Two-dimensional bilayer blue phosphorus Dirac-like material: a multi-orbital tight-binding investigation

Amine Benhaij *a and Omar Mounkachi *ab
aLaboratory of Condensed Matter and Interdisciplinary Sciences (LaMCScI), Faculty of Sciences, Mohammed V University of Rabat, BP 1014, RP Rabat, Morocco. E-mail: amine_benhaij@um5.ac.ma; omar.mounkachi@um5.ac.ma
bCollege of Computing, Mohammed VI Polytechnic University, Lot 660, Hay Moulay Rachid, Ben Guerir, 43150, Morocco

Received 13th May 2024 , Accepted 19th August 2024

First published on 19th August 2024


Abstract

This study presents a theoretical examination of the electronic band structure of AA (AB) stacked bilayer blue phosphorus system within the fifth intralayer (5NN) and second interlayer nearest-neighbor (2NN) multi-orbital tight-binding (MOTB) approach. The variation of energy levels has been investigated through the symmetrical tensile strain of the low-buckled honeycomb lattice. Here, the primary objective is to examine the existence of Dirac electronic features in hexagonal stacked bilayer geometry. Our theoretical calculations predict that the AA bilayer is a new hexagonal two-dimensional material with px,y-orbital Dirac-like states at the high-symmetry K point. Consequently, these systems can host massless (massive) Dirac fermions. In particular, the AA bilayer exhibits zero-gap Dirac-like properties and manifests distinguishable Dirac-like cones in the presence of weak spin–orbit coupling when a modest stretch of 2.30% is achieved with a remarkably high Fermi velocity of approximately vf ≈ 0.12 × 105 m s−1. The behavior of the dispersion bands aligns reasonably well with recent experimental observations. Moreover, a stretch of 7.17% breaks some of the sublattice equivalence and enhances the spin–orbit interaction, resulting in the emergence of an electronic band gap of approximately ≈ 0.27 eV in the proximity of the high-symmetry K point. Furthermore, the tiny gap induced by the spin–orbit interaction implies topological nontriviality in the electronic state (quantum anomalous Hall state) of the honeycomb lattice. These findings categorize the AA bilayer as a rare two-dimensional Dirac-like material. This work provides, to the extent of our knowledge, a pioneering investigation into the existence of Dirac electronic properties in bilayer blue phosphorus. In addition, we present the first derivation of the MOTB model. However, the identified electronic characteristics designate this two-dimensional system as an ideal candidate for high-performance nanoelectronic devices.


I. Introduction

In recent years, considerable research effort has been devoted to investigating mono-elemental two-dimensional (2D) systems. These materials originate from the nitrogen (N, Z = 7), carbon (C, Z = 6), and boron (B, Z = 5) columns of the periodic table.1 In the C group, we have graphene, which exhibits intriguing properties, notably the existence of a Dirac cone at the high-symmetry K (K′) point.2,3 This finding was theoretically corroborated through a simplified analysis focusing solely on pz-orbital NN couplings within the tight-binding (TB) model.4 Furthermore, the manifestation of Dirac cones in graphene entails numerous substantial implications and fascinating electronic features.5 It generates high electron mobility (≈ 106 m s−1), which is only two-fold lower than the speed of light.6 This renders the material conducive to various applications, such as high-performance nanoelectronic devices.5,7 Moreover, it results in, for instance, the emergence of a zero-gap. This facilitates the occurrence of ballistic transport (a phenomenon where electrons travel freely without scattering), and furthermore, the creation of extremely stable electron–hole pairs.8 However, the emergence of Dirac fermions in this hexagonal 2D material has been experimentally demonstrated due to abnormal quantum Hall effect (QHE) observations. The latter effect is a remarkable example of a quantum topological state. It occurs when time-reversal symmetry is broken, typically by an external magnetic field. This broken symmetry leads to quantized Hall conductance, which is a hallmark feature of topological materials. However, anomalous QHE (which emerges in 2D materials) does not require an external magnetic field and exhibits characteristic chiral edge states surrounding the insulating bulk. These edge states conduct current in only one direction without dissipation.9 This unusual motion of fermions in 2D configurations has led to their classification as a 2D Dirac system, which has garnered significant interest from both theoretical and experimental perspectives. Moreover, 2D Dirac materials exhibit exotic physics associated with massless Dirac states and topological phenomena.2 These include the half-integer3,6/fractional4,5/fractal QHE and quantum spin Hall effect (QSHE).8 In addition, recent studies have further shown the significance of Dirac cones in the field of topological quantum materials10 and their potential for revolutionizing quantum computation, offering promising properties for the development of next-generation qubits.11–14

Other mono-elemental 2D materials within the C group comprise silicon (Si, Z = 14), germanium (Ge, Z = 32), and tin (Sn, Z = 50) atoms, characterized by a low-buckled honeycomb formation in hybridized sp2 (sp3) mode, which have already been successfully fabricated.15–17 Transitioning to the B family, borophene emerges as an intriguing 2D boron compound, which also exhibits Dirac cones with a Fermi velocity (vf, characterizes the speed at which electrons move near the Fermi level) approximately two-fold that of graphene along the high-symmetry ΓX (YM) direction.18–22 Nevertheless, these configurations do have particular restrictions. For instance, the disparity between their on- and off-current states (on/off ratios) is insufficient for effective electronic use.23 Consequently, attaining both an adequate on- and off-current state ratio alongside superb carrier mobility in these substances presents a crucial challenge.

Nonetheless, the N column presents a compelling alternative for synthesizing 2D black phosphorus (puk-P), arsenene (As, Z = 33), antimonene (Sb, Z = 51), and bismuthene (Bi, Z = 83).24–27 In contrast, much research has concentrated on multilayer structures made solely of elemental phosphorus (P, Z = 15) among these materials.28–30 Furthermore, optical examinations29,31 have revealed that the energy band gap (Eg) of puk-P varies from 0.5 to 2 eV, contingent upon the number of sheets present.32–34 This gap is notably higher compared to its bulk counterpart, exceeding it by approximately ≈ 0.3 eV.24,32,35 In addition, the puk-P-based FET displays an adequate on/off ratio of 105 and excellent carrier mobility of approximately ≈ 103 cm2 V−1 s−1.24,36,37 Despite its potential as a 2D semiconductor, puk-P rapidly oxidizes in the presence of light, air, and humidity.38 However, previous research has confirmed that its stability can be preserved using thin films of Al2O3.39

The distinctive characteristics of monolayer puk-P have propelled it into a rapidly evolving domain within condensed matter and materials science, with particular emphasis on its novel allotrope, blue phosphorene (buk-P).40 Initially foreseen through first-principles (FP) calculations, buk-P has been successfully fabricated on an Au(111) substrate via the MBE approach.41 Furthermore, it adopts a structure akin to that of graphene, germanene, arsenene, stanene, and silicene,17,23,25,42–44 characterized by a low-buckled honeycomb geometry. Notably, the synthesis of these compounds in group-IV often involves strong interfacial interactions with substrates, which can disrupt crucial pz-orbital Dirac excitations. However, previous research has demonstrated the existence of robust Dirac states arising from the px and py orbitals (px,y-orbitals) in silicene grown on a 4H-SiC(0001) substrate.45 These states exhibit exceptional tunability and a greater degree of control over their electronic properties. Moreover, the stronger spin–orbit coupling (SOC) observed in heavier group-IV elements suggests the potential of realizing an exotic QSHE at room temperature. This result is expected to be applicable to germanene, arsenene, and stanene. Consequently, buk-P holds promise for fulfilling the expectations established for the aforementioned materials, especially regarding the existence of massless Dirac excitations.

Numerous examinations have been carried out to elucidate its mechanical, electrical, and optical properties.41,46–50 According to these theoretical analyses, the monolayer buk-P demonstrates an indirect Eg of approximately ≈ 2.0 eV. In addition, the carrier mobility is estimated to be approximately ≈ 4.6 × 102 cm2 V−1 s−1 for the armchair axis and ≈ 4.7 × 101 cm2 V−1 s−1 for the zigzag direction. Moreover, its Eg lies within the range of the visible light spectrum (1.65–3.1 eV), suggesting its promise for optoelectronic applications as a 2D system.41,48,51–53 Furthermore, recent experimental studies have delved into determining the Eg of a monolayer utilizing a set of DFT, ARPES, and STS techniques. The reported results (0.8–1.1 eV)41,51,54,55 diverge from the theoretical FP predictions.40,56 In addition, the presence of Dirac electronic features at the K point of the first Brillouin zone (BZ) has been experimentally demonstrated,57 aligning with theoretical predictions indicating a Fermi velocity of approximately vf ≈ 0.85 × 106 m s−1,58–61 which broadens the potential uses of this system.

