Optically Controlled Topological Phases in the Deformed - Lattice
Abstract
Haldane’s tight-binding model, which describes a Chern insulator in a two-dimensional hexagonal lattice, exhibits quantum Hall conductivity without an external magnetic field. Here, we explore an lattice subjected to circularly polarized off-resonance light. This lattice, composed of two sublattices (A and B) and a central site (C) per unit cell, undergoes deformation by varying the hopping parameter while keeping = = . Analytical expressions for quasi-energies in the first Brillouin zone reveal significant effects of symmetry breaking. Circularly polarized light lifts the degeneracy of Dirac points, shifting the cones from M. This deformation evolves with , breaking symmetry at , as observed in Berry curvature diagrams. In the standard case (), particle-hole and inversion symmetries are preserved for and . The system transitions from a semi-metal to a Chern insulator, with band-specific Chern numbers: , , and for shifting to , , and when For , the system enters a trivial insulating phase. These transitions, confirmed via Wannier charge centers, are accompanied by a diminishing Hall conductivity. Our findings highlight tunable topological phases in lattices, driven by light and structural deformation, with promising implications for quantum materials.
Keywords: lattices; Off-resonance light; effective Hamiltonian; topological properties; Hall conductivity.
I Introduction
Higher-order topological phases have revolutionized our understanding of quantum materials [1, 2], by extending the concept of topology beyond conventional edge or surface states to boundary modes localized at corners or hinges [3]-[6]. These phases represent a profound advancement in the study of topological phenomena, challenging traditional classifications based on symmetry and topology [7, 8]. This framework builds upon earlier discoveries such as the quantum Hall effect [9, 10], where quantized Hall conductivity is governed by topological invariants like the Chern number [11, 12]. Haldane’s seminal work on breaking time-reversal symmetry in honeycomb lattices provided a theoretical foundation for understanding these invariants and their role in phase transitions [13], inspiring experimental realizations in diverse geometries and materials including Lieb and Kagome lattices [15, 16], iron-based honeycomb ferromagnetic insulators [17], and electronic and photonic systems [18, 19, 20, 21].
Two-dimensional topological insulators are defined by time-reversal symmetry and a topological invariant [22, 23]. Breaking this symmetry triggers a topological phase transition, forming a Chern insulator characterized by chiral edge states and the quantized Hall effect [24]. The Chern number determines the system’s topological phase, with zero indicating a trivial phase. Additionally, circularly polarized off-resonant light can induce topological transitions [25], driving systems from trivial to non-trivial phases through Floquet theory [26, 27]. Second-order photon processes allow for band gap tuning, enabling effective Hamiltonians dependent on light frequency and intensity, as in the Haldane model. This framework explains transitions like the polarized light-induced transition of semi-metallic graphene into a Chern insulator [28]. Such prominent example under non-resonant polarized light involves a phenomenon corroborated by numerous analogous experiments on radiative systems [29, 30].
While most Chern insulators have a Chern number of 1, research is increasingly focused on higher Chern numbers, which have been observed experimentally in systems like thin film magnetic topological insulators and photonic crystals [31]. It have been also predicted theoretically in the Dirac-Weyl semimetals on the lattice [32]. Numerous studies have revealed remarkable properties of the system, including enhanced Hall conductivity and unconventional Berry phase effects, which are closely tied to its unique electronic structure [33, 34]. For instance, the lattice, characterized by a tunable parameter controlling the weight of its flat band, exhibits intriguing phenomena like Klein tunneling and Fabry-Perot resonances, which have been extensively studied in both single-layer and bilayer configurations [33, 34]. Additionally, the interplay between the flat band and the Berry phase in these systems leads to novel magneto-optical properties and phase transitions, such as the quantum spin Hall phase transition [35]. Further insights have been gained from detailed analyses of Floquet states in optically driven lattices. Under resonant and circularly polarized light irradiation, these systems exhibit the opening of Berry phase-dependent optical gaps, revealing unique topological signatures and symmetry-driven phenomena [36]. Moreover, the dice lattice, a special case of the model with a flat band parameter, has shown to exhibit higher Chern numbers, accompanied by a significant enhancement in unconventional Hall conductivity [37].
