Near invariance of quasi-energy spectrum of Floquet Hamiltonians
Abstract.
The spectral analysis of the unitary monodromy operator, associated with a time-periodically (paramatrically) forced Schrödinger equation, is a question of longstanding interest. Here, we consider this question for Hamiltonians of the form
where is an unperturbed autonomous Hamiltonian, , and has a period of . In particular, in the small regime, we seek a comparison between the spectral properties of the monodromy operator, the one-period flow map associated with the dynamics, and that of the autonomous (unforced) flow, . We consider which is spatially periodic on with respect to a lattice. Using the decomposition of and into their actions on spaces (Floquet-Bloch fibers) of pseudo-periodic functions, we establish a near spectral-invariance property for the monodromy operator, when acting data which are -localized in energy and quasi-momentum. Our analysis requires the following steps: (i) spectrally-localized data are approximated by band-limited (Floquet-Bloch) wavepackets; (ii) the envelope dynamics of such wavepackets is well approximated by an effective (homogenized) PDE, and (iii) an exact invariance property for band-limited Floquet-Bloch wavepackets, which follows from the effective dynamics. We apply our general results to a number of periodic Hamiltonians, , of interest in the study of photonic and quantum materials.
1. Introduction
We consider a class of -dimensional Schrödinger equations with time-periodic forcing, governing , a complex-valued function of and :
(1.1) |
where, , the operator is self-adjoint on , and for each the operator is also self-adjoint on . Furthermore, is periodic of period , i.e., for all . Hence,
(1.2) |
We consider (1.1) for and small: the regime of small and slowly varying time-periodic forcing. Very briefly, wave-packet initial-data will deform on a time-scale which depends on its spectral localization. The parameter is therefore chosen so that this time-scale and the forcing period are matched; see Section 1.1.
Since is time-dependent (non-autonomous), the spectra of the family of operators does not determine the time-dynamics (1.1). Instead, one must study the period mapping for (1.1). For initial data , the solution of the initial value problem (1.1) is defined by the unitary operator
(1.3a) | |||
The period mapping, or monodromy operator, is the unitary operator | |||
(1.3b) |
The driven and undriven problems can be compared by viewing the autonomous case , i.e., the dynamics of , as having (trivial) periodicity. In this case, and the monodromy operator is
(1.4) |
The spectrum of acting in has a simple relation to the spectrum of :
(1.5) |
This relation motivates the notion of quasi-energy: A point on the spectrum of the monodromy operator can be written as . The phase is called a Floquet exponent, and is called a quasi-energy. We ask:
Question 1.
What is the relation between the spectrum of the monodromy operator and that of , arising from the non-trivial time-periodic forcing?
In general time-periodic settings, beyond being unitary, very little is known about the spectrum of the monodromy operator; see the discussion of Section 1.2.
In this paper, we gain insight on this question for the class of operators , where the unperturbed Hamiltonian, , is periodic with respect to spatial translations in a lattice , i.e., for all and all . Since commutes with -translations, it may be decomposed into its action on distinct spaces of pseudo-periodic functions, where varies over the Brillouin zone; see Section 2.1. This decomposition allows us to formulate and address a spectrally-local version of Question 1:
Question 2.
What is the relation between the spectrum of the monodromy operator and that of , when restricted to data concentrated near an energy and quasi-momentum ?
1.1. Discussion of main results
We begin by discussing more specifically the type of spectral localization we have in mind. First, fix a general quasi-momentum and energy , and consider initial data which is “ spectrally localized” with respect to – a Bloch wave-packet of band-width (see Proposition 4.7). The (unforced) evolution of such data by , on large finite time-scales, has the structure of a slowly-varying spatial and temporal modulation of Bloch modes with energy and quasi-momentum . The slow evolution is described by an effective Hamiltonian, which is determined by the local character of the band structure (energy dispersion curves and eigenspaces) near . The effective Hamiltonian captures the transport and spreading dynamics of such data.
We choose the forcing time-scale of our time-dependent Hamiltonian, , so that there is a non-trivial interplay between the unforced transport dynamics and the effect of time-dependent forcing. 111Such balancing corresponds to what is typically done in experiments. For example, a wavepacket excitation is designed by a choice of a laser frequency and pulse band-width, and balanced with forcing to measure effects on experimentally accessible spatial and temporal scales. The parameter in (1.1) is chosen to achieve this balance of effects. For example, the choice in (1.1) is appropriate for the dynamics where our Bloch wavepacket is concentrated near a point where the dispersion is locally linear, thus allowing for simple transport or conical diffraction, for example. In Section 5 we discuss a number of examples.
A key to our analysis is the space of band-limited wavepackets (Definition 4.1): the space of Fourier band-limited envelope modulations of Bloch modes ( pseudo-periodic eigenstates) of with energy . states are good approximations to spectrally localized a Floquet-Bloch wavepackets of band-width (Proposition 4.7, also [39]).