The emergence of Dirac excitations predominantly results from the contributions of the P atom s- and px,y-orbitals at the K point. In addition, it has been confirmed that they exhibit exceptional adaptability along the ΓK direction while maintaining their degeneracy.61 Notably, the in-plane orientation of the px,y-orbitals minimizes their interaction with the substrate. Consequently, these Dirac states exhibit reduced sensitivity to the substrate, potentially leading to enhanced robustness for practical applications. Recently, numerous 2D configurations, including X-hydride/halide (X = N–Bi),62 functionalized germanene,63 hydrogenated and halogenated buk-P,59 hydrogenated borophane,21 specific organic substances,64–66 and fluoridated tin films,16,67 featuring Dirac fermions originating from px,y-orbitals have been demonstrated. These studies underscore the diversity of Dirac excitations and their prevalence across various material platforms. Despite considerable advancements in the fabrication of these compounds, the experimental validation of 2D Dirac systems has remained limited.68,69

The remarkable progress achieved in monolayer buk-P research necessitates a thorough investigation of its bilayer counterpart. The latter configuration presents a unique opportunity to explore potentially novel properties that may not be present in monolayer form. Nonetheless, the present surge of interest and anticipated importance attributed to the px,y-orbital Dirac excitations in the honeycomb configuration demand a thorough and profound understanding of these phenomena. Moreover, carrying out an in-depth investigation and attaining a profound perception of the nature of these Dirac electronic features is crucial for thoroughly elucidating their underlying physical principles and harnessing their promising attributes across a diverse spectrum of technological uses. Therefore, the emergence of Dirac-like excitations in the AA (hexagonal) stacked bilayer system within the 5NN in- and 2NN out-of-plane coupling MOTB approach has been explored. Here, we aim to improve upon the few recently presented interpretations and quantitatively enhance the multilayer buk-P overall TB picture through the incorporation of more distant interacting sites.70 The incorporation of distant neighbors beyond the 2NN in-plane and NN interlayer couplings is crucial for representing the energy dispersion throughout the BZ. Thus, the quantitative deviation of the 2NN intralayer and NN out-of-plane coupling MOTB calculations can be enhanced. Note that our models for the unstrained bilayer (monolayer) closely align with the most effective self-energy calculations based on the GW method. In addition, this effective MOTB Hamiltonian provides an intuitive understanding of the px,y-orbital Dirac states and distant transfer integral processes both in- and out-of-plane. However, for the first time, the AA (AB) stacked bilayer electronic band structure has been reported utilizing the aforementioned approximations.

This article is structured as follows: in Section II, we explore the structural configuration of the monosheet and derive its 5NN intralayer interaction MOTB dispersion bands to verify the applicability of our 5NN in- and 2NN out-of-plane coupling model. We dedicate Section III entirely to the discussion of the electronic MOTB Hamiltonian and the resulting energy dispersion representing the valence orbital electrons of the bilayer (monolayer) buk-P system. This section considers various parameters coupling atomic orbitals (AOs), distant transfer integrals, and the neglected overlap. Furthermore, the presence of Dirac electronic features in the AA bilayer configuration has been investigated through an applied symmetrical tensile strain of the low-buckled honeycomb formation. In addition, the spin–orbit interaction in the aforementioned system has been discussed. Finally, in Section IV, we draw our conclusions and final remarks. The details of the calculations conducted within the MOTB approximations are provided in the ESI.

II. Crystal structure of monolayer and bilayer blue phosphorus

A. Bravais lattice

Before investigating the electronic band structure of a bilayer (monolayer), it is essential to explore its structural formation, as it significantly influences the energy levels. As illustrated in [Fig. 1 and bottom panel, side view], the monolayer forms a periodic hexagonal lattice configuration consisting solely of P atoms arranged within a 2D plane. Furthermore, it adopts a low-buckled arrangement consisting of two non-equivalent triangular sublattices with distinct bond orientations, denoted here as Anj and Bnj sublattices (n = 1–5 and j = 1–6 or 1–3). The interatomic distance b is approximately ≈ 2.26 Å, which coincides with the distance of the first A1j-NNs (j = 1–3). The unit cell consists of two P atoms with the lattice parameter a = |a1| = |a2| = 3.28 Å.71–73 Furthermore, a1 and a2 represent the unit cell vectors, with their components relative to the central atom B0 given as a/2 (1, √3) and a/2 (1, −√3), respectively.43,61,74 Typically, within the 5NN intralayer coupling MOTB model, each P atom has three NNs (j = 1–3) belonging to the other sublattice, six 2NNs (j = 1–6) from the same sublattice, three 3NNs (j = 1–3) belonging to the other sublattice, six 4NNs (j = 1–6) from the other sublattice, and six 5NNs (j = 1–6) belonging to the same sublattice. The two sublattice planes are spaced by a distance of approximately hBA ≈ 1.23 Å [bottom panel of Fig. 1]. In addition, d11, d12, and d13 represent the lattice displacement vectors (further technical aspects of the computations are provided in the ESM). Note that the crystal lattice arrangement differs from that of puk-P, which exhibits a puckered orthorhombic layered formation.29 Nonetheless, puk-P can transform into buk-P via a specific sequence of adjustments.40
image file: d4cp01988a-f1.tif
Fig. 1 Structural geometry of the monosheet buk-P (Bravais lattice). The orange (green) atom represents the B-sublattice (A-sublattice) within the 5NN intralayer coupling MOTB model. The red rhombus is the unit cell, a1 and a2 are its lattice vectors. Furthermore, the B0 site is the origin and has three NNs (A1j, j = 1–3), six 2NNs (B2j, j = 1–6), three 3NNs (A3j, j = 1–3), six 4NNs (A4j, j = 1–6), and six 5NNs (B5j, j = 1–6).

B. Monolayer blue phosphorene multi-orbital tight-binding model

The elemental P possesses the electronic configuration [Ne] 3s2 3p3. Note that we only consider the valence orbital states (3s and 3px,y,z) that participate in covalent bonding. Initially, to verify the applicability of the model, we derive the NN coupling MOTB interpretation of the monolayer. Subsequently, we assume that the primary contribution to off-diagonal matrix element M1 arises solely from the transfer integrals between A1j-NNs. Therefore, Bloch's wave function image file: d4cp01988a-t1.tif can be written as follows:
 
image file: d4cp01988a-t2.tif(1)

The valence AOs s, x, y, and z are represented by n, while m denotes A1j (an atom in sublattice A at a given site) and B0 from the two different sublattices. In addition, N represents the number of unit cells, [small tau, Greek, vector]m represents the lattice translation vector, [k with combining right harpoon above (vector)] represents Bloch's wave vector, image file: d4cp01988a-t3.tif represents the phase factor, and ϕn([r with combining right harpoon above (vector)][small tau, Greek, vector]m) represents an atomic wave function. Moreover, Bloch's molecular orbital (MO) wave function is constructed by summing up the individual Bloch's wave functions. This summation incorporates contributions from various Bloch sums, which account for the periodic nature of the lattice. Mathematically, Bloch's MO wave function can be written as:

 
image file: d4cp01988a-t4.tif(2)

We must first determine the values of E(k) and the coefficients cn,m in order to solve eqn (3). The variational method is one way to achieve this, as demonstrated in eqn (4):

 
image file: d4cp01988a-t5.tif(3)
 
image file: d4cp01988a-t6.tif(4)

The system of equations resulting from multiplying both sides of eqn (3) yields eight equations that depend on cn,m and E(k), which can be solved in matrix form as a generalized eigenvalue problem. Note that solving the Hamiltonian matrix can be accomplished using Matlab, Python, Mathematica, GNU Octave, Fortran, or Julia. In this paper, the MOTB computations are all performed employing self-developed Matlab scripts, and the resulting energy dispersion levels are illustrated in [Fig. 2 and 4–6]. Therefore, the following can be used to represent the effective Hamiltonian matrix:

 
image file: d4cp01988a-t7.tif(5)


image file: d4cp01988a-f2.tif
Fig. 2 Electronic band levels of the monosheet buk-P along the high-symmetry points ΓMK axis within the 5NN intralayer interaction MOTB approach. The highlighted inset corresponds to the reciprocal lattice with b1 and b2 as its vectors. The red dots represent the other two energies that resemble VBM along the ΓM axis.

To derive the 4 × 4 matrices M0 and M1 within the NN coupling MOTB model, we consider the three A1j-NNs and sum over their lattice displacement d1j, as shown in [Fig. 1 upper panel, top view]. Note that superscript t in matrix M1 represents the transpose sign or operation, and the asterisk denotes the complex conjugate. The Hamiltonian in eqn (5) is represented by the 8 × 8 matrix with B3s, B3x, B3y, B3z, A3s, A3x, A3y, and A3z as the basis.

 
image file: d4cp01988a-t8.tif(6)
 
image file: d4cp01988a-t9.tif(7)

Moreover, the following can be used to define the geometrical and phase factors:

 
image file: d4cp01988a-t10.tif(8)

The diagonal entries in eqn (6) represent the on-site energy levels of the valence electrons. Here, we adopt the values reported in recent research.61,70,74,75 Conversely, the hopping integrals of the NN coupling matrix between the A1j and B0 positions are represented by the off-diagonal elements in eqn (7). The electronic band structure of the monosheet can be derived by diagonalizing the resulting 8 × 8 secular equation as follows:

 
image file: d4cp01988a-t11.tif(9)

However, by utilizing the hopping integral parameters provided by the TB-DFT approximation, we establish our 5NN intralayer interaction MOTB calculations. The selection of 5NN is based on achieving a balance between computational efficiency and accuracy. It captures the necessary interactions within the layer without introducing excessive computational complexity that comes with higher-order in-plane interactions. Therefore, the quantitative accuracy of the model depends on the next-NN hopping energies, which are intricately linked to the atomic arrangement and the underlying crystal lattice structure of the material. Note that the hopping integral parameters exponentially approach zero as the distance between the NN sites increases. Consequently, the contribution of higher-order integrals becomes negligible (6NN ≈ 0 eV).

