Arindam Mallick
[email protected]Center for Theoretical Physics of Complex Systems, Institute for Basic Science (IBS), Daejeon 34126, Korea
Nana Chang
[email protected]Center for Theoretical Physics of Complex Systems, Institute for Basic Science (IBS), Daejeon 34126, Korea
Beijing National Research Center for Information Science and Technology, Tsinghua University, Beijing 100084, China
Center for Advanced Quantum Studies, Department of Physics, Beijing Normal University, Beijing 100875, People’s Republic of China
Alexei Andreanov
[email protected]Center for Theoretical Physics of Complex Systems, Institute for Basic Science (IBS), Daejeon 34126, Korea
Basic Science Program, Korea University of Science and Technology (UST), Daejeon 34113, Korea
Sergej Flach
[email protected]Center for Theoretical Physics of Complex Systems, Institute for Basic Science (IBS), Daejeon 34126, Korea
Basic Science Program, Korea University of Science and Technology (UST), Daejeon 34113, Korea
Abstract
We consider tight-binding single particle lattice Hamiltonians which are invariant under an antiunitary antisymmetry: the anti- symmetry.
The Hermitian Hamiltonians are defined on -dimensional non-Bravais lattices.
For an odd number of sublattices, the anti- symmetry protects a flatband at energy .
We derive the anti- constraints on the Hamiltonian and use them to generate examples of generalized kagome networks in two and three lattice dimensions.
Furthermore, we show that the anti- symmetry persists in the presence of uniform DC fields
and ensures the presence of flatbands in the corresponding irreducible Wannier-Stark
band structure.
We provide examples of the Wannier-Stark band structure of generalized kagome networks in the presence of DC fields, and their implementation using Floquet engineering.
Introduction.—
Flatband systems with single particle dispersionless bands in their band structure [1, 2, 3, 4, 5, 6, 7, 8]
are important and promising platforms for exploring exotic phases and unconventional orders, due to the combined effect of macroscopic degeneracy of flatbands and applied perturbations.
Possible perturbations include disorder [9, 10, 11], nonlinear interactions [10, 12],
and various many-body interactions [13, 14, 15, 16, 17].
The presence of localized eigenstates of a flatband are argued to be useful for quantum information storage and transfer [18, 19, 20], and for observing memory effects [21].
Remarkably, the presence of a uniform DC field leads to a Wannier-Stark (WS) ladder of -dimensional irreducible band structures in a -dimensional lattice.
These irreducible band structures can again contain flatbands [22, 23].
Being fine-tuned by nature, finding flatband Hamiltonians is in general a challenging problem.
Multiple methods were developed to generate flatbands in translationally invariant systems that are based on fine-tuning [5, 24, 25], line graphs [26],
origami rules [27], repetition of miniarrays [28], and application of magnetic field [29, 6, 30, 31].
Flatbands can also emerge as a consequence of a symmetry.
Local and latent symmetries have been shown to generate flatbands [32, 33].
The other class of symmetries are global symmetries of the Hamiltonian.
A global symmetry is associated with a symmetry operator which is either unitary or antiunitary.
A single particle Hamiltonian is antisymmetric if the following relation holds: .
The antisymmetry implies that for each eigenvalue with eigenvector there exists the negative eigenvalue with eigenvector .
If the total number of eigenvalues is odd, it follows that at least one of them is zero.
Translationally invariant lattice Hamiltonians are characterized by the number of their sublattices.
Transforming the Hamiltonian into Bloch momentum space and observing results in a macroscopically degenerated symmetry-protected flatband for an odd number of sublattices.
One such example is the chiral symmetry that is realized by a unitary operator .
The chiral Hamiltonian in momentum space turns bipartite,
,
where is a null matrix and is a rectangular matrix.
Chiral flatband models and exhausting flatband generators have been reported for dimension [34, 35, 36].
In this Letter, we explore the other possibility when the symmetry operator is antiunitary and analyze the effect of the applied DC field.
Only few results are known in this case.
In , Green et al. [37] introduced a family of modified kagome lattices with three sublattices with nonzero local flux distributions which have a symmetry-protected flatband at energy despite the breaking of time-reversal symmetry.
Specific members of this modified kagome family were reported in later publications as well [38, 39].
A specific decoration of the 2D Lieb lattice was also reported to features a symmetry-protected flatband [40].
We note that all of the respective antiunitary operators consist of a spatial point reflection (inversion through a point) in lattice position space ,
followed by a time reversal operation (usually simply an antilinear complex conjugation operation in lattice position basis): .
Therefore, all of the above examples enjoy anti- Hamiltonians.
When a commensurate uniform DC field [23] is applied, the band structure of the original -dimensional Hamiltonian is modified into a Wannier-Stark ladder of irreducible -dimensional band structures with the same number of bands [41].
The particular case of the 2D dice lattice with three bands resulted in a WS flatband in the presence of a DC field, which was believed to be protected by the chiral (bipartite) symmetry of the original dice lattice [22].
