Brought to you by:
Paper

Confinement in bilayer graphene via intra- and inter-layer interactions

, and

Published 28 December 2021 © 2021 IOP Publishing Ltd
, , Citation Miguel Castillo-Celeita et al 2022 J. Phys. A: Math. Theor. 55 035202 DOI 10.1088/1751-8121/ac40e1

1751-8121/55/3/035202

Abstract

We consider confinement of Dirac fermions in AB-stacked bilayer graphene by inhomogeneous on-site interactions, (pseudo-)magnetic field or inter-layer interaction. Working within the framework of four-band approximation, we focus on the systems where the stationary equation is reducible into two stationary equations with 2 × 2 Dirac-type Hamiltonians and auxiliary interactions. We show that the localized states are given in terms of solutions of an effective Schrödinger equation with energy-dependent potential. We consider several scenarios where bilayer graphene is subject to inhomogeneous (pseudo-)magnetic field, on-site interactions or inter-layer coupling. In explicit examples, we provide analytical solutions for the states localized by local fluctuations or periodicity defects of the interactions.

Export citation and abstract BibTeX RIS

1. Introduction

In bilayer graphene, two flakes of graphene are close each other such that their electrons can mutually interact. The relative orientation of the two layers can vary. In case of Bernal (or AB-) stacking, the two layers are relatively shifted such that some bonds are parallel in the two lattices. Denoting the atoms in the two triangular sublattices of the jth layer as Aj and Bj , the atoms A2 of the upper layer sit just above the B1 atoms of the lower layer, whereas the atoms A1 are below the centers of the hexagons of the upper lattice. The effective Hamiltonian (in the four-band approximation) for the low-energy particles can be written as [1]

Equation (1)

The basis of the bispinors is (A1, B1, A2, B2). The quantities ${{\epsilon}}_{{A}_{j}({B}_{j})}$, j ∈ {1, 2}, correspond to the on-site energies that can originate from an external electric potential, spin–orbit interaction or an interaction with the substrate. The parameter γ represents interaction of the electrons on the sites A2 and B1. The term proportional to v3 is related to the interlayer trigonal warping [1], which is frequently set to zero in the literature. The four-band Hamiltonian (1) was used e.g. in the analysis of strains and their effect on electronic [24] or topological properties of bilayer graphene [5]. It was used in the analysis of confinement of Dirac fermions in quantum dots formed by doping [6] or by local variation of the interlayer coupling related to local delamination of the bilayer graphene (graphene blisters) studied recently in [7, 8]. It serves well for description of other bilayer Dirac materials, see e.g. bilayer silicene [911]. Qualitatively the same operator with v3 = 0 appears in description of spin–orbit interaction in graphene, see e.g. [12, 13].

Electrons on the binding sites A2 and B1 form dimers. For Eγ, the dynamics on the non-dimer sites A1 and B2 gets dominant, and the effective Hamiltonian can be derived from (1), see [1, 14]. The latter case is known as the two-band approximation, in which the energy operator reads as $H=\left(\right.\enspace \enspace _{{({\pi }^{{\dagger}})}^{2}}^{0}\enspace \enspace {}_{0}{}^{{\pi }^{2}}\left.\right)$. This framework proved to be useful in the analysis of various situations where electric or magnetic fields are inhomogeneous, or the bilayer is subjected to inhomogeneous deformations [1417]. Let us also mention that the exactly solvable two-band Hamiltonians with inhomogeneous magnetic field were constructed [18]. Coherent states for the system described by this operator were found in [19]. In most cases, the studied systems possessed translational or rotational symmetry so that effectively one-dimensional systems were analyzed.

In the current article, we are interested in systems where, besides the on-site interactions and (pseudo-)magnetic field, the inter-layer coupling γ can also be inhomogeneous. We are particularly interested in situations where fluctuations of the involved interactions can confine the Dirac fermions. In this quest, we prefer working within the framework of the four-band approximation. Therefore, we will consider the settings described by the following energy operator

Equation (2)

When compared to (1), this operator can be matched with the Hamiltonian of bilayer graphene for inhomogeneous ${{\epsilon}}_{{A}_{1}}={{\epsilon}}_{{B}_{2}}={{\epsilon}}_{1}(x)$, ${{\epsilon}}_{{B}_{1}}={{\epsilon}}_{{A}_{2}}={{\epsilon}}_{2}(x)$ and with the longitudinal momentum ky = 0. Therefore, (2) can describe dynamics of the fermions that bounce on the potential in normal direction. The operator (2) also contains an additional potential term Ax σ0σ1Ay σ3σ2, A = Ax + iAy . There exist numerical methods for solving the stationary equation associated with (2), such as the Fourier grid Hamiltonian [20] and its extension for matrix Hamiltonians [21]. They provide information in a finite region (local), and the intrinsic numerical error is carried on when computing other physically relevant quantities, such as expectation values. In this regard, analytic solutions prove to be more reliable as they provide global information, which becomes valid on the whole considered domain. Additionally, they are error-free by definition, which makes them outstanding candidates when it comes to implementing perturbative theories. Despite their advantages, it might be a challenging and sometimes impossible task to find them explicitly. Still, there are mathematical techniques that help to find the analytical solutions.

It was discussed recently in [22] that the Hamiltonian (2) belongs to the class of reducible operators where the solution of the associated stationary equation can be found via two, lower-dimensional, dynamical equations with auxiliary interactions. This substantially simplifies the determination of exact solutions. Indeed, let us make an ansatz for the eigenstates of (2) in the following manner,

Equation (3)

Then, the bispinors Ψ and Ξ satisfy 3

Equation (4)

provided that the spinors $\mathbf{\xi }={({\xi }_{1},{\xi }_{2})}^{\text{T}}$ and $\mathbf{\chi }={({\chi }_{1},{\chi }_{2})}^{\text{T}}$ are solutions of the following two equations,

Equation (5)

Equation (6)

with E and $\bar{E}$, in general, different from each other.