Hence, within the framework of the 2NN interaction MOTB description, the Hamiltonian in eqn (5) can be formulated as follows:

 
image file: d4cp01988a-t12.tif(10)

The new 4 × 4 matrix M2 represents the 2NN interactions between the B0 and B2j substances. Note that from here onward, we consider that the transfer parameter Vllq exponentially depends on the vector positions dnj:74,76,77

 
image file: d4cp01988a-t13.tif(11)
Here, b is the BnjAnj (adjacent) distance, l and l′ denote the AOs of Bnj and Anj, consecutively. Furthermore, u0 = (ab)/ln[thin space (1/6-em)]10 represents the decay length. The quantum number q represents the liaison type (σ and π) along the direction connecting the Anj and Bnj atoms, where V(0) represents the first NN transfer integral parameter. Moreover, when considering up to 5NN in-plane interactions, the Hamiltonian in eqn (10) can be expressed as follows:
 
image file: d4cp01988a-t14.tif(12)

The newest 4 × 4 matrices M3, M4, and M5 can be considered the third-, fourth-, and fifth-NN in-plane interactions between the central and A3j, A4j, and B5j atoms, consecutively. However, the primitive reciprocal lattice vectors are represented by b1 and b2, with their components relative to the rectangular coordinate system (kx, ky) provided by 2π/a (1, 1/√3) and 2π/a (1, −1/√3), respectively. The corresponding BZ is a Bravais honeycomb lattice, as depicted in the highlighted inset [Fig. 2] with the high-symmetry points located at Γ = (0, 0), M = 2π/a (0, 1/√3), and K = 2π/a (1/3, 1/√3).

III. Results and discussions

A. Monolayer energy dispersion

The electronic bands of the monosheet shown in [Fig. 2] accurately match the preceding TB-DFT and GW representations.40,43,71,78–82 This consistency validates the accuracy of our 5NN MOTB approach and offers a robust foundation for subsequent representations of the electronic band structure of the bilayer. Several factors contribute to the observed agreement. Fitted hopping integral parameters, particularly those capturing interactions beyond 2NN, might inadvertently account for electron–electron interactions considered in GW. Furthermore, the localized nature of AOs in MOTB models allows the inclusion of relevant physics with a gradual decay of interactions over distance. The assumed orthogonality of AOs simplifies calculations without compromising accuracy, as long as the orbitals remain orthogonal. In addition, the effective mass approximation employed in the models might be justified for longer-range interactions in specific scenarios. However, between the 5th and 6th energy levels, an indirect Eg emerges at the high-symmetry ΓK direction, where the Fermi level is located along with several possible electronic transitions (points resembling VBM) situated at the ΓM and ΓK regions. Therefore, the VBM appears to be degenerated in three different locations, showing various possibilities for other indirect electronic transitions. Moreover, when the 5NN hopping integrals are included, the overall VBM (CBM) picture is substantially enhanced, and their energies occur at ΓK and ΓM regions consecutively, with an Eg of approximately ≈ 3.00 eV. Note that the CBM is shifting from the ΓK region in the 2NN interaction MOTB dispersion to the ΓM region in the 5NN coupling MOTB representation.

B. Bilayer blue phosphorene

In an investigation conducted by Zhu et al.,40 the layered structure (stacking a few monolayers) of buk-P was predicted. Note that the sheets are stacked together through weak interlayer interactions. Unlike graphene, which only has two stacking forms, the monolayer of buk-P has additional stacking sequences between the neighboring layers due to the low-buckled form and the fact that the Anj and Bnj positions of the honeycomb lattice in the same sheet are situated at distinct planes.71,80,81,83 In [Fig. 3], we present two commonly stacking combinations, denoted as AA [Fig. 3(a) and bottom panel, side view] and AB (Bernal) [Fig. 3(b) and bottom panel, side view]. In the first configuration, the upper and lower sheets possess the same buckling sequence. On the other hand, in the AB stacking arrangement, the Anj+ atoms are situated below the center of a hexagon [as indicated by the dashed lines in Fig. 3(b) upper panel, top view]. Meanwhile, the Anj atoms are directly located over the Bnj+ atoms in the bottom layer.
image file: d4cp01988a-f3.tif
Fig. 3 (a) and (b) Plans (upper panel, top view) of the crystal structure of the AA (AB) stacked bilayer buk-P. Atoms Anj and Bnj on the upper layer are shown as green and orange circles, consecutively. Anj+ and Bnj+ on the lower layer are blue and gray circles, respectively.

This section outlines the MOTB description of the bilayer buk-P systems. To do so, the previously discussed model in eqn (12) is adopted. The unit cell depicted in [Fig. 3 (red rhombus)] includes four atoms, the A1j (B0) from the upper layer and A1j+ (B0+) from the lower sheet. The primitive lattice vectors a1 and a2, as well as the lattice constant a, are identical to those of the monosheet (the lattice constant has been preserved, independently of the stacking mode, alongside the buckling height). Thus, the reciprocal lattice and first BZ are identical to those of the monosheet. Moreover, the AA and AB interlayer distances Δint are 3.24 and 3.21 Å, respectively.70

1. AA stacking. In a basis with the elements A1j (B0) and A1j+ (B0+), the transfer integral parameter matrices can be derived as follows:
 
image file: d4cp01988a-t15.tif(13)

The top-left and bottom-right 8 × 8 blocks Hmono(k) represent the behavior within the top Anj (Bnj) and bottom Anj+ (Bnj+) sheets, consecutively. On the other hand, the top-right and bottom-left 8 × 8 blocks of the Hamiltonian describe interlayer interactions, where Hl is given by:

 
image file: d4cp01988a-t16.tif(14)

In this stacking configuration, the new 4 × 4 matrices M2′, M3′, Mzeros, and M5′ in eqn (14) can be regarded as the NN and 2NN out-of-plane interactions between B0 (B0+), B0 (A1j+), A1j (B0+), and A1j (A1j+), consecutively. Note that Mzeros, as indicated, is a matrix of zeros, and M2′ describes the NN interlayer coupling (as well as M5′). Furthermore, the resulting energy levels are illustrated in [Fig. 4(a)]. The overall electronic band levels are similar to those of the monosheet, with each monolayer energy level divided into two bands by an energy that is approximately equal to the out-of-plane couplings. A comparison with preceding theoretical calculations and experimental studies shows good matching with the findings.46,71,80,81 This approach offers a computationally efficient alternative to more complex methods.


image file: d4cp01988a-f4.tif
Fig. 4 Electronic band structure (a) of the AA bilayer system and (b) the AB stacked formation within the 5NN intra- and 2NN interlayer coupling model along the ΓMK axis.

Despite being significantly weaker than the in-plane couplings, the out-of-plane interactions present profound differences between the electronic and transport characteristics of monolayer and bilayer buk-P systems. However, the numerous VBM (CBM) with comparable energy values indicate various potentials of other indirect electronic transitions. These transitions can have a significant influence on optical and electrical features, leading to interesting phenomena (excitonic effects, photovoltaic responses, and nonlinear optical behavior). Thus, further studies are necessary to provide deeper insights. Note that the NN interlayer couplings qualitatively show good overall matching of the electronic band structure. Consequently, the incorporation of the 2NN interlayer couplings considerably enhances the overall VBM (CBM) picture with FP computations.46,70,71,80,81 Furthermore, the VBM and CBM energies occur at the ΓK and ΓM regions consecutively, with an Eg of approximately ≈ 1.44 eV. The incorporation of distant neighbors beyond the 2NN in-plane and NN interlayer couplings is crucial for representing the energy dispersion throughout the BZ.

2. AB stacking. In our AB stacking form, where the lower sheet stacks with the shift a/2 (1, −√3) in the xy plane with respect to the first one, only atoms A1j of the upper layer and atoms B0+ of the lower layer are directly connected by tunneling [as shown in Fig. 3(b), side view with solid vertical black line].
 
image file: d4cp01988a-t17.tif(15)

The new 4 × 4 matrices Mzeros, M2′, M4′, and M5′ in eqn (15) can be regarded as the NN and 2NN out-of-plane interactions between B0 (A1j+), B0 (B0+), A1j (B0+), and A1j (A1j+), consecutively. Moreover, the resulting energy dispersion band is shown in [Fig. 4(b)]. Furthermore, the VBM and CBM energies occur at ΓK and ΓM regions consecutively, with an Eg of approximately ≈ 1.48 eV.

