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

Self-duality triggered dynamical transition

Italo Guarneri,1 Chushun Tian,2 and Jiao Wang3 1Center for Nonlinear and Complex Systems, Universita`\grave{a} degli Studi dell’Insubria, via Valleggio 11, 22100 Como, Italy
2CAS Key Laboratory of Theoretical Physics and Institute of Theoretical Physics, Chinese Academy of Sciences, Beijing 100190, China
3Department of Physics and Key Laboratory of Low Dimensional Condensed Matter Physics (Department of Education of Fujian Province), Xiamen University, Xiamen 361005, Fujian, China
Abstract

A basic result about the dynamics of spinless quantum systems is that the Maryland model exhibits dynamical localization in any dimension. Here we implement mathematical spectral theory and numerical experiments to show that this result does not hold, when the 2-dimensional Maryland model is endowed with spin 1/2 – hereafter dubbed spin-Maryland (SM) model. Instead, in a family of SM models, tuning the (effective) Planck constant drives dynamical localization–delocalization transitions of topological nature. These transitions are triggered by the self-duality, a symmetry generated by some transformation in the parameter – the inverse Planck constant – space. This provides significant insights to new dynamical phenomena such as what occur in the spinful quantum kicked rotor.

-

I Introduction

The dynamics of complex quantum systems is currently a very active area of research. The quantum kicked rotor (QKR) Casati79 ; Fishman82a – a rotating particle subject to time-periodic “kicks” – is a paradigmatic model, which has played a key role in the development of this field; moreover, it has allowed for experimental realizations Fishman10 ; Raizen11 ; Garreau16 . A prototypical feature of the QKR is dynamical localization Casati79 : though the classical version of the QKR exhibits unbounded chaotic diffusion in momentum space Chirikov79 , the QKR momentum remains bounded in time. Thorough investigations of the analogy Fishman82a between dynamical localization and Anderson localization in disordered solids have given rise to a highly interdisciplinary field; see Refs. Fishman10 ; Raizen11 ; Garreau16 ; Izrailev90 ; scho for reviews of different aspects. Several variants of the QKR have been examined, including 2-dimensional (22D) versions Fishman88 and 1-dimensional (11D) versions with spin Scharf89 . Such variants did not command major changes of the general picture of dynamical localization. A surprising result has been instead observed with a recent 22D variant endowed with spin 1/21/2 Tian14 ; Tian16 . By tuning the Planck constant e\hbar_{e}, a sequence of dynamical transitions arise. At certain critical values of e\hbar_{e} the wavepacket spreads diffusively, leading to dynamical delocalization, while in between different critical values dynamical localization persists. This phenomenon is remindful of the integer quantum Hall effect (IQHE) Klitzing80 , with which it shares the phase diagram and the universality class of phase transition Khmelnitskii83 ; Pruisken07 . In particular, dynamical localization corresponds to a quantum Hall insulator, with a topological angle emerging from a supersymmetric field-theoretic description of the spinful QKR, and the dynamical localization–delocalization transition corresponds to the Hall plateau transition. This finding suggests that spin is a seed for a wealth of novel dynamical phenomena, which have deep connections to topological quantum matter but go beyond the Thouless-Kohmoto-Nightingale-den Nijs-like topological theory for various driven quantum systems Lebouf90 ; Galitski11 ; Gong12 ; Gong18 ; Dana19 .

A fundamental difficulty in assimilating dynamical localization to Anderson localization is the absence of randomnness/disorder in QKR. Appeal has been made to a pseudorandom character Brenner92 of the eigenphases of free rotation, and/or to a chaotic nature of the early QKR rotor evolution Izrailev90 ; Tian04 ; Tian05 , inherited from the classical limit. However it is now generally acknowledged, that quantum localization is a general phenomenon, of which Anderson and dynamical localization are different occurrences. Important in this sense has been a modification of the original QKR, the so-called Maryland model Fishman82 ; Fishman84 , where the free Hamiltonian is linear in momentum, and not quadratic, as it is in the QKR footnote_SM . Due to this difference, pseudo-randomness is replaced by quasi-periodicity, and, moreover, the model is classically integrable Berry84 . That notwithstanding, quantum localization still survives, motivating taking the model as a basis to study some new problems in condensed matter Sarma14 ; Galitskii16 . However this localization is in a more elementary form. In particular, it is not a quantum interference effect, so it cannot be described by the scaling theory of Anderson localization and dynamical localization in QKR Tian04 ; Tian05 ; Anderson79 ; Larkin79 ; Garreau18 .

In this paper, we present a SM model that, like its spinless ancestor, though bearing no analogies to crystals, is neither pseudo-random nor chaotic and thus is very different from spinful QKR. We show that a family of SM models preserve some aspects of the similarity between IQHE on the one side and 2D, spin-1/2 QKR on the other. Notably, dynamical localization–delocalization transitions appear at half-integer e1α/π\hbar_{e}^{-1}\equiv\alpha/\pi, indicating a deeper origin of that similarity. We find that such transition arises from a self-duality. That is, SM systems on one side of a critical point, labelled by the parameter α\alpha, can be one-to-one mapped onto SM systems on the other side via a unitary transformation, so that they are in a “duality” relation; at the critical point the system is mapped onto itself, i.e. it is self-dual, and as such it bears an emergent unitary symmetry. Thus this transition resembles a very recent discovery of topological sound waves in self-dual mechanical systems Vitelli20 .

Unlike the original Maryland model, the SM model has no classical limit (e0\hbar_{e}\to 0); nevertheless both models are equivalent to certain classical dynamical systems, which belong to the family of so-called skew products aar . In particular, the SM model is equivalent to a class of skew-product on 𝕋2×SU(2)\mathbb{T}^{2}\times SU(2), which has attracted significant mathematical attention Eliasson98 ; Eliasson02 ; Aldecoa15 ; Karaliolios18 . We also show that the observed transition finds a counterpart in the equivalent classical system: at critical α\alpha the deviation of initially close trajectories grows in time tt as t\propto\sqrt{t} asymptotically, while at noncritical α\alpha the growth of the deviation saturates at long time.

The rest of the paper is organized as follows. In Sec. II we first define the SM model. Then we show that, for any e\hbar_{e}, this model has an equivalent classical dynamical system, and thus the SM model can be viewed from both a quantum and a classical viewpoint. A summary of main reults follows. In Sec. III we present numerical results. In Sec. IV we develop the spectral theory of SM models and study its dynamical implications, which provide a theoretical basis for the numerical findings. In Sec. V we show that transport associated to the quantum picture of SM models is fully equivalent to trajectory stability associated to the classical picture of SM models, and derive a result used in Sec. IV. The technical proofs of the mathematical results claimed in Sec. IV are shuffled to Sec. VI. We conclude with Sec. VII. Some additional information are provided in Appendixes.

II SM model and summary of main results

II.1 Definitions and basic properties

We start with a description of the SM model. Consider a spin 1/21/2 quantum particle moving on a 22-torus 𝕋2{\mathbb{T}}^{2} (i.e., a 22-rotor), with a constant angular velocity 𝝎(ω1,ω2){\bm{\omega}}\equiv(\omega_{1},\omega_{2}). The particle’s angular coordinates are denoted by 𝜽(θ1,θ2){\bm{\theta}}\equiv(\theta_{1},\theta_{2}), and their conjugate momentum operators are denoted e𝑵ie(θ1,θ2){\hbar_{e}\bm{N}}\equiv-i\hbar_{e}(\partial_{\theta_{1}},\partial_{\theta_{2}}). The state vector of the particle at time tt is a spinor with two complex components,

𝝍t(𝜽)=(ψ1,t(𝜽)ψ2,t(𝜽)),\boldsymbol{\psi}_{t}({\bm{\theta}})=\left(\begin{array}[]{c}\psi_{1,t}({\bm{\theta}})\\ \psi_{2,t}({\bm{\theta}})\\ \end{array}\right), (1)

which belongs to the Hilbert space L2(𝕋2)2{\cal H}\equiv L^{2}({\mathbb{T}}^{2})\otimes{\mathbb{C}}^{2}. The particle is kicked periodically in time; with the period set to unity, at all integer times tt\in\mathbb{Z} it undergoes instantaneous pulses, or kicks, that prompt sudden jumps of its momentum and simultaneously rotate its spin in a way, that depends on the angular coordinates. Such kicks are produced by switching on an external “potential” 𝑽(𝜽)=k=13Vk(𝜽)𝝈k\boldsymbol{V}({\bm{\theta}})=\sum_{k=1}^{3}V_{k}({\bm{\theta}})\boldsymbol{\sigma}_{k}. Here 𝝈k\boldsymbol{\sigma}_{k} are the Pauli matrices and the VkV_{k} are some functions 𝕋2\mathbb{T}^{2}\rightarrow\mathbb{R}. The quantum dynamics follows the Schrödinger equation:

iet𝝍t(𝜽)=𝑯(t)𝝍t(𝜽),\displaystyle i\hbar_{e}\partial_{t}\boldsymbol{\psi}_{t}({\bm{\theta}})=\bm{H}(t)\boldsymbol{\psi}_{t}({\bm{\theta}})\,,\qquad\qquad\qquad (2)
𝑯(t)=e𝝎𝑵+V(𝜽)sδ(ts).\displaystyle\bm{H}(t)=\hbar_{e}{\bm{\omega}}\boldsymbol{\cdot}\bm{N}+V({\bm{\theta}})\sum_{s\in\mathbb{Z}}\delta(t-s).

The first term of the Hamiltonian 𝑯(t)\bm{H}(t) corresponds to rotation over the torus 𝕋2\mathbb{T}^{2} with constant angular velocity and the second to the kick. The dynamics from immediately after a kick to immediately after the next is described by:

t:𝝍t+1(𝜽)=𝑼α,𝝎𝝍t(𝜽),𝑼α,𝝎:=𝑴α𝑻𝝎,\forall t\in\mathbb{Z}:\,\boldsymbol{\psi}_{t+1}({\bm{\theta}})\;=\;{\bm{U}}_{\alpha,{\bm{\omega}}}\boldsymbol{\psi}_{t}({\bm{\theta}}),\quad{\bm{U}}_{\alpha,{\bm{\omega}}}:=\bm{M}_{\alpha}\bm{T}_{{\bm{\omega}}}, (3)

which defines the Floquet operator 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}}. Here 𝑻𝝎\bm{T}_{{\bm{\omega}}} is the translation operator: 𝑻𝝎𝝍(𝜽)𝝍(𝝉𝝎1𝜽)\bm{T}_{{\bm{\omega}}}\boldsymbol{\psi}({\bm{\theta}})\equiv\boldsymbol{\psi}(\boldsymbol{\tau}_{\boldsymbol{\omega}}^{-1}{\bm{\theta}}) where 𝝉𝝎:𝜽𝜽+𝝎\boldsymbol{\tau}_{\boldsymbol{\omega}}:{\bm{\theta}}\mapsto\boldsymbol{{\bm{\theta}}}+\boldsymbol{\omega} is translation by 𝝎\boldsymbol{\omega} along the torus 𝕋2\mathbb{T}^{2}. Instead 𝑴α=eiαπ𝑽(𝜽)\bm{M}_{\alpha}=e^{-i\frac{\alpha}{\pi}\boldsymbol{V}({\bm{\theta}})} is a smooth map 𝕋2SU(2){\mathbb{T}}^{2}\to SU(2); so 𝑽(𝜽)\boldsymbol{V}({\bm{\theta}}) is a smooth map from 𝕋2\mathbb{T}^{2} to the 2×22\times 2 self-adjoint matrices with null trace. It will be assumed to have a constant Hilbert-Schmidt norm such that tr(𝑽(𝜽)2)=2π2{\rm tr}(\boldsymbol{V}({\bm{\theta}})^{2})=2\pi^{2} , so it can be written in the general form:

𝑽(𝜽)=πk=13dk(𝜽)𝝈k,\boldsymbol{V}({\bm{\theta}})=\pi\sum_{k=1}^{3}d_{k}({\bm{\theta}})\boldsymbol{\sigma}_{k}\;, (4)

which with some abuse of notation will be concisely written in the form π𝒅(𝜽)𝝈\pi\bm{d}({\bm{\theta}})\boldsymbol{\cdot}\bm{\sigma}. Here 𝒅(𝜽)=(d1(𝜽),d2(𝜽),d3(𝜽)){\boldsymbol{d}}({\bm{\theta}})=(d_{1}({\bm{\theta}}),d_{2}({\bm{\theta}}),d_{3}({\bm{\theta}})) is a fixed map from 𝕋2\mathbb{T}^{2} to the unit real 22-sphere 𝕊2\mathbb{S}^{2}. Equation (2) or its discrete-time version Eq. (3) define our general SM model.

The following basic property will be extensively used in the following:

𝝆𝑴α𝝆=𝑴α,𝝆𝑼α,𝝎𝝆=𝑼α,𝝎{\bm{\rho}}\bm{M}_{\alpha}{\bm{\rho}}^{\dagger}\;=\;\bm{M}_{\alpha},\quad{\bm{\rho}}{\bm{U}}_{\alpha,{\bm{\omega}}}{\bm{\rho}}^{\dagger}\;=\;{\bm{U}}_{\alpha,{\bm{\omega}}}\; (5)

where 𝝆{\bm{\rho}} is the standard time-reversal operator for spin-1/21/2 particles; 𝝆:=i𝝈2C{\bm{\rho}}:=-i\boldsymbol{\sigma}_{2}C (where CC denotes complex conjugation in the coordinate representation). It is the antiunitary map that is defined in \cal H by

𝝆:(ψ1ψ2)(ψ2¯ψ1¯)\displaystyle{\bm{\rho}}:\left(\begin{array}[]{c}\psi_{1}\\ \psi_{2}\end{array}\right)\mapsto\left(\begin{array}[]{c}-\overline{\psi_{2}}\\ \overline{\psi_{1}}\end{array}\right) (10)

The ordinary Maryland model Fishman82 ; Fishman84 is recovered from the present one, whenever the potential 𝑽(𝜽)\bm{V}({\bm{\theta}}) is invariant under spin rotations, i.e., 𝑽(𝜽)\bm{V}({\bm{\theta}}) does not couple the angular and spin degrees of freedom. As shown below, thanks to this difference the SM model exhibits a much more intriguing phenomenology than the ordinary Maryland model.

On the other hand, the SM model crucially differs from the spinful QKR Tian14 ; Tian16 in that its free Hamiltonian between kicks is linear in n1n_{1}, while for the spinful QKR the free Hamiltonian is quadratic in n1n_{1}. It is well known that kicked systems with a quadratic free Hamiltonian may exhibit a chaotic behavior (in some sense), which cannot be observed with a linear one.

II.2 Equivalent classical system

Thanks to the linear momentum dependence of the free Hamiltonian, the SM model Eq. (3), though of quantum origin, can be translated into a classical dynamical system, with the usual technical meaning, i.e. a discrete time dynamics which is generated by iterating a measure preserving transformation in some phase space. Though mathematically trivial, this quantum-classical juxtaposition illustrates that in SM models wavepacket propagation and motion along phase-space trajectories are equivalent pictures of the same dynamics. The phase space is Ω𝕋2×12\Omega\equiv{\mathbb{T}}^{2}\times{\mathbb{C}}_{1}^{2}, where 12{\mathbb{C}}_{1}^{2} denotes the unit sphere in 2\mathbb{C}^{2}, and the map is defined as follows:

𝓢α,𝝎:(𝜽,ϕ)(𝝉𝝎𝜽,𝑴α(𝝉𝝎𝜽)ϕ).\displaystyle{\bm{{\cal S}}}_{\alpha,{\bm{\omega}}}:({\bm{\theta}},\boldsymbol{\phi})\mapsto(\bm{\tau}_{\boldsymbol{\omega}}{\bm{\theta}},\bm{M}_{\alpha}(\bm{\tau}_{\boldsymbol{\omega}}{\bm{\theta}})\boldsymbol{\phi})\,. (11)

This map preserves the natural measure dμd\mu that is defined in Ω\Omega by the product of the Lebesgue measure dmdm on 𝕋2\mathbb{T}^{2}, normalized to unity, and the uniform measure on 12{\mathbb{C}}^{2}_{1} [or equivalently the Haar measure on SU(2)SU(2)]. In mathematical terms the map described by Eq. (11) defines a skew product action of SU(2)SU(2) Eliasson98 ; Eliasson02 ; Aldecoa15 ; Karaliolios18 . The corresponding Perron-Frobenius operator, denoted by 𝓤α,𝝎{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}, acts on functions FL2(Ω,dμ)F\in{\mathfrak{H}}\equiv L^{2}(\Omega,d\mu) as follows:

𝓤α,𝝎F(𝜽,ϕ)\displaystyle{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}F({\bm{\theta}},\boldsymbol{\phi}) =\displaystyle= F(𝓢α,𝝎1(𝜽,ϕ))\displaystyle F({\bm{{\cal S}}}_{\alpha,{\bm{\omega}}}^{-1}({\bm{\theta}},\boldsymbol{\phi})) (12)
=\displaystyle= F(𝝉𝝎1𝜽,𝑴α(𝜽)1ϕ).\displaystyle F(\bm{\tau}^{-1}_{\boldsymbol{\omega}}{\bm{\theta}}\;,\;\boldsymbol{M}_{\alpha}({\bm{\theta}})^{-1}\boldsymbol{\phi})\;.

It preserves the scalar product in \mathfrak{H}, i.e. F,F′′\forall F^{\prime},F^{\prime\prime}\in\mathfrak{H},

F|F′′\displaystyle\langle F^{\prime}|F^{\prime\prime}\rangle_{\mathfrak{H}} :=\displaystyle:= Ω𝑑μ(𝜽,ϕ)F(𝜽,ϕ)¯F′′(𝜽,ϕ)\displaystyle\int_{\Omega}d\mu({\bm{\theta}},\boldsymbol{\phi})\;\overline{F^{\prime}({\bm{\theta}},\boldsymbol{\phi})}F^{\prime\prime}({\bm{\theta}},\boldsymbol{\phi}) (13)
=\displaystyle= 𝓤α,𝝎F|𝓤α,𝝎F′′.\displaystyle\langle{{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}F^{\prime}}|{{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}F^{\prime\prime}}\rangle_{\mathfrak{H}}\;.