Let denote the spectral measure associated with unitary operator (see Section 2); let be the projection onto the subspace of , which is the space of functions which are localized, with respect to , in quasi-momentum and energy about . In particular, note that . Our main near-invariance results are:
-
(1)
Theorem 4.4 [Near invariance on ] Let denote any arc such that the spectrum of is contained in . Then, for sufficiently small,
or equivalently This provides an answer to Question 2: when states in the range of evolve under the forced monodromy operator, , the resulting state has very small projection onto quasi-energies far from their quasi-energies under . Equivalently, denoting the spectral measure of by ,
Underlying the proof of Theorem 4.4 is the following strict invariance result on , the space of band-limited wave-packets:
- (2)
Remark 1.1.
These results are not obtainable by standard perturbation theory. Indeed, although the forcing term in (1.1) is multiplied by a small parameter, , the monodromy operator is determined by the evolution on a time-scale of length (the period of the forcing). Hence, formally, the cumulative contribution of forcing is of order one.
1.2. Relevant analytical work on temporally-forced Hamiltonian PDEs
In the present article, we give a novel perspective on the spectrum of parametrically (periodically) forced Schrödinger equations. Here, we draw connections between our result and other approaches to similar problems.
Reducibility
By translation invariance with respect to a lattice (see Section 2.1), the dynamics of (1.1) can reduce to the spectral study of the Floquet Hamiltonian
over the family of spaces . For each fixed , spectral problems of this latter type correspond to time-periodically forced wave equations on the spatial torus. By constructing a change of variables which approximately maps the original Hamiltonian to an autonomous Hamiltonian (reducibility), it is shown that has pure point spectra, with quantitative control on the effect of the forcing on the unperturbed point spectrum [4, 3, 9, 13, 21, 30]. These results hypothesize very strong rate of growth assumptions on the point eigenvalues of . For Schrödinger operators, the Weyl asymptotics imply that these growth assumptions are only satisfied for spatial dimension . An exception to that is [9], which establish analogous results for for the case where is assumed to be sufficiently small. To the best of our knowledge, beyond these works, the nature of the spectrum of remains an open problem [30].
Adiabatic Theory
Adiabatic theory studies the time-dynamics generated by Hamiltonians, which are slowly varying in time:
(1.6) |
The mapping defines a curve of Hamiltonians acting in a Hilbert space , of some specified regularity. A common setting is that the initial data is taken to be in the range of , a spectral projector associated with an isolated subset of the spectrum of . Under general assumptions, one has that for times , the solution is well-approximated by a vector, , which, for each is in the range of spectral projector , associated with the family of instantaneous Hamiltonians . The order of the approximation with respect to depends on the smoothness of ; see, for example, [2, 6, 8, 14, 16, 22, 24, 31, 32].
While slowly varying in time, the class of Hamiltonians we study, in (1.1), has a different structure, and so does not fall into the category of Hamiltonians (1.6). Moreover, in most adiabatic theorems, the spectral gap assumption used to construct the family of time-dependent projections, . In contrast, our near-invariance results are in terms of a fixed, time-independent projections and subspaces, i.e. , its range, and .
1.3. Outline of the paper.
In Section 2 we provide necessary background on Floquet-Bloch theory of periodic Hamiltonians, the spectral theorem for unitary operators, and introduce relevant notation. Section 3 presents a brief intuitive introduction to effective dynamics and homogenization of periodic Hamiltonians. The main results of this article are presented in Section 4. In Section 5 we demonstrate how these results apply to a number of specific periodic Hamiltonians, and the associated , of physical interest. The proofs of the main results are presented in Section 6. Finally, a formal derivation of an effective transport equation (see Section 5.1) is presented in Section 7.
1.4. Acknowledgments
AS would like to thank P. Kuchment and V. Rom-Kedar, whose questions inspired this research. This research was supported in part by National Science Foundation grants DMS-1620418, DMS-1908657 and DMS-1937254 (MIW) as well as Simons Foundation Math + X Investigator Award #376319 (MIW, AS) and the AMS-Simons Travel Grant (AS).
2. Mathematical Preliminaries
2.1. Floquet-Bloch theory
We consider Hamiltonians , where is periodic with respect to a lattice , where is a linearly independent set of vectors in . 222We believe our analysis can be extended to general classes of elliptic operators (scalar Hamiltonians and systems) , whose coefficients are periodic with respect to a lattice; see [29]. Since commutes with lattice translations, the Hamiltonian admits a fiber-decomposition , where denotes the operator acting in the space , consisting of functions which are pseudoperiodic, i.e. if a.e. in . The set is the Brillouin zone, a choice of fundamental cell in . For each quasimomentum, , is self-adjoint and compact resolvent. Hence, for each , has a real sequence of discrete eigenvalues of finite multiplicity
tending to infinity. The corresponding eigenfunctions, denoted by :
or equivalently . These eigenfunctions may be chosen to form an orthonormal basis for . Each function is continuous and piecewise analytic [28, Theorem 5.5], and hence Lipschitz continuous.