The overall electronic band levels are similar to those of a monolayer, where each energy level is divided into two bands. Thus, the split of the energy levels originates from the atomic positions. Note that Eg in the AA (AB) stacked bilayer reduces because of the weakening of the confinement effect. In a single-layer 2D material, the electrons are strongly confined to move in only two dimensions, leading to a large Eg and unique electronic properties. However, as more layers are added, the confinement becomes weaker, and electrons are able to spread out in the third dimension (reduced Coulomb repulsion). This increases the overlap between the electronic wave functions in adjacent layers, which leads to the formation of new electronic states that modify the band structure. Specifically, it modifies the effective mass of the electrons, which in turn affects the electronic band structure. These effects combined to produce a narrower Eg for multilayer 2D materials compared to their single-layer counterparts. Thus, the interlayer interactions in the AA (AB) stacking bilayer form can modify the electronic structure and contribute to the reduction of Eg. Note that the overall AA features resemble the AB stacking case with some differences due to the specific atomic arrangements, and the Eg in the latter (≈ 1.48 eV) is slightly larger compared to the AA stacking form (≈ 1.44 eV). Furthermore, we demonstrate that, for all examined stacking forms, bilayer buk-P retains its indirect Eg semiconducting feature, as depicted in [Fig. 4], which is, unfortunately, a serious limitation to the use of this 2D material, especially in the solar cell industry, due to the poor efficiency that characterizes indirect Eg semiconductors.49,84 In contrast, it is a suitable candidate for optoelectronic devices.41,51,85

Our findings suggest that the electronic characteristics of all the systems presented are outstanding for various applications due to other possible electronic transitions; however, we observed the existence of unique aspects in the energy dispersion, suggesting that quasi-direct electronic transitions and Dirac electronic features can occur. Therefore, the use of these systems in various electronic devices is applicable. In addition, this demonstrates the suitability of the present 5NN intra- and 2NN interlayer coupling MOTB descriptions for quantitative analysis.

However, it was found that the AA is energetically more stable than the AB stacking sequence.46,71,80,81,83 Furthermore, no experimental details regarding their stacking patterns have been reported to date. This presents a significant knowledge gap that future research should address. Therefore, we propose the AA stacked bilayer buk-P configuration as a more realistic and pertinent reference for investigating the presence of Dirac electronic features in this system. In addition, we anticipate that this concept can be extended to AB and AC configurations. Our MOTB calculations are established by employing the SKH hopping integrals (yet we do not expect accurate findings for some potential reasons, which we shall mention afterward; although this model might not be perfect for presenting highly accurate quantitative descriptions, the adopted interpretation can easily capture qualitative aspects of the energy dispersion). The resulting energy dispersion bands are illustrated in [Fig. 5(a)–(f)]. Although the theoretical introduction of substantial strain into a bilayer buk-P appears feasible, a thorough examination (validation) is paramount to establishing its practical viability. The successful convergence of these approaches will ensure the practical applicability of the proposed strained materials.


image file: d4cp01988a-f5.tif
Fig. 5 (a)–(f) Electronic band structures of AA stacked bilayer along the ΓMK axis within the 5NN intra- and 2NN interlayer coupling MOTB approximations with different lattice parameters a and buckling height hBA. A key characteristic of these electronic bands is the presence of a linear (parabolic) intersection occurring around the ΓK direction close to the Fermi line. The highlighted insets correspond to the zoom-in of the green rectangle above Ef (1.2–3.5 eV) at the high-symmetry K point. Here, Ef is set to zero and is denoted by blue dashed lines.
3. Dirac-like cones. The manifestation of Dirac electronic characteristics in 2D materials is contingent on stringent criteria pertaining to specific symmetries, parameters, Fermi level, and band overlap conditions.61,68,69 Therefore, the realization of Dirac systems is comparatively rare and is confined to a limited set of lattice configurations.86 Typically, the honeycomb emerges as the predominant platform for accommodating these distinctive features.15,42 For instance, in a periodic honeycomb arrangement, the occurrence of Dirac electronic features in the material within the NNs depends on the capability of the three transfer parameters to establish a triangular formation. These hopping integrals are subject to the influence of the atomic structure, notably the bond length, which is closely related to the crystal lattice. Nevertheless, the emergence of Dirac electronic features in monosheet buk-P can be comprehensively elucidated through our effective MOTB model, the intricacies of which are thoroughly expounded in the (ESI).61

The application of biaxial tensile strain induces a significant alteration in the geometrical characteristics of the bilayer system (lattice symmetry, electronic interactions, and the potential experienced by electrons). Consequently, shifting the positions of the VBM and CBM. This results in a strain-induced band gap tuning, with the VBM shifting upwards and the CBM downwards, leading to a reduction in the overall band gap. It is necessary to adjust intralayer and interlayer interactions, thereby inducing Dirac-like cones.12,53,59,82 Overall, the theoretical basis lies in the interplay between strain-induced symmetry breaking and band splitting. Compared to graphene, with its honeycomb lattice structure, which inherently possesses the necessary conditions for Dirac cones, strain-induced Dirac cones can be more tunable but potentially less stable.3,6,8,77 However, this transformation manifests as a notable elongation of bond length b, accompanied by a substantial decrease in buckling height hBA and interlayer distance Δint. As a result of these modifications, the a demonstrate an approximate augmentation of ≈ 21.95%, 28.05%, 29.88%, 32.32%, 40.24%, and 43.29% relative to the pristine form. Moreover, under increasing tensile strain, the bilayer material adopts a quasi-planar configuration, whereby the system angle approaches approximately ≈ 120°, closely resembles the characteristic bonding angle observed in sp2 hybridized configurations. However, this transition from a buckled form to a quasi-planar hexagonal lattice profoundly affects the electronic structure (the energy levels of the s and px,y,z orbitals can be shifted). This critical distinction leads to the emergence of Dirac-like cones that exhibit distinct characteristics for in-plane σ and out-of-plane π bonds.

Our findings, as illustrated in Fig. 5, reveal a key characteristic of the energy band structure of the AA stacking mode under tensile strain. Notably, zero Eg (≈ 1 meV) emerges as a cone-shaped dispersion (linear crossing slightly anisotropic within the model) centered around the ΓK direction upon reaching a tensile strain of approximately ≈ 2.30%, which aligns reasonably well with recent experimental observations,57 as illustrated in Fig. 5(a). This behavior of the energy dispersion bands can be mathematically described by the equation E = ħ·vf·k. Furthermore, the Dirac-like point is located approximately ≈ 2.46 eV above Ef. This upward shift can be attributed to the inherent electron deficiency within the bilayer system. Interestingly, the Dirac-like intersection in Fig. 5(b) shifts to a new position closer to its original location. Moreover, the increased a induces a downward energetic shift of the crossing point, which brings it closer to the Fermi line, as demonstrated in Fig. 5(c). This observation underscores the tunability of the Dirac-like cones along the ΓK direction with respect to Ef by varying the hopping integral parameters. In addition, the deposition of the bilayer system on a substrate enables the modulation of the energy positions of the Dirac points and compensates for the electron deficiency within it. Note that we provide compelling evidence that a tunable Eg can be realized through intralayer strain without compromising the underlying crystallographic symmetry. This translates into the ability to manipulate the energy range in which massless Dirac fermions emerge.

Nevertheless, the energy dispersion bands near the contact points of the CBM and VBM [as illustrated in Fig. 5(b)–(f)] exhibit a parabolic shape (quadratic dependence on Bloch's wave vector k). This implies that electrons residing in the vicinity of these points effectively possess features akin to massless Dirac fermions. Consequently, these points can be considered analogous to Dirac cones. Note that the energy disparity between the CBM and VBM is quantified to be approximately ≈ 0.27–0.4 eV, as demonstrated in the insets of Fig. 5(b)–(f). The energy separation can be modulated through the application of both an electric field and doping.