However, this symmetry appears to be lost in the presence of DC fields, and the flatband existence proof in Ref. [22] does not explicitly rely on it.
However, we note that the dice lattice is also invariant under anti- symmetry.
As we show below, this symmetry remains, in general, intact in the presence of a nonzero DC field.
We derive the constraints for a general Hermitian Hamiltonian on a -dimensional non-Bravais lattice to be anti- symmetric:
(1)
The anti- symmetry condition and an odd number of sublattices are sufficient to protect at least one flatband—both in the absence and presence of DC fields.
Definitions.—
We consider a Hermitian tight-binding Hamiltonian on a -dimensional non-Bravais lattice.
Every lattice site is labeled by its unit cell index vector and sublattice index .
The numbers are integers and are -dimensional unit cell basis vectors (and are, in general, neither orthogonal nor normalized).
Similar to the lattice vector , we define the sublattice vectors which locate sublattice sites relative to a unit cell: .
Consequently, we label the Hilbert space basis vectors as .
The single-particle translationally invariant Hamiltonian reads
(2)
The hopping amplitude connects site with site .
Application of the Bloch theorem on the Hamiltonian (2) block diagonalizes it in the quasimomentum basis . Each block is a matrix acting only on the sublattice space:
(3)
If is odd, an antisymmetry of the Hamiltonian results in an zero eigenvalue.
If that antisymmetry holds for all , then the Hamiltonian possesses a zero-energy flatband.
The anti- symmetry operator reads
(4)
The one-to-one map describes the swap of the sublattice indices upon lattice point inversion.
It is defined by the lattice geometry and it is its own inverse: .
For instance, with three sublattices the only choices are , and (up to a freedom of the sublattice index relabeling).
The 2D Lieb and kagome lattices implement , while the 2D dice lattice implements .
Inversion in position space results in inverting the sign of a unit cell vector .
However, the inversion can map a given sublattice point of unit cell into one of the neighboring cell of .
Therefore, we had to introduce the lattice vectors in Eq. (4).
relates the sublattice vectors: .
The gauge phases relate to the magnetic flux distributions (if present) in the models of interest.
Since we consider an odd number of sublattices, it follows that (see Section .2 of Supplemental Material for details).
This implies the following constraints: and .
For instance, in the case of three sublattices, allows for three independent gauge phases while allows for only two independent gauge phases.
Combining Eqs. (1) and (4), we arrive at the following constraints on the hoppings for an anti- symmetric Hamiltonian (2):
(5)
The above constraint on the hoppings can be used to efficiently construct anti- symmetric Hamiltonians.
For a single sublattice (e.g., a Bravais lattice) the above condition (5) reduces to .
At the same time, the Hermiticity of the Hamiltonian enforces .
Both conditions can only be satisfied for the trivial case of no hopping .
Therefore, the anti- symmetry requires two or more sublattices.
Anti- protected flatbands.— Let us project both sides of Eq. (1) onto the -space:
(6)
For a Hamiltonian satisfying Eq. (5), the anti- operator (4) transforms as
(7)
where is a complex conjugation operator and it acts only on the sublattice space.
For an odd number of sublattices , one of the eigenvalues of is zero.
As this is true for all , it follows that one of the bands must be flat with energy equal to zero.
In Fig. 1, we show an anti- symmetric generalized 2D kagome lattice with an flatband compatible with Eq. (5).
The sublattice vectors are , , and , while , , , and .
The hopping parameters are detailed in the caption of Fig. 1.
Diagonalizing the Hamiltonian for this choice of parameters, we obtain three bands
(see Section .4 of Supplemental Material).
The anti- band structure is shown in Fig. 1.
The anti- flatband supports eigenstates which are compact localized states (CLSs) occupying three unit cells as shown in Fig. 1.
The CLS amplitudes up to normalization are (black diamonds), (black filled circle), (empty big circle),
(black filled square), and (empty square).
Figure 1: (a) The anti- symmetric generalized 2D kagome lattice.
The lattice sites are shown by small empty black circles.
A single unit cell is shown within a shaded triangle with the sublattice sites (), (), and ().
The hoppings (black dashed lines),
(black dashed-dotted lines),
(black dotted lines), (yellow solid lines), (solid black lines).
The fluxes induced by anti- symmetric complex hopping choices are denoted inside each plaquette, with all fluxes computed counter-clockwise.
The compact localized eigenstate at the anti- flatband energy has nonzero wave-function amplitudes indicated by large circles, diamonds, and squares (for more details, we refer to the main text).
(b) Band structure for and .
(c) Three subsequent irreducible Wannier-Stark band structures computed using Eq. (16) for the DC field direction .
The field strength .
To arrive at a 3D version of the kagome lattice, shown in Fig. 2,
we stack the 2D kagome lattices shown in Fig. 1 on top of each other vertically with .
Two additional hoppings connect neighboring 2D kagome planes: and .
The spectrum is now a function of three reciprocal momenta .
