Previous Article in Journal
Comparison of Two- and Three-Phase Devices Generating a Rotating Magnetic Fieldfor Magnetic Hyperthermia Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Landau Levels and Electronic States for Pseudospin-1 Lattices with a Bandgap: Application to a Lieb Lattice

1
Department of Physics and Computer Science, Medgar Evers College of City University of New York, Brooklyn, NY 11225, USA
2
Department of Physics & Engineering Physics, Fordham University, Bronx, NY 10458, USA
3
Department of Physics and Astronomy, Hunter College of the City University of New York, 695 Park Avenue, New York, NY 10065, USA
4
Donostia International Physics Center (DIPC), P de Manuel Lardizabal, 4, 20018 San Sebastian, Basque Country, Spain
5
Space Vehicles Directorate, US Air Force Research Laboratory, Kirtland Air Force Base, Albuquerque, NM 87117, USA
*
Authors to whom correspondence should be addressed.
Magnetism 2025, 5(3), 22; https://doi.org/10.3390/magnetism5030022
Submission received: 13 August 2025 / Revised: 7 September 2025 / Accepted: 10 September 2025 / Published: 16 September 2025

Abstract

We have carried out detailed theoretical and numerical calculations and developed a physics-based model for quantitatively describing the Landau levels of several pseudospin-1 structures with a flat band and a finite bandgap in their electronic-energy spectrum under a strong and uniform magnetic field. We have investigated the Landau-level-based dynamics, as well as the corresponding eigenstates, for gapped graphene, a dice lattice with both a zero and finite bandgap and, eventually, for the Lieb lattice, which represents a separate type of square lattice with a very special non-symmetric (elevated) location of the flat band which intersects the conduction band at its lowest point. Exact analytical consideration of Landau-level states has been performed and explained when dealing with all types of considered lattices. Our model could be further generalized for treating cases with an arbitrary position for the flat band between the valence and conduction bands. Our current results have direct implications for a deep-level investigation of the quantum Hall effect, as well as other magnetic and topological properties of these novel materials.

1. Introduction

Since the groundbreaking transport measurements were reported on graphene in 2004 [1], all novel low-dimensional materials have received tremendous attention from researchers in condensed matter physics and relevant materials science fields. Later, some other structures with tunable spin-orbit bandgaps, such as buckled honeycomb lattices, were discovered. Their magneto-optical properties [2] were also studied thoroughly, including general collective-excitation modes [3,4,5,6,7,8] and massive hyperbolic plasmons [9]. Furthermore, other novel electron dispersions have been founded in both Kekule-distorted graphene [10,11,12,13] and semi-Dirac materials. Meanwhile, their optical properties [14,15] and topological electronic behaviors [16] were investigated extensively. Specifically, the optical conductivity of all these materials [12,17,18,19], including anisotropic and tilted [20] Dirac cones [21,22,23,24], were fully explored both theoretically and experimentally.
Among these recently discovered two-dimensional materials, considerable attention has been given to a unique group with a flat (or nondispersive) band in their low-energy band structure [25,26]. This flat band, which could be located arbitrarily between an upper-conduction and a lower-valance band, gives rise to infinite degeneracy for energies coinciding with those of the flat-band electrons and makes the electronic properties of this class of materials feature-rich [27]. Some typical examples of such flat-band materials include the dice, Lieb [28,29,30] and kagome lattices [31,32,33], as well as a few other important materials [34,35]. The latter materials also exhibit unique weak antilocalization and ferromagnetism in single crystalline magnetic Weyl semimetals, as well as quantum oscillation for the nodal line semimetal [36,37].
The α - T 3 model represents a very special type of two-dimensional structure which also exhibits a flat band in their low-energy spectrum [38,39]. The flat band in the α - T 3 model originates from the existence of an additional hub atom at the center of each hexagon of a graphene-like two-dimensional lattice. The hopping coefficients between the hub and rim (regular) atoms in the α - T 3 model could be parameterized by a chosen number α within the range 0 < α < 1 . Consequently, we obtain a physical model, which realistically describes a variety of materials with a flat band, and simultaneously, receives tremendous attention from researchers. Specifically, electronic structures [40,41,42], transport [43,44,45], collective optical [4], magnetic [46] and electronic phenomena in α - T 3 rings [47,48] and the properties of other two-dimensional lattice materials [27,49] have been extensively investigated.
Interestingly, the main characteristics of the electronic dispersions, such as bandgaps, group velocities and the anisotropy of graphene and other two-dimensional materials, will change significantly by applying an off-resonance electromagnetic dressing field [50,51,52,53]. In such a case, the modification of the energy spectrum depends strongly on the polarization of incident irradiation, in addition to the field frequency and strength. In fact, there has already been a number of crucial publications aiming to study the electronic, collective and transport properties of materials with a Dirac cone deformed by an optical-dressing field [54,55,56].
As a different and impactful issue, the magnetic properties of these emergent two-dimensional materials, as well as their corresponding electron dynamics under magnetic fields, have been explored extensively. These have received encouraging responses from researchers working in various fields. This includes the calculation and detailed analysis of Landau-level dispersion and electronic states under a magnetic field in graphene [57,58,59], silicene [60,61] and transition-metal dichalcogenides [62,63], covering their magneto-optical [64,65] and magneto-transport [66,67,68,69,70] properties as well. Meanwhile, various magnetic-field responses of the α - T 3 material were also studied [61]. Additionally, these research efforts further extend to magneto-plasmons and other collective behaviors in graphene materials [71] with a tunable spin-orbit bandgap [2,61,72,73] and in α - T 3 model [74,75] as well. The Hamiltonian, Landau-level spectrum and zero-mode degeneracies in pseudospin-1 systems in Ref. [76] have a direct connection with the lattice topology [77], which represents one of the first demonstrations of an α T 3 lattice. The low-energy spin-1 Hamiltonian and Landau levels in a Lieb lattice were initially addressed in Ref. [78]. The electronic states and the Landau levels for the magnetic field are the key building blocks for calculating the most important collective and transport properties of a two-dimensional material, such as magneto-conductivity, magneto-optical conductivity, Boltzmann transport coefficients and the corresponding inverse relaxation time. Specifically, the polarization function, which is the key ingredient for calculating the dielectric function and conductivity, is obtained as the sum of all possible transitions between the unoccupied and occupied electronic state (in our case, the transitions between Landau levels). This function can be calculated and evaluated theoretically once you know the electronic states in the new material. At the same time, magneto-conductivity, magneto-optical conductivity and magneto-plasmons have been experimentally measured in most known Dirac cone materials [79,80,81]. The structure of the Landau levels directly determines the Hall plateau and the Hall plateau-to-plateau transitions in many Dirac materials which directly affect a number of crucial physical quantities such as magneto-conductivity and their temperature dependence [82].
The Landau-level structure could be measured with the use of different techniques, including exploiting the Hall plateaus, which in the simplest approximation, represent each step at the energy of each consequent level. Integer or half-integer structures of the Hall plateaus are easily detectable in most infrared transmission measurements [83]. One can also directly measure the density-of-states in a new material using the so-called capacitance measurements [84].
The remainder of this paper is organized as follows: In Section 2, we present a general formalism for Dirac electrons in magnetic fields, including their energy eigenvalues (Landau levels) as well as their wave functions. Section 3 is devoted to presenting several novel derivations for some known results associated with Landau levels within gapped graphene, including computing wave functions for a case with a finite gap Δ 0 . Importantly, our most crucial results likewise stem from considering the finite bandgaps in several Dirac cone materials. In Section 4, we consider Landau quantization of electron energies and corresponding electronic wave functions within a dice lattice with both zero and finite bandgaps, accompanied by a flat band located midway between the valence and conduction bands. Following this, we derive in Section 5 the Landau levels and corresponding eigenstates of electrons within a Lieb lattice with an elevated flat band. Our final remarks and conclusions are presented in Section 6.

