,Q)1\displaystyle C_{\varepsilon}\cdot C\cdot A_{1}\cdot\sqrt{t}\sum_{K^{\prime}=R-K-2}^{R-1}\,\|V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\|\cdot\|V^{(K,Q)_{-1}}_{\mathcal{I}_{K^{\prime},Q^{\prime}}}\| (3.68)
\displaystyle\leq CεCA1tm=0K1tK116K2tRK+m316(RK2+m)2\displaystyle C_{\varepsilon}\cdot C\cdot A_{1}\cdot\sqrt{t}\sum_{m=0}^{K-1}\,\frac{t^{\frac{K-1}{16}}}{K^{2}}\cdot\frac{t^{\frac{R-K+m-3}{16}}}{(R-K-2+m)^{2}} (3.69)
=\displaystyle= CεCA1ttR416m=0K1tm16K2(RK2+m)2\displaystyle C_{\varepsilon}\cdot C\cdot A_{1}\cdot\sqrt{t}\cdot t^{\frac{R-4}{16}}\sum_{m=0}^{K-1}\,\frac{t^{\frac{m}{16}}}{K^{2}\cdot(R-K-2+m)^{2}} (3.70)
\displaystyle\leq Cεt516tR116K2(RK2)2,\displaystyle C^{\prime}_{\varepsilon}\cdot t^{\frac{5}{16}}\cdot\frac{t^{\frac{R-1}{16}}}{K^{2}\cdot(R-K-2)^{2}}\,, (3.71)

where CεC_{\varepsilon} (see Lemma 3.3) and CεC^{\prime}_{\varepsilon} are constants depending on ε\varepsilon.

We can bound the sum of terms in (3.67)(\ref{tilde-case}) by

(3.67)Cε′′t14tR316(R2)2=Cε′′t18tR116(R2)2.\|(\ref{tilde-case})\|\leq C^{\prime\prime}_{\varepsilon}\cdot t^{\frac{1}{4}}\cdot\frac{t^{\frac{R-3}{16}}}{(R-2)^{2}}=C^{\prime\prime}_{\varepsilon}\cdot t^{\frac{1}{8}}\cdot\frac{t^{\frac{R-1}{16}}}{(R-2)^{2}}\,. (3.72)

for some ε\varepsilon-dependent constant Cε′′C_{\varepsilon}^{\prime\prime}.

We observe that, at fixed KK, the occurrence in 2-i) takes place only twice, whereas the one described in 2-ii) happens once and only for K=R2K=R-2. In conclusion, starting from (K^,Q^)(\hat{K},\hat{Q}) and re-expanding back down to level (0,N)(0,N), the following estimate holds provided tt is sufficiently small and by using the input VR,J(0,N)=0\|V^{(0,N)}_{\mathcal{I}_{R,J}}\|=0 for R>1R>1:

VR,J(K^,Q^)\displaystyle\|V^{(\hat{K},\hat{Q})}_{\mathcal{I}_{R,J}}\| \displaystyle\leq VR,J(0,N)\displaystyle\|V^{(0,N)}_{\mathcal{I}_{R,J}}\| (3.74)
+K=1R32Cεt516tR116K2(RK2)2+Cε′′t18tR116(R2)2\displaystyle+\sum_{K=1}^{R-3}2\cdot C^{\prime}_{\varepsilon}\cdot t^{\frac{5}{16}}\cdot\frac{t^{\frac{R-1}{16}}}{K^{2}\cdot(R-K-2)^{2}}+C_{\varepsilon}^{\prime\prime}\cdot t^{\frac{1}{8}}\cdot\frac{t^{\frac{R-1}{16}}}{(R-2)^{2}}
\displaystyle\leq tR116R2,\displaystyle\frac{t^{\frac{R-1}{16}}}{R^{2}}\,, (3.75)

where, in the step from (3.74) to (3.75), we can take advantage of the extra-factors t516t^{\frac{5}{16}} and t14t^{\frac{1}{4}}.