Clearly, a undergoes a significant increase of 43.29%, transitioning from a = 3.28 Å in the pristine bilayer to a = 4.7 Å observed in our model. However, b experiences a simpler elongation from 2.26 Å to 2.71 Å in the strained representation (a = 4.7 Å), which corresponds to a stretch of approximately ≈ 19.58%. Furthermore, the Δint undergoes a significant decrease of approximately ≈ −33.95%, transitioning from Δint = 3.24 Å to Δint = 2.14 Å observed in the strained model a = 4.7 Å. The Dirac electronic features can be directly attributed to the two equivalent sublattices forming the hexagonal lattice geometry. Through spatial inversion, the sublattices interchange. Importantly, the emergence of a Dirac-like cone is achievable even with a relatively low (2.30%) elongation of the P–P bond. The key is to separate the energy levels of the in-plane s- and px,y-orbitals, and out-of-plane pz-orbitals through the sp2 configuration, yielding respective σ- and π-character Dirac-like cones. Hence, it is evident that the emergence of the distinctive Dirac-like cone spectrum is not solely attributed to strain but is significantly influenced by the geometrical phase transition from a buckled form to quasi-planar geometry. This highlights the critical role of structural modifications in shaping electronic band structures. Note that a further increase in a induces a downward energetic shift and a wider gap. This trend continues until strong correlation effects dominate (the inter-P hopping parameter becomes negligible), which leads to the disappearance of the Dirac electronic features.58 Furthermore, we unveil a key distinction between the px,y- and pz-orbital Dirac-like states. Whereas out-of-plane π Dirac fermions typically exhibit only two linearly dispersed bands near the Dirac intersection, the in-plane σ counterparts possess additional bands. Interestingly, this observation aligns with the established trends reported in prior studies on px,y-orbital Dirac states.59,61,87,88

In strained buk-P, the electronic states near Ef (1.58–2.46 eV) predominantly originate from the px,y-orbitals. Consequently, these states present exceptional tunability and a greater degree of control over their electronic properties. In addition, vf is more sensitive not only to the transfer integral parameters through the σ (π) bonding but also to the alterations of a, hBA, Δint, and buckling angles induced by the applied tensile strain. In bilayer buk-P, the Fermi velocity is generally lower than that in the monolayer. This is due to the weakening of the confinement effect and reduced Coulomb repulsion. Our calculations revealed a remarkably high Fermi velocity of approximately vf ≈ 0.12 × 105 m s−1. This is consistent with the findings of previous experimental and theoretical studies that investigated analogous geometry,89–92 thereby affirming the reliability of the methods utilized in our calculations. However, the key factor governing the formation of Dirac-like cones is the geometrical phase shift from buckled to quasi-planar geometry.58,59,87,88 Thus, bilayer buk-P can be transformed from an indirect Eg configuration into a 2D Dirac-like system by applying tensile strain, an electrical field, doping, or covalent functionalization.

The realization of Dirac cones in experimental settings poses significant challenges. In some cases, the substantial strain values predicted theoretically may not be readily achievable within the limits of practical laboratory techniques.58,59,86 The 3D VBM and CBM bands are further depicted in Fig. 6, where the crucial features of the Dirac cone are prominently displayed along the high-symmetry ΓK direction of the BZ.


image file: d4cp01988a-f6.tif
Fig. 6 3D electronic band structure plotted versus kx and ky in the range of [−2π/a, 2π/a] using Matlab. The inset corresponds to the energy levels in the vicinity of the Dirac-like intersection within the MOTB model. The cone-shaped structure located in the proximity of the ΓK region is slightly anisotropic, and the Dirac-like point was found to be positioned approximately ≈ 2.45 eV above Ef when a stretch of 2.30% is attained.
4. Spin–orbit coupling. Within the framework of the MOTB approximation, the manifestation of Dirac electronic features relies on specific combinations of the hopping integral parameters.93,94 Therefore, the quantitative accuracy of the model depends on the hopping energies, which are intricately linked to the atomic arrangement and the underlying crystal lattice structure of the material. In the case of bilayer buk-P, the application of tensile strain disrupts some of the honeycomb lattice symmetry, leading to enhanced SOC. This slight symmetry breaking at the sublattice level triggers the opening of a tiny Eg of approximately ≈ 0.27 eV and enables tuning of the inversion asymmetry, transforming the massless Dirac fermions into massive particles.62,95 Furthermore, crystal symmetry permits the existence of an intrinsic (on-site) SOC.58,96 Similar observations have been reported in silicene and germanene,42,45,97 where buckling enhances the SOC because of the direct hybridization between the out-of-plane π and in-plane σ orbitals. Moreover, the dominance of the px,y-orbitals over their pz counterparts at low energy levels is predicted to significantly influence the emergence of a large intrinsic SOC. This enhanced overlap can lead to a stronger interaction between the electron's spin and its orbital motion, contributing to the SOC. The px,y-orbitals are more sensitive to relativistic effects, which are the core reason behind SOC. This makes them more susceptible to spin–orbit interactions compared to pz-orbitals. Despite the presence of an Eg, SOC, and tunable inversion asymmetry in the buckled hexagonal structures, the potential to realize intriguing anomalous Hall effects remains.62

Intriguingly, the bilayer exhibits a zero-gap Dirac-like character and presents a distinguishable Dirac-like cone in the absence of a strong SOC, as depicted in Fig. 5(a). However, upon the incorporation of different a, hBA, and Δint values, the Dirac-like cone can be gapped, and a new tiny Eg of approximately ≈ 0.27 eV emerges at the Dirac-like points, as demonstrated in Fig. 5(b). This trend continues with the increased a until the SOC becomes strong, at which point an Eg of approximately ≈ 0.4 eV emerges, as depicted in Fig. 5(f). Notably, variations in buckling strength are expected to exert a significant influence on the magnitude of the SOC. It is important to acknowledge that these variations may also lead to subtle modifications in other parameters within the model.97 The gap opening by the SOC implies the topological nontriviality of the electronic state (anomalous QH state) of the honeycomb lattice,58,96,98 so the configuration can be effectively classified as a quantum spin Hall insulator, expanding the potential applications of this 2D system, particularly in electronic and spintronic devices.99

Although this MOTB interpretation provides a qualitative understanding of the behavior of a dispersion band under tensile strain, its predictions align reasonably well with recent experimental observations.57,100 Therefore, it serves as a valuable foundation for more complex models of advanced physical phenomena (proximity effect and spin relaxation). Note that a perfect free-standing monosheet cannot exist in reality (intrinsic structural instability). Consequently, the ideal Dirac cone-shape observed in the monolayer, as predicted by the representation and supported by the current experiment,57 can only be a theoretical construct. Realistically, the cone will inevitably distort due to symmetry changes. In simpler terms, the cone-like levels demonstrated in the experiment are not true Dirac cones because they do not correspond to massless fermions.95

In contrast to its monolayer counterpart, the bilayer buk-P exhibits greater synthetic accessibility. This, coupled with the presence of theoretically identified Dirac-like electronic features and a near-zero band gap (≈ 1 meV), positions bilayer buk-P as a promising candidate for the next generation of high-performance nanoelectronic devices. Notably, its Eg properties offer a potential solution to overcome the limitations of graphene, which lacks a native band gap. The intriguing behavior observed in bilayer buk-P raises the question of whether similar Dirac-like electronic features can be achieved in few-layer buk-P (e.g., three, four, or five sheets) with reduced strain requirements (flexible substrates). Exploring this possibility through future experimental studies could provide valuable insights into the manipulation of Dirac fermions and potentially lead to novel material properties.

IV. Conclusions

In summary, the MOTB approach offers an alternative to the limitations of FP methods. Furthermore, it provides physical knowledge and computational simplicity by combining the computational effectiveness of TB methods with a less empirical and more direct parameterization than DFT calculations. In addition, it enables sizable computation for edges, surfaces, and disorders. In contrast, ab initio or FP investigations present more accurate data with less input beyond the atomic coordinates, simulation box, and boundary conditions. However, the complex nature of the wave function makes the interpretation of the results difficult and computationally expensive. Note that within the MOTB approach, the massive amount of computational labor required to rigorously compute the numerous matrix components of energy is a challenge. Its semi-empirical nature is a drawback when dealing with complex systems dominated by diverse electronic interactions. In such cases, the method cannot accurately describe the system, and the computational demands associated with numerous integral evaluations can be overwhelming.

We have conducted a comprehensive examination of the valence states of bilayer (monolayer) buk-P configurations within MOTB interpretations. Initially, a preliminary MOTB model was employed, considering only NN couplings. Subsequently, interactions up to the 5NN were incorporated, transforming the unstrained monolayer description from a qualitative to a quantitative interpretation, enabling more rigorous calculations. To validate the reliability of the approximation, the energy dispersion bands reported in previous GW-based calculations were used as a reference. The findings exhibited excellent agreement with the established approach. Incorporating distant neighbors beyond the 2NN intralayer and NN interlayer coupling is crucial for representing the energy dispersion of the configurations. Nonetheless, it has already been demonstrated the nonnegligible influence of these couplings in bilayer (monolayer) systems. Thus, the quantitative deviation of the 2NN intralayer and NN out-of-plane interaction MOTB calculations can be enhanced by incorporating 5NN in-plane and 2NN interlayer interactions. As anticipated, however, the hopping parameters diminish exponentially as the distance between the origin and the NN interacting sites increases. Note that there is always an underestimation when the transfer parameters decrease exponentially as the distance expands. Furthermore, the shift in the interlayer distance due to the stacking sequences is critical for determining the Eg. The behavior of the energy levels for all stacking forms is extremely connected to the coupling of the interlayer orbitals of each sheet. For instance, in the AB stacking mode, the Δint is smaller in contrast to the AA, and consequently, the coupling between the sheets is maximized. Thus, the influence of confinement is stronger in this stacking mode than in the AA. This effect directly impacts the evolution of Eg of the configurations along with the NN couplings.