High Chern numbers have also been reported in decorated lattices,
multi-orbital systems, and lattices with spin-orbit coupling or ultracold
gases, broadening the understanding of topological materials. Introducing
the Haldane model into a bilayer of the lattice has resulted
in observed Chern numbers of up to 5 [38], along with a substantial
enhancement of the Hall conductivity [38, 39]. Other
lattice structures, such as the decorated honeycomb or starlike lattices
[40], and multi-orbital triangular lattices [41], also exhibit
high Chern numbers and larger jumps. High Chern numbers have also been
reported in Dirac [42] and semi-Dirac [43] systems.
Additionally, the presence of spin-orbit coupling in honeycomb lattices [44] and ultracold gases in triangular lattices [45] has been shown
to give rise to high Chern numbers. These studies contribute to a deeper
understanding of the topological properties of these advanced materials.
In this work, we explore the topological properties of the lattice, a fascinating system that generalizes the honeycomb lattice. The lattice consists of two sublattices, A and B, representing carbon atoms located at the vertices of a hexagonal structure, and a third site, C, positioned at the center of each hexagon, as illustrated in Fig. 1. Within this structure, quasiparticles can hop from site C to site A within the same hexagon. The hopping energy between sites A and B is described as , while that between sites A and C is . The parameter , defined by , governs the lattice configuration. By varying between 0 and 1, different lattice structures can be realized: when , the lattice corresponds to graphene, and when , it becomes the dice lattice. Further details about this lattice structure are provided in [46].
In this study, we investigate the topological properties of the lattice under the influence of an external off-resonant electric field. This field induces a term analogous to the Haldane model, breaking time-reversal symmetry and modulating the band gap, which is pivotal in determining the system’s topological properties. The simultaneous tunability of the parameter and the presence of a flat band create a distinctive energy dispersion. Additionally, we consider the effect of anisotropic deformations by altering the hopping energies of quasiparticles between nearest-neighbor bonds. Specifically, the hopping energy along the position vector changes to for sites A and C, and for sites A and B, while the hopping energies along vectors and remain unchanged. As the parameter increases, the Dirac points approach each other and eventually merge at the M point in the Brillouin zone when , leading to band closure and the loss of topological properties. This critical point marks a topological phase transition. Beyond this point, the energy dispersion acquires a Dirac-like form along the -axis, a phenomenon that has been previously observed in graphene [47], graphene bilayers [48], and the dice lattice [49]. Furthermore, this method has been applied to the single-layer honeycomb structure , yielding a semi-Dirac dispersion [50].
To gain deeper insights, we study the effects of deformation on the
topological properties by tuning the hopping energy to
specific values. We calculate the Berry curvature, which encapsulates key
symmetries such as particle-hole symmetry, inversion symmetry, and
symmetry in the standard case (). As
deviates from , the symmetry is broken, leading to
topological transitions. We also compute the Chern number, which indicates
the system’s transition from a non-trivial to a trivial topological phase.
To further confirm this transition, we analyze the evolution of the Wannier
charge center along a closed loop in the Brillouin zone, representing the
average charge position within the unit cell. This analysis is consistent
with the behavior of surface energy bands [51, 52, 53]. Moreover, we
investigate the quantum Hall conductivity and its response to the
deformation of the energy spectrum as varies. Our results
provide valuable insights into the interplay between lattice deformation and
topological phase transitions, shedding light on the tunable nature of
topological properties in lattices.
The originality of the present work lies in its comprehensive analysis of
the interplay between lattice deformation, external electric fields, and
topological properties in the lattice. By exploring the
combined effects of anisotropic hopping energies, Berry curvature, and
quantum Hall conductivity, this study not only advances the understanding of
topological materials but also paves the way for potential applications in
quantum technologies.
This article is organized as follows: Section II presents the Hamiltonian describing quasiparticle dynamics in irradiated and deformed lattices. Section III analyzes and discusses the quasi-energy spectrum. Section IV investigates the topological properties of this lattice, including the Berry curvature (Subsection IV.1), the Chern phase diagram (Subsection IV.2), and the evolution of the Wannier charge center (Subsection IV.3). Section V examines the anomalous Hall conductivity. Finally, Section VI provides a conclusion and summary of the work.
II Model and Hamiltonian
The rescaled tight-binding Hamiltonian describing the motion of a quasiparticle along a orbital in a lattice between its nearest neighbours is given by [46]
(1) |
where and are the vectors connecting the nearest neighbors, and is the hopping energy. In our model, we consider
along the direction and , along the and directions where is the distance of the
nearest neighbour, as shown in Fig.1. In this section we’ll study
photon-electron interactions by applying a polarised electric field
perpendicular to the plane of the lattice. The electric field is
derived from the vector potential with respect to time ,
and this vector potential depends on time, , where and are the radiation
frequency and the vector potential amplitude, respectively. The intensity of
light is characterized by a dimensionless parameter and is much smaller than 1, where and
is the electric charge.