These equations resemble the 1 + 1 dimensional Dirac equation up to the matrix coefficient in the kinetic term that contains the nonvanishing constant v3. Let us investigate how the equations (5) and (6) can be solved and what are the physical scenarios that can be studied in this way.

2. Bilayer graphene via Schrödinger equation with energy-dependent potential

For the sake of generality, let us consider the matrix Hamiltonian

Equation (7)

The stationary equation

Equation (8)

gives rise to the set of coupled equations for the spinor components ψ1 and ψ2,

Equation (9)

Equation (10)

with f' ≡ ∂f/∂x . The equations (9) and (10) can be decoupled by fixing

Equation (11)

so that (9) turns into a second-order differential equation for ψ1. In order to bring it into the Sturm–Liouville form, we make an additional energy-dependent transformation

Equation (12)

where

Equation (13)

Notice that the first term of ϕ(x) is not a pure phase so that the normalization of ψ1 and ${\tilde{\psi }}_{1}$ can differ. This should be kept in mind when imposing boundary conditions on the functions ${\tilde{\psi }}_{1}$ in general. The equation for ${\tilde{\psi }}_{1}$ reads as

Equation (14)

Therefore, ${\tilde{\psi }}_{1}$ can be found as the zero-mode of the Schrödinger equation with a potential given in terms of A(x), V1(x), and V2(x). Although it is a rather complicated task to find its solutions in general, one can identify (14) with a Schrödinger equation that possesses a solvable potential V0, $-{\tilde{\psi }}_{1}^{{\prime\prime}}+{V}_{0}{\tilde{\psi }}_{1}=0$. However, by doing so, one has to solve a nonlinear differential equation for either Im A or V2. To overcome this issue, one can fix either Re A(x) or V1(x) so that the potential in (14) coincides with V0. This way, one of the two quantities can be fixed as

Equation (15)

Then we can acquire a solution ${\tilde{\psi }}_{1}$ for a single energy level E. Indeed, changing E would alter the interaction V1(x) so that we would deal with a different physical setting. We will discuss this situation in the section 4.

For constant V2, equation (14) simplifies considerably as the third line in (14) cancels out, leading to

Equation (16)

where

Equation (17)

Equation (18)

The equation (16) corresponds to a stationary Schrödinger equation with an energy-dependent potential term VE (x). The energy-dependent part in (17) cancels out effectively when V1 is constant and either v3 vanishes or Re A is a constant. In the latter case, the term v3(EV2)A0 + (EV1)(EV2) adds a shift to the eigenvalue λE .

Notice that when E = V2, the equations (9) and (10) can be solved directly with

Equation (19)

We can recover the 2 × 2 bilayer Hamiltonians h1 and h2 in (5) and (6), respectively, through the following identification,

Equation (20)

Therefore, if we have

Equation (21)

then there also holds

Equation (22)

Equation (23)

The corresponding bispinor solutions of (4) are

Equation (24)

The bispinors are normalized provided that ψ and χ are normalized. Let us notice that if we can find a fundamental systems of solutions of (5) and (6) for $E=\varepsilon =\tilde{\varepsilon }$, then we can find the fundamental system for (4) for E = ɛ.

When epsilon2 and γ are constants, both h1 and h2 correspond to h with constant V1 and V2 (recall that the explicit form of V2 differs in h1 and h2). In this case, the potential VE and λE in (16) acquire this form

Equation (25)

Equation (26)

For V2 constant, the transformation (12) does not alter integrability, such that when ${\tilde{\psi }}_{1}$ is square-integrable, so is ${\psi }_{1}={\text{e}}^{-\text{i}\phi (x)}{\tilde{\psi }}_{1}$. Therefore, we get solutions of (5) and (6) as in (22) and (23), respectively. The two solutions (3) for the stationary equation (4) then read as in (24).

Now, if we allow epsilon1(x), epsilon2(x) and γ(x) to be inhomogeneous, we can still keep V2 constant in either h1 or h2, provided either epsilon2(x) + γ(x) or epsilon2(x) − γ(x) is constant, respectively. For convenience, let us fix that there holds 4

Equation (27)

In this case, we get constant V2 = epsilon2(x) + γ(x) in h1 whereas h2 corresponds to h with inhomogeneous V2(x) = epsilon2(x) − γ(x), see (20). Notice that V1 = epsilon1(x) can be inhomogeneous now. The equation (5) reduces into (16) whereas (6) leads to (14). It is feasible to solve analytically only one of the two equations by fixing epsilon1(x) appropriately. In either case, the analytical solutions of (4) are either Ψ or Ξ in (24) with V1epsilon1(x). It is worth noticing that the solutions of (16) depend just on epsilon2(x) + γ(x), they do not 'feel' the explicit form of the inter-layer interaction γ(x) and of the on-site coupling epsilon2(x). These interactions can be changed without altering ${\tilde{\psi }}_{1}$ provided the change complies with (27). As we can find only part of the solutions of (4) analytically, we consider the models with inhomogeneous epsilon2(x) and γ(x) that satisfy (27) as quasi-exactly solvable.