2. General Formalism for Landau Quantization

We begin with a brief review of the general mathematical formalism employed in calculating the energy eigenstates of free electrons under a spatially uniform magnetic field B, usually referred to as Landau levels. We shall employ the Landau gauge for the vector potential A = A x , A y , A z so that A x = B 0 y while A y = A z = 0 . Therefore, the magnetic field in this case is given by B = × A = B 0 z ^ , which is uniform in space domain.
Formally, the effect of a magnetic field B can be included by using a so-called Peierls substitution in any considered Hamiltonian, i.e.,
k x , y k x , y + e A x , y .
In the case of graphene, its Hamiltonian operator acquires the following explicit form [85]
H ^ g = 2 v F B 0 a ^ a ^ 0 ,
where B = / e B 0 is magnetic length, and two operators are defined by
a ^ = B k x + y B + i B k y , a ^ = B k x + y B i B k y ,
which are known as the creation and annihilation operators, respectively. Using the expression in Equation (3), one can further verify the well-known commutation relation a ^ , a ^ = 1 and meanwhile obtain the results for their actions on the electronic states in the Fock space, given by
a ^ | n = n | n 1 , a ^ | n = n + 1 | n + 1 .
Here, the results in Equations (1)–(4) are all the relations required to calculate the actual Landau levels and their corresponding wave functions in the Dirac cone materials considered in this study.

3. Magneto-Energy Levels of Gapped Graphene

To start, we consider the case of gapped graphene described by the following low-energy Hamiltonian
H ^ g ( 0 ) = Σ ^ x , y ( 2 ) · k + Δ 0 Σ ^ z ( 2 ) = Δ 0 v F ( k x i k y ) v F ( k x + i k y ) Δ 0 ,
which differs from well-known graphene by a finite gap 2 Δ 0 included as a Σ ^ z ( 2 ) term. Here, Σ ^ x , y , z ( 2 ) are regular 2 × 2 Pauli matrices.
In the presence of a magnetic field B = B 0 z ^ , the canonical or Peierls substitution in Equation (1) can be employed. This leads to the following new Hamiltonian
H ^ g ( B ) = Δ 0 γ B a ^ γ B a ^ Δ 0 ,
where γ B = 2 v F / B . Mathematically speaking, the simplest way to find energy eigenvalues of the Hamiltonian in Equation (6) is by simply taking the square on both sides of the eigenvalue equation. As a result, we get a new diagonal Hamiltonian matrix, written as
H ^ g 2 ( B ) = 2 v F B 2 γ B ( a ^ a ^ + 1 ) + Δ 0 2 0 0 γ B a ^ a ^ + Δ 0 2 .
By using the harmonic-oscillator relation a ^ a ^ | n = n | n , as well as the following eigenvalue relation
H ^ g 2 | Ψ n = ε n 2 | Ψ n ,
we immediately find ε n = ± 2 n γ B 2 + Δ 0 2 .
In a similar way, we will look for a solutions in the following general form
| Ψ n 1 , n 2 ( G ) = c 1 | n 1 c 2 | n 2 ,
then the corresponding eigenvalue equations become
Δ 0 c 1 | n 1 + γ B c 2 a ^ | n 2 = ε c 1 | n 1 , γ B c 1 a ^ | n 1 Δ 0 c 2 | n 2 = ε c 2 | n 2 .
Equation (10) can be equivalently written as
Δ 0 c 1 | n 1 + γ B c 2 n 2 | n 2 1 = ε c 1 | n 1 , γ B c 1 n 1 + 1 | n 1 + 1 Δ 0 c 2 | n 2 = ε c 2 | n 2 .
From Equation (11), we easily find that its solution in the Fock space exists only for n 1 = n 2 1 , or equivalently n 2 = n 1 + 1 . Now, let us assume n 2 = n and n 1 = n 1 . By following Equation (11), we acquire
Δ 0 c 1 | n 1 + γ B c 2 n | n 1 = ε c 1 | n 1 , γ B c 1 n | n Δ 0 c 2 | n = ε c 2 | n .
The presence of a physical solution to Equation (12) requires that its coefficient determinant must be zero. This leads to the following dispersion relation
2 v F 2 e B n ε n 2 + Δ 0 2 = 0 ,
which produces ε n = ± 2 v F 2 e B n + Δ 0 2 , and it is equivalent to a spectrum equation for the Landau levels of a gapped graphene. Additionally, its wave function for n 1 is found to be
| Ψ n ( Δ 0 ) = γ B n | n 1 ± γ B 2 n + Δ 0 2 Δ 0 | n .
If Δ 0 0 , on the other hand, the result of Equation (14) immediately turns into
| Ψ n > 0 ( Δ 0 0 ) = 1 2 | n 1 ± | n ,
which is reduced in the case of free-standing (zero-gap) graphene. In particular, for n = 0 , the wave function (14) reduces to
| Ψ n = 0 = 0 | 0 .
Here, we have finished the calculation of a closed-form analytical expression for both the electronic states and Landau energy levels of electrons within a gapped graphene under an external perpendicular magnetic field.

4. Dice Lattice with Zero Bandgap

As we mentioned previously in this paper, our present study focuses on investigating pseudospin-1 Dirac materials with a flat band in their low-energy spectrum. The schematics of energy dispersions in both gapped dice and Lieb lattices, as depicted in Figure 1, contain a flat band, which makes these materials and their electronic properties similar to each other. Interestingly, different positions of this flat band inside a bandgap result in a significant difference between the two. In the case of a gapped dice lattice, it sits exactly in the middle between the valence and conduction bands, separated equally from band edges by a bandgap parameter Δ 0 , and this makes the whole energy spectrum very symmetric. In another case, i.e., the Lieb lattice, the flat band is located at an elevated position, intersecting the bottom of a conduction band. Such an overlap of the flat band with the bottom of a conduction band has a lot of implications for the collective and many-body properties of electrons in these types of materials, e.g., in the computed polarization function, where all electronic transitions and various energy separations will be taken into account. Here, we focus on the role which the location of the flat-band plays in the presence of a perpendicular and spatially uniform magnetic field.
The low-energy Hamiltonian for a dice lattice with a zero-gap parameter Δ 0 = 0 (the actual bandgap is 2 Δ 0 ) can be written as
H ^ ( D ) ( k ) = v F 2 0 k e i Θ k 0 k e i Θ k 0 k e i Θ k 0 k e i Θ k 0 .
However, in the presence of a magnetic field, the Hamiltonian in Equation (17) is transformed into
H ^ ( D ) ( k ) = γ B ( D ) 0 a ^ 0 a ^ 0 a ^ 0 a ^ 0 ,
where a ^ and a ^ represent annihilation and creation operators, respectively, and γ B ( D ) = γ B ( G ) / 2 = v F / B . From now on, we will calculate the energy levels (Landau levels) of a dice lattice in cases with a zero bandgap Δ 0 = 0 .
Here, we first write down the wave function in the most general form as
| Ψ n 1 , n 2 , n 3 ( D ) = c 1 | n 1 c 2 | n 2 c 3 | n 3 ,
where c 1 , c 2 , c 3 are three coefficients to be determined. Inserting the wave function in Equation (19) into a Schrödinger equation with its Hamiltonian given by Equation (18), one obtains an eigenvalue equation as follows:
γ B ( D ) a ^ c 2 | n 2 ε c 1 | n 1 = 0 , γ B ( D ) a ^ c 1 | n 1 + γ B a ^ c 3 | n 3 ε c 2 | n 2 = 0 , γ B ( D ) a ^ c 2 | n 2 ε c 3 | n 3 = 0 ,
where parameter ε stands for the eigenvalue to be determined. After analyzing Equation (20), we know that the conditions of n 1 = n 2 1 and n 2 = n 3 1 must be satisfied in order to acquire a nonzero (non-trivial) solution for these coupled equations. Specifically, by writing n 3 = n , n 2 = n 1 and n 1 = n 2 , we find from Equation (20) that
ε c 1 + γ B ( D ) n 1 c 2 = 0 , γ B ( D ) n 1 c 1 ε c 2 + γ B ( D ) n c 3 = 0 , γ B ( D ) n c 2 ε c 3 = 0 .
After an analysis of Equation (21), it becomes clear that this system can support the nonzero solution [ c 1 , c 2 , c 3 ] T 0 only if its coefficient determinant is zero. In this way, we find the energy dispersions of a dice lattice, given by ε n = 0 for a flat band, as well as by another nonzero solution
ε n = ± γ B ( D ) 2 n 1 .
The solved wave functions associated with Equation (22) are
| Ψ n ( D ) = 1 2 ( 2 n 1 ) n 1 | n 2 ± | n 1 n | n ,
where n = 1 , 2 , 3 , . By choosing n = 1 in Equation (23) as an example, we have
| Ψ n = 1 ( D ) = 1 2 0 ± | 0 | 1 .
For the flat band ε n = 0 , on the other hand, its wave function is
| Ψ n ( D ) = 1 ( 2 n 1 ) n | n 2 0 n 1 | n ,
where n = 2 , 3 , 4 , . However, as n = 1 , the wave function is modified to
| Ψ n = 1 ( D ) = 1 2 0 0 | 0 .
The Landau levels for a zero-bandgap dice lattice, as shown in Figure 2, are actually a specific limiting case of the energy eigenstates of α - T 3 materials, which were discussed in Ref. [46]. In the presence of a magnetic field, the eigen-energies of valence and conduction bands reduce to an infinite set of quantized energy levels ± 2 n 1 , but the flat band ε n = 0 remains dispersionless with an infinite degeneracy. Consequently, we conclude that such a configuration of Landau levels displays both similarities and distinctions (a constant energy shift) in comparison with the energy dispersion ε n 2 n of graphene.