when an electric field is applied to electrons moving from site to the nearest site , they gain energy, represented by the hopping energy , which is transformed into , where is the phase factor gained by the electron. The Hamiltonian then becomes
(2) |
with , where is a Bessel function of the first kind. We consider that the light incident on the lattice has an off-resonance frequency, as explained in [25]. When an electron is subjected to an off-resonance frequency, it does not excite it directly but simply modifies the band structure in an effective way by absorbing and emitting virtual photons. An off-resonance state is reached when the frequency of the photons is well above the bandwidth, i.e. . We have a time-dependent and periodic Hamiltonian . Floquet’s theorem is perhaps the most convenient solution to this problem, with . In this case, we can describe the properties of the off-resonance light effect by means of the effective Hamiltonian [25, 39].
(3) |
We express the time-dependent Fourier components of the Hamiltonian as follows:
(4) |
The second term is responsible for the absorption of a virtual photon by an electron through and its emission through . We give the explicit expression of the effective Hamiltonian considering terms up to .
(5) |
and are defined as follows
(6) |
where . We consider a weak conduit , which implies taking the approximation and . is the light-induced term, identical to the near-second Haldane term with and , which satisfies the Haldane model for graphene and the dice lattice [37, 47]. This term is responsible for breaking the time reversal symmetry. This results in a gap in the Dirac points. Later we’ll discuss the -effect, which is a mesh distortion that changes the location of this gap.
As explained previously [32], if we take , satisfies the anticommutation relations when and .
(7) |
where is defined as the graphene operator and as the dice operator, given as:
(8) |
with as the complex conjugate, equation (7) implies the existence of a particle and a hole such that , as well as a flat band at zero. However, when we take and whatever is, at point (see Fig.2), satisfies the anticommutation relations.
(9) |
In this case, we can conclude that the radiation does not affect the system in the same way as it would normally, because the system no longer depends on . We will later demonstrate the importance of modifying , as this results in the deformation of a lattice, a change , and an observation of its effect on topological properties.
III Quasienergy and band structure
In this section, we study the energy bands as varies. The Hamiltonian will be diagonalized to obtain the eigenvalues. This diagonalization leads to the characteristic equation, known as the depressed cubic equation, , whose solutions can be expressed as follows:
(10) |
and
(11) | ||||
(12) | ||||
(13) |
The quasi-energy associated with the conduction band is represented by the value , while the flat band and the valence band are represented by the values and , respectively. These quasienergies correspond to normalised pseudo-vectors.
(14) |
and
(15) |

Before discussing the band structure, we determine the path of the first
Brillouin zone in the reciprocal lattice, which is in the form of a
hexagonal lattice, to construct the band structure path, as shown in Fig.2-(m). Now, we analyse and discuss the band structure by varying for three values: 0, and 1. In addition, for each
fixed value of , we vary while keeping
constant. It is known that when the lattice is not exposed
to radiation, the three bands touch at the Fermi level at points and ,
known as Dirac points. On the other hand, under the effect of active
radiation, a gap opens that is in a quasi-energy sense dependent on . As increases, we observe that for (see fig2(a)-2(c)), in the case of graphene, the gap progressively
decreases, moving away from the Dirac points and adopting a behaviour
similar to the deformed Haldane model. When reaches a
particular value, , the gap disappears at the point (see figure 2(c)), where two bands
degenerate. Similar behaviour is observed for dice lattices (),
which also have a structure close to the deformed Haldane model. In this
case the gap decreases with increasing g1 and the conduction ,flat, and
valence bands become degenerate at the point for (see fig2(d)). This happens despite the presence of
radiation that breaks the time-reversal symmetry. The spectrum then exhibits
a semi-Dirac-type dispersion, characterised by a quadratic dependence along and a linear dependence along , in the absence of radiation.
In the presence of radiation, for , electrons move
with distinct and linear velocities in both directions, thereby revealing a
linear anisotropy. When 1 passes ,
a gap opens at point . For , the gap partially
closes at points and , as illustrated in Figure2(i)-(j). At points and , the dispersion
becomes kinetic, except in the interval . Conversely, when , the gap disappears
and the bands become degenerate at point . When , the effect of radiation is unable to control the gap adjustment, leading
to a reopening of the gap.In summary, we can conclude that the effect of
radiation is significant only in the interval .