The actual link between the “quantum” SM model and its “classical” equivalent is as follows. Given a 𝝍\boldsymbol{\psi}\in\cal H, let

F𝝍(𝜽,ϕ):=𝝍(𝜽)ϕ.F_{\boldsymbol{\psi}}({\bm{\theta}},\boldsymbol{\phi}):=\boldsymbol{\psi}({\bm{\theta}})\boldsymbol{\cdot}\boldsymbol{\phi}. (14)

Here the dot denotes the canonical scalar product of vectors in 2{\mathbb{C}}^{2}: 𝝍ϕ:=ψ1¯ϕ1+ψ2¯ϕ2\boldsymbol{\psi}\boldsymbol{\cdot}\boldsymbol{\phi}:=\overline{\psi_{1}}\phi_{1}+\overline{\psi_{2}}\phi_{2}. Then

𝓤α,𝝎F𝝍(𝜽,ϕ)=𝑼α,𝝎𝝍(𝜽)ϕ=F𝝍(𝜽,ϕ),\displaystyle{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}F_{\boldsymbol{\psi}}({\bm{\theta}},\boldsymbol{\phi})\;=\;{\bm{U}}_{\alpha,{\bm{\omega}}}\boldsymbol{\psi}({\bm{\theta}})\boldsymbol{\cdot}\boldsymbol{\phi}\;=\;F_{\boldsymbol{\psi}^{\prime}}({\bm{\theta}},\boldsymbol{\phi})\;, (15)

where 𝝍=𝑼α,𝝎𝝍\boldsymbol{\psi}^{\prime}={\bm{U}}_{\alpha,{\bm{\omega}}}\,\boldsymbol{\psi}. This equation formalizes the quantum-classical juxtaposition within the same model. It is important to emphasize that the quantum system (3) and the classical dynamical system (11) are fully equivalent, and the latter is not related to any notion of classical limit whatsoever: the subscript α\alphai.e. the inverse Planck constant – in 𝓢α,𝝎{\bm{{\cal S}}}_{\alpha,{\bm{\omega}}} is a bookkeeping of this fact. The existence of a classical equivalent of the SM model implies that quantum interference does not play any role in its phenomenology. As mentioned in the introduction, the same is true for the original Maryland model, where the observed localization is not related to the Anderson-like dynamical localization that typically occurs in quantum systems exhibiting chaotic diffusion in the classical limit and results from quantum interference.

The spectral and dynamical properties of the SM models crucially depend on the arithmetic type of the triple (ω1,ω2,π)(\omega_{1},\omega_{2},\pi). In particular, the following well-known fact is of capital importance to the present work: the 𝛚{\bm{\omega}}-shift 𝛕𝛚\bm{\tau}_{\boldsymbol{\omega}} is ergodic in 𝕋2{\mathbb{T}}^{2} whenever (ω1,ω2,π\omega_{1},\omega_{2},\pi) are an incommensurate triple of real numbers. More detailed arithmetic properties of the triple play important spectral dynamical roles; a few more comments on this theme are deferred to Sec. IV.3. However, for the purposes of the present paper sheer incommensuration will be sufficient. From now on, we shall always consider incommensurate triples, unless explicitly mentioned otherwise.

II.3 Outline of main results

In this work we implement the quantum-classical juxtaposition to investigate both mathematically and numerically the dynamical properties of SM. In numerical simulations, we used two different incommensurate frequency vectors 𝝎\boldsymbol{\omega}, which are specified in Sec. III.2.1.

Upon expanding 𝝍t(𝜽)\boldsymbol{\psi}_{t}({\bm{\theta}}) in the 22D Fourier basis we move from the 𝜽{\bm{\theta}}-coordinate representation to the momentum 𝑵\bm{N}-representation, where the SM dynamics can be pictured as propagation of a quantum wavepacket over the 22D integer lattice 2\mathbb{Z}^{2}. This defines quantum transport for the SM models in the same way as for the QKR Casati79 and for the ordinary Maryland model Fishman82 ; Fishman84 . A basic probe of the quantum transport associated with the evolution of a state 𝝍t{\boldsymbol{\psi}}_{t} from an initial state 𝝍0{\boldsymbol{\psi}}_{0} is

Ej(𝝍t):=𝝍t|nj2|𝝍t=θj𝝍t2,j=1,2\displaystyle E_{j}({\boldsymbol{\psi}}_{t}):=\langle{\boldsymbol{\psi}}_{t}|n_{j}^{2}|{\boldsymbol{\psi}}_{t}\rangle=\|\partial_{\theta_{j}}\boldsymbol{\psi}_{t}\|^{2}_{\cal H},\,j=1,2
E(𝝍t)=E1(𝝍t)+E2(𝝍t),\displaystyle E({\boldsymbol{\psi}}_{t})\;=\;E_{1}({\boldsymbol{\psi}}_{t})\;+\;E_{2}({\boldsymbol{\psi}}_{t}),\; (16)

which characterizes how the wavepacket spreads in the course of time. Like QKR systems Casati79 ; Fishman82a ; Tian14 ; Tian16 , this probe provides an important link between SM models and condensed matter systems. Notably, when E(𝝍t)E({\boldsymbol{\psi}}_{t}) remains bounded as tt\to\infty, the wavepacket does not spread at long time and remains localized in a finite region in momentum space; no quantum transport arises and this simulates a quantum insulator in condensed matter physics. When either or both E1,2(𝝍t)E_{1,2}({\boldsymbol{\psi}}_{t}) display unbounded growth, the wavepacket propagation is delocalized; and if the asymptotic growth is linear in time, then quantum diffusive transport arises, and this simulates a quantum normal metal in condensed matter physics. Such different transport properties are mirrored by the spectral properties of the Floquet operators 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}}; a detailed analysis is given in Sec. IV.

We will show that quantum transport in the SM models finds a “classical” counterpart in the stability of trajectories, i.e. in the way initially close trajectories separate in time. This is a fundamental token in the theory of classical dynamical systems, and a commonly accepted definition of ”chaos” applies whenever the divergence of trajectories is exponentially fast. Specifically, the quantum transport in momentum space, probed by E1,2(𝝍t)E_{1,2}({\boldsymbol{\psi}}_{t}) defined by Eq. (II.3), has an equivalent description in terms of the instability of trajectories (𝜽t,ϕt):=𝓢α,𝝎t(𝜽,ϕ)({\bm{\theta}}_{t},\boldsymbol{\phi}_{t}):={\bm{{\cal S}}}_{\alpha,{\bm{\omega}}}^{t}({\bm{\theta}},\boldsymbol{\phi}) in the phase space Ω\Omega, in the following sense. The (linear) stability of a trajectory (𝜽t,ϕt)({\bm{\theta}}_{t},\boldsymbol{\phi}_{t}) is determined by the behavior in time of the derivatives θjϕt(𝜽,ϕ)\partial_{\theta_{j}}\boldsymbol{\phi}_{t}({\bm{\theta}},\boldsymbol{\phi}) with j=1,2j=1,2; see Sec. V, where we prove that if the initial spinor 𝝍0(𝜽){\boldsymbol{\psi}}_{0}({\bm{\theta}}) is localized at the origin of momentum space then the quantum transport which is defined by Eq. (II.3) is related to trajectory stability by:

Ej(𝝍t)\displaystyle E_{j}({\boldsymbol{\psi}}_{-t}) =\displaystyle= Ω𝑑μ(𝜽,ϕ)|θjϕt(𝜽,ϕ)|2.\displaystyle\int_{\Omega}d\mu({\bm{\theta}},\boldsymbol{\phi})\;\boldsymbol{|}\partial_{\theta_{j}}\boldsymbol{\phi}_{t}({\bm{\theta}},\boldsymbol{\phi})\boldsymbol{|}^{2}. (17)

Spectral implications of this result are presented in Sec. IV.3.

In this paper we study three different SM models. Their Floquet operators 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} share the general form described in Sec. II.1. Differences arise from the explicit form of 𝒅(𝜽)\bm{d}({\bm{\theta}}):

  • SM Model I:

    𝒅(𝜽)=p(𝜽)(sin(θ1),sin(θ2),β(1cos(θ1)cos(θ2))),\displaystyle\bm{d}({\bm{\theta}})=p({\bm{\theta}})(\sin(\theta_{1}),\sin(\theta_{2}),\beta(1-\cos(\theta_{1})-\cos(\theta_{2}))), (18)

    where p(𝜽)p({\bm{\theta}}) is the normalization factor, chosen such that 𝒅(𝜽)\bm{d}({\bm{\theta}}) is a unit vector, and β>0\beta>0 is a real parameter.

  • SM Model II:

    𝒅(𝜽)=p(𝜽)(sin(θ1),sin(θ2),const.0);\displaystyle\bm{d}({\bm{\theta}})=p({\bm{\theta}})(\sin(\theta_{1}),\sin(\theta_{2}),\mbox{\rm const.}\neq 0); (19)
  • SM Model III:

    𝒅(𝜽)=p(𝜽)(sin(θ1),sin(θ2),cos(θ1)+cos(θ2))or𝒅(𝜽)=p(𝜽)(sin(θ1),sin(θ2),0).\displaystyle\begin{array}[]{c}\bm{d}({\bm{\theta}})=p({\bm{\theta}})(\sin(\theta_{1}),\sin(\theta_{2}),\cos(\theta_{1})+\cos(\theta_{2}))\\ {\rm or}\,\bm{d}({\bm{\theta}})=p({\bm{\theta}})(\sin(\theta_{1}),\sin(\theta_{2}),0).\end{array} (22)

An important part of our study is how the transport properties of such models depend on the parameter α\alpha. In this respect, the following properties:

𝑼α+2π,𝝎=𝑼α,𝝎,𝑼α+π,𝝎=𝑼α,𝝎{\bm{U}}_{\alpha+2\pi,{\bm{\omega}}}={\bm{U}}_{\alpha,{\bm{\omega}}}\,,\,\,\,\,{\bm{U}}_{\alpha+\pi,{\bm{\omega}}}=\,-\,{\bm{U}}_{\alpha,{\bm{\omega}}} (23)

which directly follow from the definition of 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}}, hold true for all SM models; so, analysis of SM Models I, II, III can be restricted to α[0,π]\alpha\in[0,\pi]. The three models share certain symmetries, which are described in Sec. IV.2. A fundamental fact is that such symmetries in physical space give rise to a symmetry in parameter space; notably, quantum SM systems whose α[0,π]\alpha\in[0,\pi] are symmetric with respect to π/2\pi/2 are in a “duality” relation nature . This duality has strong spectral implications, as shown in Sec. IV.2. Besides, at α=π/2\alpha=\pi/2 it forces the system to be invariant under a unitary symmetry transformation and thus to be self-dual.

The dynamical behaviors of the three SM models are very different:

  • SM Model I:
    We find that for α[0,π]\alpha\in[0,\pi], on the left or on the right of the critical point α=π/2\alpha=\pi/2, both E1,2(𝝍t)E_{1,2}({\boldsymbol{\psi}}_{t}) remain bounded in time, so the system is dynamically localized, and localization is exponential in momentum space. At the critical point E1,2(𝝍t)tE_{1,2}({\boldsymbol{\psi}}_{t})\propto t approximately, so the system is dynamically delocalized. In the classical system, the deviation δt\delta_{t} of initially close trajectories grows as t\sqrt{t} (approximately) at half-integer α/π\alpha/\pi, whereas δt\delta_{t} saturates at long time away from the critical point. The transition in quantum transport is thus mirrored by a change of stability of classical trajectories. Interestingly, sublinear divergence of trajectories is what one should expect if the classical SM model were known to be ergodic; see Sec. V. We explain delocalization at the critical value, proving that self-duality is incompatible with dynamical localization. As we will show in Sec.V, this dynamical transition mirrors the transition in the spectral structure of Floquet operator 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}}.

    Moreover, it turns out that the dynamically localized phases which are separated by the critical points can be obtained from topologically distinct phases which are found in the spinful QKR Tian14 ; Tian16 by deforming that QKR continuously into the considered SM.

  • SM Model II:
    As α\alpha varies, dynamical localization is always observed.

  • SM Model III:
    As α\alpha varies, dynamical delocalization is always observed except at integer α/π\alpha/\pi, where dynamical localization trivially occurs because 𝑴α=±𝕀\bm{M}_{\alpha}=\pm\mathbb{I} there.

We see that, due to the presence of spin, SM models exhibit rich dynamical phenomena. This is in sharp contrast to the ordinary Maryland model for which only dynamical localization can occur. In addition, the Maryland model inherits dynamical localization from the integrability of the classical dynamics in the limit e0\hbar_{e}\to 0 Berry84 ; this mechanism for dynamical localization does not apply here, because the SM model is ill-defined in that limit.

III Numerical experiments

In this section we describe numerical results about SM Models I, II and III.

III.1 11D equivalents

Direct numerical simulation of the 22D quantum evolution requires a 22D Fourier basis. Thus it cannot be pushed too far in time. This difficulty is greatly softened by a trick which was first used in Ref. Casati89 . It consists in replacing the 22D model described by Eq. (3) by two 11D models, which allow to separately compute E1,2(𝝍t)E_{1,2}({\boldsymbol{\psi}}_{t}) [abbreviated as E1,2(t)E_{1,2}(t) throughout this section]. This possibility crucially rests on the 𝝎{\bm{\omega}}-shift 𝝉𝝎\bm{\tau}_{\bm{\omega}} in 𝕋2\mathbb{T}^{2}, namely, the first term of 𝑯(t)\bm{H}(t) given by Eq. (2), as we will show below.

By performing the transformation:

𝑼α,𝝎\displaystyle{\bm{U}}_{\alpha,{\bm{\omega}}} \displaystyle\rightarrow etω2θ2𝑼α,𝝎e(t1)ω2θ2\displaystyle e^{t\omega_{2}\partial_{\theta_{2}}}{\bm{U}}_{\alpha,{\bm{\omega}}}e^{-(t-1)\omega_{2}\partial_{\theta_{2}}}
=\displaystyle= 𝑴α(θ1,θ2+ω2t)eω1θ1=:𝑼~t,\displaystyle\bm{M}_{\alpha}(\theta_{1},\theta_{2}+\omega_{2}t)e^{-\omega_{1}\partial_{\theta_{1}}}=:\tilde{{\bm{U}}}_{t},
𝝍t\displaystyle\boldsymbol{\psi}_{t} \displaystyle\rightarrow etω2θ2𝝍t=:𝝍~t,\displaystyle e^{t\omega_{2}\partial_{\theta_{2}}}\boldsymbol{\psi}_{t}=:\tilde{\boldsymbol{\psi}}_{t}, (24)

we rewrite Eq. (3) as

𝝍~t=Ts=1t𝑼~s𝝍~0,\tilde{\boldsymbol{\psi}}_{t}\;=\;_{\rm T}\!\prod_{s=1}^{t}\tilde{{\bm{U}}}_{s}\tilde{\boldsymbol{\psi}}_{0}, (25)

where T{}_{\rm T}\!\prod denotes a time-ordered product, with the index ss increasing from right to left. In Eqs. (III.1) and (25), θ2\theta_{2} is no longer a dynamical variable, rather it is an angular parameter. Thus the quantum evolution (25) is 11D. When the 22D model is traded for such 11D model, we have

E1(t)=𝝍~t|θ12𝝍~t~θ2,E_{1}(t)=-\langle\langle\tilde{\boldsymbol{\psi}}_{t}|\partial_{\theta_{1}}^{2}\tilde{\boldsymbol{\psi}}_{t}\rangle_{\tilde{\cal H}}\rangle_{\theta_{2}}, (26)

where the Hilbert space ~L2(𝕋1)2\tilde{{\cal H}}\equiv L^{2}({\mathbb{T}}^{1})\otimes{\mathbb{C}}^{2} and θ2\langle\cdot\rangle_{\theta_{2}} denotes the average over θ2\theta_{2}. The whole derivation is fully symmetric with respect to the indices 1,21,2, so similar equations are obtained for E2(t)E_{2}(t) by just interchanging indices in the above equations. Such equations will share the same reference numbers.

Throughout this paper in both 11D and 22D simulations we choose initial states which have only the spin-up (s=1s=1) component and are zero momentum state.

III.2 Dynamics of SM Model I

III.2.1 Quantum transport

The 11D quantum evolutions which are described by Eqs. (25) were numerically simulated up to t=108t=10^{8}, using 11D Fast Fourier Transform (FFT) with a basis size 216=655362^{16}=65536. The parameter β\beta in Eq. (18) was set to 0.80.8. Two different incommensurate frequency vectors were used: 𝝎=(1,2π/5){\bm{\omega}}=(1,2\pi/\sqrt{5}) and 𝝎=(2πx,2πx2){\bm{\omega}}=(2\pi x,2\pi x^{2}), where xx is the real root of the equation x3x1=0x^{3}-x-1=0 footnote_x .

Refer to caption
Refer to caption
Figure 1: Numerical experiments on SM Model I indicate a dynamical localization–delocalization transition at half-integer α/π\alpha/\pi. (a) shows that E1(t)E_{1}(t) grows diffusively at half-integer α/π\alpha/\pi while it saturates at long times away from these points. (b) suggests a sharp transition at infinite time: the diffusion coefficient is finite at half-integer α/π\alpha/\pi while it vanishes away from the critical points. 𝝎=(1,2π/5){\bm{\omega}}=(1,2\pi/\sqrt{5}).
Refer to caption
Figure 2: A plot of It(n1,n2)I_{t}(n_{1},n_{2}) obtained from numerical experiments on SM Model I shows that within the inspected time range, the wavepacket propagation is extended in momentum space for critical α/π=0.5\alpha/\pi=0.5 (a-d), while it is restricted to a finite domain for noncritical α/π=0.7\alpha/\pi=0.7 (e-h). Moreover, the propagation is anisotropic in both cases. 𝝎=(1,2π/5){\bm{\omega}}=(1,2\pi/\sqrt{5}).

Results for E1(t)E_{1}(t) and E1(t)/tE_{1}(t)/t are shown in Fig. 1. Results for E2(t)E_{2}(t) are qualitatively similar. In (a) the cases with α/π=0.3\alpha/\pi=0.3 and =0.7=0.7 provide evidence of dynamical localization, and are representative of what was observed away from α/π=0.5\alpha/\pi=0.5. At α/π=0.5\alpha/\pi=0.5 or =1.5=1.5 delocalization was observed, with E1,2(t)E_{1,2}(t) increasing approximately linearly in time. The α\alpha-dependence of E1(t)/tE_{1}(t)/t at fixed time is shown in (b). The appearance of two sharp peaks indicates that delocalization is confined to a narrow vicinity of α/π=0.5\alpha/\pi=0.5 or =1.5=1.5 . Comparison of plots of E1(t)/tE_{1}(t)/t vs α/π\alpha/\pi at fixed tt, computed for different values of tt, shows that as tt increases the two peaks become sharper and sharper, suggesting that they would eventually shrink to two critical points: α/π=1/2\alpha/\pi=1/2 and 3/23/2. This indicates a dynamical localization–delocalization transition at half-integer α/π\alpha/\pi. Note that these critical values of α\alpha are independent of the choice of 𝝎{\bm{\omega}}, so long as (𝝎,π)({\bm{\omega}},\pi) is an incommensurate triple. Moreover, the height of the two peaks remains finite, so it yields a finite diffusion coefficient E1(t)/t=𝒪(1)E_{1}(t)/t={\cal O}(1). The smallness of this coefficient reflects the quantum nature of critical diffusion.