Solution of (16) with the energy-dependent potential (17) is nontrivial. Identification of a complete set of solutions to an energy-dependent Schrödinger equation is, if possible, a challenging task in most cases [23, 24]. We will focus on the cases where confinement of Dirac fermions is caused either by A(x), epsilon1(x), or γ(x). The three situations will be discussed in the following three sections separately. Our strategy will be to identify the potential (17) with the potential of a known solvable system. This way, we will be able to identify a set of square-integrable solutions and the corresponding set of eigenvalues λE .

3. Confinement by the vector potential

The potential term Ax σ0σ1Ay σ3σ2 in (1) resembles a magnetic vector potential. However, the corresponding magnetic field would have a different sign on the two layers, which might be physically unfeasible. Nevertheless, it is known that deformations of graphene layers are manifested in the form of the effective (pseudo-)magnetic vector potential [2, 25, 26]. Therefore, we can see A as a combination of the magnetic and pseudo-magnetic field that can acquire different values on the two layers, see e.g. [27].

In this section, we fix epsilon1, epsilon2 as well as γ in (2) to be constant,

Equation (28)

It allows us to convert the task of solving both (5) and (6) into solution of the Schrödinger equation with energy-dependent potential (16). We will identify the latter equation with a stationary equation of known exactly solvable model. This way, we shall find the explicit solutions ${\tilde{\psi }}_{1}$ and energies E of (16). As a consequence, the corresponding solutions of (5) and (6) are found through the relationships (22) and (23), respectively. We will discuss two settings, the harmonic oscillator and Rosen–Morse system.

3.1. Harmonic oscillator case

Let us fix the vector potential A(x) as a complex-valued function linear in x in both its real and imaginary parts, so that

Equation (29)

We can interpret the vector potential term as the consequence of external homogeneous magnetic field and a mechanical strain that gives rise to homogeneous pseudo-magnetic field. Homogeneous pseudo-magnetic field in bilayer graphene was discussed in [26], whereas Dirac fermions in bilayer graphene in presence of homogeneous magnetic field were discussed in [4, 14].

The effective potential VE (x) in (16) takes the form

Equation (30)

Clearly, the latter implies that we must solve the eigenvalue equation of the well-known stationary oscillator. Its general solution can be found once we cast the eigenvalue equation into the confluent hypergeometric equation. It reads explicitly as (for details, see [28])

Equation (31)

Here, 1 and 2 are constant coefficients, ${}_{1}F_{1}[a,b;z]$ stands for the Kummer or confluent hypergeometric function [29], and

Equation (32)

It is worth noticing that the energy-dependent terms in the potential (30) cause just shifted decentering ΩE of the harmonic oscillator and shift the energies λE .

Now, from the asymptotic behavior of the hypergeometric functions, it is straightforward to determine the physical values E for which the function ${\tilde{\psi }}_{1}$ becomes square integrable. This is achieved by imposing a polynomial behavior on the confluent hypergeometric function so that the Gaussian term (31) vanishes faster than the polynomial at x → ±. This leads to an exponentially vanishing function for large |x|. Such a polynomial behavior is achieved if a = −n in ${}_{1}F_{1}[a,b;z]$, with n being a non-negative integer or zero. To simplify the discussion we consider two cases. First, the conditions 2 = 0 and $\left(\frac{1}{4}-\frac{{\lambda }_{E}}{4m{w}_{i}}\right)=-n$ lead to λE = 2mwi (2n + 1/2), from which we obtain ${\tilde{\psi }}_{1}$ in terms of even Hermite polynomials H2n (z). Second, the conditions 1 = 0 and $\left(\frac{3}{4}-\frac{{\lambda }_{E}}{4m{w}_{i}}\right)=-n$ allow us to obtain λE = 2mwi ((2n + 1) + 1/2), and solutions for ${\tilde{\psi }}_{1}$ in terms of odd Hermite polynomials H2n+1(z).

We can thus unify both the even and odd solutions as

Equation (33)

The physical energies En are determined after comparing λE in (32) with (33). We get

Equation (34)

Here, it is worth remarking that the case n = 0 should be addressed with caution, as the exact value of ${E}_{0}^{(\pm )}$ depends on the sign of V1V2v3 A0. That is,

Equation (35)

from which we see that E = V2 for either ${E}_{0}^{(+)}$ or ${E}_{0}^{(-)}$. We can find the spinor corresponding to this energy level from (19). It reads as

Equation (36)

It is not square-integrable as its the second component diverges for x → ±. The latter means that E = V2 is not a physical energy, and we thus identify the point spectrum of h as

Equation (37)

Now, we identify the corresponding spinor to each element in Sp(h). From the general solution (31), and after some calculations, we get

Equation (38)

for n = 0, 1, ..., and ${E}_{0}^{(\pm )}\ne {V}_{2}$, where we have introduced the reparametrized coordinate and decentering shift

Equation (39)

respectively.

Although the Hermite polynomials in (38) depend explicitly on the energy, it is still possible to compute the normalization factor for each spinor. This is done by exploiting the well-known properties of the Hermite polynomials, leading to, up to a global complex-phase,

Equation (40)

which holds true only for the elements in (37).

Heretofore, we have determined the eigenvalue problem related to h, and now the corresponding information for the reduced Hamiltonians h1 and h2 may be extracted directly from the relationships given in (22) and (23). For clarity, we use the notation

Equation (41)

to denote the physical energies of h1 and h2, respectively. The corresponding bispinors Ψn and Ξn of (2) are similarly extracted via (24).