5. Gapped Dice Lattice

A dice lattice with a finite bandgap Δ 0 is obtained in terms of a 3 × 3 generalization of the Pauli matrices Σ ^ x , y , z ( 3 ) in the following way:
H ^ ( D ) ( k ) = Σ ^ x , y ( 3 ) · k + Δ 0 Σ ^ z ( 3 ) = v F 2 Δ 0 k e i Θ k 0 k e i Θ k 0 k e i Θ k 0 k e i Θ k Δ 0 ,
where
Σ ^ x ( 3 ) = 0 1 0 1 0 1 0 1 0 ,
Σ ^ y ( 3 ) = 0 i 0 i 0 i 0 i 0 ,
which are supplemented by an extra bandgap-related matrix, i.e.,
Σ ^ z ( 3 ) = 1 0 0 0 0 0 0 0 1 .
In the presence of a magnetic field, the Hamiltonian in Equation (27) is transformed into a form similar to the expression in Equation (18), and written as
H ^ ( D ) ( k ) = γ B ( D ) Δ 0 a ^ 0 a ^ 0 a ^ 0 a ^ Δ 0 .
In correspondence with Equation (21) having Δ 0 = 0 , for the current case, we find
( Δ 0 + ε ) c 1 + γ B ( D ) n 1 c 2 = 0 γ B ( D ) n 1 c 1 ε c 2 + γ B ( D ) n c 3 = 0 γ B ( D ) n c 2 + ( Δ 0 ε ) c 3 = 0 .
Compared with the case of Δ 0 = 0 , the finite value of Δ 0 here results in a quite different asymmetric dispersion equation, i.e.,
γ B ( D ) 2 Δ 0 γ B ( D ) 2 ε n + 2 n γ B ( D ) 2 ε n + Δ 0 2 ε n ε n 3 = 0 ,
which cannot be solved in any straightforward way. However, it could be solved by utilizing an analogy to a trigonometrical expansion formula, as discussed in Ref. [42].
Equation (33) could also be solved approximately by using a perturbative approach. The comparatively basic problem with a completely analytical solution is a case of zero bandgap. Therefore, the perturbation equations result in expressions for the energy eigenvalues with good precision for the small bandgap. Specifically, it should be small compared to the main characteristic energy given by the coefficient γ B ( L ) for the specific material under consideration.
The linear term of the perturbation expansion is obtained as follows:
ε n = ε n ( 0 ) + ε n ( 1 ) Δ 0 + ε n ( 2 ) Δ 0 2 + ,
where the bandgap Δ 0 is regarded as a small parameter for this expansion. Interestingly, for the flat band with ε n ( 0 ) = 0 , one obtains ε n ( 1 ) = 1 / ( 2 n 1 ) , and thus
ε n = Δ 0 2 n 1 .
We have obtained numerically a set of discrete Landau levels which are presented in Figure 3. We first notice that the valence and conduction bands are not exactly symmetric with respect to zero energy and no simple analytical formula or approximate expression exists for describing the energy locations of these levels. For the high-energy states with n 1 , the energy eigenstates are well approximated by expressions ± 2 n 1 , implying that the effect due to bandgap Δ 0 is decreased. This further supports the validity of a simple but efficient application of the WKB approximation for this case. Here, the flat band is no longer dispersionless and its degeneracy is lifted. On the other hand, we find an infinite set of separated non-degenerate Landau levels within a negative-energy ε n < 0 region. For large n 1 , the effect of an energy gap becomes negligible and these levels approach the ε n = 0 level.
We obtained a precise numerical solution and also applied perturbation theory to validate our closed-form analytical expression which, of course, had limited precision. We note that this analytical formula reproduces the most important features of the obtained exact results: it is negative (lies below the zero energy level) and is proportional to the gap. This means that it would be equal to zero for Δ 0 0 .
Next, we focus on the dependence of energy levels for a gapped dice lattice on the gap parameter Δ 0 , and present various numerically computed results in Figure 4 for a full comparison among Landau levels in the conduction, valence and flat bands all together. As found from Figure 4, these numerical results once again agree with our previous findings that the high-energy Landau levels with a large level index n ( n 1 ) will be much less sensitive to the size of bandgap Δ 0 , which can be easily verified by the fact that large-n Landau levels would like to group together as Δ 0 increases in the flat-band case. In particular, the n = 1 Landau level in panel (b) demonstrates a negative linear dependence on Δ 0 , which fully agrees with our perturbation-based solution in Equation (35).

