This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Dirac equation on a catenoid bridge: a supersymmetric approach

Ö. Yeşiltaş Department of Physics, Faculty of Science, Gazi University, 06500 Ankara, Turkey yesiltas@gazi.edu.tr    J. Furtado Universidade Federal do Cariri(UFCA), Av. Tenente Raimundo Rocha,
Cidade Universitária, Juazeiro do Norte, Ceará, CEP 63048-080, Brasil
   J.E.G. Silva Universidade Federal do Cariri(UFCA), Av. Tenente Raimundo Rocha,
Cidade Universitária, Juazeiro do Norte, Ceará, CEP 63048-080, Brasil
Abstract

In this paper we study the Dirac equation for an electron constrained to move on a catenoid surface. We decoupled the two components of the spinor and obtained two Klein-Gordon-like equations. Analytical solutions were obtained using supersymetric quantum mechanics for two cases, namely, the constant Fermi velocity and the position dependent Fermi velocity cases.

I Introduction

The geometry of the graphene layer geim ; novoselov ; katsnelson plays an important role in the electronic structure. The possibility of developing new electronic devices due to the effect of graphene’s curvature is promoting a growing interest in quantum mechanics at curved surfaces. Electronic properties of a graphene nanotorus were studied considering the dynamics governed by the Schrödinger equation GomesSilva:2020fxo , dainamle as well as by Dirac equation Yesiltas:2018zoy , dainam2 . Geometry fluctuations produce pseudomagnetic fields ribbons whose effects can be seen at ripples contijo and corrugated layers corrugated . The curvature at the tip of a conical layer produces a topological phase furtado , whereas helical strips induce chiral properties dandoloff1 ; atanasovhelicoid ; atanasov .

Another interesting geometry studied in the last years is the catenoid geometry. As a minimal surface, two dimensional wormhole geometry is equivalent to a catenoid which are possible solutions for traversable wormholes Wo . In Ref. wormhole ; picak a bridge connecting a bilayer graphene was proposed using a nanotube. In order to obtain a smooth bridge, Ref.Dandoloff , dandoloff suggested a catenoid surface to describe the bilayer and the bridge using only one surface. This can be achieved due to the catenoid curvature which is concentrated around the bridge and vanishes asymptotically spivak . For a non-relativistic electron, the surface curvature induces a geometric potential in the Schrödinger equation. The effects of the geometry and external electric and magnetic fields upon the graphene catenoid bridge was explored in Ref.euclides , where a single electron is governed by the Schrödinger equation on the surface. Incidentally, the influence of a position-dependent mass problem upon the electron on a catenoid bridge was studied in Ref. euclides2 , where it was proposed an isotropic position-dependent mass as a function of the Gaussian and mean curvatures.

The supersymmetry (SUSY) arose as an effort to obtain an unified description of all interactions of nature wess , but it was immediately realized that it provides powerful techniques to obtain analytical solutions for problems in several branches of physics s1 . The factorization method was first introduced by Dirac in order to obtain the spectrum of the quantum harmonic oscillator dirac . Later, Schrödinger applied the same method for the radial part of the Coulomb problem schrodinger . The supersymmetric quantum mechanics (SUSY QM) is an intense field of research and nowadays there is a wide set of Hamiltonians which are analytically solvable through the factorization method sukhatme . On the other hand, there are more representations of SUSY QM such as shape invariance R , Darboux transformations F and Hamilton hierarchies DJ . SUSY QM has received lot of interest for beautiful mathematical insight in both relativistic and non-relativistic quantum mechanics 1 ; 2 ; 3 .

In this paper we study some properties of a relativistic electron constrained to move on a catenoid surface. Unlike the non-relativistic electron, the Dirac electron does not couple directly to the surface curvature. Instead, the momentum operator (covariant derivative) is modified by the surface tetrads and connection 1-forms. In section II, we study the massless Dirac equation on the catenoid and we decoupled the two components of the spinor. As a result, we obtain two second-order Klein-Gordon-like equations with an effective potential depending on the surface curvature and on the effective Fermi velocity. In section III we present briefly the SUSY QM approach which was used to find analytical solutions for the Dirac equation on the catenoid in two particular cases, namely, the constant Fermi velocity case and the position-dependent Fermi velocity case. In section IV we present our final remarks.