One aspect that merits attention is that, despite the system being subjected to radiation and being fixed at 2, the gap closes and the time-reversal symmetry remains broken. This could indicate a change in the topological properties of the lattice. This is precisely what we are going to examine in the study of topological properties as varies.
IV Topological properties
In this section, we study the effect of deformation from change on Berry curvature, Chern number, and Wannier charge center.
IV.1 Berry curvature
In this subsection we analyze the Berry curvature, which plays a crucial role in topological quantum physics. It is characterized by the operations of the following discrete symmetries: time inversion symmetry , charge conjugation symmetry , and inversion symmetry , where represents the quasi-energy index conjugate of (). We are interested in the effect of the variation of on this Berry curvature symmetry. To do this, we numerically calculate the Berry curvature of the system in the z-component, which is given by [54]
(16) |
where is the effective velocity
in the axial direction . We know that the Berry curvature, if not
zero, results from the breaking of the time-reversal symmetry, or at least
from the presence of a broken symmetry, as mentioned previously. We now plot
the Berry curvature to analyze this symmetry and to study the effect of
variations of . We vary in two cases: and .
For , time-reversal symmetry is broken when Berry
curvatures differ from zero for any value of alpha and any individual Berry
curvature. Charge conjugate symmetry is present for and 1, see
Fig .3 where the Berry curvature and the associated
curvature of a flat band is zero, and when k is given, it’s
clear that the sum of the individual Berry curvatures is zero, so the local
conservation of Berry curvature and also the inversion symmetry is present and , but for and 1 the conjugate charge
and inversion symmetry are broken. The expression for the Berry curvature
can be found analytically in the case where , and is written like
this
(17) |
and
(18) |


As we saw previously, the Berry curvature is concentrated at Dirac points and for in the Brillouin zone (BZ). As increases, we observe that the Berry curvature moves slowly towards point for (see Fig.4). However, this figure does not include results for other values of . The Berry curvature becomes singular at , where quasi-energy is degenerate at point . When we vary , the system breaks the inversion symmetry and, more precisely, the symmetry is also broken in the cases where and (see Fig.4).
IV.2 Chern phase diagrams
In this subsection, we aim to obtain the phase diagram of the Chern number. As we have previously mentioned, light induces a term similar to Haldane’s term for , which is responsible for breaking the time-reversal symmetry. This term, as we have studied, generates a non-zero Berry curvature, which is concentrated at the Dirac points due to the opening of a gap caused by light. This implies the existence of a phase transition, which can be observed through the surface integral of the Berry curvature over the BZ. This integral gives an integer, called the Chern number for the -th band, which is defined by the following relationship [54].
(19) |
Before discussing the evolution of the Chern number as a function of parameter variations, we first consider the standard case where no deformation varies the hopping energy , i.e. . In the lattice there are three bands: two dispersive bands ( and ) and a flat band (), which becomes dispersive at the Dirac point when . To observe the topological evolution of the system, we study the transition between the graphene lattice ( ) and the dice lattice limit (), varying from 0 to 1. The system becomes topologically non-trivial due to a non-zero Chern number, except for the Chern numbers associated with the bands and . We see that there is a phase transition at , where the Chern numbers of the bands are in the range . At these values change from to when (see Fig.5-(a)). On the other hand, the Chern number associated with the band remains zero, indicating that it is topologically trivial.

We now consider the lattice irradiated and deformed along
the displacement vector , resulting in a change in the
hopping energy . For this study we choose a variation region of
between . We investigate the evolution of
the Chern number , which is the inverse of , while
remains topologically trivial. Fixing and
as well as , we plot as a function of . For , the Cherne number is initially equal to 1 for and 2 for , as shown in Fig.5-(a). As increases from g to , remains
constant regardless of the value of . However, as
exceeds , the system undergoes a change in topological properties
from a topologically non-trivial to a topologically trivial state. This
change, characteristic of a phase transition, occurs precisely at . As shown in Fig.5-(b), the Cherne number
changes from 1 () or 2 () for
to 0 () or 0 () for .
We also analyze variations in the Cherne number by modifying the
parameters and . To do this, we plot as a
function of these parameters by fixing in three cases: , and , as illustrated in the topological
phase diagram in Fig.6. This diagram highlights the topological
phases of the system, represented by different colors: red corresponds to , yellow to , and white indicates a topologically trivial
phase with . In the first case, where (see
Fig.6-(a)), when the polarization amplitude is
initially very low, this keeps the system in a topologically trivial phase
with in the interval .