Induction step in the proof of 𝒮2)\mathcal{S}2)

By means of 𝒮1)\mathcal{S}1) in step (K^,Q^)(\hat{K},\hat{Q}) that we have just proven, and assuming 𝒮2)\mathcal{S}2) in step (K^,Q^)1(\hat{K},\hat{Q})_{-1}, the required property is a consequence of Lemma 3.1. \Box

In the next lemma, we derive the estimate of the operator norm of the bulk potentials after the block-diagonalization. We recall that, by construction (see the algorithm in Definition 2.15), each block-diagonalized (bulk) potential does not change in the successive steps of the flow.

Lemma 3.3.

Assume that t>0t>0 is sufficiently small, VK,Q(K,Q)1tK116K2\|V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\|\leq\frac{t^{\frac{K-1}{16}}}{K^{2}}, and ΔK,Qε2\Delta_{\mathcal{I}_{K,Q}}\geq\frac{\varepsilon}{2}. Then, for arbitrary NN, K1K\geq 1, and Q1Q\geq 1, the inequalities

ZK,QA1t2εVK,Q(K,Q)1\|Z_{\mathcal{I}^{*}_{K,Q}}\|\leq A_{1}\cdot\sqrt{t}\cdot\frac{2}{\varepsilon}\,\|V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\| (3.76)
j=2tj12(VK,Q(K,Q)1)jdiagDεtVK,Q(K,Q)1\sum_{j=2}^{\infty}t^{\frac{j-1}{2}}\,\|(V^{(K,Q)_{-1}}_{\mathcal{I}^{*}_{K,Q}})^{\text{diag}}_{j}\|\leq D_{\varepsilon}\cdot\sqrt{t}\,\|V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\| (3.77)
VK,Q¯(K,Q)CεVK,Q(K,Q)1\|V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{K,Q}}}\|\leq C_{\varepsilon}\,\|V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\|\, (3.78)

hold true for some universal constant A1A_{1}, for Cε:=3+2A12εC_{\varepsilon}:=3+2\cdot A_{1}\cdot\frac{2}{\varepsilon}, and DεD_{\varepsilon} an ε\varepsilon-dependent constant.

Proof In order to state the inequalities in (3.76) and (3.77) we can essentially proceed as in [FP, Lemma A.3]. However, here we make the gap (i.e., ε\varepsilon) dependence of our constants more explicit; the sufficiently small tt is eventually ε\varepsilon-dependent. The bound in (3.78) is then obtained from (1\mathfrak{I}1)) as follows:

VK,Q¯(K,Q)VK,Q(K,Q)1+2VK,Q(K,Q)1+2tZK,Q\|V^{(K,Q)}_{\overline{\mathcal{I}^{*}_{K,Q}}}\|\leq\|V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\|\,+2\|V^{(K,Q)_{-1}}_{\mathcal{I}_{K,Q}}\|+\frac{2}{\sqrt{t}}\|Z_{\mathcal{I}^{*}_{K,Q}}\|\,

which we combine with (3.76). \Box

We can now prove the main result of the paper.

Theorem 3.4.

There exists some t¯>0\bar{t}>0 independent of NN such that, for any coupling constant tt\in\mathbb{R} with |t|<t¯|t|<\bar{t}, and for all 0<N<0<N<\infty,

  1. (i)

    the spectrum of KΛ(t)K_{\Lambda}(t) is contained in two disjoint, tt-dependent regions, σ+\sigma^{+} and σ\sigma^{-}, separated by a uniformly positive gap ΔΛ(t)ε4\Delta_{\Lambda}(t)\geq\frac{\varepsilon}{4}, with ε\varepsilon independent of NN, as specified in Theorem 1.2; i.e., EE′′>ΔΛ(t)E^{\prime}-E^{\prime\prime}>\Delta_{\Lambda}(t), for all Eσ+E^{\prime}\in\sigma^{+} and all E′′σE^{\prime\prime}\in\sigma^{-};