II Massless Dirac equation on a catenoid

The catenoid can be described by the meridian u(,)u\in(-\infty,\infty) and the parallel ϕ[0,2π)\phi\in[0,2\pi), as

r=R2+u2(cosϕi^+sinϕj^)+Rsinh1(u/R)k^,\vec{r}=\sqrt{R^{2}+u^{2}}(\cos\phi\hat{i}+\sin\phi\hat{j})+R\sinh^{-1}\left(u/R\right)\hat{k}, (1)

where RR is the radius of the catenoid bridge and u=u(z)=Rsinh(z/R)u=u(z)=R\sinh(z/R) , as depicted in Fig.(1). Using this coordinate system, the 2+12+1 spacetime interval reads

ds2=dt2du2(R2+u2)dϕ2.ds^{2}=dt^{2}-du^{2}-(R^{2}+u^{2})d\phi^{2}. (2)
Refer to caption
Figure 1: Catenoid

Consider an orthogonal moving frame defined by the tangent vectors e^1=ru\hat{e}_{1}=\frac{\partial\vec{r}}{\partial u}, e^2=1R2+u2rϕ\hat{e}_{2}=\frac{1}{\sqrt{R^{2}+u^{2}}}\frac{\partial\vec{r}}{\partial\phi} and the normal unit vector e^3=e^1×e^2\hat{e}_{3}=\hat{e}_{1}\times\hat{e}_{2}. The respective co-frame is given by ω0=dt\omega^{0}=dt, ω1=du\omega^{1}=du, ω2=R2+u2dϕ\omega^{2}=\sqrt{R^{2}+u^{2}}d\phi and ω3=0\omega^{3}=0, where ωa=eμadxμ\omega^{a}=e^{a}_{\mu}dx^{\mu}. Thus, dr=ω1e^1+ω2e^2d\vec{r}=\omega^{1}\otimes\hat{e}_{1}+\omega^{2}\otimes\hat{e}_{2} and the metric can be rewritten as ds2=ηabωaωbds^{2}=\eta_{ab}\omega^{a}\otimes\omega^{b}, being ηab=diag(+1,1,1)\eta_{ab}=diag(+1,-1,-1). Using the torsion-free Cartan equation Ta=dωa+ωbaωb=0T^{a}=d\omega^{a}+\omega^{a}_{b}\wedge\omega^{b}=0 leads to the one-form connections ωba=Γcbaωc\omega^{a}_{b}=\Gamma_{cb}^{a}\omega^{c} of form ω31=RR2+u2ω1\omega^{1}_{3}=\frac{R}{R^{2}+u^{2}}\omega^{1}, ω32=RR2+u2ω2\omega^{2}_{3}=-\frac{R}{R^{2}+u^{2}}\omega^{2} and ω12=uR2+u2ω2\omega^{2}_{1}=\frac{u}{R^{2}+u^{2}}\omega^{2}. Since de^3=ω31e1+ω32e2=Kbaωaebd\hat{e}_{3}=\omega^{1}_{3}\vec{e}_{1}+\omega^{2}_{3}\vec{e}_{2}=K_{b}^{a}\omega^{a}\otimes\vec{e}_{b}, then the second fundamental form KbaK^{a}_{b} has components

Kba\displaystyle K^{a}_{b} =\displaystyle= RR2+u2(1001).\displaystyle\frac{R}{R^{2}+u^{2}}\left(\begin{array}[]{cc}1&0\\ 0&-1\end{array}\right). (5)

Therefore, the catenoid has gaussian and mean curvatures

K=det(Kba)=R2(R2+u2)2\displaystyle K=det(K^{a}_{b})=-\frac{R^{2}}{(R^{2}+u^{2})^{2}} , M=12Kaa=0.\displaystyle M=\frac{1}{2}K^{a}_{a}=0. (6)

Accordingly, the catenoid has negative gaussian curvature vanishing far from the the bridge throat.