Particularly, we depict the energy levels structure of the bilayer Hamiltonian H in figure 1, for m = 1, A0 = 0, epsilon1 = −1, wr = 2, wi = 1.8, epsilon2 = 1.5, v3 = 0.3, and γ = 0.7. In this configuration, V1V2A0 v3 < 0, and so the energies ${\varepsilon }_{0}^{(+)}$ and ${\tilde{\varepsilon }}_{0}^{(+)}$ are both removed from the spectrum. Similarly, the corresponding bispinors are discarded. The behavior of the probability distributions related to the bispinors ${\mathbf{\Psi }}_{n+1}^{(+)}$, ${\mathbf{\Xi }}_{n+1}^{(+)}$, ${\mathbf{\Psi }}_{n}^{(-)}$, and ${\mathbf{\Xi }}_{n}^{(-)}$ are depicted in figures 1(a) and (b) for n = 0, 1.

Figure 1.

Figure 1. (a) Probability densities ${\mathcal{P}}_{n}^{(\pm )}$ related to the bispinors ${\mathbf{\Psi }}_{n}^{(\pm )}$ for n = 1 (solid-blue) and n = 2 (dashed-red). (b) Probability densities ${\mathcal{P}}_{n}^{(\pm )}$ related to the bispinors ${\mathbf{\Xi }}_{n}^{(\pm )}$ for n = 0 (solid-blue) and n = 1 (dashed-red). (c) Energy levels ${\varepsilon }_{n}^{(+)}$ (solid-blue), ${\varepsilon }_{n}^{(-)}$ (dashed-blue), ${\tilde{\varepsilon }}_{n}^{(+)}$ (solid-red), and ${\tilde{\varepsilon }}_{n}^{(-)}$ (dashed-red). In all the cases, the parameters have been fixed as m = 1, A0 = 0, wr = 2, wi = 1.8, epsilon1 = −1, epsilon2 = 1.5, v3 = 0.3, and γ = 0.7.

Standard image High-resolution image

3.2. Morse–Rosen potential

Now, let us associate A(x) with a smooth step-like profile, defined in terms of a purely imaginary function of the form,

Equation (42)

In this case, the effective potential in (16) becomes

Equation (43)

which corresponds to the MorseRosen interaction [30] (also known as the hyperbolic RosenMorse potential), one of the well-known exactly solvable models in quantum mechanics [28, 3133]. Interestingly, for = 1, the potential (43) reduces to the PöschlTeller interaction, a particular case to be discussed in detail in the upcoming sections. Let us notice that Dirac fermions in (single-layer) graphene in presence of (42) were discussed in [34].

For the rest of this section, we focus on the case ⩾ 1 and κ ≠ 1. We will consider bound states that comply with the boundary condition ${\left.{\tilde{\psi }}_{1}\right\vert }_{x\to \pm \infty }\to 0$. Thus, taking the differential equation for ${\tilde{\psi }}_{1}$ into the hypergeometric form [28], and after some calculations, we get the eigenstates

Equation (44)

Equation (45)

where n = 0, 1, ..., nmax, and ${P}_{n}^{(\alpha ,\beta )}(z)$ stands for the Jacobi polynomials, with

Equation (46)

and the normalization constant [31]

Equation (47)

From (44), it follows that ${\tilde{\psi }}_{1;n}$ is square-integrable only when both αn , βn > 0. This leads us to a condition for the existence of at least one bound state and an upper bound nmax given by

Equation (48)

respectively.

The physical energies En are then determined by comparing λE in (45) with (26), from which we obtain a polynomial equation of second-order for En . We thus get the energies

Equation (49)

The second spinor component ψ2;n is determined from (11), and the corresponding spinors take the form

Equation (50)

with zz(x) := tanh(U0 x).

The exact value of ${E}_{0}^{(\pm )}$ depends on the sign of V1V2. In analogy to the oscillator-like interaction of the previous section, we have ${E}_{0}^{(+)}={V}_{2}$ for V1 < V2, and ${E}_{0}^{(-)}={V}_{2}$ for V1 > V2. The corresponding spinors (19) are not square integrable so that these values do not belong to the point spectrum of h.

We found the following set Sp(h) of discrete energies of h,

Equation (51)

Since αn and βn do not depend on the energy upper-index (±), the upper bound nmax is the same for both energies ${E}_{n}^{(\pm )}$. For nmax = 0, the set of discrete energies is just $\left\{{E}_{0}^{(-)}\right\}$ for V1 < V2, and $\left\{{E}_{0}^{(+)}\right\}$ for V1 > V2. Moreover, the point spectrum may be empty if the inequality in (48) is not fulfilled.

Now, the discrete energy levels associated with h1 and h2 are obtained via (22) and (23),

The bispinor solutions of (4) can be obtained via (24). In figure 2, we illustrate calculated energy levels of h1, h2 and H, and probability density related to the bispinors Ψn and Ξn . In particular, we have considered U0 = 1, κ = 2.5, = 1.1, epsilon1 = −1, epsilon2 = 1.5, v3 = 0.3, and γ = 0.7. In such a case, we obtain nmax = 1, so that we generate two physical solutions for each reduced Hamiltonian. On the other hand, for both h1 and h2, we get V1 < V2, which means that ${E}_{0}^{(+)}$ is discarded from the point spectrum. Each reduced Hamiltonian contributes with three physical energies, and so the bilayer Hamiltonian H contains six energy levels. See figure 2(c).

Figure 2.