Each image, , is a subinterval of called the spectral band. The graphs of are called dispersion surfaces. The collection of all pairs and corresponding normalized eigenfunctions, , is called the band structure of . Finally, the family of Floquet-Bloch modes is complete in ; for any ,
where the sum is interpreted as a convergence of partial sums in . For simplicity, we will assume henceforward that .
2.2. , and its associated spectral measure
In (1.3a) we introduced the (unitary in ) evolution associated with the dynamics (1.1). In this section we give a brief outline of a construction of and then discuss the spectral measure of the associated monodromy operator. Under very general assumptions on and the operators , a unitary propagator can be shown to exist using semigroup methods, see [34, Section 5] and [37, Section X.12].
From and , we obtain the unitary monodromy operators and which, by the spectral theorem, are equipped with associated spectral (projection-valued) measures and , respectively. For completeness, we review the definition and properties of a spectral measure and the spectral theorem for unitary operators. We refer the reader to [7, 17, 38, 40] for details.
Let be a Hilbert space, let be a set, and a -algebra in . A map , the Banach space of bounded linear operators on , is called a projection-valued measure if the following properties hold:
-
(1)
is an orthogonal projection for every .
-
(2)
and .
-
(3)
If are disjoint then
-
(4)
for all .
Theorem 2.1.
Let be a unitary operator on . There exists a unique projection-valued measure on the Borel -algebra of , which contains the spectrum of , such that for every
2.3. Notation and conventions
-
(1)
We adopt the convention of considering all vectors as column vectors. If , .
-
(2)
Fourier transform on : For a function, , defined on , define its Fourier transform as
Furthermore, introduce
The mappings and map Schwartz class, , to itself and for all , we have the Plancherel identity and the inversion formula . Hence, on . By density, both mappings extend to bounded linear transformations on which satisfy the Plancherel identity, and the inversion formula.
-
(3)
Pauli matrices are given by , and
-
(4)
For a vector , we write .
3. Wavepackets, the geometry of dispersion surfaces, and periodic homogenization
Our spectrally local formulation concerning the quasi-energy spectrum , Question 2. is a natural relaxation of Question 1. In physical settings, a crystalline structure is experimentally probed in a narrow spectral range, e.g. a bulk material is externally excited (e.g. electrically, optically, elastically, acoustically). Such settings induce the propagation of spectrally localized wavepackets (quasi-particles), whose envelope dynamics are given by a simplified effective Hamiltonian.
To illustrate this last point and how effective Hamiltonians emerge, consider the following “toy model” of continuously translation-invariant and time-periodically forced Hamiltonian dynamics governing a wave-field :
(3.1) |
where is sufficiently smooth and localized on . The real-valued dispersion relation is, for simplicity, taken to be smooth. Clearly, an explicit solution can be given in terms of the Fourier transform, but our goal here will be to discuss the notion of effective dynamics.
Consider initial data whose Fourier transform is concentrated near :
The solution of the initial value problem (3.1) may be written as:
where
If , then by Taylor expansion of about , we obtain the following approximation of the solution of (3.1) with , which is valid on the time scale: :
where the envelope is governed by a driven transport equation
If, on the other hand, and , the Hessian matrix, is non-singular, then we obtain the following approximate solution of (3.1) with , which is valid on the time scale: :
where satisfies an (generally anisotropic) effective Schrödinger equation:
In each case, the function , which provides the slow envelope evolution, is governed by a time-dependent effective Hamiltonian:
(3.2) |
in which both the effects of deformation under and temporal forcing are captured. Note also that commutes with continuous spatial translations and therefore can be analyzed using the Fourier transform.
In general, for spatially homogeneous media and for the case of crystalline (lattice periodic) media described by , which is invariant under discrete translations in a lattice, the dispersion relation eigenvalue-branches may be degenerate. At such degeneracies the dispersion relations may not be smooth, although they are Lipschitz continuous if is self-adjoint. Furthermore, in such cases, the eigenvector maps may even be multivalued. 333 In this paper, we discuss only isolated point degeneracies. Other types of band degeneracies may arise. Examples are (i) the touching of two bands along a submanifold of quasi-momenta due the underlying symmetries and (ii) degeneracies of infinite multiplicity such as “flat bands,” as in e.g., the Landau Hamiltonian [19]. We do not treat these situations in the present work. Nevertheless, Fourier-type analysis (based on Floquet-Bloch modes) and multiple-scale / homogenization methods can be used to rigorously derive, with accompanying error bounds, effective envelope dynamics. Examples are
4. Main results
4.1. Hypotheses and definitions
Our first assumption concerns the character of the energy band structure near ; in particular if is a degeneracy, then this degeneracy is isolated:
Hypothesis 1 (Spectral separation).