Intrinsically, only the ω12\omega^{2}_{1} is non zero. Using the metric compactibility condition ωab+ωba=0\omega_{ab}+\omega_{ba}=0 and using ω12=ωμ12dxμ\omega^{12}=\omega^{12}_{\mu}dx^{\mu}, then

ωϕ12=uR2+u2.\displaystyle\omega^{12}_{\phi}=\frac{u}{\sqrt{R^{2}+u^{2}}}. (7)

The massless Dirac equation on a surface reads

ivFDμΨ=0,iv_{F}\hbar D_{\mu}\Psi=0, (8)

where vFv_{F} is the Fermi velocity. The covariant spinor derivative is Dμ=μΓμD_{\mu}=\partial_{\mu}-\Gamma_{\mu}, where Γμ=14Γμabγaγb\Gamma_{\mu}=\frac{1}{4}\Gamma^{ab}_{\mu}\gamma_{a}\gamma_{b} is the spinor connection. From Eq.(7) the only non-vanishing spinor connection component is Γϕ=14Γϕ12γ1γ2\Gamma_{\phi}=\frac{1}{4}\Gamma^{12}_{\phi}\gamma_{1}\gamma_{2}. In (2+1)-D we can adopt the following representation for the flat γa\gamma^{a} matrices γ0=σ3\gamma_{0}=\sigma_{3}, γ1=iσ2\gamma_{1}=-i\sigma_{2} and γ2=iσ1\gamma_{2}=-i\sigma_{1}. The curved γμ\gamma^{\mu} matrices are related to the flat γa\gamma^{a} matrices by γμ=eaμγa\gamma^{\mu}=e^{\mu}_{a}\gamma^{a}, so that,

γt\displaystyle\gamma^{t} =\displaystyle= e0tγ0=γ0\displaystyle e^{t}_{0}\gamma^{0}=\gamma_{0} (9)
γu\displaystyle\gamma^{u} =\displaystyle= γ1\displaystyle\gamma^{1} (10)
γϕ\displaystyle\gamma^{\phi} =\displaystyle= e2ϕγ2=1R2+u2γ2,\displaystyle e_{2}^{\phi}\gamma^{2}=\frac{1}{\sqrt{R^{2}+u^{2}}}\gamma^{2}, (11)

and the spinor connection is

Γϕ=i2uR2+u2γ0.\Gamma_{\phi}=\frac{i}{2}\frac{u}{\sqrt{R^{2}+u^{2}}}\gamma_{0}. (12)

The massless Dirac equation takes the form

ivF(γ00+γ1u+1R2+u2γ2ϕiu2(R2+u2)γ2σ3)Ψ=0.\displaystyle i\hbar v_{F}\left(\gamma^{0}\partial_{0}+\gamma^{1}\partial_{u}+\frac{1}{\sqrt{R^{2}+u^{2}}}\gamma^{2}\partial_{\phi}-i\frac{u}{2(R^{2}+u^{2})}\gamma^{2}\sigma_{3}\right)\Psi=0.

Therefore, the catenoid geometry leads to a geometric potential as V1(u)=u4(R2+u2)γ1V_{1}(u)=\frac{u}{4(R^{2}+u^{2})}\gamma_{1}, which vanishes asymptotically and exhibits a parity odd behaviour near the throat. The Dirac equation can be written as

iΨt=HDΨ,\displaystyle i\hbar\frac{\partial\Psi}{\partial t}=H_{D}\Psi, (14)

where the Hamiltonian is given by

HD=ivF(σ1(uu2(R2+u2))σ21R2+u2ϕ).H_{D}=-i\hbar v_{F}\left(\sigma_{1}\left(\partial_{u}-\frac{u}{2(R^{2}+u^{2})}\right)-\sigma_{2}\frac{1}{\sqrt{R^{2}+u^{2}}}\partial_{\phi}\right). (15)

The independence of the catenoid metric with respect to time and the angular coordinate ϕ\phi leads to the conservation of energy and of the angular momentum along the direction zz. Thus, the wave function can be written as Ψ(t,u,ϕ)=eiEteimϕψ(u)\Psi(t,u,\phi)=e^{i\frac{E}{\hbar}t}e^{im\phi}\psi(u) what leads to the Dirac hamiltonian

HD\displaystyle H_{D} =\displaystyle= ivF(0uu2(R2+u2)mR2+u2uu2(R2+u2)+mR2+u20).\displaystyle-i\hbar v_{F}\left(\begin{array}[]{cc}0&\partial_{u}-\frac{u}{2(R^{2}+u^{2})}-\frac{m}{\sqrt{R^{2}+u^{2}}}\\ \partial_{u}-\frac{u}{2(R^{2}+u^{2})}+\frac{m}{\sqrt{R^{2}+u^{2}}}&0\end{array}\right). (18)

Using the Hamiltonian in eq.(18), the Dirac equation leads to coupled equations for the two components of the spinor ψ=(ψ1ψ2)\psi=\left(\begin{array}[]{cc}\psi_{1}\\ \psi_{2}\end{array}\right), as