Figure 2. (a) Probability densities ${\mathcal{P}}_{n}^{(\pm )}$ related to the bispinors ${\mathbf{\Psi }}_{1}^{(+)}$ (upper solid-blue), ${\mathbf{\Psi }}_{0}^{(-)}$ (lower solid-blue), and ${\mathbf{\Psi }}_{1}^{(-)}$ (lower dashed-red). (b) Probability densities ${\mathcal{P}}_{n}^{(\pm )}$ related to the bispinors ${\mathbf{\Xi }}_{1}^{(+)}$ (upper blue), ${\mathbf{\Xi }}_{0}^{(-)}$ (lower solid-blue), ${\mathbf{\Xi }}_{1}^{(-)}$ (lower dashed-red). (c) Energy levels ${\varepsilon }_{n}^{(+)}$ (solid-blue), ${\varepsilon }_{n}^{(-)}$ (dashed-blue), ${\tilde{\varepsilon }}_{n}^{(+)}$ (solid-red), and ${\tilde{\varepsilon }}_{n}^{(-)}$ (dashed-red). In all the cases, the parameters have been fixed as U0 = 1, κ = 2.5, = 1.1, epsilon1 = −1, epsilon2 = 1.5, v3 = 0.3, and γ = 0.7.

Standard image High-resolution image

4. Confinement by the on-site interactions

In this section, we focus on the case where A(x) in (2) vanishes. The on-site interactions epsilon1(x), epsilon2(x) as well as the inter-layer coupling γ(x) can be inhomogeneous. Let us suppose that epsilon2(x) and γ(x) are related by (27), i.e. V2 = epsilon2(x) + γ(x) is constant. It brings the equations (5) and (6) into

Equation (52)

Equation (53)

As discussed in the section 2, the solutions of (52) can be found via the Schrödinger equation with energy-dependent potential (16). It acquires the following simple form

Equation (54)

where the spinor components of $\xi ={({\xi }_{1},{\xi }_{2})}^{\text{T}}$ are determined from ${\tilde{\psi }}_{1}$ through

Equation (55)

see (22) and (23). The equation (53) reduces into (14). With current fixing of the quantities and denoting $\mathcal{B}(x)={{\epsilon}}_{2}(x)-\gamma (x)$, the latter equation reads as

Equation (56)

where ${\chi }_{1}=\sqrt{E-\mathcal{B}(x)}{\text{e}}^{-\text{i}\frac{{v}_{3}}{2}\left(x-{\int }^{x}\mathcal{B}(s)\mathrm{d}s\right)}{\tilde{\chi }}_{1}$, see (12) and (23).

We find it physically reasonable to consider the systems where epsilon1(x) is bounded. We shall match either (54) or (56) with the stationary equation of a solvable quantum systems that meets these requirements. Let us consider the stationary equation of the Pöschl–Teller system

Equation (57)

It is worth noticing that Dirac electrons in bilayer graphene were studied in presence of Pöschl–Teller electrostatic potential in [35], see also [36]. The equation (57) is a special case ( = 1) of the Rosen–Morse equation (43) that was discussed in the previous section. Therefore, we can use (44) and write down the square integrable solutions ${\tilde{\psi }}_{1;n}$ of (57),

Equation (58)

where n = 0, 1, ..., nmax = ⌊κ − 1⌋, and αn = (κn − 1). The corresponding eigenvalues λn are

Equation (59)

Now, we shall identify either (54) or (56) with (57). Let us start with (54),

Equation (60)

There are different ways how to fix epsilon1(x), and each of them leads to different values of E. Let us discuss some of them.

4.1. Case I

To begin with, let us consider the inhomogeneous on-site interaction

Equation (61)

Then (60) is satisfied provided that we fix κepsilon κ and E such that they solve the following two equations,

Equation (62)

That is, the Pöschl–Teller amplitude κepsilon depends explicitly on the energy through

Equation (63)

On the other hand, the second equation in (62) yields to a fourth-order polynomial equation for E of the form

Equation (64)

the solutions of which become unfeasible to obtain in the general setup. Despite such complexity, we can proceed further and obtain some additional information.

The straightforward calculations show that the square-integrable condition (59) still holds in this case, with κκepsilon given in (63). Thus, with the current choice of parameters, the solutions of (57)

Equation (65)

Equation (66)

are square-integrable provided that ${\bar{\alpha }}_{n}$ is positive. In order to keep ${\bar{\alpha }}_{n}$ real, the term inside the square-root of ${\bar{\alpha }}_{n}$ must be positive. Since $\mathcal{A} > 0$, we get $E< {V}_{2}+\frac{1}{4\mathcal{A}}$. Additionally, the requirement ${\bar{\alpha }}_{n} > 0$ is satisfied provided that $E< {V}_{2}-\frac{n(n+1)}{\mathcal{A}}$. Now, the right-hand side of (64) is negative, as it a multiple of $-{\bar{\alpha }}_{n}^{2}$. Therefore, one obtains real solutions of (64) for (V2E) only if the term on the left is negative as well, which is quadratic and convex on (V2E). We obtain $0< ({V}_{2}-E)< \frac{4{V}_{2}}{4+{v}_{3}^{2}}$, for V2 > 0, and $\frac{4{V}_{2}}{4+{v}_{3}^{2}}< ({V}_{2}-E)< 0$, for V2 < 0. Therefore, any real solution of (64) has to lie inside the one of the following intervals

Equation (67)

where in the latter is clear that $\frac{{v}_{3}^{2}}{4+{v}_{3}^{2}}< 1$, for ${v}_{3}\in \mathbb{R}$. The requirement that the intersections are non-empty sets an upper bound for possible values of n. Indeed, when V2 > 0, the intersection is nonempty for nnmax where

Equation (68)

It provides us with an upper bound for the maximum number of physical solutions which is nmax + 1. It is worth noticing that the trigonal warping term acts against the confinement here; the larger is |v3|, the smaller is nmax. When V2 < 0, it is clear that only n = 0 leads to an non-empty intersection of the energy intervals. This yields to E = V2. Nevertheless, the expression (19) suggests that the corresponding spinor is not square integrable. Thus, such a solution is discarded, and no physical solutions are produced for V2 < 0.