In Fig. 2-2, we plot three different 3D intersections of the band structure .
All of them contain an anti- flatband at zero energy.
Figure 2: The anti- symmetric generalized 3D kagome lattice.
(a)-(c) Three constrained band structures:
(a) , (b) , (c) .
(d) The lattice structure. The sites are denoted by small solid red spheres.
The hopping connections within each 2D kagome plane are the same as in Fig. 1.
The intraplane hoppings and .
(e) Three subsequent irreducible Wannier-Stark band structures computed using Eq. (16)
for the field direction .
The field strength .
Anti- protected Wannier-Stark flatbands.—
We now outline and prove the survival of the anti- symmetry in the presence of a uniform DC field for an anti- symmetric Hamiltonian.
The DC field adds an on-site potential term in the Hamiltonian (2) and the full Hamiltonian reads
(8)
Here we defined the lattice position operator as .
The DC field term changes sign under the application of the anti- operator due to lattice reflection : .
Together with condition (5), this ensures
(9)
The application of the uniform DC field breaks translation invariance and eliminates the band structure for generic directions of the DC field.
However, for special field directions, translation invariance is broken only partially and a WS band structure emerges as translation invariance is preserved in the direction orthogonal to the field.
We refer to such field directions as commensurate [23].
The unit cell and sublattice coordinates along the field, and , respectively, are defined as with the scaling factor ensuring that taking integer values.
The directions perpendicular to are parametrized by a dimensional integer vector (see Section .5 of Supplemental Material for details).
The Hamiltonian is translationally invariant in .
With the use of the Bloch basis for ,
Each block is infinite dimensional due to the coupling along the -direction.
The commensurability condition for the DC field implies the persistence of a generalized translational invariance along the field direction, which goes along with an overall shift of the eigenenergies.
Following Refs. [22, 23], we Fourier transform from -space to its conjugate momentum -space
to arrive at coupled differential equations (see Sections .6 and .7 of Supplemental Material for derivation):
(12)
The resulting Hermitian Hamiltonian
(13)
is a matrix which acts on the sublattice space only, is the hopping perpendicular to the field.
Equation (12) describes a unitary evolution of in -space,
(14)
where is a -ordered exponential of the integrated :
(15)
By construction, , where the matrix is diagonal with entries .
Then, from Eq. (14) and the above periodicity condition, we arrive at the eigenvalue problem on the WS bands:
(16)
The spectrum of is obtained by solving the above eigenproblem,
(17)
where and are the eigenvalues of the unitary matrix .
The irreducible WS band structure is obtained by choosing a particular value of , e.g., .
The entire spectrum is generated by a parallel shift of the irreducible band structure and is parametrized by the band indices .
We now arrive at the formulation of our anti- theorem in the presence of the commensurate DC field:
If the original Hamiltonian has an odd number of sublattices and is anti- symmetric, the irreducible WS band structure of contains at least one flatband.
Proof:
Indeed, the anti- condition (9) translates into a similar condition for the effective Hamiltonian ,
(18)
where the unitary matrix
(19)
is the same vector function of as in Eq. (13) but its argument is replaced by .
Then, from Eq. (15) it is straightforward to establish that
(20)
We note that by definition of the commensurate DC field direction, the projection of along the field direction will be an integer and hence will be an integer as well (see Section .8 of Supplemental Material for details).
Therefore, .
Since the operator maps the sublattice vector to , it follows that
(21)
We use the relations (20) and (21) to rewrite the eigenvalue problem (16) into the following form (see Section .8 of Supplemental Material):
(22)
Equations (16) and (22) imply that the eigenvalues of the unitary operator come in pairs .
For an odd number of sublattices , the number of eigenvalues of the operator is also odd.
Therefore, at least one eigenvalue satisfies with being independent and a multiple of .
Therefore, the irreducible WS band structure contains at least one anti- symmetry protected flatband.
We check the validity of the above theorem by computing WS band structures with (16) for the 2D kagome lattice in Fig. 1 and 3D kagome lattice in Fig. 2.
Details on the field direction and strength are provided in the corresponding captions.
We observe and confirm the presence of anti- protected WS flatbands in Fig. 1 for the 2D case and in Fig. 2 for the 3D case.
Experimental realizations.—
Flatband models have already been designed in metallic systems [43], photonic lattices [44, 45, 46, 47, 48, 49], and ultracold atoms in optical lattices [50].
The unperturbed kagome lattices introduced above can be tested in similar setups [44, 50] by proper design of hopping parameters with artificial gauge fields.
To observe the Wannier-Stark effect in optical lattices with ultracold atomic gases, they can be tilted, so the gravitational field acts as a DC field source [51].
Moreover, one can study the impact of an electric DC field on centrosymmetric lattices as reported in a very recent experiment on diamond [52].
Another option for implementing the WS Hamiltonians is to use Floquet engineering following recent experiments, which implemented Floquet Hamiltonians using ultracold atoms [53, 54].
The spectrum of WS Hamiltonians can be mapped onto that of periodically driven systems [55] and vise versa.