Let be such that has an eigenvalue of multiplicity , i.e., for some
(4.1) |
Furthermore, is isolated in the band structure in the sense that
for all in an open neighborhood of the quasimomentum . Introduce an orthonormalize basis for the degenerate eigenspace:
With Question 2 in mind and assuming spectral separation as defined in Hypothesis 1, we define a projection, , associated with a subspace of consisting of states, which are superpositions of modes whose quasimomenta and energy are near :
(4.2) |
where is fixed.
We next present two additional assumptions concerning the underlying wave-packet dynamics. Let satisfy the spectral separation Hypothesis 1 with parameters and . Denote the vector of degenerate Floquet-Bloch modes:
We next introduce the subspace of , consisting of Fourier band-limited wave-packets, which are modulations .
Definition 4.1 (Band-limited wave-packets).
For fixed parameters , we define:
(4.3) |
Hypothesis 2 (Translation invariant effective dynamics).
There is a one-parameter family of unitary operators on , , with the following properties:
-
(1)
(Spatially translation invariant effective dynamics)
For , is defined by:where is a smooth mapping from into the space of unitary matrices.
-
(2)
(Approximation by effective dynamics)
Let be defined as in (1). If , then(4.4)
where is given in (1.2).
Remark 4.2.
For , Hypothesis 2 implies slow envelope effective dynamics. Indeed, let
Then, using the space-time scaling , we may write:
Equivalently, evolves under the effective Hamiltonian , which generates the unitary flow :
The effective evolution operator, , naturally gives rise to an
(4.5) |
for ,
Hypothesis 3 (Spectrum of the effective monodromy operator).
For every sufficiently small there exists such that
(4.6) |
4.2. A theorem on near-invariance of quasi-energy spectrum
Since the monodromy operator is unitary (see (1.3)), has a spectral representation as an integral with respect to a projection-valued spectral measure, , which is supported on the unit circle; see Sec. 2.2. We now state our main theorem, which addresses Question 2.
Denote by the arc .
Theorem 4.4 (Near invariance).
Consider the periodically forced Schroedinger equation (1.1). Assume that for some quasi-momentum / energy pair (see Remark 4.3) Hypotheses 1–3 are satisfied. Let , defined in (4.2), denote the projection onto Bloch modes of of energy and quasi-momentum in an neighborhood of .
Then, for every there exists such that for all and every , then
(4.8) |
Theorem 4.4 is a near-invariance (or stability) result for a spectral subspace associated with , the range of , under the perturbed dynamics . Indeed, let ; Then, under Hypothesis 1, the non-driven monodromy operator , restricted to the range of , is given by:
Since the dispersion surfaces are Lipschitz continuous, there is a constant such that for and for all we have
Denoting the spectral measure of by , we have that for a fixed and sufficiently small , by inspecting the explicitly expression above,
(4.9) |
As discussed in Remark 1.1, it is non-trivial that a form of (4.9) persists for time-periodic forcing in (1.1), due to the formally order-one cumulative effect of a perturbation of size on the time-scale .
4.3. The main result for the space of band limited wavepackets
As a step toward the proof of Theorem 4.4, we first prove its analog, Theorem 4.6, a strict invariance property for functions in (see (4.3)), a closed subspace of . Since approximates the range of (Proposition 4.7), we can then use Theorem 4.6 to prove Theorem 4.4, which concerns the range of .
Lemma 4.5 ([39]).
There exists , such that for all , defined in (4.3) is a closed subspace of . Hence, has the decomposition
with corresponding orthogonal projections on denoted
and . |
is a very natural space with which to study the effects of time-dependent forcing. In fact, the proof of Theorem 4.4, follows from its analog for the space :
Theorem 4.6 (Invariance on ).
Consider (1.1) and suppose it satisfies Hypotheses 1–3 at some quasi-momentum energy pair . Fix and such that . Then, for every there exists such that for all
(4.10) |
Equivalently,
(4.11) |
Theorem 4.6 is proved in Section 6. Here, we first use it to give a proof of the main result, Theorem 4.4 (concerning ).
Proof of the main result, Theorem 4.4.
To prove Theorem 4.4 we shall use Theorem 4.6 above and the following Proposition, which is proved in Section 6.1:
Proposition 4.7 ( approximates ).