(uu2(R2+u2)+mR2+u2)ψ1\displaystyle\left(\partial_{u}-\frac{u}{2(R^{2}+u^{2})}+\frac{m}{\sqrt{R^{2}+u^{2}}}\right)\psi_{1} =\displaystyle= iEvFψ2\displaystyle-i\frac{E}{\hbar v_{F}}\psi_{2} (20)
(uu2(R2+u2)mR2+u2)ψ2\displaystyle\left(\partial_{u}-\frac{u}{2(R^{2}+u^{2})}-\frac{m}{\sqrt{R^{2}+u^{2}}}\right)\psi_{2} =\displaystyle= iEvFψ1.\displaystyle-i\frac{E}{\hbar v_{F}}\psi_{1}. (21)

Decoupling the system of first-order equations in eq. (21) leads to

ψ1′′+uR2+u2ψ1+U1ψ1\displaystyle-\psi_{1}^{\prime\prime}+\frac{u}{R^{2}+u^{2}}\psi_{1}^{\prime}+U_{1}\psi_{1} =\displaystyle= E22vF2ψ1,\displaystyle\frac{E^{2}}{\hbar^{2}v_{F}^{2}}\psi_{1}, (22)

where U1=(uR2+u2)(uR2+u2)2+(mR2+u2)2(mR2+u2)U_{1}=\left(\frac{u}{R^{2}+u^{2}}\right)^{\prime}-\left(\frac{u}{R^{2}+u^{2}}\right)^{2}+\left(\frac{m}{\sqrt{R^{2}+u^{2}}}\right)^{2}-\left(\frac{m}{\sqrt{R^{2}+u^{2}}}\right)^{\prime} and the prime stands for the derivative with respect to uu. Employing the change on the wave function of the form ψ1=(R2+u2)1/4χ1\psi_{1}=(R^{2}+u^{2})^{1/4}\chi_{1} we obtain

χ1′′+Veff,1χ1\displaystyle-\chi_{1}^{\prime\prime}+V_{eff,1}\chi_{1} =\displaystyle= E22vF2χ1,\displaystyle\frac{E^{2}}{\hbar^{2}v_{F}^{2}}\chi_{1}, (23)

where the effective potential Veff,1V_{eff,1} is given by

Veff,1(u)=[(mR2+u2)2(mR2+u2)]\displaystyle V_{eff,1}(u)=\Bigg{[}\left(\frac{m}{\sqrt{R^{2}+u^{2}}}\right)^{2}-\left(\frac{m}{\sqrt{R^{2}+u^{2}}}\right)^{\prime}\Bigg{]} (24)

Analogously for the second component of the spinor we have:

χ2′′(u)+Veff,2(u)χ2=E22vF2χ2(u),\displaystyle-\chi_{2}^{\prime\prime}(u)+V_{eff,2}(u)\chi_{2}=\frac{E^{2}}{\hbar^{2}v_{F}^{2}}\chi_{2}(u), (25)

where the potential Veff,2(u)V_{eff,2}(u) is exactly like Veff,1(u)V_{eff,1}(u) but under the interchanging mmm\rightarrow-m. The above equations were obtained under the consideration of a constant Fermi velocity vFv_{F}. The position dependent Fermi velocity case will be addressed in the following subsection.

II.1 Position-dependent Fermi velocity

It is interesting to consider a position-dependent Fermi velocity, i.e., vF=vF(u)v_{F}=v_{F}(u). The dependence of the Fermi velocity as a function only of uu lies in the symmetry of the catenoid on the angular variable, so that no dependence of ϕ\phi is expected. Starting from (20) and (21), the process of decoupling the two equations renders for ψ1\psi_{1}