Mapping the frequency-space of a Floquet -dimensional lattice Hamiltonian to a new spatial dimension produces an effective static Hamiltonian in a -dimensional lattice with WS potential.
In this case, the single set of Floquet bands, which are periodic in energy, unfolds into an infinitely repeated tower of Wannier-Stark bands.
The details of the mapping between our WS Hamiltonians in 2D (3D) kagome networks and the Hamiltonians having Floquet Peierls phases in 1D (2D) diamond lattices are provided in Section .9 of the Supplemental Material.
Discussion and conclusions.—
We considered tight-binding lattice Hamiltonians on -dimensional non-Bravais lattices, which are invariant under the anti- symmetry.
We proved that the anti- symmetry protects a flatband at energy for odd numbers of sublattices.
We derived the precise anti- constraints on the Hamiltonian and used them to generate examples of generalized kagome networks.
Remarkably the anti- symmetry persists in the presence of uniform DC fields.
We prove that the corresponding irreducible Wannier-Stark band structures will again contain anti- protected flatbands.
We demonstrate the validity of our results by computing examples of the Wannier-Stark band structure of generalized 2D and 3D kagome networks in the presence of DC fields.
The zero-energy flatbands reported in Refs. [37, 38, 39, 40] belong to the anti- class.
They were reported for specific choices of hoppings for two-dimensional lattices.
Our results also explain the persistence of the flatband in the dice lattice [22] in the presence of the DC field.
The original proof relied on specific properties of the hopping network, and subsequent conjectures attempted to connect the proof to the bipartiteness of the unbiased lattice.
Actually, the unbiased dice lattice is both chiral and anti- symmetric.
Therefore, its flatband is protected by both the chiral and the anti- symmetries.
Adding a DC field destroys the chiral symmetry but preserves the anti- symmetry.
Therefore, the emerging WS flatbands in the irreducible WS band structure are protected by the anti- symmetry.
Anti- networks do not need to be bipartite and our proof is valid for any -dimension with arbitrary number of sublattices.
Our study focused on spinless single particle translationally invariant Hermitian Hamiltonians on non-Bravais lattices.
Our results also apply to a particle with an integer spin (or other internal degrees of freedom, e.g., orbital degrees of freedom) including spin-orbit coupling on a Bravais lattice.
The impact of disorder, many-body interactions, nonlinearities, or non-Hermiticity on our system are possible interesting directions for future investigations.
We expect that methods developed to analyze the impact of these perturbations for other flatband models might be helpful in our setting as well.
It is also interesting to study the case of incommensurate DC field directions that are expected to generate quasicrystalline structures.
Acknowledgments.— This work was supported by Institute for Basic Science in Korea (No. IBS-R024-D1). N.C. acknowledges financial support from the China Scholarship Council (No. CSC-201906040021).
We thank Jung-Wan Ryu for helpful discussions.
Supplemental Material: Anti- flatbands
.1 Anti- symmetry conditions for (sub)lattice vectors
We label every lattice site by its unit cell index vector and a sublattice index where is the number of sublattices.
Here is the lattice basis vector for unit cell, is the number of sublattices, and always take integer values.
Note that the vectors are complete and linearly independent, but they need not to be orthogonal nor normal.
We can express sublattice positions measured with respect to a unit cell vector , in terms of the basis vectors: with for all .
The combination uniquely identifies the lattice vector .
Therefore we can use them to label the Hilbert space basis vectors as .
The lattice shift vector can also be written as: where are integers.
The reflection of the position operator
under the antisymmetry operation reads
(23)
By definition the sublattice vectors satisfy and is always integer.
Therefore the components of the shift vector can only take values from for any .
.2 Implications of
The operator consists of a spatial reflection and a complex conjugation operation in position basis:
(24)
Therefore, when acts on an arbitrary state two times, it should return the same state up to an overall phase factor :
(25)
where we used the fact that is its own inverse: .
is the identity operator both in the sublattice space as well as in the unit cell position space.
Therefore for all
(26)
The condition is equivalent to the particle-hole symmetry requirement (see Ref. [56] for details).
The Hamiltonian can be classified topologically according to the value of , the presence of time-reversal and chiral symmetries.
If for at least one sublattice the condition Eq. (26) implies and therefore .
This happens for example, for an odd number of sublattices.
.3 Conditions on the hopping matrices under anti- symmetry of the Hamiltonian
Note that the derivation of the condition (28) did not impose any constraint on : its components can take any integer values.
Enforcing the anti- symmetry on a lattice, restricts the possible values of the components of as was described in Section .1.
However the above result suggests possible generalizations for Hamiltonians on generic graphs, provided a suitable generalization for the reflection operator is given.
.4 Anti- symmetric generalized kagome lattice
We choose the following unit cell basis vectors for the 3D kagome lattice:
(29)
where are the Cartesian orthonormal coordinate axes.
Therefore we have , ,
, , , .