There exists such that for every the following holds: for every there is a with (see (4.3)) such that
(4.12) |
Conversely, there exists such that for every with sufficiently small, then
(4.13) |
5. Applications of the main result, Theorem 4.4
In this section we apply Theorem 4.4 to time-periodically forced (Floquet) Hamiltonians of the form:
(5.1) |
Here, is -periodic with zero mean, i.e., . A discussion, with references, of how this class of models arises in condensed matter physics and photonics is presented in Appendix A.
The setting of Theorem 4.4 is a Floquet Hamiltonian, here (5.1), and a neighborhood of an energy quasi-momentum pair in the band structure of , in which the class of wave-packet initial data are spectrally localized. Here, we characterize the local character of the band structure at by a number of parameters. As in Hypothesis 1, we denote by the multiplicity of . The parameter in (5.1) is chosen to match the rate at which to energy, , deviates from for small. Table 1 summarizes four cases of physical interest, which are discussed in the following subsections.
Section | dispersion rate | , degeneracy order | dimension | effective equation |
---|---|---|---|---|
5.1 | Transport (5.3) | |||
5.2 | Dirac system (5.7) | |||
5.3 | Schrödinger (5.8) | |||
5.4 | Schrödinger system (5.9) |
In what follows, wavepackets are always denoted by , where the dimension of and is , the degree of the degeneracy at .
In the non-driven case, i.e., when , the effective/homogenized models, which govern the large time dynamics of wave-packet envelopes, are continuously translation-invariant PDEs of the form ; see references below. Our analysis shows that, for (5.1) with , the dynamics of wave-packet envelopes, is governed by
where the non-autonomous Hamiltonian is obtained from arising from via the formal replacement In each example below, we display .
5.1. simple and a non-critical point - ballistic transport
For a given Hamiltonian , let be a pair of a quasi-momentum and index such that is a simple eigenvalue of in with a linear dispersion relation, i.e.,
(5.2a) | |||
and | |||
(5.2b) | |||
where (since the dispersion surfaces are real-valued). |
By continuity of the energy bands, Hypothesis 1 holds. The effective Hamiltonian, governing the wave-packet envelope , is given by
(5.3) |
5.2. of multiplicity two; conical touching of dispersion surfaces at , a Dirac point
An example which plays an important role in the modeling of two-dimensional materials such as graphene is the case where , where is a honeycomb lattice potential, i.e. has the symmetries of a honeycomb tiling of . (A one-dimensional variant of such potentials, dimer potentials, was studied in [10] and the following discussion can adapted to this setting as well.) For generic honeycomb lattice potentials, conical degeneracies (Dirac points) occur in the band structure at pairs , where is any vertex (high symmetry quasimomentum) of the hexagonal Brillouin zone [11]. In a neighborhood of a Dirac point one has two consecutive dispersion surfaces, , satisfying
(5.6) |
The slope of the cone, , is referred to as the Dirac or Fermi velocity. From (5.2b) we see that Hypothesis 1 is satisfied.
5.3. simple, a non-degenerate critical (quadratic) point of a band
Suppose is such that is a simple -eigenvalue of and is a non-degenerate critical point of the band dispersion function : and . Then, , and the envelope dynamics for are given by an driven effective Schrödinger-type Hamiltonian:
(5.8) |
The validity on time scales of order , and therefore Hypothesis 2 follows along the lines of [1] or [39].
Note that such quadratic points may occur at spectral band edges, in which case the Hessian is positive or negative definite or at ; or where is interior to a spectral band, in which case the Hessian might have an indefinite signature. Finally, a similar homogenization argument can be carried in the case where and non-degenerate. In this case one gets a Schrödinger equation on the time-scales of , with a drift term on the time-scale of ; see, for example, [1].
5.4. of multiplicity two; quadratic touching of two dispersion surfaces at
Consider a two-dimensional Hamiltonian where the potential which is periodic with respect to the lattice , real-valued, even, and invariant under a rotation. We can take the Brillouin zone, , to be a square, centered at the origin in . The vertices of are high-symmetry quasi-momenta. In [25] it is proved that the band structure of has consecutive band dispersion surfaces which touch quadratically over the vertices of at an eigenvalue with a two-fold degenerate eigenvalue.
Hence, we consider (5.1) with for data near these high-symmetry -points. The effective envelope dynamics of data can be shown, in a manner analogous to the derivation in [26], to be governed by the matrix-Schrödinger effective Hamiltonian:
(5.9a) | |||
where | |||
(5.9b) |
Here, the coefficients and can be expressed as inner products involving a basis for the dimensional of ; see [26], and are Pauli matrices. As in the case of the effective Dirac equation (Section 5.2), Hypothesis 3 is verified as follows: (i) Fourier-transforming (5.9) yields a system of linear and time-periodic system of (Floquet) ODES, which is parametrized by . (ii) Since this matrix defining this system of ODEs has trace equal to and is continuous in , the Floquet multipliers are continuous functions of on unit circle. (iii) Hence, for data with a fixed band-width (see Definition 4.1), there exists a continuous function such that the data of is contained in the arc .