The AA bilayer exhibits a zero-gap Dirac-like character (Eg ≈ 1 meV) and presents a distinguishable Dirac-like cone in the absence of a strong SOC, demonstrating an approximately linear conical dispersion near the CBM and VBM, centered around the ΓK direction, with a vf ≈ 0.12 × 105 m s−1, upon reaching a tensile strain of approximately ≈ 2.30%. The increased application of symmetrical tensile strain in bilayer buk-P disrupts some of the honeycomb lattice symmetry, leading to enhanced SOC. Furthermore, the tuning of inversion asymmetry at the sublattice level triggers the opening of a tiny Eg of approximately ≈ 0.27–0.4 eV. However, the dominance of the px,y-orbitals instead of their pz counterparts within low energy levels and variations in the buckling strength are predicted to be significant factors in the emergence of a large intrinsic SOC. Note that the gap opening by the SOC implies the topological nontriviality of the electronic state (quantum anomalous Hall state) of the honeycomb lattice. In addition, the key factor governing the formation of Dirac-like cones is the geometrical phase shift from a buckled form to quasi-planar geometry. Thus, bilayer buk-P can be transformed from an indirect Eg configuration into a 2D Dirac-like system by applying tensile strain, an electrical field, doping, or covalent functionalization.

Finally, understanding the formation mechanism and electronic features of a Dirac-like cone offers valuable insights into the design principles of novel Dirac-like materials. In addition, our comprehensive method, combined with experimental findings and FP calculations, holds substantial potential for elucidating the inherent features of Dirac-like configurations beyond bilayer buk-P.

Author contributions

Conceptualization: AB; methodology: AB; software: AB; formal analysis: AB; writing – original draft: AB; writing – review & editing: OM; supervision: OM.

Data availability

All generated data and electronic dispersion derivation equations in the presence of various NNs within the MOTB model, as well as all the Matlab scripts used to generate these data, are freely provided in the ESM.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The research leading to these results has received funding from OCP grant AS70 “Towards phosphorene-based materials and devices”.