ψ1′′+[Σ(u)(1vF(u))vF(u)+Λ(u)]ψ1+[Σ(u)(1vF(u))vF(u)Σ(u)+Λ(u)Σ(u)]ψ1=E22vF(u)2ψ1\displaystyle\psi_{1}^{\prime\prime}+\left[\Sigma(u)-\left(\frac{1}{v_{F}(u)}\right)^{\prime}v_{F}(u)+\Lambda(u)\right]\psi_{1}^{\prime}+\left[\Sigma(u)^{\prime}-\left(\frac{1}{v_{F}(u)}\right)^{\prime}v_{F}(u)\Sigma(u)+\Lambda(u)\Sigma(u)\right]\psi_{1}=-\frac{E^{2}}{\hbar^{2}v_{F}(u)^{2}}\psi_{1}

where,

Σ(u)\displaystyle\Sigma(u) =\displaystyle= mR2+u2u2(R2+u2),\displaystyle\frac{m}{\sqrt{R^{2}+u^{2}}}-\frac{u}{2(R^{2}+u^{2})}, (27)
Λ(u)\displaystyle\Lambda(u) =\displaystyle= mR2+u2u2(R2+u2).\displaystyle-\frac{m}{\sqrt{R^{2}+u^{2}}}-\frac{u}{2(R^{2}+u^{2})}. (28)

In order to obtain a Klein-Gordon-like equation we must perform the following change of variables ψ1(u)=κ(u)ω1(u)\psi_{1}(u)=\kappa(u)\omega_{1}(u), imposing that ω1(u)\omega_{1}(u) obeys a Sturm-Liouville equation. Such imposition leads to a condition on the function κ(u)\kappa(u) which states that

κ(u)=exp[14ln(R2+u2)12ln(vF(u))].\kappa(u)=\exp\left[\frac{1}{4}\ln(R^{2}+u^{2})-\frac{1}{2}\ln(v_{F}(u))\right]. (29)

Hence, we obtain

ω1(u)′′+Ueff,1(u)ω1(u)=E22vF(u)2ω1(u),\displaystyle-\omega_{1}(u)^{\prime\prime}+U_{eff,1}(u)\omega_{1}(u)=\frac{E^{2}}{\hbar^{2}v_{F}(u)^{2}}\omega_{1}(u), (30)

where,

Ueff,1(u)=Veff,1(u)+V¯eff,1(u),\displaystyle U_{eff,1}(u)=V_{eff,1}(u)+\bar{V}_{eff,1}(u), (31)

being

V¯eff,1(u)=(vF(u))22vF(u)[2mvF(u)R2+u2+vF(u)′′]4vF(u)2.\bar{V}_{eff,1}(u)=-\frac{\left(v_{F}(u)^{\prime}\right)^{2}-2v_{F}(u)\left[\frac{2mv_{F}(u)^{\prime}}{\sqrt{R^{2}+u^{2}}}+v_{F}(u)^{\prime\prime}\right]}{4v_{F}(u)^{2}}. (32)

Note that the potential V¯eff,1(u)\bar{V}_{eff,1}(u) contains only the contributions from the position-dependent Fermi velocity, so that if we consider a constant Fermi velocity, V¯eff,1(u)=0\bar{V}_{eff,1}(u)=0 and we recover eq.(24). Analogously, for the second component of the spinor we have an equation similar to eq.(30) but with a potential Ueff,2(u)U_{eff,2}(u) equals to Ueff,1(u)U_{eff,1}(u) under the interchanging mmm\rightarrow-m.

III Supersymmetric Quantum Mechanics Approach

Let us give a brief review of some aspects of supersymmetric quantum mechanics (SUSY QM) junker that we use in our search of exact solutions for the effective potential models given in the previous Sections. The techniques based on the factorization method which takes place in SUSY QM helps to identify the Hamiltonians with solvable potentials. First, considering the one dimensional eigenvalue equation below

Hψ(z)=(d2dz2+U(z))ψ(z)=Eψ(z)\emph{H}~\psi(z)=\left(-\frac{d^{2}}{dz^{2}}+U(z)\right)\psi(z)=E\psi(z) (33)

where U(z)U(z) is a real scalar function. If we find two operators which are adjoint of each other and first order differental operators as

𝒜=ddz+W(z),𝒜=ddz+W(z),\mathcal{A}=\frac{d}{dz}+W(z),~~~~\mathcal{A^{{\dagger}}}=-\frac{d}{dz}+W(z), (34)

where the superpotential W(z)W(z) is a real function and element of the first order operators. Using the product of these operators, we can introduce two partner Hamiltonians defined by

H=H1\displaystyle\emph{H}=\emph{H}_{1} =\displaystyle= 𝒜𝒜=d2dz2+U1(z)\displaystyle\mathcal{A}^{{\dagger}}\mathcal{A}=-\frac{d^{2}}{dz^{2}}+U_{1}(z) (35)
H2\displaystyle\emph{H}_{2} =\displaystyle= 𝒜𝒜=d2dz2+U2(z).\displaystyle\mathcal{A}\mathcal{A}^{{\dagger}}=-\frac{d^{2}}{dz^{2}}+U_{2}(z). (36)