The anti- symmetry constrains the hopping parameters as implied by Eq. (28)
(30)
and all the other hoppings equal zero.
We consider a special class of examples which satisfy the conditions (.4) in the main text:
(31)
Diagonalizing the Hamiltonian with these choices of parameters we obtain three bands:
(32)
Therefore this choice of hoppings supports a flatband at .
For the two dimensional kagome case we set and for the three dimensional case we set .
The corresponding flatband eigenstate forms a compact localized state occupying five (three) unit cells in 3D (2D):
(33)
.5 Parametrization of DC field coordinates
In this section we define the coordinates along and perpendicular to the DC field and present a way to parametrize these new coordinates.
We express the uniform DC field in a -dimensional lattice in terms of the unit-cell basis vectors :
(34)
We define the coordinate along the DC field as
(35)
where is a proportionality factor to be specified later.
Similar to the unit cell coordinate along the field we define the sublattice coordinate along the field:
(36)
If the field direction is commensurate [23] we can find vectors perpendicular to the field,
and all of them can be expressed through unit-cell lattice vectors where .
The corresponding orthogonality condition reads
(37)
Equation (37) is a set of degenerate linear equations for real variables with integer coefficients .
Therefore one can fix one variable, for example , and determine all the other variables
as rational numbers ,
(38)
We assumed that .
If then we can pick any other index for which —its existence is guaranteed for nonzero DC field.
Thanks to the relation (38), we can fix by requiring to be integers for all and enforcing .
We conclude that takes only integer values, furthermore because of the generalized Bézout’s identity it takes all integer values upon varying the lattice unit cell indices .
Equation (35) is rewritten as
(39)
Using the relation (39) the coordinates along are expressed as a function of the coordinate and other parameters :
(40)
where we defined .
We define a rescaled coordinate along by multiplying the above equation with a factor :
(41)
For a fixed all the allowed values of form a periodic lattice structure.
Note that unlike the perpendicular coordinates are not integer in general, and cannot be made integer by applying suitable scaling factors [23].
Because of the factor the physical distance between the two neighboring coordinates for a fixed and is in general different from the nearest neighbor distance for the original unit cell coordinate .
We use independent coordinates to parametrize the set.
Here the integer components are the equivalents of in the above expressions.
We also use these integer coordinates to label the Hilbert space basis , is a -dimensional vector with integer components.
.6 The Hamiltonian in the presence of the DC field
The tight-binding single-particle Hamiltonian in the presence of a uniform DC field is
(42)
Using Eqs. (35) and (36) the potential energy term of the Hamiltonian can be simplified as follows
(43)
The action of the hopping shifts lattice vector components from to , where we expanded the hopping vector over the unit cell basis vectors: .
Therefore the total Hamiltonian expressed in the coordinates reads as
(44)
Here is a dimensional vector which for our parametrization—e.g., for the choices made in Eqs. (39) and (41)—becomes .
A different parametrization compared to the case described in Eqs. (39) and (41) might change the form of the dimensional vectors .
For a general parametrization is a linear function of all the components .
The results presented below are valid irrespective of the choice of the parametrization.
.7 Diagonalization of the Hamiltonian in the presence of a DC field
The Hamiltonian is translationally invariant in , therefore we can apply the Bloch’s theorem and partially diagonalize it using the Fourier transform over
(45)
In the Bloch basis the Hamiltonian is block diagonal, and each block corresponds to a fixed dimensional momentum and acts on the and sublattice spaces:
(46)
The eigenvector of with eigenvalue is a function of only:
(47)
It is expressed as a linear superposition of wavevectors over different and :
This is a system of linear equations that couple different values of .
For a given this system is a set of linear coupled equations.
Because of the linearity of the system we can define the following generating function (a Fourier transformation over ):
(49)
which is possible since takes only integer values for a commensurate DC field:
This turns the eigensystem (48) into a set of coupled differential equations
(50)
Equation (.7) is a Schrödinger equation describing unitary evolution under the effective Hermitian Hamiltonian , but with time replaced by the variable .
Therefore the evolve unitarily in -space:
(51)
Here is the -ordered exponential of the integrated matrix :
Here we defined the diagonal matrix associated with the potential energy at sublattice sites:
(55)
Equation (54) is an eigenvalue equation for the operator which has possible eigenvalues .
These are functions of only (because depends on only).
At the same time, for each the total number of eigenvalues of must be the total number of values that is taken by the coordinate .
All the eigenvalues of the Hamiltonian are generated using the periodicity of :
(56)
where are the eigenvalues of the unitary matrix .
The two subscripts and of the eigenvalues identify the Wannier-Stark bands.
For any , is an irreducible band structure and carries the complete information about the band structure.
For any , can be generated from by a constant shift in energy.
.8 Robustness of anti- symmetric flatbands in the presence of a DC field
Therefore since is integer by the definition of a lattice vector, and since is integer because of the choice of the commensurate DC field direction,
we conclude that can only take integer values.
The unitary matrix is diagonal, therefore .