However, when , the system undergoes a topological
phase transition: goes from 0 to 1 in the interval . In the range , the
Chern number changes from to for ,
precisely in the interval . As D increases
further (), the system reaches a phase where
holds for any value of in the interval .

In the second case, by setting , we observe that the
system undergoes a topological phase transition. It goes from a non-trivial
phase in the interval to a
topologically trivial phase. Another phase transition is identified in the
interval , which occurs at
(see Fig. 6-(b)).
In the third case, when , the system changes its
topological property between . It
moves from a topologically trivial to a non-trivial phase, unlike the case , where the values of and are
replaced by (see Fig. 6-(c)). These observations show
that varying changes the topological properties of the system,
moving it from a non-trivial to a trivial state.
To confirm whether or not the system remains in a non-trivial topological insulating state as varies, we would like to study the evolution of the Wannier charge center in the next subsection.
IV.3 Wannier charge centers
This subsection focuses on examining Wannier charge centres(WCCs) as a means to observe and analyse topological evolution, a fundamental aspect of the diagnostic process of topological band properties [51, 53]. These centres are defined as the mean charge position in a system based on Wannier functions. This mathematical conceptualisation finds application in bands of a two-dimensional system, defined as follows [52, 53]
(20) |
It is evident that these WCCs are equivalent to the Berry phase calculation. However, in order to examine the topological evolution and extract information about the system, the WCCs must be calculated along as a function of the transverse moment , based on a modified form of equation (20)
(21) |
WCCs are calculated only for , signifying the displacement of the referential unit cell’s center of charge along a closed trajectory of . By varying the hopping energy for four cases (, and ), it is observed that the WCC plots exhibit winding and discontinuity at the Dirac point as increases, particularly in the standard case (), indicating a non-trivial topological phase of the system(see Fig.7-(a)). However, when surpasses 2, as illustrated in Fig.7.(c)-(d), the movement of these centers suggests a transition from a non-trivial topological insulating state to a trivial topological state. This transition is characterized by the dissipation of the WCC windings, which become continuous and begin to oscillate around zero.

V Anomalous Hall Conductivity
In this section, we calculate and discuss the anomalous Hall conductivity (AHC) for the lattice exposed to irradiation and deformation. This conductivity is evaluated by examining the electron response to an external electric field in an anisotropic context, providing information about the system topology. To determine the AHC, we integrate the Berry curvature over all occupied states in the entire Brillouin zone (BZ), by Eq.(16). We use the following formula to perform this calculation [54].
Where is the Fermi-Dirac distribution function, is the chemical potential, T is the temperature, is the Boltzmann constant and . The AHC can be calculated numerically as
a function of the chemical potential at T=100 K by fixing and for four cases: and 1, while varying , as shown in Fig.8.
We first examine the first two cases, and , where the hopping energy varies, as shown in Fig. 8(a) and (d). A quantised plateau is observed with for and for . When the chemical potential lies within the global gap, the width of the plateau corresponds to that of the gap in the dispersion spectrum and increases proportionally with increasing . It is noteworthy that for is twice that for . In this case, the Berry curvature associated with the flat band () disappears and has no
contribution to . When intercepts the bands (whetherconduction or valence ), decreases. This decrease is explained by the fact that the integral isperformed on occupied states and has left the global band.
In addition, the width of the plateau decreases with increasing , as the overall gap between the dispersive bands narrows. At the plateau disappears completely, as does the Hall conductivity.
For cases and 1, the flat band becomes dispersive, which changes the behaviour of the Hall conductivity. In fact, due to this dispersive nature, there is no longer a smooth plateau in the Hall conductivity. Instead, exhibits a summit located at , as observed for (see Fig. 8(a)). In this
scenario, two distinct plateaus of are observed when lies within two energy gaps: one between the flat band and at point K, and the other between and at point K’. In these cases, the value of for each plateau is .
For , two similar plateaus are also observed, but the value of reaches . This increase in results from a phase transition when . As discussed earlier, the Chern number remains constant in the phase, which keeps at 2.
However, in this regime, there are no smooth plateaus in the Hall
conductivity. Instead, exhibits a dip near , as shown in Fig. 8(d).
Here, for and , the width of the plateaus in decreases with increasing as long as and disappears completely when .