  2. (ii)

    for any d[1,N2)d\in\mathbb{N}\cap[1\,,\,\frac{N}{2}), the eigenspace corresponding to the eigenvalues contained in σ\sigma^{-} is four-dimensional; the gaps between these eigenvalues coincide with the gaps between the eigenvalues of the symmetric matrix

    PΛ()(ti=1dVi,i+1+ti=NdN1Vi,i+1)PΛ(),P^{(-)}_{\Lambda}\,\Big{(}\,t\sum_{i=1}^{d}\,V_{i,i+1}+\,t\,\sum_{i=N-d}^{N-1}V_{i,i+1}\Big{)}\,P^{(-)}_{\Lambda}\,, (3.79)

    up to corrections bounded by

    |t|3(d1)+o(|t|).|t|\cdot 3^{-(d-1)}\,+\,o(|t|)\,.

Proof. As in the rest of this section, we assume that t>0t>0, without loss of generality. By using the results of Theorem 3.2 (combined with Lemma 3.3), in step (K,Q)𝐟:=((N1)t,1)1(K,Q)^{{\bf{f}}}:=((N-1)\cdot\sqrt{t},1)_{-1}, we obtain the transformed Hamiltonian

KΛ(K,Q)𝐟(t)\displaystyle K_{\Lambda}^{\,(K,Q)^{{\bf{f}}}}(t) =\displaystyle= HΛ0\displaystyle H^{0}_{\Lambda} (3.82)
+tQV1,Q¯(K,Q)𝐟++tQV(N1)t3,Q¯(K,Q)𝐟+tV(N1)t2,2¯(K,Q)𝐟\displaystyle+{\sqrt{t}}\sum_{Q^{\prime}}V^{\,(K,Q)^{{\bf{f}}}}_{\overline{\mathcal{I}^{*}_{1,Q^{\prime}}}}+\dots+{\sqrt{t}}\sum_{Q^{\prime}}V^{\,(K,Q)^{{\bf{f}}}}_{\overline{\mathcal{I}^{*}_{(N-1)\cdot\sqrt{t}-3,Q^{\prime}}}}+{\sqrt{t}}V^{\,(K,Q)^{{\bf{f}}}}_{\overline{\mathcal{I}^{*}_{(N-1)\cdot\sqrt{t}-2,2}}}
+tQW1,Q(K,Q)𝐟++tW(N1)t,1(K,Q)𝐟\displaystyle+{\sqrt{t}}\sum_{Q^{\prime}}W^{\,(K,Q)^{{\bf{f}}}}_{\mathcal{I}_{1,Q^{\prime}}}+\dots+\sqrt{t}W^{\,(K,Q)^{{\bf{f}}}}_{\mathcal{I}_{(N-1)\cdot\sqrt{t},1}}

where all the bulk potentials are block-diagonalized. As a next step, we consider the boundary terms all together, i.e., we define

t𝖶:=tQW1,Q(K,Q)𝐟++tW(N1)t,1(K,Q)𝐟,\sqrt{t}\,\mathsf{W}:={\sqrt{t}}\sum_{Q^{\prime}}W^{\,(K,Q)^{{\bf{f}}}}_{\mathcal{I}_{1,Q^{\prime}}}+\dots+{\sqrt{t}}W^{\,(K,Q)^{{\bf{f}}}}_{\mathcal{I}_{(N-1)\cdot\sqrt{t},1}}\,, (3.83)

whose norm is bounded by 𝒪(t)\mathcal{O}(\sqrt{t}) (due to statement b) in Theorem 3.2), and we implement a block-diagonalization step w.r.t. the projections