6. Modeling Lieb Lattice with Elevated Flat Band

We now turn our attention to the major issue of this paper, i.e., modeling the effect of a magnetic field on a Lieb lattice with a non-symmetric elevated flat band, as depicted in Figure 1b. In this case, the low-energy Hamiltonian is found to be [24]
H ^ ( L ) ( k | k Δ ) = v F k Δ k x 0 k x k Δ k y 0 k y k Δ ,
where the following substitution
k x , y π a 0 + k x , y
is required to obtain the low-energy spectrum for a Lieb lattice. Here, a 0 represents the lattice parameter and before making the substitution in Equation (37) into Equation (36), the three energy dispersions can be found as eigenvalues of the Hamiltonian in Equation (36) as ε γ = ± 1 ( L ) ( k | k Δ ) / v F = γ k Δ 2 + k x 2 + k y 2 and ε γ = 0 ( L ) ( k | k Δ ) / v F = k Δ , which could be combined into a single expression, yielding
ε γ ( L ) ( k | k Δ ) = v F δ γ , 0 k Δ + γ ( 1 δ γ , 0 ) k Δ 2 + k 2 ,
where γ = 0 , ± 1 and δ γ , 0 is the Kronecker delta symbol. Its nonzero values demonstrate one of the two types of energy eigenvalues: dispersive valence and conduction bands for γ = ± 1 and the flat band for γ = 0 .
In the presence of a magnetic field, however, the two components of the wave vector are modified as k x , y k x , y + r / A x , y . Therefore, we obtain that 2 k x = k + + k and 2 i k y = k + k and we can present the Hamiltonian in Equation (36) in terms of the substitutions calculated earlier for quantities k + and k . As a result, in the presence of a magnetic field, we obtain
H ^ ( L ) ( k | k Δ ) = Δ 0 γ B ( L ) 2 a ^ + a ^ 0 γ B ( L ) 2 a ^ + a ^ Δ 0 γ B ( L ) 2 i a ^ a ^ 0 γ B ( L ) 2 i a ^ a ^ Δ 0 .
Here, it is straightforward to verify that one cannot project this Hamiltonian onto the same Fock state | n representation. However, by searching for the eigen-function in terms of Equation (39), we obtain one of its eigenvalue equations, given by
Δ 0 c 1 | n 1 + γ B ( L ) 2 n 2 c 2 | n 2 1 + n 2 + 1 c 2 | n 2 + 1 = ε n c 1 | n 1 .
For the considered model, each of the three equations corresponding to the three components of the wave functions contain non-equivalent Fock states | n 1 and | n 2 . Therefore, Equation (40) cannot be satisfied by any | n 1 and | n 2 states.
On the other hand, we are also able to model the Lieb lattice by an elevated flat band, as illustrated in Figure 1. For this purpose, let us first write down the Hamiltonian, using an elevated flat band, as
H ^ ( L ) ( k ) = Δ 0 γ B ( L ) k 2 e i Θ k 0 γ B ( L ) k 2 e i Θ k Δ 0 γ B ( L ) k 2 e i Θ k 0 γ B ( L ) k 2 e i Θ k Δ 0 ,
which results in two energy dispersions, given by ε γ ( L ) ( k ) = Δ 0 as well as ε γ ( L ) ( k ) = γ k 2 + Δ 0 2 . Furthermore, in the presence of a magnetic field, we find that the Hamiltonian in Equation (39) changes into
H ^ ( L ) ( k ) = γ B ( L ) Δ 0 a ^ 0 a ^ Δ 0 a ^ 0 a ^ Δ 0 .
Our model Hamiltonian (42) still carries the most important features of the initially considered Hamiltonian of the Lieb lattice and faithfully reproduces the most crucial observables of the investigated quantum states. That is, the energy band structure and the density-of-states. This is directly verified by comparing both types of eigenvalue equations in k-space in the absence of a magnetic field.
Correspondingly, the previous Equation (21) for a dice lattice will be changed into
( Δ 0 ε ) c 1 + γ B ( L ) n 1 c 2 = 0 , γ B ( L ) n 1 c 1 ( Δ 0 + ε ) c 2 + γ B ( D ) n c 3 = 0 , γ B ( L ) n c 2 + ( Δ 0 ε ) c 3 = 0 .
Equation (43) leads to the eigenvalue equation as
γ B ( L ) 2 Δ 0 2 n γ B ( L ) 2 Δ 0 Δ 0 3 γ B ( L ) 2 ε n + 2 n γ B ( L ) 2 ε n + Δ 0 2 ε n + Δ 0 ε n 2 ε n 3 = 0 ,
which gives rise to the energy levels by
ε n = Δ 0 ,
ε n = ± γ B ( L ) 2 ( 2 n 1 ) + Δ 0 2 .
Here, the most interesting thing is the wave function corresponding to the flat band with ε n = Δ 0 . In fact, from Equation (43), we immediately conclude that c 2 = 0 . Thus, this leaves us with
n 1 c 1 = n c 3 ,
and the corresponding wave function is given by
| Ψ n ( Δ 0 ) = 1 ( 2 n 1 ) n | n 2 0 n 1 | n ,
where n = 2 , 3 , 4 , . For n = 1 , especially, this wave function is taken as
| Ψ n = 1 ( Δ 0 ) = 1 2 0 0 | 0 .
From Equations (48) and (49), we quickly find that they are identical to Equations (25) and (26) in the case of a zero-gap dice lattice.
Finally, we calculate and plot the Landau levels for our model for a Lieb lattice obtained by relatively simple analytical expressions in Equations (45) and (46). Our results are presented in Figure 5 and Figure 6. From these two figures, we find that these energy levels, corresponding to the valence and conduction bands, become symmetric with respect to the zero energy. Similarly to the previously considered case for a gapped dice lattice, the effect of an energy bandgap diminishes for higher Landau levels with n 1 . Meanwhile, the flat band remains dispersionless with a simple shift to the gap level ε n = Δ 0 and remains infinitely degenerate. The dependence of these energy levels on the gap Δ 0 , as shown in Figure 6, clearly demonstrates a linear dependence ε n = Δ 0 , as well as a smooth and monotonic dependence of all energy levels, corresponding to the valence and conduction bands, on the gap parameter Δ 0 .

7. Summary and Remarks

In conclusion, we note that the main thrust of this paper is to present a thorough theoretical and numerical investigation of the quantized energy levels under a uniform external magnetic field (i.e., Landau levels) and the corresponding electronic states for several important Dirac cone materials with a flat band in their low-energy spectra. The key research focus of this paper has been an investigation of how the combination of a flat band and a finite bandgap affects the energy levels and electronic eigenstates in the presence of a strong magnetic field. Specifically, we were interested to find out how the position of the flat band affects the magnetic field properties of a two-dimensional lattice since in the absence of the gap a specific set of degenerate Landau levels are related to the flat band. However, the situation changes significantly in the presence of the gap and depends on where exactly this flat band is located inside the bandgap.
First, we considered a well-known case of graphene with an energy bandgap and provided new derivations for Landau quantization in this system, given by ( v F k ) 2 + Δ 0 2 . Surprisingly, the square Hamiltonian of graphene with a finite bandgap under a magnetic field reduced to a diagonal matrix, which immediately gave rise to the corresponding energy eigenvalues, similarly to the case of an intrinsic graphene with a zero bandgap.
Next, we investigated the electronic states under a magnetic field in a dice lattice with a finite bandgap, which was shown as a limiting case of the general α - T 3 model. However, the presence of a finite gap turned the eigenvalue equations into a non-trivial cubic equation for computing Landau levels in this system. This cubic equation could be solved either numerically or by using an analogy between this expression and a well-known trigonometric equation, or even a perturbation theory for a small bandgap parameter Δ 0 . The result is in stark contrast with regular energy dispersions (in the absence of a magnetic field), which could be easily found from a simple algebraic equation.
Finally, we investigated the Landau levels for a realistic model of a Lieb lattice with an elevated flat band intersecting the conduction band at its lowest point. This model has both common features and differences from the dice lattice and the α - T 3 model. The main interest of such physics models is the presence of a flat band within the bandgap region, which greatly affects the electronic, magnetic and collective properties of the material. More importantly, the location of this flat band can be quite unique for the Lieb lattice, which results in substantially reduced symmetry of its low-energy spectrum. We have calculated the Landau levels of this type of energy dispersion and obtained a relatively simple analytical result 2 n γ B ( L ) 2 + Δ 0 2 , which provides a quantitative description of its dynamical feature under a quantizing magnetic field. Meanwhile, we have also calculated the wave function corresponding to the flat band and compared its similarity with other wave functions in a dice lattice.
Generally speaking, magnetic quantization as well as electronic and collective properties under a strong magnetic field, including magneto-transport and magneto-plasmon properties, are crucial for understanding the fundamental physics of any new material, such as the quantum Hall effect. We have calculated new types of energy eigenvalues and corresponding wave functions. We strongly believe that these findings for several novel materials will inevitably open up a number of opportunities for researchers to investigate their various properties: electronic, optical, collective, magnetic and quantum transport. Therefore, we believe that our work and the obtained results have a lot of potential to be expanded upon in terms of their application to specific electronic states. Our theoretical model, novel closed-form analytical expressions and numerical results are believed to become a crucial advancement in developing next-level novel electronic nanodevices.