Finally, we find that the value of plays a crucial role in increasing the Hall conductivity as well as in widening the corresponding plateaus. These plateaus are proportional to the Chern number, with as long as the system remains a non-trivial topological insulator. However, these plateaus are sensitive to the hopping energy , which leads to a progressive reduction in their width as increases. Ultimately, these plateaus disappear when the band gap closes at the critical point , marking the system’s transition to a trivial topological insulator state.
VI Conclusion
In this work, we have investigated the effect of deformation on the topological properties induced by light polarisation in lattices. We introduced this deformation by modifying the hopping energy in the lattice, particularly by modifying at the A-B and A-C sites along the direction, while remaining unchanged hopping energies along the and directions. This modification of led to a change in the band structure for three specific cases: , and . In these cases, the Dirac cones at points and move towards point when the system is subjected to off-resonance circular polarization. This polarisation induces a Haldane mass term at , breaking the time-reversal symmetry. This term is responsible for the opening of a gap. However, in the case of , the gap partially opens at points and . Moreover, the gap size decreases with increasing and disappears completely when , where the system adopts semi-Dirac behaviour. This deformation, which breaks the symmetry,leads to a shift in the concentration of Berry curvature in the Brillouin zone , from points and to point . In the standard case, , we observe particle-hole symmetry for and and inversion symmetry. We have calculated the corresponding Chern numbers for the conduction (), flat () and valence () bands: For the Chern numbers are , and . For the values change to , and . This shows a phase transition. We then focused on the Chern number corresponding to . Fixing , we plotted as a function of . It turned out that there is a phase transition with gap closure at . When , the system changes from a non-trivial topological insulator () to a trivial topological insulator (). We have also plotted a phase diagram in the parameter space of and by fixing . When , in the interval of and , the system transforms into a trivial topological insulator. This transition was confirmed by plotting the Wannier charge centre. Fixing , we observed that the system changes state from a non-trivial topological insulator to another trivial topological insulator state when . We have also calculated the anomalous Hall conductivity for four values of : 0, 0.45, 0.8, and 1. For and , the conductivity shows quantized behaviour in the form of plateaus with a value of . For and , the conductivity value increases to , characterized by a plateau at . This plateau decreases with increasing and disappears when . These plateaus correspond to the Chern number, where . The disappearance of these plateaus indicates that the system is transitioning from a non-trivial topological insulator to a trivial topological state.
Acknowledgements
The authors would like to acknowledge the ” Hassan II Academy of Sciences and Techologies-Morocco for its financial support. The authors also thank the LPHE-MS, Faculty of Sciences, Mohammed V University in Rabat, Morocco for the technical support through facilities.
References
- [1] N. Bultinck, B. A. Bernevig, M. P. Zaletel, Three dimensional superconductors with hybrid higher-order topology, Phys. Rev. B 99 (2019) 125149.
- [2] L. B. Drissi, S. Lounis, E. H. Saidi, Higher order topological matter and fractional chiral states, The European Physical Journal Plus 137, Article number: 796 (2022).
- [3] Nagaosa, N., et Tokura, Y. (2013). Topological properties and dynamics of magnetic skyrmions. Nature nanotechnology, 8(12), 899-911.
- [4] L. B. Drissi, E. H. Saidi, and M. Bousmina, Four-dimensional graphene. Physical Review D, 2011, vol. 84, no 1, p. 014504.
- [5] C. C. Liu, W. Feng, Y. Yao, spin Hall effect in silicene and two-dimensional germanium. Physical review letters, 2011, vol. 107, no 7, p. 076802.
- [6] L. B. Drissi, E. H. Saidi, Domain Walls in Topological Tri-hinge Matter, European Physical Journal Plus 136 (2021) 68.
- [7] F. Schindler, A. M. Cook, M. G. Vergniory, Z. Wang, S. S. Parkin, B. A. Bernevig, and T. Neupert, Higher order topological insulators, Science advances 4 (2018) eaat0346.
- [8] Hiromu Araki, Tomonari Mizoguchi, Yasuhiro Hatsugai, ZQ Berry Phase for Higher-Order Symmetry-Protected Topological Phases, Phys. Rev. Research 2 (2020) 012009.
- [9] Klitzing, K. V., Dorda, G., et Pepper, M. (1980). New method for high-accuracy determination of the fine-structure constant based on quantized Hall resistance. Physical review letters, 45(6), 494.
- [10] Thouless, D. J., Kohmoto, M., Nightingale, M. P., et den Nijs, M. (1982). Quantized Hall conductance in a two-dimensional periodic potential. Physical review letters, 49(6), 405.