We can call the eigenfunctions of the Hamiltonians H1\emph{H}_{1}, H2\emph{H}_{2} as ψn(1)\psi^{(1)}_{n} and ψn(2)\psi^{(2)}_{n} while eigenvalues as En(1)E^{(1)}_{n} and En(2)E^{(2)}_{n} respectively. Then, one can introduce the partner potentials in terms of superpotentials as

U1(z)=W(z)2W(z),U2(z)=W(z)2+W(z).U_{1}(z)=W(z)^{2}-W^{\prime}(z),~~~~U_{2}(z)=W(z)^{2}+W^{\prime}(z). (37)

These partner Hamiltonians are sharing the same energy eigenvalues except the groundstate. In case of unbroken SUSY, the ground state is not degenerate with zero energy E0(1)E^{(1)}_{0} which also leads to

𝒜ψ0(1)(z)=0.\mathcal{A}\psi^{(1)}_{0}(z)=0. (38)

One can express the groundstate wavefunction in terms of the superpotential using (38)

ψ0(n)(z)=Cnexp(zW(t)𝑑t).\psi^{(n)}_{0}(z)=C_{n}\exp\left(\int^{z}W(t)dt\right). (39)

When it comes to the discrete spectrum of H2H_{2}, the relationship between the energies of these Hamiltonians are given by

En(2)=En+1(1),E0(1)=0.E^{(2)}_{n}=E^{(1)}_{n+1},~~~~E^{(1)}_{0}=0. (40)

The eigenfunctions of the partner Hamiltonians are linked as below

ψn(2)=1En+1(1)𝒜ψn+1(1),\psi^{(2)}_{n}=\frac{1}{\sqrt{E^{(1)}_{n+1}}}\mathcal{A}\psi^{(1)}_{n+1}, (41)
ψn+1(1)=1En(2)𝒜ψn(2).\psi^{(1)}_{n+1}=\frac{1}{\sqrt{E^{(2)}_{n}}}\mathcal{A}^{{\dagger}}\psi^{(2)}_{n}. (42)

We remind that if ψn+1(1)\psi^{(1)}_{n+1} is normalized, then, ψn(2)\psi^{(2)}_{n} is also normalized(non-relativistic case). The annihilation of the groundstate of the operator H1H_{1} means that there will be no SUSY partner for this state only sukhatme .

III.1 Constant Fermi velocity case

Let us consider the effective potential model for the constant Fermi velocity given below

H1\displaystyle\emph{H}_{1} =\displaystyle= d2du2+V1,eff(u)\displaystyle-\frac{d^{2}}{du^{2}}+V_{1,eff}(u) (43)
V1,eff(u)\displaystyle V_{1,eff}(u) =\displaystyle= m2R2+u2+mu(R2+u2)3/2,\displaystyle\frac{m^{2}}{R^{2}+u^{2}}+\frac{mu}{(R^{2}+u^{2})^{3/2}}, (44)

and

H1χn(1)=ϵ2χn(1)(u),\displaystyle\emph{H}_{1}\chi^{(1)}_{n}=\epsilon^{2}\chi^{(1)}_{n}(u), (45)

where ϵ=EvF\epsilon=\frac{E}{v_{F}}. The corresponding superpotential of the effective potential of the Hamiltonian H1\emph{H}_{1} is

W(u)=mR2+u2.W(u)=\frac{m}{\sqrt{R^{2}+u^{2}}}. (46)

Then, one can obtain the groundstate wavefunction as

χ0(1)=2m(u+R2+u2)m.\chi^{(1)}_{0}=2^{-m}(u+\sqrt{R^{2}+u^{2}})^{-m}. (47)

On the other hand, the partner Hamiltonian is given by

H2\displaystyle\emph{H}_{2} =\displaystyle= d2du2+V2,eff(u)\displaystyle-\frac{d^{2}}{du^{2}}+V_{2,eff}(u) (48)
V2,eff(u)\displaystyle V_{2,eff}(u) =\displaystyle= m2R2+u2mu(R2+u2)3/2,\displaystyle\frac{m^{2}}{R^{2}+u^{2}}-\frac{mu}{(R^{2}+u^{2})^{3/2}}, (49)

where we can write the first order operators 𝒜,𝒜\mathcal{A},\mathcal{A}^{{\dagger}} as