Author Contributions

Conceptualization, LJ., L.Z., A.I., G.G. and D.H.; Methodology, L.J. and L.Z.; Software, L.Z. and A.I.; Validation, L.J., G.G. and D.H.; Formal analysis, L.J., L.Z., A.I. and G.G.; Investigation, A.I., G.G. and D.H. All authors have read and agreed to the published version of the manuscript.

Funding

A.I. was supported by the funding received from TradB-56-75, PSC-CUNY Award # 68386-00 56. G.G. gratefully acknowledges funding from the U.S. National Aeronautics and Space Administration (NASA) via the NASA-Hunter College Center for Advanced Energy Storage for Space under cooperative agreement 80NSSC24M0177. D.H. would like to acknowledge the Air Force Office of Scientific Research (AFOSR) and state that the views expressed are those of the authors and do not reflect the official guidance or position of the United States Government, the Department of Defense or the United States Air Force. We acknowledge the support from CUNY Research Scholars Program (CRSP); Lovely Joseph received a fellowship from CRSP.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Geim, A.K.; Novoselov, K.S. The rise of graphene. Nat. Mater. 2007, 6, 183–191. [Google Scholar] [CrossRef]
  2. Tabert, C.J.; Nicol, E.J. Valley-spin polarization in the magneto-optical response of silicene and other similar 2D crystals. Phys. Rev. Lett. 2013, 110, 197402. [Google Scholar] [CrossRef]
  3. Malcolm, J.; Nicol, E. Frequency-dependent polarizability, plasmons, and screening in the two-dimensional pseudospin-1 dice lattice. Phys. Rev. B 2016, 93, 165433. [Google Scholar] [CrossRef]
  4. Iurov, A.; Zhemchuzhna, L.; Gumbs, G.; Huang, D.; Fekete, P.; Anwar, F.; Dahal, D.; Weekes, N. Tailoring plasmon excitations in α-T3 armchair nanoribbons. Sci. Rep. 2021, 11, 20577. [Google Scholar] [CrossRef] [PubMed]
  5. Apalkov, V.; Wang, X.F.; Chakraborty, T. Collective excitations of Dirac electrons in graphene. Int. J. Mod. Phys. B 2007, 21, 1165–1179. [Google Scholar] [CrossRef]
  6. Gumbs, G.; Balassis, A.; Iurov, A.; Fekete, P. Strongly localized image states of spherical graphitic particles. Sci. World J. 2014, 2014, 726303. [Google Scholar] [CrossRef]
  7. Guinea, F.; Le Doussal, P.; Wiese, K.J. Collective excitations in a large-d model for graphene. Phys. Rev. B 2014, 89, 125428. [Google Scholar] [CrossRef]
  8. Rodin, A.; Trushin, M.; Carvalho, A.; Castro Neto, A. Collective excitations in 2D materials. Nat. Rev. Phys. 2020, 2, 524–537. [Google Scholar] [CrossRef]
  9. Mojarro, M.; Carrillo-Bastos, R.; Maytorena, J.A. Hyperbolic plasmons in massive tilted two-dimensional Dirac materials. Phys. Rev. B 2022, 105, L201408. [Google Scholar] [CrossRef]
  10. Andrade, E.; Carrillo-Bastos, R.; Naumis, G.G. Valley engineering by strain in Kekulé-distorted graphene. Phys. Rev. B 2019, 99, 035411. [Google Scholar] [CrossRef]
  11. Andrade, E.; Carrillo-Bastos, R.; Naumis, G.G. Topical Review: Electronic and optical properties of Kekulé and other short wavelength spatialmodulated textures of graphene. J. Phys. Condens. Matter 2025, 37, 193003. [Google Scholar] [CrossRef]
  12. Herrera, S.A.; Naumis, G.G. Electronic and optical conductivity of kekulé-patterned graphene: Intravalley and intervalley transport. Phys. Rev. B 2020, 101, 205413. [Google Scholar] [CrossRef]
  13. Iurov, A.; Zhemchuzhna, L.; Gumbs, G.; Huang, D. Application of the WKB theory to investigate electron tunneling in Kek-Y graphene. Appl. Sci. 2023, 13, 6095. [Google Scholar] [CrossRef]
  14. Carbotte, J.; Bryenton, K.; Nicol, E. Optical properties of a semi-Dirac material. Phys. Rev. B 2019, 99, 115406. [Google Scholar] [CrossRef]
  15. Mojarro, M.; Carrillo-Bastos, R.; Maytorena, J.A. Optical properties of massive anisotropic tilted Dirac systems. Phys. Rev. B 2021, 103, 165415. [Google Scholar] [CrossRef]
  16. Mondal, S.; Ganguly, S.; Basu, S. Topology and applications of 2D Dirac and semi-Dirac materials. Phys. Sci. Rev. 2022, 9, 497–527. [Google Scholar] [CrossRef]
  17. Xiong, Q.Y.; Ba, J.Y.; Duan, H.J.; Deng, M.X.; Wang, Y.M.; Wang, R.Q. Optical conductivity and polarization rotation of type-II semi-Dirac materials. Phys. Rev. B 2023, 107, 155150. [Google Scholar] [CrossRef]
  18. Stauber, T.; San-Jose, P.; Brey, L. Optical conductivity, Drude weight and plasmons in twisted graphene bilayers. New J. Phys. 2013, 15, 113050. [Google Scholar] [CrossRef]
  19. Dey, B.; Ghosh, T.K. Dynamical polarization, optical conductivity and plasmon mode of a linear triple component fermionic system. J. Phys. Condens. Matter 2022, 34, 255701. [Google Scholar] [CrossRef]
  20. Mojarro, M.; Carrillo-Bastos, R.; Maytorena, J.A. Thermal difference reflectivity of tilted two-dimensional Dirac materials. Phys. Rev. B 2023, 108, L161401. [Google Scholar] [CrossRef]
  21. Tan, C.Y.; Yan, C.X.; Zhao, Y.H.; Guo, H.; Chang, H.R. Anisotropic longitudinal optical conductivities of tilted Dirac bands in 1T-MoS2. Phys. Rev. B 2021, 103, 125425. [Google Scholar] [CrossRef]
  22. Wild, A.; Mariani, E.; Portnoi, M.E. Optical absorption in two-dimensional materials with tilted Dirac cones. Phys. Rev. B 2022, 105, 205306. [Google Scholar] [CrossRef]
  23. Iurov, A.; Zhemchuzhna, L.; Gumbs, G.; Huang, D. Optical conductivity of gapped α-T3 materials with a deformed flat band. Phys. Rev. B 2023, 107, 195137. [Google Scholar] [CrossRef]
  24. Oriekhov, D.; Gusynin, V. Optical conductivity of semi-Dirac and pseudospin-1 models: Zitterbewegung approach. Phys. Rev. B 2022, 106, 115143. [Google Scholar] [CrossRef]
  25. Regnault, N.; Xu, Y.; Li, M.R.; Ma, D.S.; Jovanovic, M.; Yazdani, A.; Parkin, S.S.; Felser, C.; Schoop, L.M.; Ong, N.P.; et al. Catalogue of flat-band stoichiometric materials. Nature 2022, 603, 824–828. [Google Scholar] [CrossRef] [PubMed]
  26. Hase, I.; Yanagisawa, T.; Kawashima, K. Computational design of flat-band material. Nanoscale Res. Lett. 2018, 13, 63. [Google Scholar] [CrossRef]
  27. Checkelsky, J.G.; Bernevig, B.A.; Coleman, P.; Si, Q.; Paschen, S. Flat bands, strange metals and the Kondo effect. Nat. Rev. Mater. 2024, 9, 509–526. [Google Scholar] [CrossRef]
  28. Slot, M.R.; Gardenier, T.S.; Jacobse, P.H.; Van Miert, G.C.; Kempkes, S.N.; Zevenhuizen, S.J.; Smith, C.M.; Vanmaekelbergh, D.; Swart, I. Experimental realization and characterization of an electronic Lieb lattice. Nat. Phys. 2017, 13, 672–676. [Google Scholar] [CrossRef]