The dimensional vector defined in Eq. (44) is a linear function of and therefore
Comparing Eqs. (54) and (63) we conclude that within each irreducible Wannier-Stark band structure for each eigenvalue of the unitary operator there exists another eigenvalue .
If the number of sublattices per unit cell is odd, then there exists at least one eigenvalue such that , implying ,
which implies in turn a constant for all , e.g., a flatband in the presence of a DC field.
.9 Reproducing the anti- Wannier-Stark band structure with a Floquet Hamiltonian
Here we demonstrate how a band structure of an anti- symmetric Hamiltonian on a -dimensional tight-binding network can be reproduced by a Floquet Hamiltonian on a -dimensional tight-binding network with time-periodic Peierls phases.
This provides a possibility for an experimental implementations of our theory in the state-of-art setups.
For example, see the experiments in Refs. [53, 54], which implement Floquet (periodic in time) Hamiltonians using ultracold atoms.
The starting point is a driven Hamiltonian on a -dimensional non-Bravais lattice with time-periodic hopping parameters:
(64)
Here is time; -dimensional vector is a linear function of hopping vector as defined in Eq. (44);
is a potential energy at sublattice , independent of the unit cell .
The associated Hilbert space is spanned by .
The Peierls phase parameter has the same form as in Eq. (44) and consequently it is always an integer.
Therefore the time period of the Hamiltonian is ,
(65)
and is a circular frequency.
Using Floquet’s Theorem we can write the eigenfunction as a product of a time-periodic wavefunction having the same period of the Hamiltonian and an additional phase factor :
(66)
Here the time-periodic part of the wavefunction is expanded in a Fourier series.
In this case the number corresponds to frequencies, not spatial coordinates.
However, thanks to the above representation of an eigenstate, the time-dependent Schrödinger equation becomes
(67)
This is transformed into a static Schrödinger equation by looking at equations for individual Fourier components in the above equation, e.g., for fixed , and
(68)
Introducing a new extended Hilbert space spanned additionally by frequency multiples: , we can rewrite the above eigen-equation as
with the redefinition of the eigenstate in the extended Hilbert space
The above equation is identical to a time-independent Schrödinger equation for a static Hamiltonian on a -dimensional lattice with DC field having strength (see Eq. (44) for further details)
(69)
Therefore both Hamiltonians of Eqs. (64) and (69) have the same spectrum.
Note that we need the same number of sublattices per unit cell for both the -dimensional Wannier-Stark problem and the -dimensional Floquet problem.
It is important to point out, that the repeating set of Wannier-Stark bands folds into a single set of Floquet bands, which are periodic in energy.
.9.1 Example: 2D and 3D kagome lattices
In the main text we introduced the tight-binding anti- Hamiltonians for the kagome lattice and its 3D version, which have Wannier-Stark flatbands.
From the above derivation it follows that the same spectrum can be obtained for Floquet Hamiltonians on 1D and 2D diamond lattices, respectively—with an appropriate choice of hopping parameters.
We use the same parametrization of field coordinates as in the previous sections: assuming we find , .
Figure 3:
Part of the 2D diamond lattice containing six unit cells. Unit cells are shown in yellow shaded ellipses.
Square shaped sites denote the first sublattice, sphere shaped sites denote the second sublattice,
and triangle shaped sites denote the third sublattice. They are represented, respectively, by the vectors , and in the Floquet Hamiltonian (70). The potential energies at first, second and third sublattice sites are , and respectively, for any unit cell.
Hopping parameters within a unit cell: black solid thin line , black dashed line ,
black dotted line ; hopping between unit cells along axis: black dash-dotted line ,
yellow thick solid line ; hopping between unit cells along axis: blue solid line ,
blue dashed line . See the Hamiltonian in Eq. (70) for more details.
It is convenient to start with the 2D diamond lattice case.
The unit cell coordinates are indexed by = .
From the Hamiltonian (31) we can write the Floquet Hamiltonian [see Eq. (64)]
(70)
Figure 3 shows this hopping network on the 2D diamond lattice, that has the same spectrum as the 3D kagome Hamiltonian discussed in the main text.
The 1D diamond Floquet Hamiltonian can be extracted from the above Hamiltonian by eliminating the contributions along coordinate.
(71)
References
Flach et al. [2014]S. Flach, D. Leykam,
J. D. Bodyfelt, P. Matthies, and A. S. Desyatnikov, Detangling flat bands into Fano lattices, Europhys. Lett. 105, 30001 (2014).
Derzhko et al. [2015]O. Derzhko, J. Richter, and M. Maksymenko, Strongly correlated flat-band systems:
The route from heisenberg spins to hubbard electrons, Int. J. Mod. Phys. B 29, 1530007 (2015).
Leykam et al. [2018]D. Leykam, A. Andreanov, and S. Flach, Artificial flat band systems: from lattice models
to experiments, Advances in Physics: X 3, 1473052 (2018).