Results of numerical simulations of the full 22D dynamics, using 22D FFT, are shown in Fig. 2 by contour plots of the quantity It(n1,n2)𝝍^t(n1,n2)𝝍^t(n1,n2)I_{t}(n_{1},n_{2})\equiv\hat{\boldsymbol{\psi}}_{t}(n_{1},n_{2})\boldsymbol{\cdot}\hat{\boldsymbol{\psi}}_{t}(n_{1},n_{2}), with 𝝍^t(n1,n2)\hat{\boldsymbol{\psi}}_{t}(n_{1},n_{2}) the 22D Fourier coefficients of 𝝍t(𝜽){\boldsymbol{\psi}}_{t}({\bm{\theta}}) . In both the delocalized (a-d) and the localized (e-h) case the wavepacket is seen to propagate in all directions in the momentum space. However, neither diffusion nor localization are isotropic over the inspected range of times. As the quantity ItI_{t} is plotted in a logarithmic scale, the approximate equidistance of the level curves in (h) implies exponential localization. Anisotropy of diffusion is confirmed over much longer times by the plots in Fig. 3, which show both E1E_{1} and E2E_{2} vs time for two different choices of 𝝎\boldsymbol{\omega}. Diffusion in the n1n_{1} and in the n2n_{2} direction follows different rates, which in turn depend on 𝝎\boldsymbol{\omega}.

To explore the nature of the dynamical localization–delocalization transition, we further carry out numerical experiments on an auxiliary 11D quantum dynamical system defined as follows:

𝝍ˇt=Ts=1t𝑼ˇs𝝍ˇ0,\displaystyle\check{{\boldsymbol{\psi}}}_{t}\;=\;_{\rm T}\!\prod_{s=1}^{t}\check{{{\bm{U}}}}_{s}\check{\boldsymbol{\psi}}_{0},
𝑼ˇt:=eiαpˇ(θ1,θ2+ω2t)𝒅(θ1,θ2+ω2t)𝝈eiα(hen1)γ\displaystyle\check{{{\bm{U}}}}_{t}:=e^{-i\alpha\check{p}(\theta_{1},\theta_{2}+\omega_{2}t)\bm{d}(\theta_{1},\theta_{2}+\omega_{2}t)\boldsymbol{\cdot}\bm{\boldsymbol{\sigma}}}e^{-i\alpha(h_{e}n_{1})^{\gamma}} (27)

with ω2\omega_{2} the second component of the frequency vector 𝝎{\bm{\omega}} in SM models. Here 𝒅(𝜽)\bm{d}(\boldsymbol{\theta}) is the same as for SM Model I with β=0.8\beta=0.8, and pˇ(𝜽)=2πarctanKp(𝜽)\check{p}(\boldsymbol{\theta})=\frac{2}{\pi}\arctan\frac{K}{p(\boldsymbol{\theta})} with KK a strictly positive parameter and p(𝜽)p(\boldsymbol{\theta}) given by Eq. (18). In addition to KK, this auxiliary model has one more strictly positive parameter γ\gamma. The 11D equivalent (25) of SM Model I adopted in the above numerical experiments corresponds to K=+,γ=1K=+\infty,\gamma=1, while a special spinful QKR Beenakker11 that can exhibit a e\hbar_{e}-driven IQHE Tian14 ; Tian16 corresponds to finite KK and γ=2\gamma=2. Like the SM models, the quantum dynamics of this system can be probed by E1(t)=𝝍ˇt|θ12𝝍ˇt~θ2E_{1}(t)=-\langle\langle\check{\boldsymbol{\psi}}_{t}|\partial_{\theta_{1}}^{2}\check{\boldsymbol{\psi}}_{t}\rangle_{\tilde{\cal H}}\rangle_{\theta_{2}}.

Then we study how the α\alpha-profile of the diffusion coefficient limtE1(t)/t\lim_{t\rightarrow\infty}E_{1}(t)/t depend on K,γK,\gamma. For this purpose we choose a continuous contour in the parameter (K,γK,\gamma) space along which the spinful QKR (K=2,γ=2K=2,\gamma=2) adopted in previous numerical experiments Tian14 ; Tian16 is deformed into the SM Model I (K=+,γ=1K=+\infty,\gamma=1). The contour consists of two pieces: (i) γ\gamma is fixed to be 22 while KK increases from 22 to ++\infty; then (ii) KK is fixed to be ++\infty while γ\gamma decreases from 22 to 11. We compute the α\alpha profile of E1(t)/tE_{1}(t)/t at t=106t=10^{6} for different (K,γK,\gamma) along the contour. Results are shown in Fig. 4. Obviously, along the piece (i) of the contour every peak signalling the topological dynamical transition in the e\hbar_{e}-driven IQHE is pushed continuously towards the origin and eventually the mmth peak (counting from the left) arrives at α/π=m12\alpha/\pi=m-\frac{1}{2} when K=+K=+\infty; along the piece (ii) the peaks are pinned and their heights evolves continuously into those of the SM Model I. As such, though the SM models and the spinful QKR are very different, and though the supersymmetric field theory which was developed for the spinful QKR Tian14 ; Tian16 does not apply here, because of the absence of fast and slow mode separation Altland15 , the sequential dynamical localization–delocalization transitions exhibited by SM Model I evolve – via the continuous change in (K,γK,\gamma) along the contour – smoothly from the e\hbar_{e}-driven IQHE exhibited by the spinful QKR. Because in the latter a critical point separates two topologically distinct phases, whose topological angles differ by 2π2\pi, the transition observed in SM Model I is of topological nature. However, it should be noted that the topological theory for the spinful QKR cannot be applied here, because of the lack of chaoticity in SM models.

Refer to caption
Refer to caption
Figure 3: Numerical experiments on SM Model I for different frequency vectors 𝝎=(1,2π/5){\bm{\omega}}=(1,2\pi/\sqrt{5}) (a) and 𝝎=(2πx,2πx2){\bm{\omega}}=(2\pi x,2\pi x^{2}) (see text for the value of xx) (b) show that the critical diffusion is anisotropic and the diffusion coefficients depend on 𝝎{\bm{\omega}}.
Refer to caption
Figure 4: The dynamical transitions exhibited by SM Model I evolve smoothly from the e\hbar_{e}-driven IQHE exhibited by the spinful QKR. 𝝎=(1,2π/5){\bm{\omega}}=(1,2\pi/\sqrt{5}).

III.2.2 Stability of trajectories of the equivalent classical dynamics

The equivalent classical dynamics described by Eq. (11) was numerically simulated up to t=106t=10^{6} for different values of α\alpha. To probe the stability of phase space trajectories we compute the time profile of the deviation δt\delta_{t} of two trajectories which are initially close. Results are shown in Fig. 5. Time profiles of (δt/δ0)2(\delta_{t}/\delta_{0})^{2} corresponding to different α\alpha are shown in (a), for which the average over initial conditions is performed, and the α\alpha profiles of (δt/δ0)2/t(\delta_{t}/\delta_{0})^{2}/t at different long times are shown in (b). We see that, like E1,2(t)E_{1,2}(t), the long-time behavior of (δt/δ0)2(\delta_{t}/\delta_{0})^{2} is very sensitive to the value of α/π\alpha/\pi: for half-integer α/π\alpha/\pi it grows unboundedly, implying that the classical trajectory is unstable, whereas away from half-integer α/π\alpha/\pi it saturates at long times, implying that the classical trajectory is stable.

Moreover, for half-integer α/π\alpha/\pi the growth of (δt/δ0)2(\delta_{t}/\delta_{0})^{2} is linear (a). Corresponding to this, as time increases the α\alpha profile of (δt/δ0)2/t(\delta_{t}/\delta_{0})^{2}/t displays a sharp peak at half-integer α/π\alpha/\pi (b): limt(δt/δ0)2t\lim_{t\to\infty}\frac{(\delta_{t}/\delta_{0})^{2}}{t} is finite for half-integer α/π\alpha/\pi and vanishes otherwise. Therefore, the dynamical localization–delocalization transition in terms of quantum transport is translated into a transition in the stability of classical trajectories. Remarkably, at the critical point the deviation asymptotically grows approximately like t\sqrt{t}, which is dramatically different from both the exponential instability of trajectories in dynamical chaos and the linear instability that is generic for Hamiltonian integrable systems linstab .

Refer to caption
Refer to caption
Figure 5: Numerical experiments on SM Model I shows that the dynamical localization–delocalization transition shown in Fig. 1 is translated into a change in stability of classical trajectories at half-integer α/π\alpha/\pi. (a) The deviation of initially close trajectories grows unboundedly at half-integer α/π\alpha/\pi and it saturates at long times away from these points. (b) As tt tends to infinity (δt/δ0)2/t(\delta_{t}/\delta_{0})^{2}/t tends to a finite value at half-integer α/π\alpha/\pi while it tends to zero away from the critical point; 𝝎=(1,2π/5){\bm{\omega}}=(1,2\pi/\sqrt{5}).

III.3 Dynamics of SM Model II

The same numerical procedures were repeated with the constant in Eq. (19) set to unity. The observed dynamical behaviors of SM Model II are totally different from SM Model I. In particular, regardless of the value of α\alpha, E1(t)E_{1}(t) always saturates at long times and the long-time diffusion coefficient limtE1(t)t\lim_{t\rightarrow\infty}\frac{E_{1}(t)}{t} always vanishes (not shown). Thus, dynamical localization is always observed.

III.4 Dynamics of SM Model III

The same numerical procedures were repeated for SM Model III with d3=p(𝜽)(cos(θ1)+cos(θ2))d_{3}=p({\bm{\theta}})(\cos(\theta_{1})+\cos(\theta_{2})) and d3=0d_{3}=0, respectively. The observed dynamics are totally opposite to those of SM Model II. In particular, dynamical delocalization is observed regardless of the value of α\alpha, with the one exception of the points α/π\alpha/\pi\in\mathbb{Z}, for which 𝑼α,𝝎=±𝑻𝝎{\bm{U}}_{\alpha,{\bm{\omega}}}=\pm\bm{T}_{{\bm{\omega}}} leading to trivial localization.

IV Spectral theory of SM models, and its dynamical implications

In this section, spectral theory is used to explore the mechanisms of various quantum dynamical phenomena which were observed in numerical experiments on SM models. Some rigorous mathematical results are presented, and their dynamical implications are discussed. To prevent mathematical technicalities from obscuring the overall physical picture, some proofs are shuffled to Sec. VI.

IV.1 Spectral properties of general SM models

We first discuss the spectral properties of general SM models, without imposing any special features on the Vk(𝜽)V_{k}({\bm{\theta}}). Such properties crucially depend on arithmetic properties of the frequency vector 𝝎{\bm{\omega}}. We always require (ω1,ω2,π)(\omega_{1},\omega_{2},\pi) to be an incommensurate triple. A general result about the spectrum of the Floquet operator 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} is stated below, and is proven in Sec.VI. Exact operator-theoretic terminology will be confined to those proofs, so here we fix a few terms of a more informal language. The spectrum of a unitary operator 𝑼{\bm{U}} in a Hilbert space \cal H consists of complex numbers of the form eiχe^{i\chi} with χ\chi\in{\mathbb{R}}, and it cannot be empty. The point spectrum of 𝑼{\bm{U}} is that part of the spectrum which consists of proper eigenvalues, i.e., of eigenvalues which are associated with normalizable eigenfunctions – if there are any. Such eigenvalues are a countable set at most, and the corresponding eigenfunctions span a subspace p{\cal H}_{p} of \cal H, which is invariant under 𝑼{\bm{U}}; so its orthogonal complement subspace c{\cal H}_{c} is also invariant under 𝑼{\bm{U}}. If p{\cal H}_{p} is not the whole of \cal H, then 𝑼{\bm{U}}, restricted to c{\cal H}_{c}, is still unitary, so it has a spectrum, which by construction cannot contain any proper eigenvalue. This is, in fact, the continuous spectrum of 𝑼{\bm{U}}, and it can be informally pictured as the set of generalized eigenvalues, which are associated with eigenfunctions that are not normalizable, because they are, in some sense, extended. The spectral type of 𝑼{\bm{U}} is said to be pure, if either p{\cal H}_{p} or c{\cal H}_{c} is the whole \cal H; in the former case it is called pure point, and in the latter case it is called purely continuous. The set of all proper eigenvalues of 𝑼{\bm{U}} will be denoted Eig(𝑼){\rm Eig}({\bm{U}}).

Proposition 1.

The spectral type of 𝐔α,𝛚{\bm{U}}_{\alpha,{\bm{\omega}}} is pure. If it is pure point, then there is λ[0,π]\lambda\in[0,\pi] such that

Eig(𝑼α,𝝎)={ei(±λ+𝒏𝝎)}𝒏2.{\rm Eig}\,({\bm{U}}_{\alpha,{\bm{\omega}}})\,=\,\left\{e^{i(\pm\lambda+\boldsymbol{n}\boldsymbol{\cdot}\boldsymbol{\omega})}\right\}_{\boldsymbol{n}\in{\mathbb{Z}}^{2}}\;. (28)

In all cases, the spectrum is dense in the unit circle.

Remarks: (i) The point spectrum (28) can be degenerate only if it contains the eigenvalue ei𝒑𝝎/2+inπe^{i\boldsymbol{p}\boldsymbol{\cdot}\boldsymbol{\omega}/2+in\pi} for some nn\in{\mathbb{Z}}, 𝒑2\boldsymbol{p}\in{\mathbb{Z}}^{2}, and then it is doubly degenerate. (ii) The phase λ\lambda is not uniquely defined, because the same set (28) can be generated with infinitely many different choices of λ\lambda.

When a SM model is viewed in the classical picture, as described by Eq. (12), its spectral analysis is based on the Perron-Frobenius operator 𝓤α,𝝎{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}. The functions on the phase space Ω\Omega which depend only on 𝜽{\bm{\theta}} are a closed subspace 0{\mathfrak{H}}_{0} of \mathfrak{H}, that can be identified with L2(𝕋2,dm)L^{2}(\mathbb{T}^{2},dm). This subspace is invariant under 𝓤α,𝝎{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}. In Sec. VI we prove the following result:

Proposition 2.

(i) The spectrum of 𝓤α,𝛚{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}} in 0{\mathfrak{H}}_{0} is simple and pure point, with eigenphases 𝐧𝛚{\boldsymbol{n}}\boldsymbol{\cdot}{\boldsymbol{\omega}} where 𝐧(n1,n2)2{\boldsymbol{n}}\equiv(n_{1},n_{2})\in{\mathbb{Z}}^{2}. (ii) If the spectrum of 𝐔α,𝛚{\bm{U}}_{\alpha,{\bm{\omega}}} is pure point then the spectrum of 𝓤α,𝛚{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}} in {\mathfrak{H}} is pure point, and

Eig(𝓤α,𝝎)={ei(mλ+𝒏𝝎)}m,n1,n2,{\rm Eig}\,({\bm{{\cal U}}}_{\alpha,{\bm{\omega}}})\,=\,\left\{e^{i(m\lambda+{\boldsymbol{n}}\boldsymbol{\cdot}\boldsymbol{\omega})}\right\}_{m,n_{1},n_{2}\in{\mathbb{Z}}}\;, (29)

where the parameter λ\lambda is the same as that in the spectrum (28). (iii) If the spectrum of 𝐔α,𝛚{\bm{U}}_{\alpha,{\bm{\omega}}} is continuous (hence purely continuous by Proposition 1) then the spectrum of 𝓤α,𝛚{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}} in {\mathfrak{H}} has a continuous component.

Pure continuity of the spectrum of 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} does not by necessity entail pure continuity of the spectrum of 𝓤α,𝝎{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}} in the orthogonal complement 0{\mathfrak{H}}_{0}^{\bot} of 0{\mathfrak{H}}_{0}; the latter may still have a pure point component. This is somehow related to ergodicity of the classical description of SM models. Ergodicity is a capital notion in the theory of classical dynamical systems, and can be stated in purely spectral terms: 𝓢α,𝝎{\bm{{\cal S}}}_{\alpha,{\bm{\omega}}} is ergodic, if and only if 11 is a simple proper eigenvalue of 𝓤α,𝝎{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}} Sinai . Since SM models bear a quantum–classical juxtaposition, a natural question arises, how would ergodicity in the classical description of SM models be mirrored in the quantum description of SM models. The following result presents a partial answer. Loosely speaking, an operator in \cal H is fibered, if it does not couple different position eigenstates; more properly, if it acts as in 𝝍(𝜽)𝑶(𝜽)𝝍(𝜽)\boldsymbol{\psi}(\boldsymbol{\theta})\mapsto\bm{O}(\boldsymbol{\theta})\boldsymbol{\psi}(\boldsymbol{\theta}), where 𝑶\bm{O} is a map from 𝕋2\mathbb{T}^{2} to the 2×22\times 2 matrices.

Proposition 3.

If the dynamical system (Ω,𝓢α,𝛚)(\Omega,{\bm{{\cal S}}}_{\alpha,{\bm{\omega}}}) is ergodic, then the spectrum of 𝐔α,𝛚{\bm{U}}_{\alpha,{\bm{\omega}}} is continuous and ”fiberwise simple”; i.e., 𝐔α,𝛚{\bm{U}}_{\alpha,{\bm{\omega}}} does not commute with any nontrivial fibered linear operator.

Remark: This result cannot be reversed.

Though not necessary for the elaborations that follow, it can be shown that the spectral properties of the Floquet operator 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} are related to the cohomology of the classical dynamical system (Ω,𝓢α,𝝎)(\Omega,{\bm{{\cal S}}}_{\alpha,{\bm{\omega}}}); see Appendix A.

IV.2 Unitary symmetry 𝑨\bm{A} and duality

We now focus on a special class of SM models, which are exemplified by those considered in numerical experiments. A SM model belongs to this class, if it is defined by maps 𝒅\bm{d} that enjoy the following special symmetries:

  1. (i)

    d1(𝜽)d_{1}({\bm{\theta}}) is odd wrt θ1\theta_{1} and even wrt θ2\theta_{2};

  2. (ii)

    d2(𝜽)d_{2}({\bm{\theta}}) is even wrt θ1\theta_{1} and odd wrt θ2\theta_{2};

  3. (iii)

    d3(𝜽)d_{3}({\bm{\theta}}) is even wrt both θ1\theta_{1} and θ2\theta_{2}.

It has been previously found Tian14 ; Tian16 that such symmetries are crucial to the IQHE-like phenomenon in the spinful QKR. SM models endowed with such symmetries will be termed class-* models. The SM Models I, II and III, as given in Eqs. (18), (19) and (22), are class-* models.

A key role will be played by the following Proposition 4, that crucially rests on the above conditions (i), (ii) and (iii), and its corollaries.

Proposition 4.