References

  1. A. J. Mannix, B. Kiraly, M. C. Hersam and N. P. Guisinger, Synthesis and chemistry of elemental 2D materials, Nat. Rev. Chem., 2017, 1(2), 0014 CrossRef CAS .
  2. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. E. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva and A. A. Firsov, Electric field effect in atomically thin carbon films, Science, 2004, 306(5696), 666–669 CrossRef CAS PubMed .
  3. Y. Zhang, Y. W. Tan, H. L. Stormer and P. Kim, Experimental observation of the quantum Hall effect and Berry's phase in graphene, Nature, 2005, 438(7065), 201–204 CrossRef CAS PubMed .
  4. P. R. Wallace, The band theory of graphite, Phys. Rev., 1947, 71(9), 622 CrossRef CAS .
  5. A. H. Castro Neto, F. Guinea, N. M. Peres, K. S. Novoselov and A. K. Geim, The electronic properties of graphene, Rev. Mod. Phys., 2009, 81(1), 109–162 CrossRef CAS .
  6. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, M. I. Katsnelson, I. V. Grigorieva, S. V. Dubonos and A. A. Firsov, Two-dimensional gas of massless Dirac fermions in graphene, Nature, 2005, 438(7065), 197–200 CrossRef CAS PubMed .
  7. X. Wang and X. Gan, Graphene integrated photodetectors and opto-electronic devices—a review, Chin. Phys. B, 2017, 26(3), 034203 CrossRef .
  8. Y. P. Bliokh, V. Freilikher and F. Nori, Ballistic charge transport in graphene and light propagation in periodic dielectric structures with metamaterials: A comparative study, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 87(24), 245134 CrossRef .
  9. P. Li and Z. X. Guo, The Dirac half-semimetal and quantum anomalous Hall effect in two-dimensional Janus Mn2X3Y3 (X, Y = F, Cl, Br, I), Phys. Chem. Chem. Phys., 2021, 23(35), 19673–19679 RSC .
  10. B. Yang, X. Zhang, A. Wang and M. Zhao, Dirac cones in a snub trihexagonal tiling lattice with reflective symmetry breaking, J. Phys.: Condens. Matter, 2019, 31(15), 155001 CrossRef CAS PubMed .
  11. H. Luo, P. Yu, G. Li and K. Yan, Topological quantum materials for energy conversion and storage, Nat. Rev. Phys., 2022, 4(9), 611–624 CrossRef CAS .
  12. G. Yang, Z. Xu, Z. Liu, S. Jin, H. Zhang and Z. Ding, Strain-and fluorination-induced quantum spin Hall insulators in blue phosphorene: A first-principles study, J. Phys. Chem. C, 2017, 121(23), 12945–12952 CrossRef CAS .
  13. P. Ball, Quantum materials: Where many paths meet, MRS Bull., 2017, 42(10), 698–705 CrossRef .
  14. F. Wilczek, Majorana returns, Nat. Phys., 2009, 5(9), 614–618 Search PubMed .
  15. S. Cahangirov, M. Topsakal, E. Aktürk, H. Şahin and S. Ciraci, Two-and one-dimensional honeycomb structures of silicon and germanium, Phys. Rev. Lett., 2009, 102(23), 236804 CrossRef CAS PubMed .
  16. Y. Xu, B. Yan, H. J. Zhang, J. Wang, G. Xu, P. Tang, W. Duan and S. C. Zhang, Large-gap quantum spin Hall insulators in tin films, Phys. Rev. Lett., 2013, 111(13), 136804 CrossRef PubMed .
  17. M. Ezawa, Monolayer topological insulators: silicene, germanene, and stanene, J. Phys. Soc. Jpn., 2015, 84(12), 121003 CrossRef .
  18. J. E. Padilha, R. H. Miwa and A. Fazzio, Directional dependence of the electronic and transport properties of 2D borophene and borophane, Phys. Chem. Chem. Phys., 2016, 18(36), 25491–25496 RSC .
  19. A. J. Mannix, X. F. Zhou, B. Kiraly, J. D. Wood, D. Alducin, B. D. Myers, X. Liu, B. L. Fisher, U. Santiago, J. R. Guest and M. J. Yacaman, Synthesis of borophenes: Anisotropic, two-dimensional boron polymorphs, Science, 2015, 350(6267), 1513–1516 CrossRef CAS PubMed .
  20. B. Feng, O. Sugino, R. Y. Liu, J. Zhang, R. Yukawa, M. Kawamura, T. Iimori, H. Kim, Y. Hasegawa, H. Li and L. Chen, Dirac fermions in borophene, Phys. Rev. Lett., 2017, 118(9), 096401 CrossRef PubMed .
  21. M. Nakhaee, S. A. Ketabi and F. M. Peeters, Tight-binding model for borophene and borophane, Phys. Rev. B, 2018, 97(12), 125424 CrossRef CAS .
  22. S. Gupta, A. Kutana and B. I. Yakobson, Dirac cones and nodal line in borophene, J. Phys. Chem. Lett., 2018, 9(11), 2757–2762 CrossRef CAS PubMed .
  23. S. Balendhran, S. Walia, H. Nili, S. Sriram and M. Bhaskaran, Elemental analogues of graphene: silicene, germanene, stanene, and phosphorene, Small, 2015, 11(6), 640–652 CrossRef CAS PubMed .
  24. L. Li, Y. Yu, G. J. Ye, Q. Ge, X. Ou, H. Wu, D. Feng, X. H. Chen and Y. Zhang, Black phosphorus field-effect transistors, Nat. Nanotechnol., 2014, 9(5), 372–377 CrossRef CAS PubMed .
  25. H. S. Tsai, S. W. Wang, C. H. Hsiao, C. W. Chen, H. Ouyang, Y. L. Chueh, H. C. Kuo and J. H. Liang, Direct synthesis and practical bandgap estimation of multilayer arsenene nanoribbons, Chem. Mater., 2016, 28(2), 425–429 CrossRef CAS .
  26. J. Ji, X. Song, J. Liu, Z. Yan, C. Huo, S. Zhang, M. Su, L. Liao, W. Wang, Z. Ni and Y. Hao, Two-dimensional antimonene single crystals grown by van der Waals epitaxy, Nat. Commun., 2016, 7(1), 13352 CrossRef CAS PubMed .
  27. F. Reis, G. Li, L. Dudy, M. Bauernfeind, S. Glass, W. Hanke, R. Thomale, J. Schäfer and R. Claessen, Bismuthene on a SiC substrate: A candidate for a high-temperature quantum spin Hall material, Science, 2017, 357(6348), 287–290 CrossRef CAS PubMed .
  28. C. O. Hugh and J. H. Pablo, Two-dimensional crystals: phosphorus joins the family, Nat. Nanotechnol., 2014, 9, 330–331 CrossRef PubMed .
  29. H. Liu, A. T. Neal, Z. Zhu, Z. Luo, X. Xu, D. Tománek and P. D. Ye, Phosphorene: an unexplored 2D semiconductor with a high hole mobility, ACS Nano, 2014, 8(4), 4033–4041 CrossRef CAS PubMed .
  30. A. Castellanos-Gomez, L. Vicarelli, E. Prada, J. O. Island, K. L. Narasimha-Acharya, S. I. Blanter, D. J. Groenendijk, M. Buscema, G. A. Steele, J. V. Alvarez and H. W. Zandbergen, Isolation and characterization of few-layer black phosphorus, 2D Mater., 2014, 1(2), 025001 CrossRef CAS .
  31. S. Zhang, J. Yang, R. Xu, F. Wang, W. Li, M. Ghufran, Y. W. Zhang, Z. Yu, G. Zhang, Q. Qin and Y. Lu, Extraordinary photoluminescence and strong temperature/angle-dependent Raman responses in few-layer phosphorene, ACS Nano, 2014, 8(9), 9590–9596 CrossRef CAS PubMed .
  32. Y. Takao, H. Asahina and A. Morita, Electronic structure of black phosphorus in tight binding approach, J. Phys. Soc. Jpn., 1981, 50(10), 3362–3369 CrossRef CAS .
  33. A. N. Rudenko and M. I. Katsnelson, Quasiparticle band structure and tight-binding model for single-and bilayer black phosphorus, Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 89(20), 201408 CrossRef .
  34. J. Dai and X. C. Zeng, Bilayer phosphorene: effect of stacking order on bandgap and its potential applications in thin-film solar cells, J. Phys. Chem. Lett., 2014, 5(7), 1289–1293 CrossRef CAS PubMed .
  35. C. Q. Han, M. Y. Yao, X. X. Bai, L. Miao, F. Zhu, D. D. Guan, S. Wang, C. L. Gao, C. Liu, D. Qian and Y. Liu, Electronic structure of black phosphorus studied by angle-resolved photoemission spectroscopy, Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 90(8), 085101 CrossRef CAS .
  36. Y. Ma, C. Shen, A. Zhang, L. Chen, Y. Liu, J. Chen, Q. Liu, Z. Li, M. R. Amer, T. Nilges and A. N. Abbas, Black phosphorus field-effect transistors with work function tunable contacts, ACS Nano, 2017, 11(7), 7126–7133 CrossRef CAS PubMed .
  37. F. Yang, Z. Zhang, N. Z. Wang, G. J. Ye, W. Lou, X. Zhou, K. Watanabe, T. Taniguchi, K. Chang, X. H. Chen and Y. Zhang, Quantum hall effect in electron-doped black phosphorus field-effect transistors, Nano Lett., 2018, 18(10), 6611–6616 CrossRef CAS PubMed .
  38. A. Favron, E. Gaufrès, F. Fossard, A. L. Phaneuf-L’Heureux, N. Y. Tang, P. L. Lévesque, A. Loiseau, R. Leonelli, S. Francoeur and R. Martel, Photooxidation and quantum confinement effects in exfoliated black phosphorus, Nat. Mater., 2015, 14(8), 826–832 CrossRef CAS PubMed .
  39. J. S. Kim, Y. Liu, W. Zhu, S. Kim, D. Wu, L. Tao, A. Dodabalapur, K. Lai and D. Akinwande, Toward air-stable multilayer phosphorene thin-films and transistors, Sci. Rep., 2015, 5(1), 8989 CrossRef CAS PubMed .
  40. Z. Zhu and D. Tománek, Semiconducting layered blue phosphorus: a computational study, Phys. Rev. Lett., 2014, 112(17), 176802 CrossRef PubMed .
  41. W. Zhang, H. Enriquez, Y. Tong, A. Bendounan, A. Kara, A. P. Seitsonen, A. J. Mayne, G. Dujardin and H. Oughaddou, Epitaxial synthesis of blue phosphorene, Small, 2018, 14(51), 1804066 CrossRef PubMed .
  42. C. C. Liu, W. Feng and Y. Yao, Quantum spin Hall effect in silicene and two-dimensional germanium, Phys. Rev. Lett., 2011, 107(7), 076802 CrossRef PubMed .
  43. J. M. Galicia Hernandez, H. N. Fernandez-Escamilla, J. Guerrero Sanchez and N. Takeuchi, Electronic and optical properties of the buckled and puckered phases of phosphorene and arsenene, Sci. Rep., 2022, 12(1), 20979 CrossRef CAS PubMed .
  44. D. Dragoman and M. Dragoman, Analogies for Dirac fermions physics in graphene, Solid-State Electron., 2023, 108818 Search PubMed .
  45. P. Li, X. Li, W. Zhao, H. Chen, M. X. Chen, Z. X. Guo, J. Feng, X. G. Gong and A. H. MacDonald, Topological Dirac states beyond π-orbitals for silicene on SiC (0001) surface, Nano Lett., 2017, 17(10), 6195–6202 CrossRef CAS PubMed .
  46. B. Ghosh, S. Nahas, S. Bhowmick and A. Agarwal, Electric field induced gap modification in ultrathin blue phosphorus, Phys. Rev. B: Condens. Matter Mater. Phys., 2015, 91(11), 115433 CrossRef .
  47. J. Zhang and H. Zhang, Tuning of the optical properties of monolayer blue phosphorene, Plasmonics, 2021, 16(4), 1213–1221 CrossRef CAS .
  48. J. Guo, Z. Zhou, T. Wang, Z. Lu, Z. Yang and C. Liu, Electronic structure and optical properties for blue phosphorene/graphene-like GaN van der Waals heterostructures, Curr. Appl. Phys., 2017, 17(12), 1714–1720 CrossRef .
  49. Q. Peng, Z. Wang, B. Sa, B. Wu and Z. Sun, Electronic structures and enhanced optical properties of blue phosphorene/transition metal dichalcogenides van der Waals heterostructures, Sci. Rep., 2016, 6(1), 31994 CrossRef CAS PubMed .