Maimaiti et al. [2017]W. Maimaiti, A. Andreanov,
H. C. Park, O. Gendelman, and S. Flach, Compact localized states and flat-band generators in one
dimension, Phys. Rev. B 95, 115135 (2017).
Khomeriki and Flach [2016]R. Khomeriki and S. Flach, Landau-zener bloch
oscillations with perturbed flat bands, Phys. Rev. Lett. 116, 245301 (2016).
Rhim et al. [2020]J.-W. Rhim, K. Kim, and B.-J. Yang, Quantum distance and anomalous landau levels of
flat bands, Nature 584, 59 (2020).
Goda et al. [2006]M. Goda, S. Nishino, and H. Matsuda, Inverse anderson transition caused by flatbands, Phys. Rev. Lett. 96, 126401 (2006).
Leykam et al. [2013]D. Leykam, S. Flach,
O. Bahat-Treidel, and A. S. Desyatnikov, Flat band states: Disorder and
nonlinearity, Phys. Rev. B 88, 224203 (2013).
Roy et al. [2020]N. Roy, A. Ramachandran, and A. Sharma, Interplay of disorder and interactions
in a flat-band supporting diamond chain, Phys. Rev. Research 2, 043395 (2020).
Danieli et al. [2021a]C. Danieli, A. Andreanov,
T. Mithun, and S. Flach, Nonlinear caging in all-bands-flat lattices, Phys. Rev. B 104, 085131 (2021a).
Danieli et al. [2021b]C. Danieli, A. Andreanov,
T. Mithun, and S. Flach, Quantum caging in interacting many-body all-bands-flat
lattices, Phys. Rev. B 104, 085132 (2021b).
Nunes and Smith [2020]L. H. Nunes and C. M. Smith, Flat-band superconductivity
for tight-binding electrons on a square-octagon lattice, Phys. Rev. B 101, 224514 (2020).
Orito et al. [2021]T. Orito, Y. Kuno, and I. Ichinose, Nonthermalized dynamics of flat-band many-body
localization, Phys. Rev. B 103, L060301 (2021).
Röntgen et al. [2019]M. Röntgen, C. Morfonios, I. Brouzos,
F. Diakonos, and P. Schmelcher, Quantum network transfer and storage with compact
localized states induced by local symmetries, Phys. Rev. Lett. 123, 080504 (2019).
Taie et al. [2020]S. Taie, T. Ichinose,
H. Ozawa, and Y. Takahashi, Spatial adiabatic passage of massive quantum particles in
an optical lieb lattice, Nature Communications 11, 257 (2020).
Röntgen et al. [2020]M. Röntgen, N. Palaiodimopoulos, C. Morfonios, I. Brouzos,
M. Pyzh, F. Diakonos, and P. Schmelcher, Designing pretty good state transfer via isospectral
reductions, Phys. Rev. A 101, 042304 (2020).
Lai and Chien [2016]C.-Y. Lai and C.-C. Chien, Geometry-induced memory
effects in isolated quantum systems: Cold-atom applications, Phys. Rev. Applied 5, 034001 (2016).
Kolovsky et al. [2018]A. Kolovsky, A. Ramachandran, and S. Flach, Topological flat
Wannier-Stark bands, Phys. Rev. B 97, 045120 (2018).
Mallick et al. [2021a]A. Mallick, N. Chang,
W. Maimaiti, S. Flach, and A. Andreanov, Wannier-stark flatbands in bravais lattices, Phys. Rev. Research 3, 013174 (2021a).
Maimaiti et al. [2019]W. Maimaiti, S. Flach, and A. Andreanov, Universal flat band generator
from compact localized states, Phys. Rev. B 99, 125129 (2019).
Maimaiti et al. [2021]W. Maimaiti, A. Andreanov, and S. Flach, Flat-band generator in two
dimensions, Phys. Rev. B 103, 165116 (2021).
Dias and Gouveia [2015]R. Dias and J. Gouveia, Origami rules for the construction of
localized eigenstates of the Hubbard model in decorated lattices, Sci. Rep. 5, 16852 EP (2015).
Morales-Inostroza and Vicencio [2016]L. Morales-Inostroza and R. A. Vicencio, Simple
method to construct flat-band lattices, Phys. Rev. A 94, 043831 (2016).
Möller and Cooper [2018]G. Möller and N. R. Cooper, Synthetic gauge fields for
lattices with multi-orbital unit cells: routes towards a -flux dice
lattice with flat bands, New J. Phys. 20, 073025 (2018).
Yu et al. [2020]D. Yu, L. Yuan, and X. Chen, Isolated photonic flatband with the effective
magnetic flux in a synthetic space including the frequency dimension, Laser & Photonics Reviews 14, 2000041 (2020).
Röntgen et al. [2018]M. Röntgen, C. Morfonios, and P. Schmelcher, Compact localized
states and flat bands from local symmetry partitioning, Phys. Rev. B 97, 035161 (2018).