P(N1)t,1()PΛ(),P(N1)t,1(+)PΛ(+)P^{(-)}_{\mathcal{I}_{(N-1)\cdot\sqrt{t},1}}\equiv P^{(-)}_{\Lambda}\,,\,P^{(+)}_{\mathcal{I}_{(N-1)\cdot\sqrt{t},1}}\equiv P^{(+)}_{\Lambda} (3.84)

associated with the whole chain; in this operation the “bulk" operator

𝖦:=HΛ0+tQV2,Q¯(K,Q)𝐟++tQV(N1)t3,Q¯(K,Q)𝐟+tV(N1)t2,2¯(K,Q)𝐟\mathsf{G}:=H^{0}_{\Lambda}+{\sqrt{t}}\sum_{Q^{\prime}}V^{\,(K,Q)^{{\bf{f}}}}_{\overline{\mathcal{I}^{*}_{2,Q^{\prime}}}}+\dots+{\sqrt{t}}\sum_{Q^{\prime}}V^{\,(K,Q)^{{\bf{f}}}}_{\overline{\mathcal{I}^{*}_{(N-1)\cdot\sqrt{t}-3,Q^{\prime}}}}+{\sqrt{t}}V^{\,(K,Q)^{{\bf{f}}}}_{\overline{\mathcal{I}^{*}_{(N-1)\cdot\sqrt{t}-2,2}}} (3.85)

plays the role of the unperturbed Hamiltonian, and we make use of the result 𝒮2)\mathcal{S}2) (see Theorem 3.2) in step ((N1)t,1)1((N-1)\cdot\sqrt{t},1)_{-1}. Upon this standard perturbation, the resulting block-diagonalized Hamiltonian is

K~Λ(t):=𝖦+t𝖶\tilde{K}_{\Lambda}(t):=\mathsf{G}+\sqrt{t}\,\mathsf{W}^{\prime} (3.86)

where 𝖶\mathsf{W}^{\prime} is expressed in terms of operators (𝖶)j(\mathsf{W})_{j}, (Z)j(Z)_{j} by means of the formulae from (2.35) to (2.39), starting from the interaction (𝖶)1=𝖶(\mathsf{W})_{1}=\mathsf{W}, from 𝖦\mathsf{G}, and from its ground-state energy 𝖤\mathsf{E}. By standard estimates, K~Λ(t)\tilde{K}_{\Lambda}(t) enjoys the spectral features described in the statement, as explained below.

i) For the claim concerning the bound

ΔΛ(t)ε4,\Delta_{\Lambda}(t)\geq\frac{\varepsilon}{4}\,,

it is enough to consider the argument used to prove Lemma 3.1 by adding the new operator t𝖶\sqrt{t}\,\mathsf{W}^{\prime}.

ii) Concerning the 4×44\times 4 matrix describing the restriction

(K~Λ(t)𝖤):P(N1)t,1()(N)P(N1)t,1()(N),(\tilde{K}_{\Lambda}(t)-\mathsf{E})\,:\,P^{(-)}_{\mathcal{I}_{(N-1)\cdot\sqrt{t},1}}\mathcal{H}^{(N)}\,\to\,P^{(-)}_{\mathcal{I}_{(N-1)\cdot\sqrt{t},1}}\mathcal{H}^{(N)}\,, (3.87)

we observe that, up to a remainder bounded in norm by o(t)o(t), we can replace the Lie Schwinger series

(t𝖶=)j=1tj2(𝖶)jdiag(\sqrt{t}\,\mathsf{W}^{\prime}=)\sum_{j=1}^{\infty}t^{\frac{j}{2}}\,(\mathsf{W})^{diag}_{j}

by the leading term t(𝖶)1diag\sqrt{t}\,(\mathsf{W})^{diag}_{1}, since (𝖶)jdiag𝒪(t)\|(\mathsf{W})^{diag}_{j}\|\leq\mathcal{O}(\sqrt{t}) for j2j\geq 2; here diagdiag stands for the diagonal part w.r.t. the projections in (3.84). Hence we can restrict the study to the matrix elements of the operator

PΛ(){tQW1,Q(K,Q)𝐟++tW(N1)t,1(K,Q)𝐟}PΛ().P^{(-)}_{\Lambda}\Big{\{}{\sqrt{t}}\sum_{Q^{\prime}}W^{\,(K,Q)^{{\bf{f}}}}_{\mathcal{I}_{1,Q^{\prime}}}+\dots+{\sqrt{t}}W^{\,(K,Q)^{{\bf{f}}}}_{\mathcal{I}_{(N-1)\cdot\sqrt{t}},1}\Big{\}}P^{(-)}_{\Lambda}\,. (3.88)