  50. J. P. Duan, J. M. Zhang, X. M. Wei and Y. H. Huang, Electronic, magnetic and optical properties of blue phosphorene doped with Y, Zr, Nb and Mo: A first-principles study, Thin Solid Films, 2021, 720, 138523 CrossRef CAS .
  51. J. L. Zhang, S. Zhao, C. Han, Z. Wang, S. Zhong, S. Sun, R. Guo, X. Zhou, C. D. Gu, K. D. Yuan and Z. Li, Epitaxial growth of single layer blue phosphorus: a new phase of two-dimensional phosphorus, Nano Lett., 2016, 16(8), 4903–4908 CrossRef CAS PubMed .
  52. H. Li, L. Zhang, X. Cai, X. Li, B. Wang, W. Yu and R. Zhao, Metal-nonmetal oscillations in doped blue phosphorene: a first-principles study, Mater. Res. Express, 2018, 5(5), 055007 CrossRef .
  53. J. Zhou, T. Y. Cai and S. Ju, Unusual strain dependence of quasiparticle electronic structure, exciton, and optical properties in blue phosphorene, Phys. Rev. Appl., 2021, 15(2), 024045 CrossRef CAS .
  54. E. Golias, M. Krivenkov, A. Varykhalov, J. Sánchez-Barriga and O. Rader, Band renormalization of blue phosphorus on Au (111), Nano Lett., 2018, 18(11), 6672–6678 CrossRef CAS PubMed .
  55. J. Zhuang, C. Liu, Q. Gao, Y. Liu, H. Feng, X. Xu, J. Wang, J. Zhao, S. X. Dou, Z. Hu and Y. Du, Band gap modulated by electronic superlattice in blue phosphorene, ACS Nano, 2018, 12(5), 5059–5065 CrossRef CAS PubMed .
  56. J. Guan, Z. Zhu and D. Tománek, Phase coexistence and metal-insulator transition in few-layer phosphorene: a computational study, Phys. Rev. Lett., 2014, 113(4), 046804 CrossRef CAS PubMed .
  57. Y. Kaddar, W. Zhang, H. Enriquez, Y. J. Dappe, A. Bendounan, G. Dujardin, O. Mounkachi, A. El Kenz, A. Benyoussef, A. Kara and H. Oughaddou, Dirac fermions in blue phosphorene monolayer, Adv. Funct. Mater., 2023, 33(21), 2213664 CrossRef CAS .
  58. Y. Li and X. Chen, Dirac fermions in blue-phosphorus, 2D Mater., 2014, 1(3), 031002 CrossRef CAS .
  59. M. Sun, S. Wang, J. Yu and W. Tang, Hydrogenated and halogenated blue phosphorene as Dirac materials: a first principles study, Appl. Surf. Sci., 2017, 392, 46–50 CrossRef CAS .
  60. B. Li, C. C. Ren, S. F. Zhang, W. X. Ji, C. W. Zhang, P. Li and P. J. Wang, Electronic Structural and Optical Properties of Multilayer Blue Phosphorus: A First-Principle Study, J. Nanomater., 2019, 2019(1), 4020762 Search PubMed .
  61. A. Benhaij and O. Mounkachi, Dirac fermions in monolayer blue phosphorus: A multi-orbital tight-binding investigation, Mater. Today Commun., 2024, 38, 107917 CrossRef CAS .
  62. S. A. Yang, H. Pan and F. Zhang, Buckled honeycomb lattice materials and unconventional magnetic responses, RSC Adv., 2015, 5(101), 83350–83360 RSC .
  63. C. Si, J. Liu, Y. Xu, J. Wu, B. L. Gu and W. Duan, Functionalized germanene as a prototype of large-gap two-dimensional topological insulators, Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 89(11), 115429 CrossRef .
  64. Z. F. Wang, Z. Liu and F. Liu, Quantum anomalous Hall effect in 2D organic topological insulators, Phys. Rev. Lett., 2013, 110(19), 196801 CrossRef CAS PubMed .
  65. Z. F. Wang, Z. Liu and F. Liu, Organic topological insulators in organometallic lattices, Nat. Commun., 2013, 4(1), 1471 CrossRef CAS PubMed .
  66. Z. Liu, Z. F. Wang, J. W. Mei, Y. S. Wu and F. Liu, Flat Chern band in a two-dimensional organometallic framework, Phys. Rev. Lett., 2013, 110(10), 106804 CrossRef PubMed .
  67. S. C. Wu, G. Shan and B. Yan, Prediction of near-room-temperature quantum anomalous Hall effect on honeycomb materials, Phys. Rev. Lett., 2014, 113(25), 256401 CrossRef PubMed .
  68. L. D. Landau and E. M. Lifshitz, Quantum mechanics: non-relativistic theory, Elsevier, 2013 Search PubMed .
  69. K. Asano and C. Hotta, Designing Dirac points in two-dimensional lattices, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 83(24), 245125 CrossRef .
  70. Y. E. Mogulkoc, M. Modarresi, A. Mogulkoc and Y. O. Ciftci, Electronic and optical properties of bilayer blue phosphorus, Comput. Mater. Sci., 2016, 124, 23–29 CrossRef CAS .
  71. R. B. Pontes, R. H. Miwa, A. J. da Silva, A. Fazzio and J. E. Padilha, Layer-dependent band alignment of few layers of blue phosphorus and their van der Waals heterostructures with graphene, Phys. Rev. B, 2018, 97(23), 235419 CrossRef CAS .
  72. Y. Ding and Y. Wang, Structural, electronic, and magnetic properties of adatom adsorptions on black and blue phosphorene: a first-principles study, J. Phys. Chem. C, 2015, 119(19), 10610–10622 CrossRef CAS .
  73. J. P. Xu, J. Q. Zhang, H. Tian, H. Xu, W. Ho and M. Xie, One-dimensional phosphorus chain and two-dimensional blue phosphorene grown on Au (111) by molecular-beam epitaxy, Phys. Rev. Mater., 2017, 1(6), 061002 CrossRef .
  74. T. Tani, M. Hitomi, T. Kawakami and M. Koshino, Topological edge and corner states and fractional corner charges in blue phosphorene, Phys. Rev. B, 2022, 105(7), 075407 CrossRef CAS .
  75. J. C. Slater and G. F. Koster, Simplified LCAO method for the periodic potential problem, Phys. Rev., 1954, 94(6), 1498 CrossRef CAS .
  76. B. Tsim, N. N. Nam and M. Koshino, Perfect one-dimensional chiral states in biased twisted bilayer graphene, Phys. Rev. B, 2020, 101(12), 125409 CrossRef CAS .
  77. Z. Gong, X. Shi, J. Li, S. Li, C. He, T. Ouyang, C. Zhang, C. Tang and J. Zhong, Theoretical prediction of low-energy Stone-Wales graphene with an intrinsic type-III Dirac cone, Phys. Rev. B, 2020, 101(15), 155427 CrossRef CAS .
  78. C. E. Villegas, A. S. Rodin, A. Carvalho and A. R. Rocha, Two-dimensional exciton properties in monolayer semiconducting phosphorus allotropes, Phys. Chem. Chem. Phys., 2016, 18(40), 27829–27836 RSC .
  79. G. Schusteritsch, M. Uhrin and C. J. Pickard, Single-layered hittorf's phosphorus: a wide-bandgap high mobility 2D material, Nano Lett., 2016, 16(5), 2975–2980 CrossRef CAS PubMed .
  80. J. Arcudia, R. Kempt, M. E. Cifuentes-Quintal, T. Heine and G. Merino, Blue phosphorene bilayer is a two-dimensional metal and an unambiguous classification scheme for buckled hexagonal bilayers, Phys. Rev. Lett., 2020, 125(19), 196401 CrossRef CAS PubMed .
  81. F. İyikanat, E. Torun, R. T. Senger and H. Sahin, Stacking-dependent excitonic properties of bilayer blue phosphorene, Phys. Rev. B, 2019, 100(12), 125423 CrossRef .
  82. Y. Nie, M. Rahman, D. Wang, C. Wang and G. Guo, Strain induced topological phase transitions in monolayer honeycomb structures of group-V binary compounds, Sci. Rep., 2015, 5(1), 17980 CrossRef CAS PubMed .
  83. D. A. Ospina, C. A. Duque, J. D. Correa and E. S. Morell, Twisted bilayer blue phosphorene: A direct band gap semiconductor, Superlattices Microstruct., 2016, 97, 562–568 CrossRef CAS .
  84. H. J. Xiang, B. Huang, E. Kan, S. H. Wei and X. G. Gong, Towards direct-gap silicon phases by the inverse band structure design approach, Phys. Rev. Lett., 2013, 110(11), 118702 CrossRef CAS PubMed .
  85. Y. An, Y. Sun, M. Zhang, J. Jiao, D. Wu, T. Wang and K. Wang, Tuning the electronic structures and transport properties of zigzag blue phosphorene nanoribbons, IEEE Trans. Electron Devices, 2018, 65(10), 4646–4651 CAS .
  86. J. Wang, S. Deng, Z. Liu and Z. Liu, The rare two-dimensional materials with Dirac cones, Natl. Sci. Rev., 2015, 2(1), 22–39 CrossRef CAS .
  87. S. R. Song, J. H. Yang, S. X. Du, H. J. Gao and B. I. Yakobson, Dirac states from px, y orbitals in the buckled honeycomb structures: A tight-binding model and first-principles combined study, Chin. Phys. B, 2018, 27(8), 087101 CrossRef .
  88. C. Wu, D. Bergman, L. Balents and S. Das Sarma, Flat bands and Wigner crystallization in the honeycomb optical lattice, Phys. Rev. Lett., 2007, 99(7), 070401 CrossRef PubMed .
  89. M. E. Dávila and G. Le Lay, Few layer epitaxial germanene: a novel two-dimensional Dirac material, Sci. Rep., 2016, 6(1), 20714 CrossRef PubMed .
  90. D. Coello-Fiallos, T. Tene, J. L. Guayllas, D. Haro, A. Haro and C. V. Gomez, DFT comparison of structural and electronic properties of graphene and germanene: Monolayer and bilayer systems, Mater. Today: Proc., 2017, 4(7), 6835–6841 Search PubMed .
  91. Q. Xiong, W. Cen, X. Wu, C. Chen and L. Lv, Effect of stacking mode on electronic structure and optical properties of bilayer germanene, Solid State Commun., 2023, 360, 115036 CrossRef CAS .
  92. R. Chegel, A. Feyzi and R. Moradian, Electrical and optical conductivities of bilayer silicene: Tight-binding calculations, Int. J. Mod. Phys. B, 2017, 31(22), 1750158 CrossRef CAS .
  93. Y. Hasegawa, R. Konno, H. Nakano and M. Kohmoto, Zero modes of tight-binding electrons on the honeycomb lattice, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 74(3), 033413 CrossRef .
  94. Z. Liu, J. Wang and J. Li, Dirac cones in two-dimensional systems: from hexagonal to square lattices, Phys. Chem. Chem. Phys., 2013, 15(43), 18855–18862 RSC .
  95. I. N. Yakovkin, Dirac cones in graphene, interlayer interaction in layered materials, and the band gap in MoS2, Crystals, 2016, 6(11), 143 CrossRef .
  96. C. L. Kane and E. J. Mele, Quantum spin Hall effect in graphene, Phys. Rev. Lett., 2005, 95(22), 226801 CrossRef CAS PubMed .
  97. C. C. Liu, H. Jiang and Y. Yao, Low-energy effective Hamiltonian involving spin–orbit coupling in silicene and two-dimensional germanium and tin, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84(19), 195430 CrossRef .
  98. C. L. Kane and E. J. Mele, Z2 topological order and the quantum spin Hall effect, Phys. Rev. Lett., 2005, 95(14), 146802 CrossRef CAS PubMed .
  99. Y. F. Zhang, J. Pan, H. Banjade, J. Yu, H. Lin, A. Bansil, S. Du, Q. Yan and M. X. Two-dimensional, Dirac materials and quantum spin Hall insulators with tunable electronic and topological properties, Nano Res., 2021, 14, 584–589 CrossRef CAS .
  100. W. F. Tsai, C. Y. Huang, T. R. Chang, H. Lin, H. T. Jeng and A. Bansil, Gated silicene as a tunable source of nearly 100% spin-polarized electrons, Nat. Commun., 2013, 4(1), 1500 CrossRef PubMed .

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4cp01988a

This journal is © the Owner Societies 2024