- [11] Laughlin, R. B. (1981). Quantized Hall conductivity in two dimensions. Physical Review B, 23(10), 5632.
- [12] Fu, L., Kane, C. L., et Mele, E. J. (2007). Topological insulators in three dimensions. Physical review letters, 98(10), 106803.
- [13] Haldane, F. D. M. (1988). Model for a quantum Hall effect without Landau levels: Condensed-matter realization of the” parity anomaly”. Physical review letters, 61(18), 2015.
- [14] Elbahri, A., Ragragui, M., Drissi, L.B., Saidi, E.H., Exploring topological phases in superconducting transition metal (Sc, Ti, V)-carbides, Materials Science in Semiconductor Processing 2025, 186, 108993.
- [15] Zhang, D. W., Zhu, Y. Q., Zhao, Y. X., Yan, H., et Zhu, S. L. (2018). Topological quantum matter with cold atoms. Advances in Physics, 67(4), 253-402.
- [16] Jiang, W., Kang, M., Huang, H., Xu, H., Low, T., et Liu, F. (2019). Topological band evolution between Lieb and kagome lattices. Physical Review B, 99(12), 125131.
- [17] Kim, H. S., et Kee, H. Y. (2017). Realizing Haldane model in Fe-based honeycomb ferromagnetic insulators. npj Quantum Materials, 2(1), 20
- [18] Lu, L., Joannopoulos, J. D., et Soljačić, M. (2014). Topological photonics. Nature Photonics, 8(11), 821-829.
- [19] Zhang, X., Zangeneh-Nejad, F., Chen, Z. G., Lu, M. H., et Christensen, J. (2023). A second wave of topological phenomena in photonics and acoustics. Nature, 618(7966), 687-697.
- [20] Li, S., Tian, R., Liu, M., Arzamasovs, M., Chen, L., et Liu, B. (2024). Topological flat band with higher winding number in a superradiance lattice. Annals of Physics, 471, 169837.
- [21] Jacak, J. E. (2021). Topological approach to electron correlations at fractional quantum Hall effect. Annals of Physics, 430, 168493.
- [22] Kane, C. L., et Mele, E. J. (2005). topological order and the quantum spin Hall effect. Physical review letters, 95(14), 146802.
- [23] Bentaibi, B., Drissi, L. B., Saidi, E. H., Bousmina,M., New room-temperature 2D hexagonal topological insulator OsC: First Principle Calculations, Materials Science in Semiconductor Processing 151 (2022) 107009.
- [24] Kou, L., Ma, Y., Sun, Z., Heine, T., et Chen, C. (2017). Two-dimensional topological insulators: Progress and prospects. The journal of physical chemistry letters, 8(8), 1905-1919.
- [25] Kitagawa, T., Oka, T., Brataas, A., Fu, L., et Demler, E. (2011). Transport properties of nonequilibrium systems under the application of light: Photoinduced quantum Hall insulators without Landau levels. Physical Review B-Condensed Matter and Materials Physics, 84(23), 235108.
- [26] McIver, J. W., Schulte, B., Stein, F. U., Matsuyama, T., Jotzu, G., Meier, G., et Cavalleri, A. (2020). Light-induced anomalous Hall effect in graphene. Nature physics, 16(1), 38-41.
- [27] Bao, C., Tang, P., Sun, D., et Zhou, S. (2022). Light-induced emergent phenomena in 2D materials and topological materials. Nature Reviews Physics, 4(1), 33-48.
- [28] Usaj, G., Perez-Piskunow, P. M., Foa Torres, L. E. F., et Balseiro, C. A. (2014). Irradiated graphene as a tunable Floquet topological insulator. Physical Review B, 90(11), 115423.
- [29] Ünal, F. N., Seradjeh, B., et Eckardt, A. (2019). How to directly measure floquet topological invariants in optical lattices. Physical Review Letters, 122(25), 253601.
- [30] Nag, T., et Roy, B. (2021). Anomalous and normal dislocation modes in Floquet topological insulators. Communications Physics, 4(1), 157.
- [31] Skirlo, S. A., Lu, L., Igarashi, Y., Yan, Q., Joannopoulos, J., et Soljačić, M. (2015). Experimental observation of large Chern numbers in photonic crystals. Physical Review Letters, 115(25), 253901.
- [32] Dey, B., et Ghosh, T. K. (2019). Floquet topological phase transition in the lattice. Physical Review B, 99(20), 205429.