For any class-* model a unitary symmetry 𝐀\boldsymbol{A} note_unitary exists, such that

𝑼πα,𝝎=𝑨𝑼α,𝝎𝑨,𝑼2πα,𝝎=𝑨𝑼α,𝝎𝑨.\boldsymbol{U}^{\dagger}_{{\pi-\alpha,\boldsymbol{\omega}}}\;=\;-\;\boldsymbol{A}\;\boldsymbol{U}_{{\alpha,\boldsymbol{\omega}}}\;\boldsymbol{A}\;,\,\boldsymbol{U}^{\dagger}_{{2\pi-\alpha,\boldsymbol{\omega}}}\;=\;\boldsymbol{A}\;\boldsymbol{U}_{{\alpha,\boldsymbol{\omega}}}\;\boldsymbol{A}\;. (30)

Proof : The first identity in Eq. (30) will be proven first; the proof of the second is essentially identical. First of all, from the general definition of 𝑴α\bm{M}_{\alpha} it follows that

𝑴2πα=𝑴α,𝑴π+α=𝑴α,𝑴α=𝑴πα,\displaystyle{\boldsymbol{M}}_{2\pi-\alpha}={\boldsymbol{M}}_{\alpha}^{\dagger},\,\,\boldsymbol{M}_{\pi+\alpha}=-{\boldsymbol{M}}_{\alpha},\,\,{\boldsymbol{M}}_{\alpha}^{\dagger}=-\boldsymbol{M}_{\pi-\alpha}, (31)

which hold true independently of the symmetries of 𝒅\bm{d}. Let 𝑹𝜼\boldsymbol{R}_{{\boldsymbol{\eta}}} denote the reflection operator in a point 𝜼𝕋2\boldsymbol{\eta}\in\mathbb{T}^{2}: that is, (𝑹𝜼𝝍)(𝜽)=𝝍(2𝜼𝜽)(\boldsymbol{R}_{{\boldsymbol{\eta}}}\boldsymbol{\psi})({\bm{\theta}})=\boldsymbol{\psi}(2\boldsymbol{\eta}-{\bm{\theta}}). From 𝑼α,𝝎=𝑻𝝎𝑴α\boldsymbol{U}^{\dagger}_{\alpha,{\bm{\omega}}}=\boldsymbol{T}_{-\boldsymbol{\omega}}\boldsymbol{M}_{\alpha}^{\dagger} and the third identity in Eq. (31), we obtain that

𝑼α,𝝎=𝑻𝝎𝑼πα,𝝎𝑻𝝎.\boldsymbol{U}^{\dagger}_{\alpha,\boldsymbol{\omega}}\,=\,-\;\boldsymbol{T}_{-\boldsymbol{\omega}}\,\boldsymbol{U}_{\pi-\alpha,-\boldsymbol{\omega}}\,\boldsymbol{T}_{\boldsymbol{\omega}}. (32)

Thanks to the assumed symmetries (i), (ii) and (iii), the unitary operator 𝑩=𝝈3𝑹𝟎\boldsymbol{B}=\boldsymbol{\sigma}_{3}\boldsymbol{R}_{\boldsymbol{0}} commutes with 𝑴α\boldsymbol{M}_{\alpha}, and 𝑩𝑻𝝎𝑩=𝑻𝝎\boldsymbol{B}\boldsymbol{T}_{\boldsymbol{\omega}}\boldsymbol{B}^{\dagger}=\boldsymbol{T}_{-\boldsymbol{\omega}}. As a result,

𝑼α,𝝎=𝑩𝑼α,𝝎𝑩.\boldsymbol{U}_{\alpha,-\boldsymbol{\omega}}\,=\,\boldsymbol{B}\,\boldsymbol{U}_{\alpha,\boldsymbol{\omega}}\,\boldsymbol{B}^{\dagger}\,. (33)

Applying this to the right-hand side of Eq. (32) gives

𝑼α,𝝎=(𝑻𝝎𝑩)𝑼πα,𝝎(𝑻𝝎𝑩).\boldsymbol{U}^{\dagger}_{\alpha,\boldsymbol{\omega}}\,=\,-(\boldsymbol{T}_{-\boldsymbol{\omega}}\boldsymbol{B})\,\boldsymbol{U}_{\pi-\alpha,\boldsymbol{\omega}}\,(\boldsymbol{T}_{-\boldsymbol{\omega}}\boldsymbol{B})^{\dagger}\;. (34)

Therefore, the first identity in Eq. (30) is true with 𝑨\bm{A} given by:

𝑨=𝑻𝝎𝑩=𝝈3𝑻𝝎𝑹𝟎=𝝈3𝑹𝝎/2.\boldsymbol{A}\;=\;\boldsymbol{T}_{-\boldsymbol{\omega}}\boldsymbol{B}\;=\;\boldsymbol{\sigma}_{3}\boldsymbol{T}_{-\boldsymbol{\omega}}\boldsymbol{R}_{\boldsymbol{0}}\;=\;\boldsymbol{\sigma}_{3}\boldsymbol{R}_{{-\boldsymbol{\omega}/2}}\;\,\,\,. (35)

To prove the second identity in Eq. (30), we use the first identity in Eq. (31) instead of the third. \Box

Remark: Together with 𝑨2=1\bm{A}^{2}=1, Eq. (30) implies, in particular, that the evolution operator i𝑼α,𝝎i\boldsymbol{U}_{\alpha,\,\boldsymbol{\omega}} at α=π/2\alpha=\pi/2, 3π/23\pi/2, 0 and π\pi are time-reversal invariant. It is worth noting that also the associated classical map which is defined in Eq. (11) is time-reversal invariant; the time-reversing involution is (𝜽,ϕ)(𝜽𝝎,𝝈3ϕ)({\bm{\theta}},\,\boldsymbol{\phi})\rightarrow(-{\bm{\theta}}-\boldsymbol{\omega},\,-\boldsymbol{\sigma}_{3}\boldsymbol{\phi}).

The symmetry 𝑨\bm{A} establishes a ”duality”  nature between models whose α\alpha values in [0,π][0,\pi] are symmetric with respect to π/2\pi/2. In other words, the model at a value α0\alpha_{0} of the parameter α\alpha is dual to the model at α=πα0\alpha=\pi-\alpha_{0}. The most immediate consequence of this duality is:

Eig(𝑼πα,𝝎)=Eig(𝑼α,𝝎),{\rm Eig}(\boldsymbol{U}_{{\pi-\alpha,\,\boldsymbol{\omega}}})\;=\;-\;{\rm Eig}(\boldsymbol{U}_{{\alpha,\,\boldsymbol{\omega}}})\,, (36)

because Eig(𝑼α,𝝎){\rm Eig}(\boldsymbol{U}_{{\alpha,\,\boldsymbol{\omega}}}) is closed under complex conjugation, thanks to the symmetry relation (5).

Models with α=π/2\alpha=\pi/2 are self-dual. This self-dual symmetry, i.e. the aforementioned time-reversal invariance has nontrivial consequences:

Corollary 1.

If the operator 𝐔π2,𝛚\boldsymbol{U}_{\frac{\pi}{2},\,\boldsymbol{\omega}} of a class-* model has a non-empty point spectrum, then there is 𝐫2\boldsymbol{r}\in{\mathbb{Z}}^{2} so that

Eig(𝑼π2,𝝎)={±iei(𝒏+𝒓/2)𝝎}𝒏2.{\rm Eig}(\boldsymbol{U}_{\frac{\pi}{2},\,\boldsymbol{\omega}})\;=\;\bigl{\{}\pm i\;e^{i(\boldsymbol{n}+\boldsymbol{r}/2)\boldsymbol{\cdot}\boldsymbol{\omega}}\bigr{\}}_{\boldsymbol{n}\in{\mathbb{Z}}^{2}}\;. (37)

The corresponding eigenfunctions 𝐯𝐧±\boldsymbol{v}^{\pm}_{\boldsymbol{n}} satisfy:

𝑨𝒗𝒏±(𝜽)=c𝒏±ei(2𝒏+𝒓)𝜽𝒗𝒏±(𝜽)with(c𝒏±)2=ei(2𝒏+𝒓)𝝎.\boldsymbol{A}\boldsymbol{v}^{\pm}_{\boldsymbol{n}}({\bm{\theta}})=c^{\pm}_{\boldsymbol{n}}e^{i(2\boldsymbol{n}+\boldsymbol{r})\boldsymbol{\cdot}{\bm{\theta}}}\boldsymbol{v}^{\pm}_{\boldsymbol{n}}({\bm{\theta}})\;\;\mbox{\rm with}\;\;(c^{\pm}_{\boldsymbol{n}})^{2}=e^{i(2\boldsymbol{n}+\boldsymbol{r})\boldsymbol{\cdot}\boldsymbol{\omega}}\;. (38)

This corollary imposes quite demanding conditions on self-dual eigenfunctions (if any). In the case of SM Model III, they force the absence of proper eigenfunctions, so pure spectral continuity follows:

Corollary 2.

The operators 𝐔π2,𝛚{\bm{U}}_{\frac{\pi}{2},\,\boldsymbol{\omega}} of SM Model III have purely continuous spectra.

For SM Model I we cannot assess the spectral type of 𝑼π2,𝝎{\bm{U}}_{\frac{\pi}{2},\,\boldsymbol{\omega}}. However the following weaker result has crucial dynamical consequences:

Corollary 3.

If 𝐔π2,𝛚{\bm{U}}_{\frac{\pi}{2},\,\boldsymbol{\omega}} of SM Model I has any proper eigenfunctions 𝐮\boldsymbol{u}, then such eigenfunctions are “delocalized”, in the sense that E1,2(𝐮)=+E_{1,2}(\boldsymbol{u})=+\infty, where EjE_{j} are defined in Eq. (II.3).

A proper mathematical statement should be: 𝒖\boldsymbol{u} does not belong to the domain of the momentum operators.

IV.3 Spectral properties vs dynamical delocalization

The exact spectral results presented above provide a basis for the study of the dynamical delocalization which is observed in numerical experiments. It is important to point out that while such exact results hold for any incommensurate triple (ω1,ω2,π)(\omega_{1},\omega_{2},\pi), the same is not true of other more detailed results to be discussed in this subsection, which strongly rely on the value of 𝝎\boldsymbol{\omega} chosen as in Sec. III.2.1, and can reasonably be taken as representatives of a wide class of strongly incommensurate situations.

Based on the definition (II.3) of EjE_{j} (j=1,2j=1,2) and EE, we give a precise definition of dynamical delocalization/localization in momentum space. For given α,𝝎\alpha,{\bm{\omega}}, the quantum dynamics 𝝍t=𝑼α,𝝎t𝝍{\boldsymbol{\psi}}_{t}={\bm{U}}^{t}_{\alpha,{\bm{\omega}}}\,{\boldsymbol{\psi}} (tt\in\mathbb{Z}) will be said to exhibit dynamical delocalization if for any 𝝍{\boldsymbol{\psi}} such that E(𝝍)<+E({\boldsymbol{\psi}})<+\infty the following relation holds:

lim suptE(𝝍t)=+\limsup_{t\to\infty}\,E({\boldsymbol{\psi}}_{t})\,=\,+\infty (39)

On the opposite, it will be said to exhibit dynamical localization whenever the right-hand side is finite. It is important to note that if Ej(𝝍)<+E_{j}({\boldsymbol{\psi}})<+\infty , then Ej(𝝍t)<+E_{j}({\boldsymbol{\psi}}_{t})<+\infty at all times whenever the map 𝑴α(𝜽)\boldsymbol{M}_{\alpha}(\boldsymbol{\theta}) that defines a SM model is smooth. This is not the case, e.g. with SM Model III, where dynamical delocalization is somewhat trivial, as it will be shown below in Sec.  VI.5. Next we recall that dynamical localization implies pure point spectra, and, equivalently, continuous spectra imply dynamical delocalization scho . So, for SM Model I (away from critical points) and Model II, we have strong numerical evidence of pure point spectra, because dynamical localization is clearly observed numerically. However, we have no mathematical proof (except, of course, in the trivial cases when α=nπ,n\alpha=n\pi,n\in\mathbb{Z}). Delocalization is produced by continuous spectra, but it can also occur with pure point spectra, for various reasons, e.g. non-uniform localization of proper eigenfunctions in momentum space, or else ”delocalization” of eigenfunctions: meaning that proper eigenfunctions, although normalizable (i.e. not extended) do not yield a finite expectation value for the squared momentum operator. This follows from a general fact, which in our case can be stated as follows:

Proposition 5.

Let the spectrum of a unitary operator 𝐔\boldsymbol{U} in \cal H be simple and pure point, with delocalized proper eigenfunctions. Then the quantum evolution that is generated by 𝐔\boldsymbol{U} in \cal H is dynamically delocalized.

A compact proof of this standard result is given in Sec. VI. In view of Corollary 3, this may be the case with SM Model I at the critical point α=π/2\alpha=\pi/2, which is self-dual. So, even though we cannot exactly assess the spectral type of that model, we can nevertheless state the following exact result, that provides a rigorous confirmation for numerical observations:

Proposition 6.

SM Model I is dynamically delocalized at the self-dual critical point α=π/2\alpha=\pi/2.

Proof : Should the spectrum of 𝑼π2,𝝎\boldsymbol{U}_{\frac{\pi}{2},\,\boldsymbol{\omega}} be pure point, the claim would immediately follow from Corollary 3 and Proposition 5. The one possible alternative is pure continuity, by spectral dichotomy. If that were the case, then the claim would be trivial. \Box

For SM Model III the situation is simpler:

Proposition 7.

SM Model III is dynamically delocalized at all αnπ,n\alpha\neq n\pi,n\in\mathbb{Z}. At least at the critical values of α\alpha, the spectrum is purely continuous.

Proof : Delocalization follows from the fact that 𝑴α(𝜽)\bm{M}_{\alpha}(\boldsymbol{\theta}) exhibits points of essential discontinuity, i.e. points where it has infinitely many distinct limit values. Such are the points where the normalization factor p(𝜽)p(\boldsymbol{\theta}) vanishes. Whatever the initial state 𝝍{\boldsymbol{\psi}}, due to the very definition of 𝑼α,𝝎{\bm{U}}_{\alpha,\boldsymbol{\omega}} such discontinuities are sooner or later transferred to 𝝍t{\boldsymbol{\psi}}_{t} whenever αnπ\alpha\neq n\pi, and they cannot be removed by re-defining 𝝍t{\boldsymbol{\psi}}_{t} in a set of measure zero. In contrast, a state for which EjE_{j} has a finite value must be defined by functions which are, almost everywhere at least, equal to everywhere continuous functions. Continuity of the spectrum is proven in Sec. VI.5. \Box

This result receives intuitive support in the tight-binding-like formulation of the SM model in Appendix B, because in the presence of discontinuities the hopping terms in the lattice Hamiltonian (B) quite slowly decay at long distances in momentum space.

Though not proven, continuity of the spectrum of SM Model I at the critical point is numerically supported, as follows. For integer tt and 𝝍\boldsymbol{\psi}\in\cal H, let us define:

P(𝝍,t)1ts=0t1|𝝍|𝝍s|2,P(\boldsymbol{\psi},t)\;\equiv\;\frac{1}{t}\sum\limits_{s=0}^{t-1}|\langle\boldsymbol{\psi}|\boldsymbol{\psi}_{s}\rangle_{\cal H}|^{2}\;,

where α,𝝎\alpha,\boldsymbol{\omega} are left understood, and as usual 𝝍s=𝑼α,𝝎s𝝍\boldsymbol{\psi}_{s}={\bm{U}}_{\alpha,{\bm{\omega}}}^{s}\boldsymbol{\psi}. This is the time-averaged probability of survival in the initial state 𝝍\boldsymbol{\psi}. Wiener’s theorem scho states that

limtP(𝝍,t)=k|𝒖k|𝝍|4,\lim\limits_{t\to\infty}P(\boldsymbol{\psi},t)\;=\;\sum\limits_{k}|\langle\boldsymbol{u}_{k}|\boldsymbol{\psi}\rangle_{\cal H}|^{4}\;, (40)

where the sum on the right-hand side is over all the proper eigenvectors 𝒖k\boldsymbol{u}_{k} of the Floquet operator 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} (with an arbitrary numbering). Vanishing of the limit in the left-hand side of Eq. (40) is thus a necessary and sufficient condition for pure continuity of the spectrum. Numerical results of 22D simulations for P(𝝍,t)P(\boldsymbol{\psi},t) are shown in Fig. 6 for SM Models I and III, with 𝝍\boldsymbol{\psi} chosen to be the n1=n2=0n_{1}=n_{2}=0 state, which is constant over 𝕋2\mathbb{T}^{2}. As shown in (a), for SM Model I and noncritical α\alpha, P(𝝍,t)P(\boldsymbol{\psi},t) clearly stabilizes at large tt, consistently with the pure point spectrum that is implied by the observed dynamical localization. At critical α\alpha, P(𝝍,t)P(\boldsymbol{\psi},t) appears to decay tν\propto t^{-\nu} with ν\nu equal or close to 11 (a final slight deviation is probably due to finite-basis effects). Such decay is almost identical to the one which is observed for SM Model III at same α\alpha [shown in (b)], where continuous spectrum is exactly proven. On such grounds spectral continuity of the critical model I is convincingly supported.

Next we address the subtler question, whether the spectrum of the critical model is absolutely continuous (a.c.) or singular continuous (s.c.). The latter type of spectrum is characterized by the fact that, despite continuity, the spectral expansions of states 𝝍\boldsymbol{\psi}\in\cal H concede positive (nonzero) weight to regions of the spectrum which have zero measure as subsets of the unit circle {eiχ,χ}\{e^{i\chi},\chi\in{\mathbb{R}}\}. It is frequently associated with Cantor-like structures and multifractality. No exact analysis will be attempted here. However, the absence of s.c. spectrum – hence, pure absolute continuity of the spectrum – in the critical model I is numerically supported as follows. It is an exact result dtwo that the decay exponent ν\nu of P(𝝍,t)P(\boldsymbol{\psi},t) is equal to the multifractal dimension D2D_{2} of the local density of states (LDOS) associated with the state 𝝍\boldsymbol{\psi}; and D2=1D_{2}=1 is a strong indication footnote_ph that the continuous spectrum is actually a.c.. This appears to be the case with SM Model I at criticality, because D2=ν1D_{2}=\nu\approx 1 according to Fig. 6.

Refer to caption
Refer to caption
Figure 6: 22D simulation of the quantum evolution of an initial state 𝝍{\boldsymbol{\psi}} localized at n1=n2=0n_{1}=n_{2}=0 for different α\alpha. The results for the decay of the corresponding survival probabilities for SM Models I and III (with d3=0d_{3}=0) are shown in (a) and (b), respectively. In both cases, 𝝎=(2πx,2πx2){\bm{\omega}}=(2\pi x,2\pi x^{2}) (see text for the value of xx).