Next we show that in (3.88) the sum of all the terms corresponding to intervals of length R2R\geq 2 is, up to a multiple of the identity operator, a matrix that can be estimated in norm less than o(t)o(t). This can be explained thinking of the growth processes yielding potentials of type WR,Q(K,Q)𝐟W^{\,(K,Q)^{{\bf{f}}}}_{\mathcal{I}_{R,Q^{\prime}}}. First of all we recall that, by construction, for R,Jb.dry\mathcal{I}_{R,J}\in\mathfrak{I}_{\text{b.dry}} with R2R\geq 2,

WR,J(0,N)=0.W_{\mathcal{I}_{R,J}}^{(0,N)}=0. (3.89)

Hence all the potentials WR,J(K,Q)W_{\mathcal{I}_{R,J}}^{(K,Q)}, with R2R\geq 2, result from successive growth processes described in c-2) of Definition 2.15. In this respect, notice that all operators from (2.72) down to (2.72) are surely supported at a distance larger than say t1/2\sqrt{t^{-1}}/2 from the boundaries, since K,Q\mathcal{I}_{K,Q} belongs to bulk\mathfrak{I}_{bulk} by hypothesis. Then, taking also (3.89) into account, we can conclude that the operator WR,J(K,Q)W_{\mathcal{I}_{R,J}}^{(K,Q)}, for R2R\geq 2, is in fact supported at distance larger than say t1/2\sqrt{t^{-1}}/2 from the boundaries. By using the LTQO property in (1.8) we conclude that

PΛ()WR,J(K,Q)PΛ()=ω(PR,J()WR,J(K,Q))P(N1)t,1()+ΔWR,J(K,Q)P^{(-)}_{\Lambda}\,W_{\mathcal{I}_{R,J}}^{(K,Q)}\,P^{(-)}_{\Lambda}=\omega(P^{(-)}_{\mathcal{I}_{R,J}}\,W_{\mathcal{I}_{R,J}}^{(K,Q)})\,P^{(-)}_{\mathcal{I}_{(N-1)\cdot\sqrt{t},1}}+\Delta W_{\mathcal{I}_{R,J}}^{(K,Q)} (3.90)

where

ΔWR,J(K,Q)𝒪(3t12WR,J(K,Q)).\|\Delta W_{\mathcal{I}_{R,J}}^{(K,Q)}\|\leq\mathcal{O}(3^{-\frac{\sqrt{t^{-1}}}{2}}\|W_{\mathcal{I}_{R,J}}^{(K,Q)}\|)\,.

It is then clear that, up to a multiple of the identity operator, the matrix in (3.88) corresponds to

PΛ(){tQW1,Q(K,Q)𝐟+QΔW2,Q(K,Q)𝐟+tΔW(N1)t,1(K,Q)𝐟}PΛ()P^{(-)}_{\Lambda}\Big{\{}{\sqrt{t}}\sum_{Q^{\prime}}W^{\,(K,Q)^{{\bf{f}}}}_{\mathcal{I}_{1,Q^{\prime}}}+\sum_{Q^{\prime}}\Delta W^{\,(K,Q)^{{\bf{f}}}}_{\mathcal{I}_{2,Q^{\prime}}}+\dots\sqrt{t}\Delta W^{\,(K,Q)^{{\bf{f}}}}_{\mathcal{I}_{(N-1)\cdot\sqrt{t}},1}\Big{\}}P^{(-)}_{\Lambda} (3.91)

where for tt sufficiently small

QΔW2,Q(K,Q)𝐟+tΔW(N1)t,1(K,Q)𝐟𝒪(3t12)\|\sum_{Q^{\prime}}\Delta W^{\,(K,Q)^{{\bf{f}}}}_{\mathcal{I}_{2,Q^{\prime}}}+\dots\sqrt{t}\Delta W^{\,(K,Q)^{{\bf{f}}}}_{\mathcal{I}_{(N-1)\cdot\sqrt{t}},1}\|\leq\mathcal{O}(3^{-\frac{\sqrt{t^{-1}}}{2}}) (3.92)

thanks to statement b) in Theorem 3.2. By collecting all the error terms, and by using Weyl inequalities for hermitian matrices, we can conclude that the differences between the eigenvalues of the 4×44\times 4 matrix corresponding to (3.87) coincide with the shifts between the eigenvalues of the matrix