- [33] Iurov, A., Zhemchuzhna, L., Gumbs, G., Huang, D., Fekete, P., Anwar, F., … et Weekes, N. (2021). Tailoring plasmon excitations in armchair nanoribbons. Scientific reports, 11(1), 20577.
- [34] Benhaida, O., Drissi, L. B., Saidi, E. H., et Laamara, R. A. (2024). Klein tunneling and Fabry-Perot resonances in the bilayer with aligned stacking. Physica Scripta, 99(8), 085958.
- [35] Wang, J., et Liu, J. F. (2021). Quantum spin Hall phase transition in the lattice. Physical Review B, 103(7), 075419.
- [36] Tamang, L., et Biswas, T. (2023). Probing topological signatures in an optically driven lattice. Physical Review B, 107(8), 085408.
- [37] Dey, B., Kapri, P., Pal, O., et Ghosh, T. K. (2020). Unconventional phases in a Haldane model of dice lattice. Physical Review B, 101(23), 235406.
- [38] Parui, P., Ghosh, S., et Chittari, B. L. (2024). Topological properties of nearly flat bands in bilayer lattice. Physical Review B, 109(16), 165118.
- [39] Benhaida, O., Saidi, E. H., Drissi, L. B., et Laamara, R. A. (2024). Topological Properties of Bilayer Lattice Induced by Polarized Light. arXiv preprint arXiv:2412.17763.
- [40] Chen, W. C., Liu, R., Wang, Y. F., et Gong, C. D. (2012). Topological quantum phase transitions and topological flat bands on the star lattice. Physical Review B—Condensed Matter and Materials Physics, 86(8), 085311.
- [41] Yang, S., Gu, Z. C., Sun, K., et Das Sarma, S. (2012). Topological flat band models with arbitrary Chern numbers. Physical Review B—Condensed Matter and Materials Physics, 86(24), 241112.
- [42] Sticlet, D., et Piéchon, F. (2013). Distant-neighbor hopping in graphene and Haldane models. Physical Review B—Condensed Matter and Materials Physics, 87(11), 115402.
- [43] Mondal, S., et Basu, S. (2022). Vanishing of the quantum spin Hall phase in a semi-Dirac Kane-Mele model. Physical Review B, 105(23), 235409.
- [44] Alase, A., et Feder, D. L. (2021). Generating and detecting topological phases with higher Chern number. Physical Review A, 103(5), 053305.
- [45] Yang, Y., Li, X., et Xing, D. (2016). Topological phases and phase transitions on the honeycomb lattice. The European Physical Journal B, 89, 1-5.
- [46] Raoux, A., Morigi, M., Fuchs, J. N., Piéchon, F., et Montambaux, G. (2014). From dia-to paramagnetic orbital susceptibility of massless fermions. Physical review letters, 112(2), 026402.
- [47] Lahiri, S., et Basu, S. (2024). Second order topology in a band engineered Chern insulator. Scientific Reports, 14(1), 1880.
- [48] Mondal, S., et Basu, S. (2023). Band-engineered bilayer Haldane model: Evidence of multiple topological phase transitions. Physical Review B, 108(4), 045307.
- [49] Mondal, S., et Basu, S. (2023). Topological features of the Haldane model on a dice lattice: Flat-band effect on transport properties. Physical Review B, 107(3), 035421.
- [50] Zhong, C., Chen, Y., Xie, Y., Sun, Y. Y., et Zhang, S. (2017). Semi-Dirac semimetal in silicene oxide. Physical Chemistry Chemical Physics, 19(5), 3820-3825.
- [51] Taherinejad, M., Garrity, K. F., et Vanderbilt, D. (2014). Wannier center sheets in topological insulators. Physical Review B, 89(11), 115102.
- [52] Gresch, D., Autes, G., Yazyev, O. V., Troyer, M., Vanderbilt, D., Bernevig, B. A., et Soluyanov, A. A. (2017). Z2Pack: Numerical implementation of hybrid Wannier centers for identifying topological materials. Physical Review B, 95(7), 075146.
- [53] Tyner, A. C. (2024). BerryEasy: a GPU enabled python package for diagnosis of nth-order and spin-resolved topology in the presence of fields and effects. Journal of Physics: Condensed Matter, 36(32), 325902.
- [54] Xiao, D., Chang, M. C., et Niu, Q. (2010). Berry phase effects on electronic properties. Reviews of modern physics, 82(3), 1959-2007.