Interestingly, even if E has to be found by numerical means, we have been able to extract general information about the spectrum and number of physically allowed solutions. Furthermore, the spinor may be computed explicitly from (55) and (65) as

Equation (69)

where z(x) := tanh(U0 x), and n = 0, 1, ..., nmax. The corresponding set of bispinors follow from (3), which in this case are given through

Equation (70)

To illustrate our results, we consider V2 = 2.3, v3 = 0.1, U0 = 1, and $\mathcal{A}=2.7$, so that real energies lie inside the interval E ∈ (0.005 735, 2.3). Moreover, from (68), the maximum number of physical energies is nmax + 1 = 3. The numerical values for En and ${\bar{\alpha }}_{n}$ are shown in table 1. For each n, we obtain two energies En , which are all real for n = 0, 1, 2, and complex for n ⩾ 3. The physical energies are identified as E0 = 1.425 92, E1 = 0.368 25, and E2 = 0.006 3725 as they render ${\bar{\alpha }}_{n}$ positive. The associated probability distributions are depicted in figure 3(a).

Table 1. Numerical solutions of the characteristic equation (64). We have fixed the parameters as $\mathcal{A}=2.7$, v3 = 0.1, U0 = 1. For h1 we used V2 = epsilon2 + γ = 2.3, whereas for h2 we have V2 = epsilon2γ = 1.7.

  h1 h2
n En ${\bar{\alpha }}_{n}$ En ${\bar{\alpha }}_{n}$
02.301.70
1.425 921.115 551.185 070.780 747
12.112 15−0.629 8281.475 78−0.575 127
0.368 250.837 8840.217 0730.562 499
21.615 15−1.051 170.779 102−0.845 786
0.006 37250.038 2660.079 2946−0.349 208
30.462 606 + 0.648 775i−1.186 08 − 0.378 511i0.089 0298 + 1.035 22i−1.265 95 − 0.625 57i
0.462 606 − 0.648 775i−1.186 08 + 0.378 511i0.089 0298 − 1.035 22i−1.265 95 + 0.625 57i
Figure 3.

Figure 3. Normalized probability distributions ${\mathcal{P}}_{n}={\mathbf{\Psi }}_{n}^{{\dagger}}{\mathbf{\Psi }}_{n}$ (a) and ${\mathcal{P}}_{n}={\mathbf{\Xi }}_{n}^{{\dagger}}{\mathbf{\Xi }}_{n}$ (b) for n = 0 (solid-blue), n = 1 (dashed-red), and n = 2 (dotted-green). In both cases, the parameters have been fixed as in table 1.

Standard image High-resolution image

4.1.1. Particular setup for h2

As we have remarked, the solutions of (54) are insensitive with respect to the explicit form of γ and epsilon2, and thus we can impose some further restrictions in order to obtain information about the reduced Hamiltonian h2. Particularly, if

Equation (71)

the equation (56) brings us back to the system (62), where now ${V}_{2}\equiv \mathcal{B}={{\epsilon}}_{2}-\gamma $. Therefore, we can solve the equation in exactly the same manner as we did in table 1 for different values of V2 now. This is illustrated in the fourth and fifth columns of table 1, where the energy levels are determined for V2 = epsilon2γ = 1.6, v3 = 0.1, U0 = 1, and $\mathcal{A}=2.7$. The corresponding probability distributions associated with the bispinors Ξn are depicted in figure 3(b).

4.2. Case II

Fixing of epsilon1(x) in (61) allowed us to have the on-site interaction independent on n. The price we paid was that the equation (64) had to be solved numerically. Let us consider the other choice of parameters such that (60) is satisfied. We fix

Equation (72)

Additionally, there must hold

Equation (73)

We can see that inclusion of κ into epsilon1(x) lowered the order of κ in the second equation in (73) when compared to (62). We can solve (73) for E and either κ, U0 or V2. As we require epsilon1(x) to be independent of n, we solve the equation for V2, and E,

Equation (74)

${V}_{2}(\mathcal{A},\kappa ,n)$ is a parabola in n with minimum ${V}_{2}({n}_{0})=\frac{(\kappa -1)}{\mathcal{A}}\left(1+\frac{{v}_{3}^{2}}{4}\right)$ at n0 = κ − 1. For each n from the allowed interval n ∈ {0, ..., ⌊κ − 1⌋}, the on-site interaction epsilon1(x) remains the same. Nevertheless, the value of V2 gets changed correspondingly.

For each of this specific configurations, we are able to find a localized solution ${\tilde{\psi }}_{1;n}$. The bispinor solution of (4) is then

Equation (75)

The solution is invariant with respect to the changes of epsilon2(x) and γ(x) that preserve V2, including the case where both epsilon2(x) and γ(x) are constant. When this is the case, the equation (56) reduces into the equation that coincides with (54), yet for V2 = epsilon2γ. If epsilon2 and γ are such that

Equation (76)

then we can get bound state solutions for each equations (52) and (53) with energies ${\epsilon}=E(\mathcal{A},\kappa ,n)$ and $\tilde{\varepsilon }=E(\mathcal{A},\kappa ,\tilde{n})$. The bispinor Ξ corresponding to the latter energy is

Equation (77)

We show density of states of the corresponding bispinors Ψ and Ξ in figure 4.

Figure 4.

Figure 4. Density of probability of Ψ (dashed-red) and Ξ (blue) from (75) and (77), respectively, for n = 0, and $\tilde{n}=1$, κ = 2.1, $\mathcal{A}=1$, U0 = 1, v3 = 0.1.