PΛ()QW1,Q(K,Q)𝐟PΛ()=PΛ()t(V1,1+V1,(N1)t)PΛ(),P^{(-)}_{\Lambda}\sum_{Q^{\prime}}W^{\,(K,Q)^{{\bf{f}}}}_{\mathcal{I}_{1,Q^{\prime}}}\,P^{(-)}_{\Lambda}\,=\,P^{(-)}_{\Lambda}\,{\sqrt{t}}\,\Big{(}V_{\mathcal{I}_{1,1}}+V_{\mathcal{I}_{1,(N-1)\cdot\sqrt{t}}}\Big{)}\,P^{(-)}_{\Lambda}\,, (3.93)

up to o(t)o(t) corrections. By rewriting V1,1V_{\mathcal{I}_{1,1}} and V1,(N1)tV_{\mathcal{I}_{1,(N-1)\cdot\sqrt{t}}} in terms of the nearest-neighbor interaction terms Vi,i+1V_{i,i+1}, the r-h-s of (3.93) reads

PΛ()(ti=1iVi,i+1+ti=i′′N1Vi,i+1)PΛ()P^{(-)}_{\Lambda}\,\Big{(}\,t\sum_{i=1}^{i^{\prime}}\,V_{i,i+1}+\,t\,\sum_{i=i^{\prime\prime}}^{N-1}V_{i,i+1}\Big{)}\,P^{(-)}_{\Lambda} (3.94)

where i=t1i^{\prime}=\sqrt{t^{-1}} and i′′=Nt1i^{\prime\prime}=N-\sqrt{t^{-1}}; recall Definition 2.2. Next, we observe that the gaps between the eigenvalues of the matrix in (3.94) do not change if we subtract a multiple of the identity matrix, namely

PΛ()(ti=d+1iω(Vi,i+1)+ti=i′′Nd1ω(Vi,i+1))PΛ(),P^{(-)}_{\Lambda}\,\Big{(}\,t\sum_{i=d+1}^{i^{\prime}}\,\omega(V_{i,i+1})+\,t\,\sum_{i=i^{\prime\prime}}^{N-d-1}\omega(V_{i,i+1})\Big{)}\,P^{(-)}_{\Lambda}\,,

where di1d\leq i^{\prime}-1, so as to study the matrix

PΛ()(ti=1dVi,i+1+ti=NdN1Vi,i+1)PΛ()\displaystyle P^{(-)}_{\Lambda}\,\Big{(}\,t\sum_{i=1}^{d}\,V_{i,i+1}+\,t\,\sum_{i=N-d}^{N-1}V_{i,i+1}\Big{)}\,P^{(-)}_{\Lambda} (3.95)
+PΛ()(ti=d+1i[Vi,i+1ω(Vi,i+1)]+ti=i′′Nd1[Vi,i+1ω(Vi,i+1)])PΛ().\displaystyle+P^{(-)}_{\Lambda}\,\Big{(}\,t\sum_{i=d+1}^{i^{\prime}}\,[V_{i,i+1}-\omega(V_{i,i+1})]+\,t\,\sum_{i=i^{\prime\prime}}^{N-d-1}[V_{i,i+1}-\omega(V_{i,i+1})]\Big{)}\,P^{(-)}_{\Lambda}\,. (3.96)