6. Proof of Theorem 4.6
Let us first recall the following centering lemma for unitary operators. Intuitively, it says that if a unitary operator acts on a function which is spectrally localized, it is approximately the same as acting as a multiplication operator. We proved a weaker version of this lemma in [39], and include the proof here for completeness.
Lemma 6.1.
Let such that and let be the mid-point of the arch . Then
where is the arclength of .
Proof.
Let . Then
Since , we only need to bound . By the spectral theorem (see Sec. 2.2), we have that
where we have used the expression for the arc length: for any with , combined with the fact that is the mid-point of . ∎
Proof of Theorem 4.6 .
To prove (4.11), let and let
for , where is defined in Hypothesis 3. We will now show that . Lemma 6.1 implies that, since is the midpoint of the arch ,
Hence
(6.1) |
To bound from below, we will prove the following lemma:
Lemma 6.2.
For any there exists such that for all , ,444An analogous formula holds if . and any ,
Let us first use lemma 6.2 with to prove the main result, Theorem 4.6, and then return to its proof. Combined with (6.1), we have that
On the other hand, since , Hypothesis 2 regarding the effective dynamics provides an upper bound on . When combined this yields that
Since , the difference on the left-hand side above is always positive. Therefore, for sufficiently small , the above inequality is only possible if . ∎
Proof of Lemma 6.2 .
By the explicit form of given in Hypothesis 2, we can write for every ,
(6.2) |
where
(6.3) |
Next, we recall the following averaging lemma from [39, Lemma 4.5]:
Lemma 6.3.
Let such that for some , and let be -periodic. Then, there exists which depends on , such that for any fixed ,
(6.4) |
Applying Lemma 6.3 to (6.2) yields, using the orthonormality of (for brevity, set without loss of generality)
where in applying Lemma 6.3, we used the fact that, while the support of the Fourier transform of might not be , it is still compact.
Hence, to prove Lemma 6.2, we need to bound the norm of from below. We now note that for every , the Fourier-transformed monodromy is an unitary matrix (where is the degree of the degeneracy in Hypothesis 1). Let be the unitary matrix which diagonlizes the monodromy, i.e.,
Hence, using Plancharel theorem and the orthogonality of , we have that
where the last inequality is derived from the arc-length formula between two angels, as well as from Hypothesis 3 on the spectrum of .
∎
6.1. Proof of Proposition 4.7
We note here that the proof of Proposition 4.7 is very similar to that which appears in [39]. However, due to many changes in the notation and change in dimensionality, we include it here for completeness.
6.1.1. From projections to wavepackets; proof of (4.12)
Let be taken sufficiently small, and let . Express acting in as a direct integral , where denotes the operator acting in . Then, taking without loss of generality, we can rewrite (4.2)
(6.5) | ||||
(6.6) |
where the the factor in the radius in the contour integral is not necessarily sharp. In order to expand for near , we next express the operators in terms of operators which acts in the fixed space . Note that , where acts in . Furthermore, .
Substitution into (6.6) yields
The contour integral inside the square brackets is smooth -valued function of , and so by Taylor expansion:
(6.7) |
The last term in (6.7) is linear in and easily seen to be bounded in by since the domain of integration is over a disc of radius .
The dominant term in (6.7) may be re-expressed as
In all, we have that
(6.8) |
where
(6.9) |
where denotes the inverse Fourier transform and denotes convolution. We next show that with by showing that . Indeed by the convolution rule,
which is supported in . This completes the proof of (4.12).
Remark 6.4.
This proof shows that, more generally, if the definition of would have been changed to a project onto the disc with , then the Proposition would have carried through with a different value of .
6.1.2. From wavepackets to projections; proof of (4.13)
Consider for some and sufficiently small, then by definition of (4.3), there exists such that
On the other hand, by (6.8), for any function and sufficiently small, there exists a function such that
(6.10) |
To prove (4.13) it suffices to show that . Substitution of into (6.9) yields
We next compute the Fourier transform of . For ,
(6.11) |
Consider the expression being summed in (6.11). Since with periodic, we have
We expand for each in a Fourier series with respect to the lattice : for every ,
Substituting the Fourier series into (6.11) yields
(6.12) |
Note that by definition, has compact support in the disc of radius around the origin. In the expansion above in (6.12), for sufficiently small, the only term that does not vanish only if (i) (with ) and (ii) due to the term, if . Hence, the only non-zero term in (6.12) arises from the lattice point . Then, by definition of the Fourier coefficient and the orthogonality of the different ’s, we have that
7. Effective transport dynamics
Consider (5.1) with and initial data of the form
(7.1) |
for sufficiently high , and where the factor keeps the overall norm of independent of . Initial data in is then a sub-class of (7.1). In this subsection, we formally derive the effective transport equation and its propagator , as given in (LABEL:eq:transport).