IV.4 Quantum transport from a.c. spectrum

We finally address quantum transport and show that numerical results are consistent with a.c. spectrum. Whenever 𝑴α\boldsymbol{M}_{\alpha} is a smooth function, the propagation over the 22D 𝑵\boldsymbol{N}-lattice is subject to a ballistic bound; that is, the asymptotic growth of E(𝝍t)E(\boldsymbol{\psi}_{t}) cannot be faster than quadratic in time. This follows from Eq. (II.3), and is equivalent to the fact that classical trajectories can be linearly unstable at most [see the comments after Eq. (56)]. The transport on the 11D momentum lattice which is associated with the standard Maryland model and the QKR model are also subject to a ballistic bound. The following lower bound is valid for transport over discrete lattices of arbitrary dimension DD induced by a unitary evolution group with a.c. spectrum scho ; jmp :

1ts=0t1E(𝝍s)const.t2D,(t).\frac{1}{t}\sum\limits_{s=0}^{t-1}E(\boldsymbol{\psi}_{s})\;\geq\;\mbox{\rm const.}\,t^{\frac{2}{D}}\;,\,\,\,(t\to\infty)\,. (41)

where E(𝝍)E(\boldsymbol{\psi}) is the DD-dimensional generalization of the definition in Eq. (II.3). With D=1D=1, and in the presence of a ballistic bound, this inequality shows that a.c. spectra enforce ballistic transport footnote_bound . Not so in higher dimension; for example, with D=3D=3 it has been proved JSP that a.c. spectra can coexist with different diffusion exponents. In SM models the ballistic bound is attained, whenever (ω1,ω2,π)(\omega_{1},\omega_{2},\pi) is a commensurate triple. Then the quantum evolution enjoys translation invariance in momentum space, leading to band a.c. spectra and quadratic growth of E(𝝍t)E(\boldsymbol{\psi}_{t}). This scenario resembles the well-known ”quantum resonances” of the QKR qkres1 ; qkres2 , but the band structure is likely to be much more complex. This issue will not be investigated in the present paper, which is restricted to the incommensurate case. In the strongly incommensurate situations which were numerically simulated, SM Model I at the critical value of α\alpha exhibits diffusive (or close to diffusive) transport, i.e. the long-time growth of E(𝝍t)E(\boldsymbol{\psi}_{t}) and E1,2(𝝍t)E_{1,2}(\boldsymbol{\psi}_{t}) is linear in time. It is not isotropic in momentum space, and the degree of anisotropy depends on the choice of 𝝎\boldsymbol{\omega}; however, it should be noted that diffusion is not restricted to special directions. Diffusive growth is consistent with the bound (41) with D=2D=2. The following heuristic argument indicates that indeed it is compatible with a.c. spectrum (in this case). We start with the equation:

Ej(𝝍t)=s=0t1𝓤α,𝝎sHj2+s=0t1𝓤α,𝝎sKj2,E_{j}({\boldsymbol{\psi}}_{t})=\bigl{\|}\,\sum_{s=0}^{t-1}{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}^{s}H_{j}\,\bigr{\|}_{\mathfrak{H}}^{2}\,+\,\bigl{\|}\,\sum_{s=0}^{t-1}{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}^{s}K_{j}\,\bigr{\|}^{2}_{\mathfrak{H}}, (42)

which is derived in Sec. V exploiting the classical picture of SM models. The functions Hj=Hj(𝜽,ϕ)H_{j}=H_{j}(\boldsymbol{\theta},\boldsymbol{\phi}), Kj=Kj(𝜽,ϕ)K_{j}=K_{j}(\boldsymbol{\theta},\boldsymbol{\phi}) are defined on the phase space Ω\Omega of the classical model and are specified immediately after Eq. (47). Their explicit forms are not important here, beyond the easily checked fact that they are orthogonal to the subspace of the spin-independent functions, where 𝓤α,𝝎{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}} has a pure point spectrum according to Proposition 2. If the spectrum of 𝑼α,𝝎\boldsymbol{U}_{\alpha,{\bm{\omega}}} is a.c., then the spectrum of 𝓤α,𝝎{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}} has an a.c. component too (see the statement in the beginning of Sec.  VI.2). Then a formal elaboration can be implemented. Let eiχe^{i\chi} denote generalized eigenvalues of 𝓤α,𝝎{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}, and |χ,ξ|\chi,\xi\rangle the corresponding generalized eigenvectors in Dirac notation; ξ\xi collectively denotes any additional quantum numbers. Then

s=0t1𝓤α,𝝎sHj2=02π𝑑χsin2(χt2)sin2(χ2)ΛHj(χ).\|\,\sum\limits_{s=0}^{t-1}\;{\bm{{\cal U}}}_{\alpha,\,\boldsymbol{\omega}}^{s}H_{j}\,\|_{\mathfrak{H}}^{2}\;=\;\int_{0}^{2\pi}d\chi\,\frac{\sin^{2}\bigl{(}\frac{\chi t}{2}\bigr{)}}{\sin^{2}\bigl{(}\frac{\chi}{2}\bigr{)}}\Lambda_{H_{j}}(\chi)\,. (43)

where ΛHj(χ)ξ|χ,ξ|Hj|2\Lambda_{H_{j}}(\chi)\equiv\sum_{\xi}|\langle\chi,\xi|H_{j}\rangle|^{2} is the LDOS of |Hj|H_{j}\rangle wrt 𝓤α,𝝎{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}. Under the assumption of a.c. spectrum, this is an ordinary function (i.e. it is not Dirac δ\delta-like or fractal-like), which will be further assumed to be smooth. A fully similar equation holds for the functions KjK_{j}. The fraction under the integral sign in the right-hand side of Eq. (43), divided by 2πt2\pi t, in the limit tt\to\infty tends to a Dirac δ\delta. Then Eq. (42) shows that the asymptotic growth of E(𝝍t)E(\boldsymbol{\psi}_{t}) is linear in time. This result for SM models is not generic for quantum dynamics on 22D lattices, as it rests on the classical description of SM models which underlies Eq. (42).

V Equivalent classical description of quantum transport

In Sec. II it was shown that SM models can be viewed from either a ”quantum” viewpoint or from a ”classical”. Combined with spectral theory, this offers advantages for exploring the SM dynamics from different perspectives. In particular, in this section we show that the quantum transport in momentum space is translated into instability of trajectories (𝜽t,ϕt):=𝓢α,𝝎t(𝜽,ϕ)({\bm{\theta}}_{t},\boldsymbol{\phi}_{t}):={\bm{{\cal S}}}_{\alpha,{\bm{\omega}}}^{t}({\bm{\theta}},\boldsymbol{\phi}) in the phase space Ω\Omega. It is assumed throughout this section that 𝑴α(𝜽)\boldsymbol{M}_{\alpha}({\bm{\theta}}) is everywhere differentiable in 𝕋2\mathbb{T}^{2}.

V.1 Instability of classical trajectories

The linear stability of a trajectory (𝜽t,ϕt)({\bm{\theta}}_{t},\boldsymbol{\phi}_{t}) is determined by the behavior in time of the derivatives θjϕt(𝜽,ϕ)\partial_{\theta_{j}}\boldsymbol{\phi}_{t}({\bm{\theta}},\boldsymbol{\phi}) with j=1,2j=1,2. From the classical dynamical equation (11) it follows that

θjϕt+1=𝑴α(𝜽t+1)θjϕt+(θj𝑴α(𝜽t+1))ϕt.\partial_{\theta_{j}}\boldsymbol{\phi}_{t+1}\;=\;\boldsymbol{M}_{\alpha}({\bm{\theta}}_{t+1})\partial_{\theta_{j}}\boldsymbol{\phi}_{t}\;+\;(\partial_{\theta_{j}}\boldsymbol{M}_{\alpha}({\bm{\theta}}_{t+1}))\;\boldsymbol{\phi}_{t}\;. (44)

Define vectors 𝝃j,t2\boldsymbol{\xi}_{j,t}\in{\mathbb{C}}^{2} as follows:

θjϕt=Ts=0t𝑴α(𝜽s)𝝃j,t,\partial_{\theta_{j}}\boldsymbol{\phi}_{t}\;=\;_{\rm T}\prod\limits_{s=0}^{t}\;\boldsymbol{M}_{\alpha}({\bm{\theta}}_{s})\;\boldsymbol{\xi}_{j,t}\;, (45)

where 𝝃j,0=0\boldsymbol{\xi}_{j,0}=0. Then by performing the summation: s=0t1θjϕs\sum_{s=0}^{t-1}\partial_{\theta_{j}}\boldsymbol{\phi}_{s}, for which θjϕs\partial_{\theta_{j}}\boldsymbol{\phi}_{s} is substituted by Eq. (45), we obtain that

ϕ0𝝃j,t=s=0t1Hj(𝜽s,ϕs)\boldsymbol{\phi}_{0}\,\boldsymbol{\cdot}\;\boldsymbol{\xi}_{j,t}\;=\;\sum\limits_{s=0}^{t-1}H_{j}({\bm{\theta}}_{s},\boldsymbol{\phi}_{s})\; (46)

and, similarly,

𝝆ϕ0𝝃j,t=s=0t1Kj(𝜽s,ϕs),\boldsymbol{\rho}\boldsymbol{\phi}_{0}\,\boldsymbol{\cdot}\;\boldsymbol{\xi}_{j,t}\;=\;\sum\limits_{s=0}^{t-1}K_{j}({\bm{\theta}}_{s},\boldsymbol{\phi}_{s})\;, (47)

where Hj(𝜽,ϕ)=ϕ𝑴α(𝜽+𝝎)θj𝑴α(𝜽+𝝎)ϕH_{j}({\bm{\theta}},\boldsymbol{\phi})=\boldsymbol{\phi}\boldsymbol{\cdot}\boldsymbol{M}_{\alpha}^{\dagger}({\bm{\theta}}+\boldsymbol{\omega})\partial_{\theta_{j}}\boldsymbol{M}_{\alpha}({\bm{\theta}}+\boldsymbol{\omega})\boldsymbol{\phi} and Kj(𝜽,ϕ)=𝝆ϕ𝑴α(𝜽+𝝎)θj𝑴α(𝜽+𝝎)ϕK_{j}({\bm{\theta}},\boldsymbol{\phi})=\boldsymbol{\rho}\boldsymbol{\phi}\boldsymbol{\cdot}\boldsymbol{M}_{\alpha}^{\dagger}({\bm{\theta}}+\boldsymbol{\omega})\partial_{\theta_{j}}\boldsymbol{M}_{\alpha}({\bm{\theta}}+\boldsymbol{\omega})\boldsymbol{\phi}. Because of 𝝆ϕ0ϕ0=0\boldsymbol{\rho}\boldsymbol{\phi}_{0}\boldsymbol{\cdot}\boldsymbol{\phi}_{0}=0, taking squared moduli in both equations, and summing them together, we obtain that

|θjϕt|2=|𝝃j,t|2=|s=0t1Hj(𝜽s,ϕs)|2+|s=0t1Kj(𝜽s,ϕs)|2,\boldsymbol{|}\partial_{\theta_{j}}\boldsymbol{\phi}_{t}\boldsymbol{|}^{2}=\boldsymbol{|}\boldsymbol{\xi}_{j,t}\boldsymbol{|}^{2}=\bigl{|}\sum\limits_{s=0}^{t-1}H_{j}({\bm{\theta}}_{s},\boldsymbol{\phi}_{s})\bigr{|}^{2}+\bigl{|}\sum\limits_{s=0}^{t-1}K_{j}({\bm{\theta}}_{s},\boldsymbol{\phi}_{s})\bigr{|}^{2}, (48)

where ||\boldsymbol{|}\cdot\boldsymbol{|} stands for the 2{\mathbb{C}}^{2} norm. As Hj,KjH_{j},K_{j} are by assumption bounded functions, this result shows that trajectories (𝜽t,ϕt)({\bm{\theta}}_{t},\boldsymbol{\phi}_{t}) are, at worst, linearly unstable, in sharp contrast to the exponential instability of trajectories for dynamical chaos Chirikov79 .

To proceed we let

G𝝍,t(𝜽,ϕ)\displaystyle G_{{\boldsymbol{\psi}},t}({\bm{\theta}},\boldsymbol{\phi}) :=\displaystyle:= (𝓤α,𝝎tG𝝍)(𝜽,ϕ)=G𝝍(𝓢α,𝝎t(𝜽,ϕ))\displaystyle({\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}^{-t}G_{{\boldsymbol{\psi}}})({\bm{\theta}},\boldsymbol{\phi})=G_{{\boldsymbol{\psi}}}({\bm{{\cal S}}}^{t}_{\alpha,{\bm{\omega}}}({\bm{\theta}},\boldsymbol{\phi})) (49)
=\displaystyle= G𝝍(𝜽t,ϕt)\displaystyle G_{{\boldsymbol{\psi}}}({\bm{\theta}}_{t},\boldsymbol{\phi}_{t})

with the function G𝝍G_{{\boldsymbol{\psi}}} defined as in Eq. (14). Taking the derivative wrt θj\theta_{j} on both sides, we obtain that

θjG𝝍,t(𝜽,ϕ)=𝝍(𝜽t)θjϕt+θj𝝍(𝜽t)ϕt.\partial_{\theta_{j}}G_{{\boldsymbol{\psi}},t}({\bm{\theta}},\boldsymbol{\phi})\;=\;{\boldsymbol{\psi}}({\bm{\theta}}_{t})\boldsymbol{\cdot}\partial_{\theta_{j}}\boldsymbol{\phi}_{t}\;+\;\partial_{\theta_{j}}{\boldsymbol{\psi}}({\bm{\theta}}_{t})\boldsymbol{\cdot}\boldsymbol{\phi}_{t}\;. (50)

On the other hand, thanks to Eq. (15) G𝝍,t(𝜽,ϕ)=𝝍t(𝜽)ϕG_{{\boldsymbol{\psi}},t}({\bm{\theta}},\boldsymbol{\phi})={\boldsymbol{\psi}}_{-t}({\bm{\theta}})\boldsymbol{\cdot}\boldsymbol{\phi}, where the initial spinor 𝝍0𝝍{\boldsymbol{\psi}}_{0}\equiv{\boldsymbol{\psi}}. Taking the derivative wrt θj\theta_{j} on both sides of this equation entails

θjG𝝍,t(𝜽,ϕ)=θj𝝍t(𝜽)ϕ.\partial_{\theta_{j}}G_{{\boldsymbol{\psi}},t}({\bm{\theta}},\boldsymbol{\phi})\;=\;\partial_{\theta_{j}}{\boldsymbol{\psi}}_{-t}({\bm{\theta}})\,\boldsymbol{\cdot}\,\boldsymbol{\phi}\;. (51)

Choosing the initial spinor 𝝍{\boldsymbol{\psi}} to be independent of 𝜽{\bm{\theta}} and equating Eqs. (50) and (51), we obtain that

θjϕt(𝜽,ϕ)𝝍=ϕθj𝝍t(𝜽).\partial_{\theta_{j}}\boldsymbol{\phi}_{t}({\bm{\theta}},\boldsymbol{\phi})\boldsymbol{\cdot}{\boldsymbol{\psi}}\;=\;\boldsymbol{\phi}\boldsymbol{\cdot}\partial_{\theta_{j}}{\boldsymbol{\psi}}_{-t}({\bm{\theta}})\;. (52)

Replacing 𝝍{\boldsymbol{\psi}} by 𝝆𝝍\boldsymbol{\rho}\,{\boldsymbol{\psi}}, a short calculation shows that

θjϕt(𝜽,ϕ)𝝆𝝍=θj𝝍t(𝜽)𝝆ϕ.\partial_{\theta_{j}}\boldsymbol{\phi}_{t}({\bm{\theta}},\boldsymbol{\phi})\boldsymbol{\cdot}\boldsymbol{\rho}{\boldsymbol{\psi}}\;=\;-\partial_{\theta_{j}}{\boldsymbol{\psi}}_{-t}({\bm{\theta}})\boldsymbol{\cdot}\boldsymbol{\rho}\boldsymbol{\phi}. (53)

Summing squared moduli in the last two equations yields

|θjϕt(𝜽,ϕ)|2=|θj𝝍t(𝜽)|2.\boldsymbol{|}\partial_{\theta_{j}}\boldsymbol{\phi}_{t}({\bm{\theta}},\boldsymbol{\phi})\boldsymbol{|}^{2}\;=\;\boldsymbol{|}\partial_{\theta_{j}}{\boldsymbol{\psi}}_{-t}({\bm{\theta}})\boldsymbol{|}^{2}\;. (54)

This identity has the important consequence, that neither the left-hand side nor the right-hand side depends on the choice of ϕ\boldsymbol{\phi} and of the initial constant spinor 𝝍\boldsymbol{\psi}.

V.2 Derivation of Eq. (42)

Combining Eq. (54) with Eq. (II.3), we obtain:

Ej(𝝍t)\displaystyle E_{j}({\boldsymbol{\psi}}_{-t}) =\displaystyle= 𝕋2𝑑m(𝜽)|θjϕt(𝜽,ϕ)|2\displaystyle\int_{\mathbb{T}^{2}}dm({\bm{\theta}})\;\boldsymbol{|}\partial_{\theta_{j}}\boldsymbol{\phi}_{t}({\bm{\theta}},\boldsymbol{\phi})\boldsymbol{|}^{2} (55)
=\displaystyle= Ω𝑑μ(𝜽,ϕ)|θjϕt(𝜽,ϕ)|2,\displaystyle\int_{\Omega}d\mu({\bm{\theta}},\boldsymbol{\phi})\;\boldsymbol{|}\partial_{\theta_{j}}\boldsymbol{\phi}_{t}({\bm{\theta}},\boldsymbol{\phi})\boldsymbol{|}^{2}\,,

where the second line is obtained by performing an additional integral wrt ϕ\boldsymbol{\phi}, which is legitimate because, as noted above, the integrand is independent of ϕ\boldsymbol{\phi}. With the substitution of Eq. (48), Eq. (55) reduces to

Ej(𝝍t)=s=0t1𝓤α,𝝎sHj2+s=0t1𝓤α,𝝎sKj2\displaystyle E_{j}({\boldsymbol{\psi}}_{-t})=\bigl{\|}\,\sum_{s=0}^{t-1}{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}^{-s}H_{j}\,\bigr{\|}_{\mathfrak{H}}^{2}\,+\,\bigl{\|}\,\sum_{s=0}^{t-1}{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}^{-s}K_{j}\,\bigr{\|}^{2}_{\mathfrak{H}}\,\,\, (56)

or equivalently, Eq. (42). Equation (55) establishes the announced juxtaposition between quantum transport and instability of classical trajectories. Moreover, from Eq. (56) or (42) it follows that, for the SM model, the quantum transport in momentum space is never faster than ballistic. It is worth mentioning that if the equivalent classical dynamical systems of SM models were known to be ergodic then Eq. (56) or (42) would entail sub-ballistic transport, because the phase space averages of both HjH_{j} and KjK_{j} vanish, and the same would be true for their time averages.

VI Proofs of some mathematical results

Here we present proofs of some mathematical results presented in Sec. IV. The readers who are not interested in mathematical details can skip this section.

VI.1 Proof of Proposition 1

Here an extended version of Proposition 1 will be proven. It includes the statement, that if the spectrum is continuous, then it is either purely absolutely continuous, or purely singular continuous.