  29. Mukherjee, S.; Spracklen, A.; Choudhury, D.; Goldman, N.; Ohberg, P.; Andersson, E.; Thomson, R.R. Observation of a localized flat-band state in a photonic Lieb lattice. Phys. Rev. Lett. 2015, 114, 245504. [Google Scholar] [CrossRef]
  30. Qiu, W.X.; Li, S.; Gao, J.H.; Zhou, Y.; Zhang, F.C. Designing an artificial Lieb lattice on a metal surface. Phys. Rev. B 2016, 94, 241409. [Google Scholar] [CrossRef]
  31. Lee, J.H.; Kim, G.W.; Song, I.; Kim, Y.; Lee, Y.; Yoo, S.J.; Cho, D.Y.; Rhim, J.W.; Jung, J.; Kim, G.; et al. Atomically Thin Two-Dimensional Kagome Flat Band on the Silicon Surface. ACS Nano 2024, 18, 25535–25541. [Google Scholar] [CrossRef]
  32. Guo, H.M.; Franz, M. Topological insulator on the kagome lattice. Phys. Rev. B Condens. Matter Mater. Phys. 2009, 80, 113102. [Google Scholar] [CrossRef]
  33. Wang, Q.; Lei, H.; Qi, Y.; Felser, C. Topological quantum materials with kagome lattice. Accounts Mater. Res. 2024, 5, 786–796. [Google Scholar] [CrossRef]
  34. Tang, L.; Song, D.; Xia, S.; Xia, S.; Ma, J.; Yan, W.; Hu, Y.; Xu, J.; Leykam, D.; Chen, Z. Photonic flat-band lattices and unconventional light localization. Nanophotonics 2020, 9, 1161–1176. [Google Scholar] [CrossRef]
  35. Nishino, S.; Goda, M. Three-dimensional flat-band models. J. Phys. Soc. Jpn. 2005, 74, 393–400. [Google Scholar] [CrossRef]
  36. Sharma, M.; Chhetri, S.K.; Acharya, G.; Graf, D.; Upreti, D.; Dahal, S.; Nabi, M.R.U.; Rahman, S.; Sakon, J.; Churchill, H.O.; et al. Quantum oscillation studies of the nodal line semimetal Ni3In2S2-xSex. Acta Mater. 2025, 289, 120884. [Google Scholar] [CrossRef]
  37. Kumar, K.; Sharma, M.; Awana, V. Weak antilocalization and ferromagnetism in magnetic Weyl semimetal Co3Sn2S2. J. Appl. Phys. 2023, 133, 023901. [Google Scholar] [CrossRef]
  38. Raoux, A.; Morigi, M.; Fuchs, J.N.; Piéchon, F.; Montambaux, G. From dia-to paramagnetic orbital susceptibility of massless fermions. Phys. Rev. Lett. 2014, 112, 026402. [Google Scholar] [CrossRef]
  39. Illes, E. Properties of the α-T3 Model. Ph.D. Thesis, University of Guelph, Guelph, ON, Canada, 2017. [Google Scholar]
  40. Iurov, A.; Gumbs, G.; Huang, D. Peculiar electronic states, symmetries, and berry phases in irradiated α-T3 materials. Phys. Rev. B 2019, 99, 205135. [Google Scholar] [CrossRef]
  41. Tamang, L.; Biswas, T. Probing topological signatures in an optically driven α-T3 lattice. Phys. Rev. B 2023, 107, 085408. [Google Scholar] [CrossRef]
  42. Dey, B.; Ghosh, T.K. Floquet topological phase transition in the α-T3 lattice. Phys. Rev. B 2019, 99, 205429. [Google Scholar] [CrossRef]
  43. Urban, D.F.; Bercioux, D.; Wimmer, M.; Häusler, W. Barrier transmission of Dirac-like pseudospin-one particles. Phys. Rev. B Condens. Matter Mater. Phys. 2011, 84, 115136. [Google Scholar] [CrossRef]
  44. Illes, E.; Nicol, E. Klein tunneling in the α-T3 model. Phys. Rev. B 2017, 95, 235432. [Google Scholar] [CrossRef]
  45. Ding, Z.; Wang, D.; Li, M.; Tao, Y.; Wang, J. Spin-resolved and charge Josephson diode effects in α-T3 lattice junctions. Phys. Rev. B 2024, 110, 155405. [Google Scholar] [CrossRef]
  46. Illes, E.; Nicol, E. Magnetic properties of the α-T3 model: Magneto-optical conductivity and the Hofstadter butterfly. Phys. Rev. B 2016, 94, 125435. [Google Scholar] [CrossRef]
  47. Islam, M.; Biswas, T.; Basu, S. Effect of magnetic field on the electronic properties of an α-T3 ring. Phys. Rev. B 2023, 108, 085423. [Google Scholar] [CrossRef]
  48. Islam, M.; Basu, S. Spin and charge persistent currents in a Kane Mele α-T3 quantum ring. J. Phys. Condens. Matter 2023, 36, 135301. [Google Scholar] [CrossRef]
  49. Leykam, D.; Andreanov, A.; Flach, S. Artificial flat band systems: From lattice models to experiments. Adv. Phys. X 2018, 3, 1473052. [Google Scholar] [CrossRef]
  50. Morina, S.; Kibis, O.; Pervishko, A.; Shelykh, I. Transport properties of a two-dimensional electron gas dressed by light. Phys. Rev. B 2015, 91, 155312. [Google Scholar] [CrossRef]
  51. Kibis, O.; Dini, K.; Iorsh, I.; Shelykh, I. All-optical band engineering of gapped Dirac materials. Phys. Rev. B 2017, 95, 125401. [Google Scholar] [CrossRef]
  52. Iurov, A.; Zhemchuzhna, L.; Gumbs, G.; Huang, D.; Tse, W.K.; Blaise, K.; Ejiogu, C. Floquet engineering of tilted and gapped Dirac bandstructure in 1T-MoS2. Sci. Rep. 2022, 12, 21348. [Google Scholar] [CrossRef]
  53. Kibis, O. Metal-insulator transition in graphene induced by circularly polarized photons. Phys. Rev. B 2010, 81, 165433. [Google Scholar] [CrossRef]
  54. Iurov, A.; Mattis, M.; Zhemchuzhna, L.; Gumbs, G.; Huang, D. Floquet Modification of the Bandgaps and Energy Spectrum in Flat-Band Pseudospin-1 Dirac Materials. Appl. Sci. 2024, 14, 6027. [Google Scholar] [CrossRef]
  55. Kristinsson, K.; Kibis, O.V.; Morina, S.; Shelykh, I.A. Control of electronic transport in graphene by electromagnetic dressing. Sci. Rep. 2016, 6, 20082. [Google Scholar] [CrossRef]
  56. Iurov, A.; Gumbs, G.; Huang, D. Exchange and correlation energies in silicene illuminated by circularly polarized light. J. Mod. Opt. 2017, 64, 913–920. [Google Scholar] [CrossRef]
  57. Ho, J.; Lai, Y.; Chiu, Y.H.; Lin, M.F. Landau levels in graphene. Phys. E Low Dimens. Syst. Nanostruct. 2008, 40, 1722–1725. [Google Scholar] [CrossRef]
  58. Guinea, F.; Castro Neto, A.; Peres, N. Electronic states and Landau levels in graphene stacks. Phys. Rev. B Condens. Matter Mater. Phys. 2006, 73, 245426. [Google Scholar] [CrossRef]
  59. Gumbs, G.; Iurov, A.; Huang, D.; Zhemchuzhna, L. Revealing Hofstadter spectrum for graphene in a periodic potential. Phys. Rev. B 2014, 89, 241407. [Google Scholar] [CrossRef]
  60. Ezawa, M. Quantum Hall effects in silicene. J. Phys. Soc. Jpn. 2012, 81, 064705. [Google Scholar] [CrossRef]
  61. Tabert, C.J.; Nicol, E.J. Magneto-optical conductivity of silicene and other buckled honeycomb lattices. Phys. Rev. B Condens. Matter Mater. Phys. 2013, 88, 085434. [Google Scholar] [CrossRef]
  62. Kuc, A.; Heine, T. The electronic structure calculations of two-dimensional transition-metal dichalcogenides in the presence of external electric and magnetic fields. Chem. Soc. Rev. 2015, 44, 2603–2614. [Google Scholar] [CrossRef]