Standard image High-resolution image

4.3. Case III

Let us relax the condition (27), i.e. both epsilon2 + γ and epsilon2γ can be inhomogeneous. In this case, the equations (5) and (6) lead to (14), yet with different form of V2(x) in each case. As we mentioned in section 2, we can find configuration of epsilon1(x), epsilon2(x) and γ(x) such that (14) is partially solvable. Let us consider the case with V2(x) = γ(x) + epsilon2(x). We fix V1(x) ≡ epsilon1(x) as

Equation (78)

Then the equation (14) reduces into

Equation (79)

Let us identify (79) with the stationary equation of the Pöschl–Teller system again, ${V}_{0}(x)=-\kappa (\kappa -1){U}_{0}^{2}\enspace {sech}^{2}\enspace {U}_{0}x+{U}_{0}^{2}{(1+n-\kappa )}^{2}$. When n is a positive integer n ∈ {0, ⌊κ − 1⌋}, the equation (79) has a localized solution ${\tilde{\psi }}_{1;n}$, see (58). Notice that this solution is independent on the explicit choice of V2(x) and E as they do not appear in (79). Nevertheless, both V2(x) = γ(x) + epsilon2(x) and E affect the form of the bispinor solution (24) via (12),

Equation (80)

In the current setting, E plays rather the role of an interaction parameter. We can tune the interaction epsilon1(x) by changing E such that it confines a bound state with energy equal to E. In order to keep Ψ(x) square-integrable, we require that $\sqrt{E-{V}_{2}(x)}$ is a bounded function.

We fix V2(x) = epsilon2(x) + γ, $\gamma \in \mathbb{R}$. When epsilon2(x) is periodic, epsilon1(x) shares its periodicity up to the last term in (78) that represents a periodicity defect. For instance, if we fix

Equation (81)

then epsilon1(x) reads explicitly

Equation (82)

Equation (83)

We illustrate the interactions for different values of parameters in figure 5 together with density of probability of the bound state.

Figure 5.

Figure 5. Inhomogeneous on-site energy epsilon1(x) (dotted-blue) in (83), epsilon2(x) (dashed-red) in (81), and the probability distribution ${\mathcal{P}}_{\mathbf{\Psi }}={{\Psi}}^{{\dagger}}{\Psi}$ (filled-green). We fixed κ = 2.2, v3 = 0.2, U0 = 0.2, c = 0.1, n = 0, γ = 0.5, and E = 0.5 (left) and E = 1 (right).

Standard image High-resolution image

5. Confinement by the inter-layer interaction

Up to now, the inter-layer coupling γ had rather implicit influence on considered solutions as it was 'hidden' in V2. Let us see whether we can get an analytical solution of confined Dirac fermions by inhomogeneous γ. It is worth noticing in this context that confinement by inhomogeneous γ with rotational symmetry was analyzed numerically in [7, 8]. We focus on the situation where epsilon1 and epsilon2 are constant (in order to eliminate localization by on-site potentials) and γ is inhomogeneous. We fix

Equation (84)

Then the equations (5) and (6) for the spinor components ξj and χj , for j = 1, 2, can be decoupled through the relationships

Equation (85)

leading to the effective energy-dependent Schrödinger equation

Equation (86)

Notice that the difference between the latter equations relies on the sign of the energy-dependent potential, which both coincide qualitatively with (54), therefore, we can follow the same steps as in the previous section. We shall identify both equations in (86) with (57). We set 5

Equation (87)

It has similar form to epsilon1(x) in (61), however, it acquires nonvanishing constant value γ0 asymptotically now. In analogy to the previous section, we identify the set of solutions as

Equation (88)

where ${\mathcal{C}}_{2}$ and ${\mathcal{D}}_{2}$ are the respective normalization factors, and

Equation (89)

together with

Equation (90)

On the other hand, from the relationship

Equation (91)

we extract the energies of h1 and h2 after choosing δ = +1 and δ = −1, respectively.

It is worth to recall that an immediate solution for the energy equation (91) can be found for n = 0 and E = epsilon1 in both cases δ = ±1. Nevertheless, such a solution is discarded as it is not square-integrable.

Likewise in (67), we can obtain the energy intervals in which E takes real values for h1 and h2. First, we should guarantee that ηn and ${\tilde{\eta }}_{n}$ in (89) are real and positive in order to get square-integrable solutions. Next, the left term in (91) should be negative as the equation would have no real solutions otherwise. Combining both results, we get the intervals where the real roots of (91) have to lie,

Equation (92)

Equation (93)

One cannot get square-integrable eigenstates of either h1 or h2 corresponding to real energies for other values of epsilon1. Now, the intervals (93) are nonempty for some values of n only. This way, we get an upper bounds nmax and ${\tilde{n}}_{\mathrm{max}}\enspace $ for the number of bound states of h1 and h2, respectively, that we can obtain this way. They are

Equation (94)

The corresponding bispinors are given by

Equation (95)

for epsilon1 > epsilon2 + γ0 and epsilon1 < epsilon2γ0, respectively, with ψ2;n and χ2;n given in (88).

To illustrate the results presented in this section, let us fix the parameters as ${U}_{0}=\mathcal{A}=1$, epsilon1 = 1.5, epsilon2 = −1.5, and γ0 = 0.3. Since epsilon1 > epsilon2 + γ0, we would expect bound states only for h1, besides the non-physical solution E = epsilon1 = 1.5. Moreover, from (94), one may see that only two bound states can be generated. Such an information may be verified in table 2, where we obtain two physical energies for h1 as E0 = −0.685 308 and E1 = −1.183 15. Although there are more real energies, they do not satisfy the finite-norm condition ηn > 0. The behavior for the corresponding probability densities ${\mathcal{P}}_{n}$ is depicted in figure 6.