(i) Let 𝒖(𝜽)1:={𝝍:𝝍(𝜽)12,𝜽}\boldsymbol{u}({\bm{\theta}})\in{\cal H}_{1}:=\{\boldsymbol{\psi}\in{\cal H}:\boldsymbol{\psi}({\bm{\theta}})\in{\mathbb{C}}^{2}_{1}\;,\forall{\bm{\theta}}\}, and for any 𝒏2{\boldsymbol{n}}\in{\mathbb{Z}}^{2} define vectors 𝒖+,𝒏:=ei𝒏𝜽𝒖{\boldsymbol{u}}_{+,\boldsymbol{n}}:=e^{i{\boldsymbol{n}}\boldsymbol{\cdot}{{\bm{\theta}}}}{\boldsymbol{u}} and 𝒖,𝒏:=𝝆𝒖+,𝒏=ei𝒏𝜽𝝆𝒖{\boldsymbol{u}}_{-,\boldsymbol{n}}:={\bm{\rho}}{\boldsymbol{u}}_{+,\boldsymbol{n}}=e^{-i{\boldsymbol{n}}\boldsymbol{\cdot}{{\bm{\theta}}}}{\bm{\rho}}\,{\boldsymbol{u}}. Then the vectors {𝒖±,𝒏}𝒏2\{{\boldsymbol{u}}_{\pm,\boldsymbol{n}}\}_{\boldsymbol{n}\in\mathbb{Z}^{2}} are a total set of vectors in \cal H. To see this, let Ψ\Psi\in{\cal H} be orthogonal to all vectors 𝒖±,𝒏{\boldsymbol{u}}_{\pm,\boldsymbol{n}}. In particular,

𝕋2𝑑m(𝜽)ei𝒏𝜽Ψ(𝜽)𝒖(𝜽)= 0,𝒏2.\int_{{\mathbb{T}}^{2}}dm({\bm{\theta}})\;e^{i{\boldsymbol{n}}\boldsymbol{\cdot}{{\bm{\theta}}}}\;\Psi({\bm{\theta}})\boldsymbol{\cdot}{\boldsymbol{u}}({\bm{\theta}})\;=\;0,\,\;\forall{\boldsymbol{n}}\in{\mathbb{Z}}^{2}\;. (57)

As 𝒖(𝜽)\boldsymbol{u}({\bm{\theta}}) never vanishes, completeness of the Fourier basis entails that Ψ(𝜽)\Psi({\bm{\theta}}) is almost everywhere in 𝕋2{\mathbb{T}}^{2} orthogonal to 𝒖(𝜽)\boldsymbol{u}({\bm{\theta}}). Hence Ψ\Psi is a multiple of 𝝆𝒖{\bm{\rho}}\,\boldsymbol{u}, because the latter vector, together with 𝒖\boldsymbol{u}, makes a basis for 2{\mathbb{C}}^{2}. So Ψ=c(𝜽)𝝆𝒖\Psi=c({\bm{\theta}}){\bm{\rho}}\,\boldsymbol{u} where c(𝜽)c({\bm{\theta}}) is some measurable function. By the same argument, one has

𝕋2𝑑m(𝜽)ei𝒏𝜽Ψ(𝜽)𝝆𝒖(𝜽)= 0,𝒏2,\int_{{\mathbb{T}}^{2}}dm({\bm{\theta}})\;e^{-i{\boldsymbol{n}}\boldsymbol{\cdot}{{\bm{\theta}}}}\;\Psi({\bm{\theta}})\boldsymbol{\cdot}{\bm{\rho}}{\boldsymbol{u}}({\bm{\theta}})\;=\;0,\,\;\forall{\boldsymbol{n}}\in{\mathbb{Z}}^{2}\;, (58)

from which it follows that Ψ\Psi must also be a scalar multiple of 𝒖\boldsymbol{u}: this is possible if and only if c(𝜽)0c({\bm{\theta}})\equiv 0. Thus Eqs. (57) and (58) can be satisfied simultaneously, if and only if Ψ(𝜽)0\Psi({\bm{\theta}})\equiv 0 .

Next, for any tt\in{\mathbb{Z}}, consider the correlations:

±,𝒏(t)=𝒖±,𝒏|𝑼α,𝝎t𝒖±,𝒏.{\cal R}_{\pm,\boldsymbol{n}}(t)\;=\;\langle{\boldsymbol{u}}_{\pm,\boldsymbol{n}}\;|\;{\bm{U}}_{\alpha,{\bm{\omega}}}^{t}{\boldsymbol{u}}_{\pm,\boldsymbol{n}}\rangle_{\mathcal{H}}\;. (59)

It is a basic notion in spectral theory Simon1 that such correlations are Fourier transforms of the spectral measures of the vectors 𝒖±,𝒏{\boldsymbol{u}}_{\pm,\boldsymbol{n}} (also known as LDOS). From the definition of 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} and from Eq. (5), it follows that

+,𝒏(t)=,𝒏(t)=ei𝒏𝝎t+,𝟎(t).{\cal R}_{+,\boldsymbol{n}}(t)\;=\;{\cal R}_{-,\boldsymbol{n}}^{*}(t)\;=\;e^{-i{\boldsymbol{n}}\boldsymbol{\cdot}{\boldsymbol{\omega}}t}{\cal R}_{+,\boldsymbol{0}}(t). (60)

This shows that the spectral measures of all vectors 𝒖±,𝒏{\boldsymbol{u}}_{\pm,\boldsymbol{n}} can be obtained from the spectral measure of any single one of them, say 𝒖+,𝟎{\boldsymbol{u}}_{+,\boldsymbol{0}}, by irrational rotations and conjugation in the unit circle. As these two operations cannot change the nature of the spectral measure (pure point, absolutely or singular continuous), and {𝒖±,𝒏}𝒏2\{{\boldsymbol{u}}_{\pm,\boldsymbol{n}}\}_{\boldsymbol{n}\in\mathbb{Z}^{2}} is a total set, purity of the spectrum follows. Furthermore, since the set {ei𝒏𝝎}𝒏2\{e^{i{\boldsymbol{n}}\boldsymbol{\cdot}{\boldsymbol{\omega}}}\}_{\boldsymbol{n}\in\mathbb{Z}^{2}} is dense in 1\mathbb{C}_{1}, the closure of the spectrum is always the full unit circle.

(ii) If the spectrum is pure point, then let 𝒖(𝜽)\boldsymbol{u}({\bm{\theta}}) be an eigenvector, with eigenvalue eiλe^{i\lambda}. Its 2{\mathbb{C}}^{2} norm is constant, as it can be seen on taking norms of both sides of 𝑴α𝒖(𝜽𝝎)=eiλ𝒖(𝜽)\bm{M}_{\alpha}\boldsymbol{u}({\bm{\theta}}-\boldsymbol{\omega})=e^{i\lambda}\boldsymbol{u}({\bm{\theta}}) and then using ergodicity of the 𝝎\boldsymbol{\omega}-shift 𝝉𝝎\bm{\tau_{\bm{\omega}}}. From the definition of 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}}, it immediately follows that 𝒖±,𝒏\boldsymbol{u}_{\pm,\boldsymbol{n}} is an eigenvector, with eigenvalue e±iλi𝒏𝝎e^{\pm i\lambda\mp i\boldsymbol{n}\boldsymbol{\cdot}\boldsymbol{\omega}}. Since {𝒖±,𝒏}𝒏2\{{\boldsymbol{u}}_{\pm,\boldsymbol{n}}\}_{\boldsymbol{n}\in\mathbb{Z}^{2}} is a total set, Eq. (28) follows. \Box

VI.2 Proof of Proposition 2

Here an extended version of the Proposition is proven. It includes a proof that the continuous component of the spectrum of 𝓤α,𝝎{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}} shares the type (s.c. or a.c.) of the spectrum of 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}}. The functions ei𝒏𝜽e^{i\boldsymbol{n}\boldsymbol{\cdot}\boldsymbol{\theta}}, (𝒏2{\boldsymbol{n}}\in{\mathbb{Z}}^{2}) are a complete basis of 0\mathfrak{H}_{0}, so claim (i) follows from 𝓤α,𝝎ei𝒏𝜽=ei𝒏𝝎ei𝒏𝜽{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}e^{i\boldsymbol{n}\boldsymbol{\cdot}\boldsymbol{\theta}}=e^{-i\boldsymbol{n}\boldsymbol{\cdot}\boldsymbol{\omega}}e^{i\boldsymbol{n}\boldsymbol{\cdot}\boldsymbol{\theta}}.

Next, let p,qp,q be nonnegative integers, j,k=±1j,k=\pm 1, and 𝝍(𝜽){\boldsymbol{\psi}}(\boldsymbol{\theta})\in\cal H. The functions

F𝒏,j,k,p,q,𝝍(𝜽,ϕ)=ei𝒏𝜽[𝝍(j)(𝜽)ϕ]p[ϕ𝝍(k)(𝜽)]q\displaystyle F_{\boldsymbol{n},j,k,p,q,{\boldsymbol{\psi}}}(\boldsymbol{\theta},{\boldsymbol{\phi}})\;=\;e^{i{\boldsymbol{n}}\boldsymbol{\cdot}\boldsymbol{\theta}}\;[{\boldsymbol{\psi}}^{(j)}(\boldsymbol{\theta})\boldsymbol{\cdot}{\boldsymbol{\phi}}]^{p}[{\boldsymbol{\phi}}\boldsymbol{\cdot}{\boldsymbol{\psi}}^{(k)}(\boldsymbol{\theta})]^{q} (61)

with 𝝍(1)=𝝍{\boldsymbol{\psi}}^{(-1)}={\boldsymbol{\psi}}, 𝝍(+1)=𝝆𝝍{\boldsymbol{\psi}}^{(+1)}=\boldsymbol{\rho}\,{\boldsymbol{\psi}}, p+q0p+q\neq 0, are a total set in 0{\mathfrak{H}}_{0}^{\bot} whenever 𝝍(𝜽){\boldsymbol{\psi}}(\boldsymbol{\theta}) is almost everywhere nonzero; this directly follows, e.g. from the complex Stone-Weierstrass theorem Simon1 . Next, thanks to Eq. (15) we note that

𝓤α,𝝎(F𝒏,j,k,p,q,𝝍)=ei𝒏𝝎F𝒏,j,k,p,q,𝝍,\displaystyle{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}}(F_{\boldsymbol{n},j,k,p,q,{\boldsymbol{\psi}}})\;=\;e^{-i\boldsymbol{n}\boldsymbol{\cdot}\boldsymbol{\omega}}F_{\boldsymbol{n},j,k,p,q,\,{\boldsymbol{\psi}}^{\prime}}\,, (62)

where 𝝍=𝑼α,𝝎𝝍{\boldsymbol{\psi}}^{\prime}={\bm{U}}_{\alpha,{\bm{\omega}}}{\boldsymbol{\psi}}.

On one branch of the dichotomy, the spectrum of 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} is pure point. Let 𝝍{\boldsymbol{\psi}} be the eigenfunction of 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} with eigenvalue eiλe^{i\lambda}. From Eq. (62) it then follows that F𝒏,j,k,p,q,𝝍F_{\boldsymbol{n},j,k,p,q,{\boldsymbol{\psi}}} is the eigenfunction of 𝓤α,𝝎{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}} with eigenvalue ei[λ(jpkq)𝒏𝝎]e^{i[\lambda(jp-kq)-\boldsymbol{n}\boldsymbol{\cdot}\boldsymbol{\omega}]}. This holds true for all 𝒏,j,k,p,q\boldsymbol{n},j,k,p,q; moreover, eigenfunctions of 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} are almost everywhere nonzero (owing to ergodicity of 𝝎{\bm{\omega}}-shift in 𝕋2{\mathbb{T}}^{2}), so the pure point spectral subspace of 𝓤α,𝝎{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}} restricted on 0\mathfrak{H}^{\perp}_{0} includes a total set. Thus the spectrum of 𝓤α,𝝎{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}} in 0\mathfrak{H}^{\perp}_{0} is pure point also. In combination with (i) this leads to (ii).

On the other branch of the dichotomy, the spectrum of 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} is purely continuous. Denote

G𝝍(𝜽,ϕ)=𝝍(𝜽)ϕ,G_{{\boldsymbol{\psi}}}({\bm{\theta}},{\boldsymbol{\phi}})={\boldsymbol{\psi}}({\bm{\theta}})\boldsymbol{\cdot}{\boldsymbol{\phi}}\,, (63)

which is the function (61) with 𝒏=0,j=1,p=1,q=0\boldsymbol{n}=0,j=-1,p=1,q=0. Then for any tt\in\mathbb{Z}:

G𝝍|𝓤α,𝝎tG𝝍\displaystyle\langle G_{{\boldsymbol{\psi}}}\,|\,{\bm{{\cal U}}}^{t}_{\alpha,{\bm{\omega}}}G_{{\boldsymbol{\psi}}}\,\rangle_{\mathfrak{H}} =\displaystyle= Ω𝑑μ(𝜽,ϕ)[𝝍(𝜽)ϕ][ϕ𝑼α,𝝎t𝝍(𝜽)]\displaystyle\int_{\Omega}d\mu({\bm{\theta}},{\boldsymbol{\phi}})\;[{\boldsymbol{\psi}}({\bm{\theta}})\boldsymbol{\cdot}{\boldsymbol{\phi}}][{\boldsymbol{\phi}}\boldsymbol{\cdot}{\bm{U}}^{t}_{\alpha,{\bm{\omega}}}{\boldsymbol{\psi}}({\bm{\theta}})] (64)
=\displaystyle= 12𝕋2𝑑m(𝜽)𝝍(𝜽)𝑼α,𝝎t𝝍(𝜽)\displaystyle\tfrac{1}{2}\int_{\mathbb{T}^{2}}dm({\bm{\theta}})\;{\boldsymbol{\psi}}({\bm{\theta}})\boldsymbol{\cdot}{\bm{U}}^{t}_{\alpha,{\bm{\omega}}}{\boldsymbol{\psi}}({\bm{\theta}})
=\displaystyle= 12𝝍|𝑼α,𝝎t𝝍.\displaystyle\tfrac{1}{2}\langle{\boldsymbol{\psi}}\,|\,{\bm{U}}^{t}_{\alpha,{\bm{\omega}}}{\boldsymbol{\psi}}\rangle_{\mathcal{H}}\,.

Therefore, the spectral measure of GG wrt 𝓤α,𝝎{\bm{{\cal U}}}_{\alpha,{\bm{\omega}}} is – apart from normalization – the same as the spectral measure of 𝝍{\boldsymbol{\psi}} wrt 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}}. So (iii) and its extended version, claimed at the beginning of this subsection, follow. \Box

VI.3 Proof of Proposition 3

It will be shown that if 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} has a proper eigenfunction 𝒖\boldsymbol{u}, and (or) 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} commutes with a nontrivial (linear) fibered operator 𝓑\boldsymbol{{\cal B}}, then (Ω,𝓢α,𝝎)(\Omega,{\bm{{\cal S}}}_{\alpha,{\bm{\omega}}}) is not ergodic.

By the very definition of ergodicity, (Ω,𝓢α,𝝎)(\Omega,{\bm{{\cal S}}}_{\alpha,{\bm{\omega}}}) cannot be ergodic, if invariant functions exist, which are not almost everywhere constant. If 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} has an eigenfunction 𝒖\boldsymbol{u}, then F(𝜽,𝒖)=|𝒖(𝜽)ϕ|F({{\bm{\theta}}},\boldsymbol{u})=|\boldsymbol{u}({\bm{\theta}})\boldsymbol{\cdot}\boldsymbol{\phi}| is a nonconstant invariant function on Ω\Omega. On the other hand, if 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} commutes with a fibered operator 𝓑\boldsymbol{{\cal B}} then the following “covariance relation” holds:

𝓑(𝜽)=𝑴α(𝜽)𝓑(𝜽𝝎)𝑴α(𝜽)1,𝜽𝕋2.\boldsymbol{{\cal B}}({\bm{\theta}})=\boldsymbol{M}_{\alpha}({\bm{\theta}})\,\boldsymbol{{\cal B}}({\bm{\theta}}-{\bm{\omega}})\,\boldsymbol{M}_{\alpha}({\bm{\theta}})^{-1}\;,\,\,\,\forall{\bm{\theta}}\in\mathbb{T}^{2}\;. (65)

Then a simple calculation shows that the function that is defined on Ω\Omega by F(𝜽,ϕ)=𝓑(𝜽)ϕϕF({\bm{\theta}},\boldsymbol{\phi})=\boldsymbol{{\cal B}}({\bm{\theta}})\boldsymbol{\phi}\boldsymbol{\cdot}\boldsymbol{\phi} is not constant (because 𝓑\boldsymbol{{\cal B}} is nontrivial), yet it is invariant under the map 𝓢α,𝝎{\bm{{\cal S}}}_{\alpha,{\bm{\omega}}}. \Box

VI.4 Proof of Corollary 1

The proof consists of two parts.

(i) Let 𝒖\boldsymbol{u} be an eigenvector of 𝑼α,𝝎{\bm{U}}_{{\alpha,\boldsymbol{\omega}}}, and eiχe^{i\chi} the corresponding eigenvalue. Thanks to the first identity in Eq. (30), 𝑨𝒖\bm{A}\boldsymbol{u} is an eigenvector of 𝑼πα,𝝎{\bm{U}}_{{\pi-\alpha,\boldsymbol{\omega}}}, with eigenvalue eiχ+iπe^{-i\chi+i\pi}. Since (ω1,ω2,π)(\omega_{1},\omega_{2},\pi) is an incommensurate triple, it easily follows from Eq. (28) that, if Eig(𝑼π2,𝝎){\rm Eig}({\bm{U}}_{\frac{\pi}{2},\boldsymbol{\omega}}) is not empty, then there are some 𝒑,𝒒2\bm{p},\bm{q}\in\mathbb{Z}^{2} so that χ=λ+𝒑𝝎\chi=\lambda+\bm{p}\boldsymbol{\cdot}{\bm{\omega}} and χ+π=λ+𝒒𝝎-\chi+\pi=\lambda+\bm{q}\boldsymbol{\cdot}{\bm{\omega}}. Adding these two equalities together gives π=2λ+(𝒑+𝒒)𝝎\pi=2\lambda+(\bm{p}+\bm{q})\boldsymbol{\cdot}{\bm{\omega}}. Thus there is some 𝒓2\boldsymbol{r}\in{\mathbb{Z}}^{2} so that λ\lambda in Eq. (28) can be expressed as λ=π/2+𝒓𝝎/2\lambda=\pi/2+\boldsymbol{r}\boldsymbol{\cdot}\boldsymbol{\omega}/2. Thus the spectrum is as declared in Eq. (37) and it is not degenerate, see remark (i) following Proposition 1.