Morfonios et al. [2021]C. Morfonios, M. Röntgen, M. Pyzh, and P. Schmelcher, Flat bands by latent symmetry, Phys. Rev. B 104, 035105 (2021).
Sutherland [1986]B. Sutherland, Localization of
electronic wave functions due to local topology, Phys. Rev. B 34, 5208 (1986).
Mur-Petit and Molina [2014]J. Mur-Petit and R. A. Molina, Chiral bound states in the
continuum, Phys. Rev. B 90, 035434 (2014).
Ramachandran et al. [2017]A. Ramachandran, A. Andreanov, and S. Flach, Chiral flat bands:
Existence, engineering, and stability, Phys. Rev. B 96, 161104(R) (2017).
Green et al. [2010]D. Green, L. Santos, and C. Chamon, Isolated flat bands and spin-1 conical bands in
two-dimensional lattices, Phys. Rev. B 82, 075104 (2010).
Koch et al. [2010]J. Koch, A. A. Houck,
K. L. Hur, and S. Girvin, Time-reversal-symmetry breaking in circuit-qed-based
photon lattices, Phys. Rev. A 82, 043811 (2010).
Wei and Sedrakyan [2021]C. Wei and T. A. Sedrakyan, Optical lattice
platform for the sachdev-ye-kitaev model, Phys. Rev. A 103, 013323 (2021).
Chen et al. [2014]L. Chen, T. Mazaheri,
A. Seidel, and X. Tang, The impossibility of exactly flat non-trivial chern bands
in strictly local periodic tight binding models, J. Phys. A: Math. Theor. 47, 152001 (2014).
Maksimov et al. [2015]D. N. Maksimov, E. N. Bulgakov, and A. R. Kolovsky, Wannier-stark states in
double-periodic lattices. ii. two-dimensional lattices, Phys. Rev. A 91, 053632 (2015).
Mallick et al. [2021b]A. Mallick, N. Chang,
A. Andreanov, and S. Flach, Supplemental material: Anti- flatbands
(2021b).
Kajiwara et al. [2016]S. Kajiwara, Y. Urade,
Y. Nakata, T. Nakanishi, and M. Kitano, Observation of a nonradiative flat band for spoof surface
plasmons in a metallic Lieb lattice, Phys. Rev. B 93, 075126 (2016).
Zong et al. [2016]Y. Zong, S. Xia, L. Tang, D. Song, Y. Hu, Y. Pei, J. Su, Y. Li, and Z. Chen, Observation of localized flat-band states in kagome photonic
lattices, Opt. Express 24, 8877 (2016).
Mukherjee and Thomson [2017]S. Mukherjee and R. R. Thomson, Observation of robust
flat-band localization in driven photonic rhombic lattices, Opt. Lett. 42, 2243 (2017).
Xia et al. [2016]S. Xia, Y. Hu, D. Song, Y. Zong, L. Tang, and Z. Chen, Demonstration of flat-band image transmission in optically induced lieb
photonic lattices, Opt. Lett. 41, 1435 (2016).
Xia et al. [2020]S. Xia, C. Danieli,
W. Yan, D. Li, S. Xia, J. Ma, H. Lu, D. Song, L. Tang, S. Flach, and Z. Chen, Observation of quincunx-shaped and dipole-like flatband states in photonic
rhombic lattices without band-touching, APL Phot. 5, 016107 (2020).
Yan et al. [2020]W. Yan, H. Zhong, D. Song, Y. Zhang, S. Xia, L. Tang, D. Leykam, and Z. Chen, Flatband line states in photonic super-honeycomb
lattices, Advanced Optical Materials 8, 1902174 (2020).
Tang et al. [2020]L. Tang, D. Song, S. Xia, S. Xia, J. Ma, W. Yan, Y. Hu, J. Xu, D. Leykam, and Z. Chen, Photonic flat-band lattices and unconventional light localization, Nanophotonics 9, 1161 (2020).
Jo et al. [2012]G.-B. Jo, J. Guzman, C. K. Thomas, P. Hosur, A. Vishwanath, and D. M. Stamper-Kurn, Ultracold atoms in a tunable optical kagome lattice, Phys. Rev. Lett. 108, 045305 (2012).
Anderson and Kasevich [1998]B. Anderson and M. Kasevich, Macroscopic Quantum
Interference from Atomic Tunnel Arrays, Science 282, 1686 (1998).
De Santis et al. [2021]L. De Santis, M. E. Trusheim, K. C. Chen, and D. R. Englund, Investigation of the stark effect on a
centrosymmetric quantum emitter in diamond, Phys. Rev. Lett. 127, 147402 (2021).
Bordia et al. [2017]P. Bordia, H. Lüschen,
U. Schneider, M. Knap, and I. Bloch, Periodically driving a many-body localized quantum system, Nature Physics 13, 460 (2017).
Chiu et al. [2016]C.-K. Chiu, J. C. Teo,
A. P. Schnyder, and S. Ryu, Classification of topological quantum matter with
symmetries, Rev. Mod. Phys. 88, 035005 (2016).