Table 2. Energy solutions of (91), together with the finite-norm condition ηn > 0 and ${\tilde{\eta }}_{n} > 0$, for the reduced Hamiltonians h1 and h2. The parameters have been fixed as ${U}_{0}=\mathcal{A}=1$, epsilon1 = 1.5, epsilon2 = −1.5, and γ0 = 0.3.

  h1
n E ηn
0−0.685 3081.060 55
1.50
1−1.183 150.212 643
1.245 11−0.789 447
2−0.873 539−0.880 266
0.381 221−1.330 05
Figure 6.

Figure 6. Probability density related to the bispinors Ψn of (95) for n = 0 (solid-blue) and n = 1 (dashed-red). The parameters has been fixed as in table 2.

Standard image High-resolution image

6. Discussion

In the article, we focused on the systems described by Dirac Hamiltonians of the form (2) that appear in the analysis of bilayer Dirac materials. We were interested in analytical treatment of confined states that can appear due to local fluctuations (61), (72), (87) or periodicity defects (83) of the involved interactions.

We have made use of the fact that equation (4) is reducible in terms of equations (5) and (6) with 2 × 2 Hamiltonians. In section 2, we showed that both equations could be solved through the solutions of the Schrödinger equation (14), whose potential is a nonlinear function of the interactions and their derivatives. We focused on the specific case where (14) can be significantly simplified into a Schrödinger equation with energy-dependent potential (16).

We considered confinement by a combination of external magnetic field and mechanical deformations in section 3 where the energy-dependent Schrödinger equation was identified with the stationary equation of the harmonic oscillator or the Rosen–Morse system. In section 4, we focused on confinement by inhomogeneities of the on-site and inter-layer interactions. We showed that Dirac fermions can be confined by a local fluctuation or periodicity defect of the on-site interaction epsilon1. We demonstrated this fact on the systems with Pöschl–Teller-type interactions (61), (72), or periodic interactions with a localized defect (83). Finally, in section 5, we considered a situation where only the interlayer interaction was inhomogeneous. Here we fixed the trigonal warping term vanishing. It allowed us analytical treatment of decoupled equations with energy-dependent potential (86).

In all the scenarios, we faced the need to solve a Schrödinger equation with energy-dependent potentials. In section 3, the latter equation occurred due to the presence of the trigonal warping term, v3 ≠ 0. If such a term were absent, decoupling of (9) would produce a Schrödinger equation with energy-independent potential. In sections 4 and 5, epsilon1(x), epsilon2(x) or γ(x) were considered inhomogeneous functions to account for the presence of an electrostatic component in the potential term of the reduced equations (5) and (6). It is known [37] that decoupling of components in the stationary equation for 2 × 2 Dirac Hamiltonian with electric potential leads to Schrödinger equation with energy-dependent potential. The problems related to the solution of such an equation are avoided when bound states with zero energy are of interest, see e.g. [3841]. The zero modes in presence of inhomogeneous electric potential and an effective mass were discussed recently in [42]. In this context, the Hamiltonian (2) can be understood as two, coupled 2 × 2 Dirac Hamiltonians with electrostatic potential accompanied by an effective mass term. For these systems, we found localized states with energies distinct from zero.

The solvability of the presented models relies on the specific form of the interactions. As much as it is rather impossible to prepare such fields in the experiments, we can see a broader application of the analytically solvable models. Besides representing a technically feasible setting that demonstrates a specific physical phenomenon (confinement of Dirac fermions in our case), they can be used to analyze physically more relevant interactions via perturbative approach where they can serve as the initial (unperturbed) system. They can also provide reference models for application of numerical methods.

In the light of the computational complexities mentioned, complete analysis of the spectral properties of the presented models would necessarily involve the use of numerical methods. Our aim in this article was different. We demonstrated confinement of Dirac fermions in bilayer graphene by different types of interactions in terms of analytical solutions. Our analytic approach is not restricted to the problems addressed in this paper and can be extended to a broader class of interactions. Indeed, in the presented models, we used exactly solvable models (e.g. harmonic oscillator, Rosen–Morse model) to find localized solutions of (16). However, to demonstrate the confinement, it would be sufficient to consider broader class of systems, where one bound state is known analytically at least. It would be also possible to analyze existence of the confined states via qualitative (e.g. asymptotic) properties of the interactions, following the same steps. Nevertheless, it goes beyond the scope of the present article and it should be addressed in the future.

Acknowledgments

MC-C thanks Department of Physics of the Nuclear Physics Institute of CAS for hospitality. MC-C acknowledges the support of CONACYT, Project FORDECYT-PRONACES/61533/2020. MC-C also acknowledges the Conacyt fellowship 301117. VJ was supported by GAČR Grant No. 19-07117S. KZ acknowledges the support from the project 'Physicists on the move II' (KINEÓ II), Czech Republic, Grant No. CZ.02.2.69/0.0/0.0/18_053/0017163.

Data availability statement

All data that support the findings of this study are included within the article (and any supplementary files).

Footnotes

  • Notice that we denote a* as complex conjugate of a in the article.

  • In the case V2 = epsilon2(x) − γ(x), ${V}_{2}\in \mathbb{R}$, all the analysis below is applicable with minor changes.

  • We have set, without loss of generality, $\mathcal{A} > 0$ to simplify the conditions for the reality of the spectrum. Nevertheless, similar conclusions can be withdrawn if we allow $\mathcal{A}< 0$.

Please wait… references are loading.
10.1088/1751-8121/ac40e1