(ii) Let 𝒗𝒏±\boldsymbol{v}^{\pm}_{\boldsymbol{n}} be an eigenvector of 𝑼π2,𝝎{\bm{U}}_{{\frac{\pi}{2},\boldsymbol{\omega}}} that corresponds to the eigenvalue ±iei(𝒏+𝒓/2)𝝎\pm ie^{i(\boldsymbol{n}+\boldsymbol{r}/2)\boldsymbol{\cdot}\boldsymbol{\omega}}; then thanks to the first identity in Eq. (30), 𝑨𝒗𝒏±\bm{A}\boldsymbol{v}^{\pm}_{\bm{n}} is an eigenvector that corresponds to the eigenvalue ±iei(𝒏+𝒓/2)𝝎\pm ie^{-i(\boldsymbol{n}+\boldsymbol{r}/2)\boldsymbol{\cdot}\boldsymbol{\omega}}. On the other hand, by the following relation:

𝑼π2,𝝎𝒗𝒏±(𝜽)ei(2𝒏+𝒓)𝜽=𝑴π2𝒗𝒏±(𝜽𝝎)ei(2𝒏+𝒓)(𝜽𝝎)\displaystyle{\bm{U}}_{\frac{\pi}{2},{\bm{\omega}}}\boldsymbol{v}_{\boldsymbol{n}}^{\pm}({\bm{\theta}})e^{i(2\boldsymbol{n}+\boldsymbol{r})\boldsymbol{\cdot}{\bm{\theta}}}=\bm{M}_{\frac{\pi}{2}}\boldsymbol{v}_{\boldsymbol{n}}^{\pm}({\bm{\theta}}-{\bm{\omega}})e^{i(2\boldsymbol{n}+\boldsymbol{r})\boldsymbol{\cdot}({\bm{\theta}}-{\bm{\omega}})} (66)
=\displaystyle= ±iei(𝒏+𝒓/2)𝝎𝒗𝒏±(𝜽)ei(2𝒏+𝒓)(𝜽𝝎)\displaystyle\pm ie^{i(\boldsymbol{n}+\boldsymbol{r}/2)\boldsymbol{\cdot}{\bm{\omega}}}\boldsymbol{v}_{\boldsymbol{n}}^{\pm}({\bm{\theta}})e^{i(2\boldsymbol{n}+\boldsymbol{r})\boldsymbol{\cdot}({\bm{\theta}}-{\bm{\omega}})}
=\displaystyle= ±iei(𝒏+𝒓/2)𝝎𝒗𝒏±(𝜽)ei(2𝒏+𝒓)𝜽\displaystyle\pm ie^{-i(\boldsymbol{n}+\boldsymbol{r}/2)\boldsymbol{\cdot}{\bm{\omega}}}\boldsymbol{v}_{\boldsymbol{n}}^{\pm}({\bm{\theta}})e^{i(2\boldsymbol{n}+\boldsymbol{r})\boldsymbol{\cdot}{\bm{\theta}}}

one finds that 𝒗𝒏±ei(2𝒏+𝒓)𝜽\boldsymbol{v}^{\pm}_{\boldsymbol{n}}e^{i(2\boldsymbol{n}+\boldsymbol{r})\boldsymbol{\cdot}{\bm{\theta}}} is an eigenvector that has the same eigenvalue as 𝑨𝒗𝒏±\bm{A}\boldsymbol{v}^{\pm}_{\bm{n}}. Since, as it was noted above, the spectrum is not degenerate, the two eigenvectors must coincide up to a constant phase factor cc. Equation (38) directly follows. The statement about the factor c=c𝒏±c=c^{\pm}_{\boldsymbol{n}} in Eq. (38) is obtained by operating with 𝑨\bm{A} on both sides of the first equation in (38), and thereafter using the same Eq. (38) in the thus obtained equation. \Box

VI.5 Proof of Corollary 2

The operator 𝑼π2,𝝎{\bm{U}}_{\frac{\pi}{2},{\bm{\omega}}} of SM Model III is denoted by 𝕌\mathbb{U}. We shall prove that the spectrum of 𝕌\mathbb{U} is continuous, first with d3=0d_{3}=0, and then with d3=cosθ1+cosθ2d_{3}=\cos\theta_{1}+\cos\theta_{2}.

(i) d3=0d_{3}=0. The corresponding operator 𝑴π2\boldsymbol{M}_{\frac{\pi}{2}} is

𝑴π2=(0ieiΦ(𝜽)ieiΦ(𝜽)0)=:𝕄,\displaystyle{\boldsymbol{M}}_{\frac{\pi}{2}}=\left(\begin{array}[]{cc}0&-ie^{-i\Phi(\bm{\theta})}\\ -ie^{i\Phi(\bm{\theta})}&0\\ \end{array}\right)=:{\boldsymbol{\mathbb{M}}}, (69)

where Φ(𝜽)\Phi(\bm{\theta}) is the argument of the complex number sin(θ1)+isin(θ2)\sin(\theta_{1})+i\sin(\theta_{2}). Continuity of the spectrum of 𝕌\mathbb{U} will be proven by contradiction. Assume the spectrum to be pure point. Due to Eq. (37), 𝕌\mathbb{U} has an eigenvalue iei𝒓𝝎/2ie^{i\boldsymbol{r}\boldsymbol{\cdot}\boldsymbol{\omega}/2} with a corresponding eigenfunction denoted by 𝒖=(u1,u2)T=𝒗𝟎+\boldsymbol{u}=(u_{1},u_{2})^{T}=\boldsymbol{v}^{+}_{\boldsymbol{0}} (TT denotes the transpose). The latter satisfies

𝜽𝕋2:ei𝒓𝝎/2𝒖(𝜽)=i𝕄(𝜽)𝒖(𝜽𝝎)\displaystyle\forall\,{\bm{\theta}}\in\mathbb{T}^{2}:e^{i\boldsymbol{r}\boldsymbol{\cdot}\boldsymbol{\omega}/2}\boldsymbol{u}({\bm{\theta}})=-i\;\boldsymbol{\mathbb{M}}({\bm{\theta}})\boldsymbol{u}({\bm{\theta}}-\boldsymbol{\omega}) (70)
{ei𝒓𝝎/2u1(𝜽)=i𝕄12(𝜽)u2(𝜽𝝎)ei𝒓𝝎/2u2(𝜽)=i𝕄21(𝜽)u1(𝜽𝝎).\displaystyle\Leftrightarrow\left\{\begin{array}[]{c}e^{i\boldsymbol{r}\boldsymbol{\cdot}\boldsymbol{\omega}/2}u_{1}({\bm{\theta}})=-i\;\boldsymbol{\mathbb{M}}_{12}({\bm{\theta}})u_{2}({\bm{\theta}}-\boldsymbol{\omega})\\ e^{i\boldsymbol{r}\boldsymbol{\cdot}\boldsymbol{\omega}/2}u_{2}({\bm{\theta}})=-i\;\boldsymbol{\mathbb{M}}_{21}({\bm{\theta}})u_{1}({\bm{\theta}}-\boldsymbol{\omega})\end{array}\right..\quad (73)

So from 𝕄12𝕄21=1\boldsymbol{\mathbb{M}}_{12}\boldsymbol{\mathbb{M}}_{21}=-1 it follows that

u1(𝜽)u2(𝜽)=ei𝒓𝝎u1(𝜽𝝎)u2(𝜽𝝎).u_{1}({\bm{\theta}})u_{2}({\bm{\theta}})=e^{-i\boldsymbol{r}\boldsymbol{\cdot}\boldsymbol{\omega}}u_{1}({\bm{\theta}}-\boldsymbol{\omega})u_{2}({\bm{\theta}}-\boldsymbol{\omega}). (74)

This entails

u1(𝜽)u2(𝜽)=c1ei𝒓𝜽,u_{1}({\bm{\theta}})u_{2}({\bm{\theta}})=c_{1}e^{-i\boldsymbol{r}\boldsymbol{\cdot}{\bm{\theta}}}, (75)

where c10c_{1}\neq 0 is a constant.

On the other hand, from Eq. (38) it follows that

u2(𝝎𝜽)=ei𝒓𝝎/2ei𝒓𝜽u2(𝜽).u_{2}(-{\bm{\omega}}-{\bm{\theta}})=-e^{i\boldsymbol{r}\boldsymbol{\cdot}\boldsymbol{\omega}/2}e^{i\boldsymbol{r}\boldsymbol{\cdot}{\bm{\theta}}}u_{2}({\bm{\theta}}). (76)

Combining this with the first equation in (73) gives

u2(𝜽)=ei𝒓𝜽𝕄21(𝜽)u1(𝜽).u_{2}({\bm{\theta}})=-e^{-i\boldsymbol{r}\boldsymbol{\cdot}{\bm{\theta}}}\boldsymbol{\mathbb{M}}_{21}({\bm{\theta}})u_{1}(-{\bm{\theta}}). (77)

With this substitution Eq. (75) reduces to

u1(𝜽)u1(𝜽)=ic1𝕄12(𝜽).u_{1}({\bm{\theta}})u_{1}(-{\bm{\theta}})\;=\;-ic_{1}\;{\boldsymbol{\mathbb{M}}}_{12}({\bm{\theta}})\;. (78)

This is impossible, because the left-hand side is even wrt 𝜽{\bm{\theta}}, while the right-hand side is odd for 𝕄12(𝜽)=𝕄12(𝜽)\boldsymbol{\mathbb{M}}_{12}(-{\bm{\theta}})=-\boldsymbol{\mathbb{M}}_{12}({\bm{\theta}}).

(ii) d3=cos(θ1)+cos(θ2)d_{3}=\cos(\theta_{1})+\cos(\theta_{2}). We generalize the proof of case (i). Assume the spectrum to be pure point. Then one can easily check that for 𝜽{\bm{\theta}} in the contour 𝒞{\cal C} defined as {𝜽𝕋2:cos(θ1)+cos(θ2)=0}\{{\bm{\theta}}\in\mathbb{T}^{2}:\cos(\theta_{1})+\cos(\theta_{2})=0\}, Eq. (78) still follows (the constant c1c_{1} is in general different). On the other hand, because the contour 𝒞{\cal C} has the inversion symmetry, if 𝜽𝒞{\bm{\theta}}\in{\cal C} then 𝜽𝒞-{\bm{\theta}}\in{\cal C}. So, by the same arguments as above, Eq. (78) cannot be true for 𝜽𝒞{\bm{\theta}}\in{\cal C}. This contradiction proves the continuity of the spectrum of 𝕌\mathbb{U}. \Box

VI.6 Proof of Corollary 3

It is sufficient to consider the eigenfunction 𝒗𝟎+\boldsymbol{v}^{+}_{\bm{0}}, which will be denoted by 𝒗\boldsymbol{v} to make all formulae compact. From Eq. (37) it follows that

𝑴π2(𝜽)𝒗(𝜽𝝎)=iei𝒓𝝎/2𝒗(𝜽).\boldsymbol{M}_{\frac{\pi}{2}}({\bm{\theta}})\,\boldsymbol{v}({\bm{\theta}}-\boldsymbol{\omega})\;=\;ie^{i\boldsymbol{r}\boldsymbol{\cdot}\boldsymbol{\omega}/2}\;\boldsymbol{v}({\bm{\theta}}). (79)

Moreover, applying Eqs. (35) and (38) gives

𝒗(𝜽𝝎)=c𝟎+ei𝒓𝜽𝝈3𝒗(𝜽).\boldsymbol{v}({\bm{\theta}}-\boldsymbol{\omega})\;=\;c^{+}_{\bm{0}}\;e^{-i\boldsymbol{r}\boldsymbol{\cdot}{\bm{\theta}}}\;\boldsymbol{\sigma}_{3}\;\boldsymbol{v}(-{\bm{\theta}})\;. (80)

With the substitution of Eq. (80), Eq. (79) reduces to

i𝑴π2(𝜽)𝝈3𝒗(𝜽)=(c𝟎+)1ei𝒓𝝎/2ei𝒓𝜽𝒗(𝜽).-i\boldsymbol{M}_{\frac{\pi}{2}}({\bm{\theta}})\boldsymbol{\sigma}_{3}\boldsymbol{v}(-{\bm{\theta}})\;=\;(c^{+}_{\bm{0}})^{-1}e^{i\boldsymbol{r}\boldsymbol{\cdot}\boldsymbol{\omega}/2}e^{i\boldsymbol{r}\boldsymbol{\cdot}{\bm{\theta}}}\boldsymbol{v}({\bm{\theta}})\;. (81)

Let this equation be written at four points: ϑ0(0,0)\boldsymbol{\vartheta}_{0}\equiv(0,0), ϑ1(0,π)\boldsymbol{\vartheta}_{1}\equiv(0,\pi), ϑ3(π,0)\boldsymbol{\vartheta}_{3}\equiv(\pi,0), and ϑ3(π,π)\boldsymbol{\vartheta}_{3}\equiv(\pi,\pi). At each such point, Eq. (18) yields that i𝑴π2(ϑj)𝝈3=d3(ϑj)𝕀i\boldsymbol{M}_{\frac{\pi}{2}}(\boldsymbol{\vartheta}_{j})\boldsymbol{\sigma}_{3}=d_{3}(\boldsymbol{\vartheta}_{j})\mathbb{I}. In addition,

d3(ϑ0)=1,d3(ϑj)= 1,(j=1,2,3).d_{3}(\boldsymbol{\vartheta}_{0})\;=\;-1\;,\;\;\;d_{3}(\boldsymbol{\vartheta}_{j})\;=\;1\;,\;\;(j=1,2,3)\;. (82)

So, if 𝒗\boldsymbol{v} is continuous at ϑ0\boldsymbol{\vartheta}_{0}, then c𝟎+𝒗(ϑ0)=ei𝒓𝝎/2𝒗(ϑ0)c^{+}_{\bm{0}}\,\boldsymbol{v}(\boldsymbol{\vartheta}_{0})\;=\;e^{i\boldsymbol{r}\boldsymbol{\cdot}\boldsymbol{\omega}/2}\boldsymbol{v}(\boldsymbol{\vartheta}_{0}). This forces the choice c𝟎+=ei𝒓𝝎/2c^{+}_{\bm{0}}=e^{i\boldsymbol{r}\boldsymbol{\cdot}\boldsymbol{\omega}/2} [note that from Eq. (38) c𝟎+=±ei𝒓𝝎/2c^{+}_{\bm{0}}=\pm e^{i\boldsymbol{r}\boldsymbol{\cdot}\boldsymbol{\omega}/2}], because 𝒗\boldsymbol{v} does not vanish at any point in 𝕋2\mathbb{T}^{2}. With this choice and the substitution of Eq. (82), Eq. (81) entails:

𝒗(ϑj)=ei𝒓ϑj𝒗(ϑj),j=1,2,3,\boldsymbol{v}(\boldsymbol{\vartheta}_{j})\,=\,-e^{i\boldsymbol{r}\boldsymbol{\cdot}\boldsymbol{\vartheta}_{j}}\,\boldsymbol{v}(\boldsymbol{\vartheta}_{j}),\;\;\;j=1,2,3, (83)

provided that 𝒗\boldsymbol{v} is continuous at such points. If true for j=1,2j=1,2 , this equality requires both r1r_{1} and r2r_{2} [𝒓(r1,r2)\boldsymbol{r}\equiv(r_{1},r_{2})] to be odd integers; but then it cannot be satisfied for j=3j=3. This contradiction implies that the eigenfunction has a non-removable discontinuity at one at least of the four points ϑj\boldsymbol{\vartheta}_{j} (j=0,1,2,3j=0,1,2,3). Using ergodicity of 𝝉𝝎\boldsymbol{\tau}_{\boldsymbol{\omega}} it is not difficult to prove that this forces the eigenfunction to be actually discontinuous at all points. This will not be shown here footnote_sing . Anyway, non-removable discontinuities cause delocalization of the eigenfunction, as they forbid finite expectation values for the momentum operators, which involve derivatives footnote_so .

VI.7 Proof of Proposition 5

By assumption, no eigenfunction belongs in the domain of the momentum operators, that is defined as {𝝍:Ej(𝝍)<+}\{{\boldsymbol{\psi}}\in{\cal H}:E_{j}({\boldsymbol{\psi}})<+\infty\} where j=1or 2j=1\,{\rm or}\,2 [recall Eq. (II.3) for the definition of Ej(𝝍)E_{j}({\boldsymbol{\psi}})]. Let the initial state 𝝍{\boldsymbol{\psi}} be in this domain and 𝝍t=𝑼t𝝍{\boldsymbol{\psi}}_{t}={\bm{U}}^{t}\,{\boldsymbol{\psi}}. For any a>0a>0 we construct a subset of {\cal H}, defined as Ka:={𝝍:Ej(𝝍)a}K_{a}:=\{{\boldsymbol{\psi}}\in{\cal H}:E_{j}({\boldsymbol{\psi}})\leq a\}. This subset is compact. Assume – in contradiction to Eq. (39) – that dynamical localization follows. Then aa can be chosen so that 𝝍tKa{\boldsymbol{\psi}}_{t}\in K_{a}, t\forall t\in\mathbb{Z}. Let eiχe^{i\chi} be any eigenvalue of 𝑼{\bm{U}}; obviously, eiχt𝝍tKae^{-i\chi t}{\boldsymbol{\psi}}_{t}\in K_{a}. Now Ej12(𝝍)E_{j}^{\frac{1}{2}}({\boldsymbol{\psi}}) is subaddittive, so we have:

Ej12(1ts=0t1eiχt𝝍t)1ts=0t1Ej12(eiχt𝝍t).\displaystyle E_{j}^{\frac{1}{2}}(\tfrac{1}{t}\sum_{s=0}^{t-1}e^{-i\chi t}{\boldsymbol{\psi}}_{t})\leq\tfrac{1}{t}\sum_{s=0}^{t-1}E_{j}^{\tfrac{1}{2}}(e^{-i\chi t}{\boldsymbol{\psi}}_{t}). (84)

Thus Ej12(1ts=0t1eiχt𝝍t)aE_{j}^{\frac{1}{2}}(\frac{1}{t}\sum_{s=0}^{t-1}e^{-i\chi t}{\boldsymbol{\psi}}_{t})\leq a, and so

1ts=0t1eiχt𝝍tKa.\tfrac{1}{t}\sum_{s=0}^{t-1}e^{-i\chi t}{\boldsymbol{\psi}}_{t}\in K_{a}. (85)

Thanks to von Neumann’s mean ergodic theorem Simon1 , in the limit t+t\to+\infty the left-hand side of the above equation tends in {\cal H} to the projection of the initial state 𝝍{\boldsymbol{\psi}} on the eiχe^{-i\chi} eigenspace. So, since the spectrum is nondegenerate, the limit of Eq. (85) is a scalar multiple of the eigenfunction. On the other hand, the limit has to be in KaK_{a} due to compactness; and this is a contradiction, because no eigenfunction of 𝑼{\bm{U}} can be in KaK_{a}. \Box

VII Conclusion

In this paper, we have shown that, when the 22D Maryland model is endowed with spin 1/2, rich dynamical localization/delocalization phenomena arise. In particular, in the family of SM Model I, dynamical localization–delocalization transitions are found to appear at half-integer e1\hbar_{e}^{-1}. Moreover, as a spinful QKR exhibiting e\hbar_{e}-driven IQHE is deformed continuously into the considered SM model, its topological phases are deformed continuously into the localized phases of the latter, although we have not been able to identify the topological invariant for the latter systems. Because dynamical properties of SM models and spinful QKR are very different and so the topological theory Tian14 ; Tian16 developed for spinful QKR does not apply here, the present findings imply that the striking similarity between IQHE on the one side and spinful QKR on the other has a deeper origin, and may be carried to a broader class of dynamical spin systems.

We have uncovered a self-duality underlying the observed dynamical transition: the dynamically localized phases on both sides of a critical point are dual in accordance with Proposition 4, while the critical delocalized phase is self-dual and as such bears an emergent unitary symmetry. This scenario of dynamical transition is conceptually different from Anderson-like dynamical transition observed in high-dimensional spinless QKR systems Casati89 ; Garreau08 ; Tian11 ; it goes far beyond the Landau-Ginzburg paradigm of phase transition which finds its origin in symmetry breaking. It remains a prominent issue to explore the applications of this new transition scenario in other quantum dynamical systems and its relations to topological transitions in spinful QKR.