𝒜\displaystyle\mathcal{A} =\displaystyle= ddu+mR2+u2\displaystyle\frac{d}{du}+\frac{m}{\sqrt{R^{2}+u^{2}}} (50)
𝒜\displaystyle\mathcal{A}^{{\dagger}} =\displaystyle= ddu+mR2+u2.\displaystyle-\frac{d}{du}+\frac{m}{\sqrt{R^{2}+u^{2}}}. (51)

Next, we use a point transformation which is u=Rtanxu=R\tan x so that the equations (50) and (51) become

𝒜\displaystyle\mathcal{A} =\displaystyle= 1Rcos2xddx+mRcosx\displaystyle\frac{1}{R}\cos^{2}x\frac{d}{dx}+\frac{m}{R}\cos x (52)
𝒜\displaystyle\mathcal{A}^{{\dagger}} =\displaystyle= 1Rcos2xddx+mRcosx+sin2xR.\displaystyle-\frac{1}{R}\cos^{2}x\frac{d}{dx}+\frac{m}{R}\cos x+\frac{\sin 2x}{R}. (53)

We can obtain the transformed Hamiltonian H1\emph{H}_{1} as

H1=cos4xR2d2dx2+2cos2xsin2xR2ddx+mcos2x(m+3sinx)R2.\emph{H}_{1}=-\frac{\cos^{4}x}{R^{2}}\frac{d^{2}}{dx^{2}}+\frac{2\cos^{2}x\sin 2x}{R^{2}}\frac{d}{dx}+\frac{m\cos^{2}x(m+3\sin x)}{R^{2}}. (54)

Using χn(1)(x)=sec2xχ¯n(1)(x)\chi^{(1)}_{n}(x)=\sec^{2}x\bar{\chi}^{(1)}_{n}(x) in (54), we obtain

χ¯1′′(x)+ξχ¯1(x)=0,-\bar{\chi}^{\prime\prime}_{1}(x)+\xi\bar{\chi}_{1}(x)=0, (55)

with

ξ=3msecxtanx+(m2+2)sec2x4E2R2vF2sec4x,\xi=3m\sec x\tan x+(m^{2}+2)\sec^{2}x-4-\frac{E^{2}R^{2}}{v^{2}_{F}}\sec^{4}x, (56)

where we use χn(1)(x)χ1(x)\chi^{(1)}_{n}(x)\rightarrow\chi_{1}(x) for the sake of simplicity. Let us analyze now two important cases, namely, the zero energy case and the case around the throat of the catenoid.

III.1.1 Case I

In case of zero energy, i.e. ϵ=ERvF=0\epsilon=\frac{ER}{v_{F}}=0, we are lead to

χ¯1′′(x)+(3msecxtanx+(m2+2)sec2x4)χ¯1(x)=0,-\bar{\chi}^{\prime\prime}_{1}(x)+(3m\sec x\tan x+(m^{2}+2)\sec^{2}x-4)\bar{\chi}_{1}(x)=0, (57)

where we can relate it to the system known as trigonometric Scarf-I potential in the literature sukhatme whose eigenvalues and eigenfunctions are well known. Thus, one can get the solutions for zero energy as:

χn(1)(x)\displaystyle\chi^{(1)}_{n}(x) =\displaystyle= N1[1+exp(2ix)]2exp{2i[xmarctanexp(ix)]}.\displaystyle N_{1}~[1+\exp(2ix)]^{2}\exp\left\{-2i[x-m\arctan\exp(ix)]\right\}. (58)

Let us turn back to (55) and apply r=sinxr=\sin x and χ¯1(r)=(r21)1/4χ¯¯1(r)\bar{\chi}_{1}(r)=(r^{2}-1)^{1/4}\bar{\bar{\chi}}_{1}(r) to get

(1r2)χ¯¯1′′(r)+2rχ¯¯1(r)+(r224(1r2)+3mr(1r2)+m2+2(1r2)4ϵ2(1r2)2)χ¯¯1(r)=0.\displaystyle-(1-r^{2})\bar{\bar{\chi}}_{1}^{\prime\prime}(r)+2r\bar{\bar{\chi}}^{\prime}_{1}(r)+\left(\frac{r^{2}-2}{4(1-r^{2})}+3m\frac{r}{(1-r^{2})}+\frac{m^{2}+2}{(1-r^{2})}-4-\frac{\epsilon^{2}}{(1-r^{2})^{2}}\right)\bar{\bar{\chi}}_{1}(r)=0. (59)