To construct a solution, we assume separation of scales, with slow time variables
and introduce the expansion
where for every
By expanding
and substituting into (1.1), we solve for each power of .
Order .
and so .
Order .
To invert in and solve for , we need to verify that the right hand side is orthogonal to the kernel, i.e., to (from here on, we suppress the and dependence for brevity).
Here, it is useful to note that (5.2) is equivalent to a statement on the Bloch mode :
Lemma 7.1.
Given (5.2), then
(7.2) |
Combining Lemma 7.1 and normalizing , we get the desired result
Proof of Lemma 7.1.
By definition, satisfies
Write
(7.3) |
which transforms the TISE to
(7.4) |
We now take on both sides of (7.4). By noting that
we get that
We rearrange some of the terms and take the inner product from the left with
Here we note that, by definition . Moreover, since is self-adjoint, combined with (7.4), the first inner product on the right-hand side vanishes. By differentiating (7.3), we get
∎
Appendix A Physical interpretations of the model
An example of physical interest is the case of (5.1), i.e., (1.1) with . Note here that (5.1) can be transformed to an equivalent “magnetic” form
where is a vector potential. This class of PDEs arises in physical settings, such as:
-
(a)
The modeling of time periodic conductors (e.g., graphene), excited by a time-varying electric field [35, 41]. Here, is a single-electron Hamiltonian for graphene and time-dependence in models the excitation of the graphene sheet by an external electric field with no magnetic field (by Maxwell’s equations, since is constant in space, see e.g., [27]).
-
(b)
For , the propagation of light in a periodic array of helically coiled optical fiber waveguides [5, 23, 33, 36]. Here, the Schrödinger equation describes the propagation in the time-like longitudinal direction of a continuous-wave (CW) laser beam propagating through a hexagonal or triangular transverse array of optical fiber waveguides. Beginning with Maxwell’s equations, under the nearly monochromatic and paraxial approximations, one obtains for the longitudinal evolution of the slowly varying envelope of the classical electric field. Suppose the fibers are longitudinally coiled. Then, in a rotating coordinate frame, we obtain (1.1) where the time-periodic perturbation, , captures effect of periodic coiling.
References
- [1] G. Allaire and A. Piatnitski. Homogenization of the schrödinger equation and effective mass theorems. Communications in mathematical physics, 258(1):1–22, 2005.
- [2] J. Avron, R. Seiler, and L. Yaffe. Adiabatic theorems and applications to the quantum hall effect. Communications in Mathematical Physics, 110(1):33–49, 1987.
- [3] D. Bambusi. Reducibility of 1-d Schrödinger equation with time quasiperiodic unbounded perturbations. I. Trans. Amer. Math. Soc., 370:1823–1865, 2017.
- [4] D. Bambusi and S. Grafi. Time quasi-periodic unbounded perturbations of Schrödinger operators and KAM methods. Comm. Math. Phys., 219:465–480, 2001.
- [5] M. Bellec, C. Michel, H. Zhang, S. Tzortzakis, and P. Delplace. Non-diffracting states in one-dimensional floquet photonic topological insulators. EPL (Europhysics Letters), 119(1):14003, 2017.
- [6] E. Davies and H. Spohn. Open quantum systems with time-dependent hamiltonians and their linear response. Journal of Statistical Physics, 19(5):511–523, 1978.
- [7] E. B. Davies. Spectral theory and differential operators, volume 42. Cambridge University Press, 1996.
- [8] W. De Roeck, A. Elgart, and M. Fraas. Locobatic theorem for disordered media and validity of linear response. arXiv preprint arXiv:2203.03786, 2022.
- [9] H. Eliasson and S. Kuksin. On reducibility of Schrödinger equations with quasiperiodic in time potentials. Comm. Math. Phys., 289:125–135, 2008.
- [10] C. L. Fefferman, J. P. Lee-Thorp, and M. I. Weinstein. Topologically protected states in one-dimensional systems. Memoirs of the American Mathematical Society, 247(1173), 2017.
- [11] C. L. Fefferman and M. I. Weinstein. Honeycomb lattice potentials and Dirac points. J. Amer. Math. Soc., 25(4):1169–1220, 2012.
- [12] C. L. Fefferman and M. I. Weinstein. Wave packets in honeycomb lattice structures and two-dimensional Dirac equations. Commun. Math. Phys., 326:251–286, 2014.
- [13] R. Feola, B. Grébert, and T. Nguyen. Reducibility of Schrödinger equation on a zoll manifold with unbounded potential. Journal of Mathematical Physics, 61(7):071501, 2020.