Finally, we note that the SM model has an equivalent classical dynamical system for any e\hbar_{e} values. This classical system belongs to a special class of skew product on 𝕋D×G\mathbb{T}^{D}\times G with GG a compact Lie group in general, which are currently investigated by mathematicians: SM models correspond to D=2D=2 and G=SU(2)G=SU(2). In the language of skew-product systems, the transition in quantum dynamics is translated into the transition in the stability of classical trajectories in phase space. In-depth investigations of this aspect may allow one to view topological quantum phenomena such as IQHE from the perspectives of skew products on 𝕋D×G\mathbb{T}^{D}\times G and vice versa.

Acknowledgements

In memory of Shmuel Fishman, who first proposed investigation of SM models. Project Nos. 11535011, 11925507 and 11947302 supported by NSFC.

Appendix A Spectral theory and cohomology

Two maps 𝑴,𝑴:𝕋2SU(2)\bm{M},\bm{M}^{\prime}:{\mathbb{T}}^{2}\to SU(2) can be said to be 𝝎{\bm{\omega}}-cohomologous aar , if there is a map 𝑽:𝕋2SU(2)\bm{V}:{\mathbb{T}}^{2}\to SU(2) so that the classical dynamical systems on Ω\Omega corresponding to 𝑴,𝑴\bm{M},\bm{M}^{\prime} are isomorphic under the fiber-preserving isomorphism (𝜽,ϕ)(𝜽,𝑽(𝜽)ϕ)({\bm{\theta}},\bm{\phi})\mapsto({\bm{\theta}},\bm{V}({{\bm{\theta}}})\bm{\phi}). Equivalently, if, 𝜽\forall{{\bm{\theta}}}:

𝑴(𝜽)=𝑽(𝜽)𝑴(𝜽)𝑽1(𝜽𝝎).\bm{M}({\bm{\theta}})=\bm{V}({\bm{\theta}})\bm{M}^{\prime}({\bm{\theta}})\bm{V}^{-1}({\bm{\theta}}-{\bm{\omega}}). (86)

In other words, 𝑴\bm{M}, 𝑴\bm{M}^{\prime} are 𝝎{\bm{\omega}}-cohomologous if, and only if, the corresponding operators 𝑼M,𝝎{\bm{U}}_{M,{\bm{\omega}}}, 𝑼M,𝝎{\bm{U}}_{M^{\prime},{\bm{\omega}}} are unitarily equivalent: 𝑼M,𝝎=𝑽𝑼M,𝝎𝑽1{\bm{U}}_{M,{\bm{\omega}}}=\bm{V}{\bm{U}}_{M^{\prime},{\bm{\omega}}}\bm{V}^{-1}, where 𝑽\bm{V} is a unitary fiber-preserving operator and 𝑼M,𝝎:=𝑴𝑻𝝎{\bm{U}}_{M,{\bm{\omega}}}:=\bm{M}\bm{T}_{{\bm{\omega}}} with the Floquet operator 𝑼α,𝝎=𝑴α𝑻𝝎{\bm{U}}_{\alpha,{\bm{\omega}}}=\bm{M}_{\alpha}\bm{T}_{{\bm{\omega}}} as a special case.

It is then easy to see that 𝑼α,𝝎{\bm{U}}_{\alpha,{\bm{\omega}}} has a pure point spectrum (28) if, and only if, 𝑴α\bm{M}_{\alpha} is 𝝎{\bm{\omega}}-cohomologous to the constant SU(2)SU(2) matrix that is diagonal on the canonical basis, with eigenvalues eiχ,eiχe^{i\chi},e^{-i\chi}; where 𝑽(𝜽)\bm{V}({\bm{\theta}}) is the SU(2)SU(2) matrix that changes the local (𝒖(𝜽),𝝆𝒖(𝜽))(\boldsymbol{u}({\bm{\theta}}),\bm{\rho}\,\boldsymbol{u}({\bm{\theta}})) basis into the canonical basis. This leads to the following formulation of the quantum dynamics:

(𝑼α,𝝎𝝍)(𝜽)\displaystyle({\bm{U}}_{\alpha,{\bm{\omega}}}\boldsymbol{\psi})({\bm{\theta}}) =\displaystyle= eiχ[𝒖(𝜽𝝎)𝝍(𝜽𝝎)]𝒖(𝜽)\displaystyle e^{i\chi}[\boldsymbol{u}({\bm{\theta}}-{\bm{\omega}})\boldsymbol{\cdot}\boldsymbol{\psi}({\bm{\theta}}-{\bm{\omega}})]\boldsymbol{u}({\bm{\theta}}) (87)
+\displaystyle+ eiχ[𝝆𝒖(𝜽𝝎)𝝍(𝜽𝝎)]𝝆𝒖(𝜽).\displaystyle e^{-i\chi}[\bm{\rho}\boldsymbol{u}({\bm{\theta}}-{\bm{\omega}})\boldsymbol{\cdot}\boldsymbol{\psi}({\bm{\theta}}-{\bm{\omega}})]\bm{\rho}\boldsymbol{u}({\bm{\theta}}).\quad

Appendix B Fishman-Grempel-Prange formulation of SM models

Let 𝑼{\boldsymbol{U}}^{\diamond} denote the inverse Cayley transform of a unitary operator 𝑼\boldsymbol{U} in \cal H: that is, the self-adjoint operator which is defined by

𝑼=i(𝕀𝑼)(𝕀+𝑼)1,\displaystyle{\boldsymbol{U}}^{\diamond}\;=\;i\,(\mathbb{I}\,-\,\boldsymbol{U})\,(\mathbb{I}\,+\,\boldsymbol{U})^{-1}\;,
𝑼=(i𝑼)(i+𝑼)1.\displaystyle\boldsymbol{U}\,=\;(i\;-\;{\boldsymbol{U}}^{\diamond})\,(i\;+\;{\boldsymbol{U}}^{\diamond})^{-1}\,. (88)

The eigenvalue equation 𝑴α𝑻𝝎𝒖=eiλ𝒖\boldsymbol{M}_{\alpha}\boldsymbol{T}_{\boldsymbol{\omega}}\boldsymbol{u}\;=\;e^{i\lambda}\boldsymbol{u} is easily seen to be equivalent to 𝑯𝒖=0\boldsymbol{H}\boldsymbol{u}=0, where 𝑯=𝑴α+(𝑻𝝎eiλ)\boldsymbol{H}={\boldsymbol{M}}_{\alpha}^{\diamond}+(\boldsymbol{T}_{\boldsymbol{\omega}}e^{-i\lambda})^{\diamond}. This formulation was first worked out by Fishman, Grempel and Prange for standard QKR Fishman82a . In the momentum-spin representation, with basis vectors |𝑵,s|\bm{N},s\rangle (𝑵2\bm{N}\in{\mathbb{Z}}^{2}, s=1,2s=1,2), matrix elements of 𝑯\boldsymbol{H} are easily computed:

𝑵,s|𝑯|𝑵,s=δssδ𝑵𝑵tan𝑵𝝎/2λ/2)+\displaystyle\langle\bm{N},s|\boldsymbol{H}|\bm{N}^{\prime},s^{\prime}\rangle\;=\;\delta_{ss^{\prime}}\delta_{\bm{N}}\bm{N}^{\prime}\tan\bm{N}\boldsymbol{\cdot}\boldsymbol{\omega}/2-\lambda/2)\;+\;
tan(α/2)k=13s|𝝈k|sd^k(𝑵𝑵),\displaystyle-\;\tan(\alpha/2)\,\sum\limits_{k=1}^{3}\langle s|\boldsymbol{\sigma}_{k}|s^{\prime}\rangle\;\hat{d}_{k}(\bm{N}-\bm{N}^{\prime})\,,\quad (89)

where d^k(.)\hat{d}_{k}(.) are the 2D Fourier coefficients of the function dk(𝜽)d_{k}(\boldsymbol{\theta}). This can be read as a Hamiltonian for a spin 1/21/2 particle on a 22D discrete lattice, and resembles a tight-binding model in solid state physics. Like in the spinless Maryland model, the first term is a spin-independent, on-site potential in the 𝑵\boldsymbol{N}-space. It is quasi-periodic, so long as 𝝎\boldsymbol{\omega} is incommensurate. The second term describes hopping as well as spin coupling between different sites. It decays with the distance: |𝑵𝑵||\bm{N}-\bm{N}^{\prime}| the faster, the smoother the functions dk(𝜽)d_{k}({\bm{\theta}}) are.

References

  • (1) G. Casati, B. V. Chirikov, J. Ford, and F. M. Izrailev, Stochastic behavior of classical and quantum Hamiltonian systems, Lecture Notes in Physics Vol. 93 G. Casati and J. Ford Eds. (Springer, New York, 1979).
  • (2) S. Fishman, D. R. Grempel, and R. E. Prange, Chaos, quantum recurrences, and Anderson localization, Phys. Rev. Lett. 49, 509 (1982).
  • (3) S. Fishman, Anderson localization and quantum chaos maps, Scholarpedia 5, 9816 (2010).
  • (4) M. Raizen and D. A. Steck, Cold atom experiments in quantum chaos Scholarpedia 6, 10468 (2011).
  • (5) J.-C. Garreau, Quantum simulation of disordered systems with cold atoms, C. R. Phys. 18, 31 (2017).
  • (6) B. V. Chirikov, A universal instability of many-dimensional oscillator systems, Phys. Rep. 52, 263 (1979).
  • (7) F. M. Izrailev, Simple models of quantum chaos: Spectrum and eigenfunctions, Phys. Rep. 196, 299 (1990).
  • (8) I. Guarneri, Spectral properties of quantum diffusion, Scholarpedia 5, 10463 (2010).
  • (9) E. Doron and S. Fishman, Anderson localization for a two dimensional kicked rotor, Phys. Rev. Lett. 60, 867 (1988).
  • (10) R. Scharf, Kicked rotator for a spin-1/2 particle, J. Phys. A: Math. Gen. 22, 4223 (1989).
  • (11) Y. Chen and C. Tian, Planck’s quantum-driven integer quantum Hall effect in chaos, Phys. Rev. Lett. 113, 216802 (2014).
  • (12) C. Tian, Y. Chen, and J. Wang, Emergence of integer quantum Hall effect from chaos, Phys. Rev. B 93, 075403 (2016).
  • (13) K. v. Klitzing, G. Dorda, and M. Pepper, New method for high-accuracy determination of the fine-structure constant based on quantized Hall resistance, Phys. Rev. Lett. 45, 494 (1980).
  • (14) I. S. Burmistrov and A. M. M. Pruisken, θ\theta renormalization, electron Celectron interactions and super universality in the quantum Hall regime, Ann. Phys. (NY) 322, 1265 (2007).
  • (15) D. E. Khmel’nitskii, Quantization of Hall conductivity, Pis ma Zh. Eksp. Teor. Fiz. 38, 454 (1983) [JETP Lett. 38, 552 (1983)].
  • (16) P. Leboeuf, J. Kurchan, M. Feingold, and D. P. Arovas, Phase-space localization: Topological aspects of quantum chaos, Phys. Rev. Lett. 65, 3076 (1990).
  • (17) N. H. Lindner, G. Refael, and V. Galitski, Floquet topological insulator in semiconductor quantum wells, Nat. Phys. 7, 490 (2011).
  • (18) D. Y. H. Ho and J. Gong, Quantized adiabatic transport in momentum space, Phys. Rev. Lett. 109, 010601 (2012).
  • (19) L. Zhou and J. Gong, Floquet topological phases in a spin-1/21/2 double kicked rotor, Phys. Rev. A 67, 063603 (2018).
  • (20) I. Dana and K. Kubo, Floquet systems with Hall effect: Topological properties and phase transitions, Phys. Rev. B 100, 040107 (2019) and references therein.
  • (21) N. Brenner and S. Fishman, Pseudorandomness and localization, Nonlinearity 5, 211 (1992).
  • (22) C. Tian, A. Kamenev, and A. I. Larkin, Weak dynamical localization in periodically kicked cold atomic gases, Phys. Rev. Lett. 93, 124101 (2004).
  • (23) C. Tian, A. Kamenev, and A. Larkin, Ehrenfest time in the weak dynamical localization, Phys. Rev. B 72, 045108 (2005).
  • (24) D. R. Grempel, S. Fishman, and R. E. Prange, Localization in an incommensurate potential: An exactly solvable model, Phys. Rev. Lett. 49, 833 (1982).
  • (25) S. Fishman, D. R. Grempel, and R. E. Prange, Localization in a dd-dimensional incommensurate structure, Phys. Rev. B 29, 4272 (1984).
  • (26) The Maryland models studied in the present work are a special class of quantum kicked systems. However this denotation is more frequently reserved to certain tight-binding-like models in a quasi-periodic potential. Such models are related to the kicked systems by a construction Fishman82a , that can be replicated for the SM models, as shown in Appendix B.
  • (27) M. V. Berry, Incommensurability in an exactly-soluble quantal and classical model for a kicked rotator, Physica 10D, 369 (1984).
  • (28) S. Ganeshan, K. Kechedzhi, and S. Das Sarma, Critical integer quantum hall topology and the integrable maryland model as a topological quantum critical point. Phys. Rev. B 90, 041405 (2014).
  • (29) A. C. Keser, S. Ganeshan, G. Refael, and V. Galitski, Dynamical many-body localization in an integrable model, Phys. Rev. B 94, 085210 (2016).
  • (30) E. Abrahams, P. W. Anderson, D. Licciardello, and T.V. Ramakrishnan, Scaling theory of localization: Absence of quantum diffusion in two dimensions, Phys. Rev. Lett. 42, 673 (1979).
  • (31) L. P. Gor’kov, A. I. Larkin, and D. E. Khmelnitskii, Particle conductivity in a two-dimensional random potential, Pis’ma Zh. Eksp. Teor. Fiz. 30, 248 (1979) [JETP Lett. 30, 228 (1979)].
  • (32) C. Hainaut, I. Manai, J.-F. Clément, J.-C. Garreau, P. Szriftgiser, G. Lemarié, N. Cherroret, D. Delande, and R. Chircireau, Controlling symmetry and localization with an artificial gauge field in a disordered quantum system, Nat. Commun. 9, 1382 (2018).
  • (33) M. Fruchart, Y. Zhou, and V. Vitelli, Dualities and non-Abelian mechanics, Nature 577, 636 (2020).
  • (34) J. Aaronson, An Introduction to Infinite Ergodic Theory, Math. Surveys and Monographs vol. 50, chapter 8 (AMS, 1997).
  • (35) L. H. Eliasson, Reducibility and point spectrum for linear quasi-periodic skew-products, in Proc. ICM (Berlin, 1998). Documenta Mathematica Extra vol. 2.
  • (36) L. H. Eliasson, Ergodic skew-systems on 𝕋d×SO(3,)\mathbb{T}^{d}\times SO(3,\mathbb{R}), Ergodic Theory Dynam. Systems 22, 1429 (2002).
  • (37) R. T. de Aldecoa, The absolute continuous spectrum of skew products of compact Lie groups, Isc. J. Math. 208, 323 (2015).
  • (38) N. Karaliolios, Continuous spectrum or measurable reducibility for quasiperiodic cocycles in 𝕋d×SU(2)\mathbb{T}^{d}\times SU(2), Commun. Math. Phys. 358, 741 (2018).
  • (39) The term “duality” is here used in a sense which is similar, though not identical, to Ref. Vitelli20 . Not so for the term “symmetry”.
  • (40) G. Casati, I. Guarneri, and D. L. Shepelyansky, Anderson transition in a one-dimensional system with three incommensurate frequencies, Phys. Rev. Lett. 62, 345 (1989).
  • (41) The triple (1,x,x2)(1,x,x^{2}) is the “most incommensurate” triple of real numbers Casati89 .
  • (42) J. P. Dahlhaus, J. M. Edge, J. Tworzydlo, and C. W. J. Beenakker, Quantum Hall effect in a one-dimensional dynamical system, Phys. Rev. B 84, 115133 (2011).
  • (43) A. Altland, S. Gnutzmann, F. Haake, and T. Micklitz, A review of sigma models for quantum chaotic dynamics, Rep. Prog. Phys. 78, 086001 (2015).
  • (44) It is worth recalling that in classical integrable systems instability is generically linear in time.
  • (45) I. P. Kornfeld and Ya. G. Sinai, Spectral theory of dynamical systems, in Dynamical Systems, Ergodic Theory and Applications Ya. G. Sinai Eds. (Springer, 2000).
  • (46) By “unitary symmetry” we refer to the standard definition in quantum mechanics, i.e. an operator AA which is unitary 𝑨𝑨=𝑨𝑨=𝕀\bm{A}\bm{A}^{\dagger}=\bm{A}^{\dagger}\bm{A}=\mathbb{I} and at the same time is a symmetry, 𝑨2=𝕀\bm{A}^{2}=\mathbb{I}. It should not be confused with the “unitary class” in the jargon of the random matrix theory and the theory of Anderson localization or topological insulators.
  • (47) R. Ketzmerick, G. Petschel, and T. Geisel, Slow decay of temporal correlations in quantum systems with Cantor spectra, Phys. Rev. Lett. 69, 695 (1992).
  • (48) In this discussion we are excluding pathological exceptions. These may be expected to occur when 𝝎\boldsymbol{\omega} is close to commensurate.
  • (49) I. Guarneri, Singular Continuous spectra and discrete wave packet dynamics, J. Math. Phys. 37, 5195 (1996) .
  • (50) Though formally 11D, the dynamics (25) are not subject to this bound, because their one-period evolution operators explicitly depend on time tt.
  • (51) J. Bellissard and H. Schulz-Baldes, Subdiffusive quantum transport for 3D Hamiltonians with absolutely continuous spectra, J. Stat. Phys. 99, 1 (2000).
  • (52) F. M. Izrailev and D. L. Shepelyansky, Quantum resonances for a rotor in a nonlinear force field, Theor. Math. Phys. 43, 353 (1980).
  • (53) I. Guarneri, On the spectrum of the resonant quantum kicked rotor, Ann. Henri Poincaré 10, 1097 (2009) and references therein.
  • (54) M. Reed and B. Simon, Methods of Modern Mathematical Physics (Vol. 1) (Academic Press, 1981).
  • (55) It follows from minimality of the shift 𝝉ω\boldsymbol{\tau}_{\omega}, because the set of the points of discontinuity is closed and invariant.
  • (56) Though intuitive, a rigorous justification of this conclusion rests on Sobolev’s lemma Simon2 .
  • (57) M. Reed and B. Simon, Methods of Modern Mathematical Physics (Vol. 2) (Academic Press, 1981).
  • (58) J. Chabé, G. Lemarié, B. Grémaud, D. Delande, P. Szriftgiser, and J.-C. Garreau, Experimental observation of the Anderson metal Cinsulator transition with atomic matter waves, Phys. Rev. Lett. 101, 255702 (2008).
  • (59) C. Tian, A. Altland, and M. Garst, Theory of the Anderson transition in the quasiperiodic kicked rotor, Phys. Rev. Lett. 107, 074101 (2011).