  63. He, W.Y.; Zhou, B.T.; He, J.J.; Yuan, N.F.; Zhang, T.; Law, K.T. Magnetic field driven nodal topological superconductivity in monolayer transition metal dichalcogenides. Commun. Phys. 2018, 1, 40. [Google Scholar] [CrossRef]
  64. Malcolm, J.D.; Nicol, E.J. Magneto-optics of massless Kane fermions: Role of the flat band and unusual Berry phase. Phys. Rev. B 2015, 92, 035118. [Google Scholar] [CrossRef]
  65. Tabert, C.; Carbotte, J.; Nicol, E. Optical and transport properties in three-dimensional Dirac and Weyl semimetals. Phys. Rev. B 2016, 93, 085426. [Google Scholar] [CrossRef]
  66. Rappoport, T.G.; Uchoa, B.; Castro Neto, A. Magnetism and magnetotransport in disordered graphene. Phys. Rev. B Condensed Matter Mater. Phys. 2009, 80, 245408. [Google Scholar] [CrossRef]
  67. Shakouri, K.; Vasilopoulos, P.; Vargiamidis, V.; Peeters, F. Spin-and valley-dependent magnetotransport in periodically modulated silicene. Phys. Rev. B 2014, 90, 125444. [Google Scholar] [CrossRef]
  68. Islam, S.F. Magnetotransport properties of 8-Pmmn borophene: Effects of Hall field and strain. J. Phys. Condens. Matter 2018, 30, 275301. [Google Scholar] [CrossRef]
  69. Islam, S.F.; Dutta, P. Valley-polarized magnetoconductivity and particle-hole symmetry breaking in a periodically modulated α-T3 lattice. Phys. Rev. B 2017, 96, 045418. [Google Scholar] [CrossRef]
  70. Nguyen, H.T.; Dinh, L.; Vu, T.V.; Hoa, L.T.; Hieu, N.N.; Nguyen, C.V.; Nguyen, H.V.; Kubakaddi, S.; Phuc, H.V. Quantum magnetotransport properties of silicene: Influence of the acoustic phonon correction. Phys. Rev. B 2021, 104, 075445. [Google Scholar] [CrossRef]
  71. Tymchenko, M.; Nikitin, A.Y.; Martin-Moreno, L. Faraday rotation due to excitation of magnetoplasmons in graphene microribbons. ACS Nano 2013, 7, 9780–9787. [Google Scholar] [CrossRef]
  72. Do, T.N.; Gumbs, G.; Shih, P.H.; Huang, D.; Chiu, C.W.; Chen, C.Y.; Lin, M.F. Peculiar optical properties of bilayer silicene under the influence of external electric and magnetic fields. Sci. Rep. 2019, 9, 624. [Google Scholar] [CrossRef]
  73. Tahir, M.; Vasilopoulos, P. Electrically tunable magnetoplasmons in a monolayer of silicene or germanene. J. Phys. Condens. Matter 2015, 27, 075303. [Google Scholar] [CrossRef]
  74. Balassis, A.; Dahal, D.; Gumbs, G.; Iurov, A.; Huang, D.; Roslyak, O. Magnetoplasmons for the α-T3 model with filled Landau levels. J. Phys. Condens. Matter 2020, 32, 485301. [Google Scholar] [CrossRef] [PubMed]
  75. Illes, E.; Carbotte, J.; Nicol, E. Hall quantization and optical conductivity evolution with variable Berry phase in the α-T3 model. Phys. Rev. B 2015, 92, 245410. [Google Scholar] [CrossRef]
  76. Bercioux, D.; Urban, D.; Grabert, H.; Häusler, W. Massless Dirac-Weyl fermions in a T3 optical lattice. Phys. Rev. A At. Mol. Opt. Phys. 2009, 80, 063603. [Google Scholar] [CrossRef]
  77. Bercioux, D.; Goldman, N.; Urban, D. Topology-induced phase transitions in quantum spin Hall lattices. Phys. Rev. A At. Mol. Opt. Phys. 2011, 83, 023609. [Google Scholar] [CrossRef]
  78. Goldman, N.; Urban, D.; Bercioux, D. Topological phases for fermionic cold atoms on the Lieb lattice. Phys. Rev. A At. Mol. Opt. Phys. 2011, 83, 063601. [Google Scholar] [CrossRef]
  79. Zuev, Y.M.; Chang, W.; Kim, P. Thermoelectric and magnetothermoelectric transport measurements of graphene. Phys. Rev. Lett. 2009, 102, 096807. [Google Scholar] [CrossRef]
  80. Crassee, I.; Orlita, M.; Potemski, M.; Walter, A.L.; Ostler, M.; Seyller, T.; Gaponenko, I.; Chen, J.; Kuzmenko, A. Intrinsic terahertz plasmons and magnetoplasmons in large scale monolayer graphene. Nano Lett. 2012, 12, 2470–2474. [Google Scholar] [CrossRef]
  81. Yan, H.; Li, Z.; Li, X.; Zhu, W.; Avouris, P.; Xia, F. Infrared spectroscopy of tunable Dirac terahertz magneto-plasmons in graphene. Nano Lett. 2012, 12, 3766–3771. [Google Scholar] [CrossRef]
  82. Giesbers, A.; Zeitler, U.; Ponomarenko, L.; Yang, R.; Novoselov, K.; Geim, A.; Maan, J. Scaling of the quantum Hall plateau-plateau transition in graphene. Phys. Rev. B Condens. Matter Mater. Phys. 2009, 80, 241411. [Google Scholar] [CrossRef]
  83. Jiang, Z.; Henriksen, E.A.; Tung, L.; Wang, Y.J.; Schwartz, M.; Han, M.Y.; Kim, P.; Stormer, H.L. Infrared spectroscopy of Landau levels of graphene. Phys. Rev. Lett. 2007, 98, 197403. [Google Scholar] [CrossRef]
  84. Ponomarenko, L.; Yang, R.; Gorbachev, R.; Blake, P.; Mayorov, A.; Novoselov, K.; Katsnelson, F.M.; Geim, A. Density of states and zero Landau level probed through capacitance of graphene. Phys. Rev. Lett. 2010, 105, 136801. [Google Scholar] [CrossRef]
  85. Luican, A.; Li, G.; Andrei, E.Y. Quantized Landau level spectrum and its density dependence in graphene. Phys. Rev. B Condens. Matter Mater. Phys. 2011, 83, 041405. [Google Scholar] [CrossRef]
Figure 1. (Color online) Comparison of schematics for low-energy electronic dispersions ε ( D , L ) ( k | Δ 0 ) of a gapped dice lattice (D, (a)) and a Lieb lattice (L, (b)). The common feature of the two materials is the presence of three energy bands, including one dispersionless flat band. However, the relative position of the flat band is quite different for these two lattices. In the left panel for a dice lattice, the flat band sits exactly in the middle between a valence and conduction band, leading to additional symmetry in this band structure. However, the flat band of the Lieb lattice in the right panel is located at an elevated position intersecting with a conduction band at its lowest point.
Figure 1. (Color online) Comparison of schematics for low-energy electronic dispersions ε ( D , L ) ( k | Δ 0 ) of a gapped dice lattice (D, (a)) and a Lieb lattice (L, (b)). The common feature of the two materials is the presence of three energy bands, including one dispersionless flat band. However, the relative position of the flat band is quite different for these two lattices. In the left panel for a dice lattice, the flat band sits exactly in the middle between a valence and conduction band, leading to additional symmetry in this band structure. However, the flat band of the Lieb lattice in the right panel is located at an elevated position intersecting with a conduction band at its lowest point.
Magnetism 05 00022 g001