Using the LTQO property in (1.8) once again, we prove the bound

PΛ()(ti=d+1i[Vi,i+1ω(Vi,i+1)]+ti=i′′Nd1[Vi,i+1ω(Vi,i+1)])PΛ()\displaystyle\Big{\|}\,P^{(-)}_{\Lambda}\,\Big{(}\,t\sum_{i=d+1}^{i^{\prime}}\,[V_{i,i+1}-\omega(V_{i,i+1})]+\,t\,\sum_{i=i^{\prime\prime}}^{N-d-1}[V_{i,i+1}-\omega(V_{i,i+1})]\Big{)}\,P^{(-)}_{\Lambda}\,\Big{\|}
\displaystyle\leq ti=d+1iPΛ()[Vi,i+1ω(Vi,i+1)]PΛ()+ti=i′′Nd1PΛ()[Vi,i+1ω(Vi,i+1)]PΛ()\displaystyle\,t\sum_{i=d+1}^{i^{\prime}}\Big{\|}\,P^{(-)}_{\Lambda}\,[V_{i,i+1}-\omega(V_{i,i+1})]P^{(-)}_{\Lambda}\,\Big{\|}+\,t\sum_{i=i^{\prime\prime}}^{N-d-1}\Big{\|}\,P^{(-)}_{\Lambda}\,[V_{i,i+1}-\omega(V_{i,i+1})]P^{(-)}_{\Lambda}\,\Big{\|}
\displaystyle\leq  2ti=d+13(i1)=t3(d1).\displaystyle\,2\cdot t\cdot\sum_{i=d+1}^{\infty}3^{-(i-1)}=\,t\cdot 3^{-(d-1)}\,.

One can easily generalize the argument to the range d<N2d<\frac{N}{2} by subtracting a suitable multiple of the identity matrix, and finally get the result (ii) in the statement of the theorem as a consequence of Weyl inequalities for hermitian matrices. In concrete applications only small values of dd are interesting, in particular did\ll i^{\prime}.