- [14] L. Garrido. Generalized adiabatic invariance. Journal of Mathematical Physics, 5(3):355–362, 1964.
- [15] J. Guglielmon, M. C. Rechtsman, and M. I. Weinstein. Landau levels in strained two-dimensional photonic crystals. Phys. Rev. A, 103:013505, 2021.
- [16] G. A. Hagedorn and A. Joye. Elementary exponential error estimates for the adiabatic approximation. Journal of mathematical analysis and applications, 267(1):235–246, 2002.
- [17] B. C. Hall. Quantum theory for mathematicians. Springer, 2013.
- [18] S. N. Hameedi, A. Sagiv, and M. I. Weinstein. Radiative decay of edge states in floquet media. arXiv preprint arXiv:2201.11219, 2022.
- [19] M. Z. Hasan and C. L. Kane. Colloquium: Topological Insulators. Reviews of Modern Physics, 82:3045, 2010.
- [20] M. A. Hoefer and M. I. Weinstein. Defect modes and homogenization of periodic Schrödinger operators. SIAM journal on mathematical analysis, 43(2):971–996, 2011.
- [21] J. Howland. Floquet operators with singular spectrum. II. Ann. de l’I.H.P. Sec. A, 50(3):325–334, 1989.
- [22] A. Joye. Adiabatic lindbladian evolution with small dissipators. Communications in Mathematical Physics, 391(1):223–267, 2022.
- [23] M. Jürgensen, S. Mukherjee, and M. C. Rechtsman. Quantized nonlinear thouless pumping. Nature, 596(7870):63–67, 2021.
- [24] T. Kato. On the adiabatic theorem of quantum mechanics. Journal of the Physical Society of Japan, 5(6):435–439, 1950.
- [25] R. T. Keller, J. L. Marzuola, B. Osting, and M. I. Weinstein. Spectral band degeneracies of -rotationally invariant periodic schrodinger operators. Multiscale Modeling & Simulation, 16(4):1684–1731, 2018.
- [26] R. T. Keller, J. L. Marzuola, B. Osting, and M. I. Weinstein. Erratum: Spectral band degeneracies of -rotationally invariant periodic schrödinger operators. Multiscale Modeling & Simulation, 18(3):1371–1373, 2020.
- [27] J. Krieger and G. Iafrate. Time evolution of Bloch electrons in a homogeneous electric field. Physical Review B, 33(8):5494, 1986.
- [28] P. Kuchment. Floquet Theory for Partial Differential Equations, volume 60. Birkhauser, Basel, 2012.
- [29] P. Kuchment. An overview of periodic elliptic operators. Bull. Amer. Math. Soc., 53(3):343–414, 2016.
- [30] R. Montalto and M. Procesi. Linear Schrödinger equation with an almost periodic potential. SIAM Journal on Mathematical Analysis, 53(1):386–434, 2021.
- [31] G. Nenciu. On the adiabatic theorem of quantum mechanics. Journal of Physics A: Mathematical and General, 13(2):L15, 1980.
- [32] G. Nenciu. Linear adiabatic theory. exponential estimates. Communications in mathematical physics, 152(3):479–496, 1993.
- [33] T. Ozawa, H. M. Price, A. Amo, N. Goldman, M. Hafezi, L. Lu, M. C. Rechtsman, D. Schuster, J. Simon, and O. Zilberberg. Topological photonics. Reviews of Modern Physics, 91(1):015006, 2019.
- [34] A. Pazy. Semigroups of linear operators and applications to partial differential equations, volume 44. Springer Science & Business Media, 2012.
- [35] P. M. Perez-Piskunow, G. Usaj, C. A. Balseiro, and L. F. Torres. Floquet chiral edge states in graphene. Physical Review B, 89(12):121401, 2014.
- [36] M. C. Rechtsman, Y. Plotnik, J. M. Zeuner, D. Song, Z. Chen, A. Szameit, and M. Segev. Topological creation and destruction of edge states in photonic graphene. Phys. Rev. Lett., 111:103901, 2013.
- [37] M. Reed and B. Simon. II: Fourier Analysis, Self-Adjointness, volume 2. Elsevier, 1975.
- [38] M. Reed and B. Simon. Methods of Modern Mathematical Physics: Analysis of Operators, Volume IV. Academic Press, 1978.
- [39] A. Sagiv and M. I. Weinstein. Effective gaps in continuous floquet hamiltonians. SIAM Journal on Mathematical Analysis, 54(1):986–1021, 2022.
- [40] M. Taylor. Partial differential equations II: Qualitative studies of linear equations, volume 116. Springer Science & Business Media, 2013.
- [41] Y. Wang, H. Steinberg, P. Jarillo-Herrero, and N. Gedik. Observation of Floquet-Bloch states on the surface of a topological insulator. Science, 342(6157):453–457, 2013.