Figure 2. (Color online) Calculated scaled Landau levels ε n / γ B ( D ) as a function of level index n for a zero-gap ( Δ 0 = 0 ) dice lattice: the scaled energy levels ε n / γ B ( D ) of a dice lattice with Δ 0 = 0 under a uniform magnetic field are obtained from analytical expressions in Equation (22). Here, the flat band (i.e., the long red line at the bottom of the right panel) corresponds to a set of degenerate Landau levels ε n = 0 . As labeled, the dashed curves in the two panels show two functions ± 2 n 1 for the left (valence band, −) and the right (conduction band, +) panels, respectively.
Figure 2. (Color online) Calculated scaled Landau levels ε n / γ B ( D ) as a function of level index n for a zero-gap ( Δ 0 = 0 ) dice lattice: the scaled energy levels ε n / γ B ( D ) of a dice lattice with Δ 0 = 0 under a uniform magnetic field are obtained from analytical expressions in Equation (22). Here, the flat band (i.e., the long red line at the bottom of the right panel) corresponds to a set of degenerate Landau levels ε n = 0 . As labeled, the dashed curves in the two panels show two functions ± 2 n 1 for the left (valence band, −) and the right (conduction band, +) panels, respectively.
Magnetism 05 00022 g002
Figure 3. (Color online) The scaled Landau levels ε n / γ B ( D ) as a function of level index n for a finite-gap dice lattice: the energy eigenstates of a dice lattice with a finite bandgap Δ 0 = 0.7 γ B ( D ) under a uniform magnetic field, which are calculated by using Equation (35). As labeled, three panels (ac) represent the energy levels of the valence, flat and conduction bands, separately. The flat band (long black line at the top of panel (b)) splits into a set of non-equidistant Landau levels with a lifted degeneracy depending on bandgap Δ 0 .
Figure 3. (Color online) The scaled Landau levels ε n / γ B ( D ) as a function of level index n for a finite-gap dice lattice: the energy eigenstates of a dice lattice with a finite bandgap Δ 0 = 0.7 γ B ( D ) under a uniform magnetic field, which are calculated by using Equation (35). As labeled, three panels (ac) represent the energy levels of the valence, flat and conduction bands, separately. The flat band (long black line at the top of panel (b)) splits into a set of non-equidistant Landau levels with a lifted degeneracy depending on bandgap Δ 0 .
Magnetism 05 00022 g003
Figure 4. (Color online) The scaled Landau levels ε n / γ B ( D ) of a finite-gap dice lattice as a function of the scaled energy bandgap Δ 0 / γ B ( D ) for a gapped dice lattice under a uniform magnetic field, which are obtained by numerically solving the root ε n of Equation (33). As labeled, three panels (ac) display the dependence of Landau levels on Δ 0 / γ B ( D ) for valence, flat and conduction bands, respectively. The flat band case in panel (b) reveals a set of non-equivalent Landau levels with a lifted degeneracy depending on Δ 0 . Here, the bandgap in panels (a,c) is chosen as Δ 0 = 0.7 γ B ( D ) .
Figure 4. (Color online) The scaled Landau levels ε n / γ B ( D ) of a finite-gap dice lattice as a function of the scaled energy bandgap Δ 0 / γ B ( D ) for a gapped dice lattice under a uniform magnetic field, which are obtained by numerically solving the root ε n of Equation (33). As labeled, three panels (ac) display the dependence of Landau levels on Δ 0 / γ B ( D ) for valence, flat and conduction bands, respectively. The flat band case in panel (b) reveals a set of non-equivalent Landau levels with a lifted degeneracy depending on Δ 0 . Here, the bandgap in panels (a,c) is chosen as Δ 0 = 0.7 γ B ( D ) .
Magnetism 05 00022 g004
Figure 5. (Color online) Landau levels for a Lieb lattice: the energy eigenstates of a Lieb lattice with a finite bandgap Δ 0 = 0.7 γ B ( D ) in the presence of a uniform magnetic field, and calculated from analytical expressions in Equations (45) and (46). Two panels (a,b) represent the energy levels corresponding to the valence (a) and conduction (b) bands, respectively.
Figure 5. (Color online) Landau levels for a Lieb lattice: the energy eigenstates of a Lieb lattice with a finite bandgap Δ 0 = 0.7 γ B ( D ) in the presence of a uniform magnetic field, and calculated from analytical expressions in Equations (45) and (46). Two panels (a,b) represent the energy levels corresponding to the valence (a) and conduction (b) bands, respectively.
Magnetism 05 00022 g005
Figure 6. (Color online) Landau levels as a function of energy bandgap Δ 0 for a Lieb lattice in the presence of a uniform magnetic field, and calculated from analytical expressions in Equations (45) and (46). Two panels (a,b) represent the energy levels corresponding to the valence (a) and conduction (b) bands. The flat band ε n = Δ 0 in panel (b) corresponds to a set of non-equivalent Landau levels with a lifted degeneracy depending on the energy gap Δ 0 .
Figure 6. (Color online) Landau levels as a function of energy bandgap Δ 0 for a Lieb lattice in the presence of a uniform magnetic field, and calculated from analytical expressions in Equations (45) and (46). Two panels (a,b) represent the energy levels corresponding to the valence (a) and conduction (b) bands. The flat band ε n = Δ 0 in panel (b) corresponds to a set of non-equivalent Landau levels with a lifted degeneracy depending on the energy gap Δ 0 .
Magnetism 05 00022 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhemchuzhna, L.; Joseph, L.; Iurov, A.; Gumbs, G.; Huang, D. Landau Levels and Electronic States for Pseudospin-1 Lattices with a Bandgap: Application to a Lieb Lattice. Magnetism 2025, 5, 22. https://doi.org/10.3390/magnetism5030022

AMA Style

Zhemchuzhna L, Joseph L, Iurov A, Gumbs G, Huang D. Landau Levels and Electronic States for Pseudospin-1 Lattices with a Bandgap: Application to a Lieb Lattice. Magnetism. 2025; 5(3):22. https://doi.org/10.3390/magnetism5030022

Chicago/Turabian Style

Zhemchuzhna, Liubov, Lovely Joseph, Andrii Iurov, Godfrey Gumbs, and Danhong Huang. 2025. "Landau Levels and Electronic States for Pseudospin-1 Lattices with a Bandgap: Application to a Lieb Lattice" Magnetism 5, no. 3: 22. https://doi.org/10.3390/magnetism5030022

APA Style

Zhemchuzhna, L., Joseph, L., Iurov, A., Gumbs, G., & Huang, D. (2025). Landau Levels and Electronic States for Pseudospin-1 Lattices with a Bandgap: Application to a Lieb Lattice. Magnetism, 5(3), 22. https://doi.org/10.3390/magnetism5030022

Article Metrics

Back to TopTop