References

  • [AKLT] I. Affleck, T. Kennedy, E.H. Lieb, H. Tasaki Rigorous results on valence-bond ground states in antiferromagnets Physical Review Letters. 59(7): 799-802 (1987).
  • [BN] S. Bachmann, B. Nachtergaele. On gapped phases with a continuous symmetry and boundary operators J. Stat. Phys. 154(1-2): 91-112 (2014)
  • [BH] S. Bravyi and M.B. Hastings. A short proof of stability of topological order under local perturbations Commun. Math. Physics. 307, 609 (2011).
  • [BHM] S. Bravyi, M.B. Hastings and S. Michalakis. Topological quantum order: stability under local perturbations J. Math. Phys. 51, 093512, (2010).
  • [DFFR] N. Datta, R. Fernandez, J. Fröhlich, L. Rey-Bellet. Low-Temperature Phase Diagrams of Quantum Lattice Systems. II. Convergent Perturbation Expansions and Stability in Systems with Infinite Degeneracy Helvetica Physica Acta 69, 752–820 (1996)
  • [DFPR1] S. Del Vecchio, J. Fröhlich, A. Pizzo, S. Rossi. Lie-Schwinger block-diagonalization and gapped quantum chains with unbounded interactions, Comm. Math. Phys. https://link.springer.com/article/10.1007/s00220-020-03878-y
  • [DFPR2] S. Del Vecchio, J. Fröhlich, A. Pizzo, S.Rossi. Lie-Schwinger block-diagonalization and gapped quantum chains: analyticity of the ground-state energy, J. Funct. Anal. V. 279 Issue 8
    https://www.sciencedirect.com/science/article/abs/pii/S0022123620302469
  • [DFPR3] S. Del Vecchio, J. Fröhlich, A. Pizzo, S. Rossi. Local iterative block-diagonalization of gapped Hamiltonians: a new tool in singular perturbation theory J. Math. Phys. 63 (2022) 073503, https://doi.org/10.1063/5.0084552.
  • [DFP] S. Del Vecchio, J. Fröhlich, A. Pizzo. Block-diagonalization of infinite-volume lattice Hamiltonians with unbounded interactions J. Funct. Anal. V. 284, Issue 1, 1 January 2023, 109734
  • [DFPRa] S. Del Vecchio, J. Fröhlich, A. Pizzo, A. Ranallo Low energy spectrum of the XXZ model coupled to a magnetic field arxiv preprint
  • [FFU] R. Fernandez, J. Fröhlich, D. Ueltschi. Mott Transitions in Lattice Boson Models. Comm. Math. Phys. 266, 777-795 (2006)
  • [FP] J. Fröhlich, A. Pizzo. Lie-Schwinger block-diagonalization and gapped quantum chains Comm. Math. Phys. 375, 2039-2069 (2020)
  • [KT] T. Kennedy, H. Tasaki. Hidden symmetry breaking and the Haldane phase in S = 1 quantum spin chains Comm.. Math. Phys. 147, 431-484 (1992)
  • [LSW] M. Lemm, A. W. Sandvik, L. Wang. Existence of a spectral gap in the Affleck-Kennedy-Lieb-Tasaki model on the hexagonal lattice. Phys. Rev. Lett., 124(17) (2020).
  • [LR] E. H. Lieb, D. W. Robinson. The finite group velocity of quantum spin systems Comm. in Math. Phys. 28 (3) 251-257 (1972).
  • [LMY] A. Lucia, A. Moon, A. Young. Stability of the spectral gap and grounds state indistinguishability for a decorated AKLT model. preprint arXiv:2209.01141
  • [MZ] S. Michalakis, J.P. Zwolak Stability of frustration-free Hamiltonians Comm. Math. Phys. 322, 277-302, (2013)
  • [MN] A. Moon, B. Nachtergaele. Stability of Gapped Ground State Phases of Spins and Fermions in One Dimension J. Math. Phys. 59, 091415 (2018)
  • [NS] B. Nachtergaele, R. Sims. Locality Estimates for Quantum Spin Systems in New Trends in Mathematical Physics 591-614 (2009). Spinger Netherlands.
  • [NSY1] B. Nachtergaele, R. Sims, A. Young. Stability of the bulk gap for frustration-free topologically ordered quantum lattice systems Lett Math Phys 114, 24 (2024)
  • [NSY2] B. Nachtergaele, R. Sims, A. Young, Quasi-Locality Bounds for Quantum Lattice Systems. Part II. Perturbations of Frustration-Free Spin Models with Gapped Ground States Ann. Henri Poincaré 23, 393-511 (2022).
  • [NSY] B. Nachtergaele, R. Sims, A. Young. Lieb-Robinson bounds, the spectral flow, and stability of the spectral gap for lattice fermion systems Mathematical Problems in Quantum Physics, pp.93-115
  • [AYLLN] H. Abdul-Rahman, A. Young, A. Lucia, M. Lemm, B. Nachtergaele. A class of two-dimensional AKLT models with a gap Analytic Trends in Mathematical Physics 2020, https://doi.org/10.1090/conm/741/14917
  • [O1] Y. Ogata. A Z2-Index of Symmetry Protected Topological Phases with Time Reversal Symmetry for Quantum Spin Chains. Comm. Math. Phys., 374 (2):705-734, July 2019.
  • [O2] Y. Ogata. A H3(G, T)-valued index of symmetry protected topological phases with on-site finite group symmetry for two-dimensional quantum spin systems. 2021. arXiv:2101.00426.
  • [O3] Y. Ogata. A Z2-index of symmetry protected topological phases with reflection symmetry for quantum spin chains. Comm. in Math. Phys., 385:1245-1272, 2021.
  • [PW1] N. Pomata, T.-C. Wei Demonstrating the Affleck-Kennedy-Lieb-Tasaki spectral gap on 2D degree-3 lattices. Phys. Rev. Lett., 124(17), April 2020.
  • [PW2] N. Pomata, T.-C. Wei. AKLT models on decorated square lattices are gapped. Phys. Rev. B, 100:094429, 2019.
  • [Y] D.A. Yarotsky. Ground States in Relatively Bounded Quantum Perturbations of Classical Systems Comm. Math. Phys. 261, 799-819 (2006)