Because (59) is not the type of hypergeometric differential equation, we shall look at the approximation methods.
Singularity analysis:
Let us consider (59) expressed below:

χ¯1′′(r)2r1r2χ¯(r)1(1r2)2(r224+3mr+(m2+2)4(1r2)ϵ21r2)χ¯=0\bar{\chi}^{\prime\prime}_{1}(r)-\frac{2r}{1-r^{2}}\bar{\chi}^{\prime}(r)-\frac{1}{(1-r^{2})^{2}}\left(\frac{r^{2}-2}{4}+3mr+(m^{2}+2)-4(1-r^{2})-\frac{\epsilon^{2}}{1-r^{2}}\right)\bar{\chi}=0 (60)

where p(r)=2r1r2p(r)=-\frac{2r}{1-r^{2}} and q(r)=1(1r2)2(r224+3mr+m2+24(1r2)ϵ21r2)q(r)=-\frac{1}{(1-r^{2})^{2}}\left(\frac{r^{2}-2}{4}+3mr+m^{2}+2-4(1-r^{2})-\frac{\epsilon^{2}}{1-r^{2}}\right). The singular points of (60) are r0=±1r_{0}=\pm 1. If (rr0)p(r)(r-r_{0})p(r) and (rr0)2q(r)(r-r_{0})^{2}q(r) are both analytic at r=r0r=r_{0}, then, r0r_{0} is a ’regular singular point’, if not, it becomes ’irregular singular point’. Hence, r0=1r_{0}=1 and r0=1r_{0}=-1 are irregular singular points because q(r)q(r) is analytic at r0=±1r_{0}=\pm 1, but q(r)q(r) is not analytic due to the term ϵ21r2\frac{\epsilon^{2}}{1-r^{2}}. If r0r_{0} is an irregular singular point, (rr0)kp(x)(r-r_{0})^{k}p(x) and (rr0)2kq(r)(r-r_{0})^{2k}q(r) are analytic where kk is the least integer number, then, the irregular singular point at r=r0r=r_{0} has a rank of k1k-1. In our problem, k=2k=2 and the rank is 11.
Next, we can continue to look for the solutions of (59).

III.1.2 Case II: r0r\rightarrow 0 limit:

Let us analyze the behavior of the solution of Eq.59 near the origin. By expanding the coefficients of Eq.59 in power-series up to first order in rr, we obtain

χ1′′2rχ1+(104m2+4ϵ243mr)χ1=0,\chi^{\prime\prime}_{1}-2r\chi^{\prime}_{1}+\left(\frac{10-4m^{2}+4\epsilon^{2}}{4}-3mr\right)\chi_{1}=0, (61)

whose solution can be written as

χ1(r)=c1e3mr2Hα(3m2+r)+c2e3mr2M(α/2,1/2,(3m2+r)2),\chi_{1}(r)=c_{1}e^{-\frac{3mr}{2}}H_{\alpha}\left(\frac{3m}{2}+r\right)+c_{2}e^{-\frac{3mr}{2}}M\left(-\alpha/2,1/2,\left(\frac{3m}{2}+r\right)^{2}\right), (62)

where α=10+5m2+4ϵ28\alpha=\frac{10+5m^{2}+4\epsilon^{2}}{8}, HαH_{\alpha} is the Hermite polynomial of degree α\alpha and M(α/2,1/2,(3m2+r)2)M\left(-\alpha/2,1/2,\left(\frac{3m}{2}+r\right)^{2}\right) is the Kummer confluent hypergeometric function. The polynomial degree α\alpha is a positive integer α=n\alpha=n provided that,

ϵ=8n5(2+m2)2.\epsilon=\frac{\sqrt{8n-5(2+m^{2})}}{2}. (63)

A deep analysis on the spectrum near the catenoid throat can be accomplished as the following. Let us look at the power series of ϵ2(1r2)2\frac{\epsilon^{2}}{(1-r^{2})^{2}} about the point r=0r=0. Up to first-order in rr, we have

ϵ2(1r2)2ϵ2.\frac{\epsilon^{2}}{(1-r^{2})^{2}}\approx\epsilon^{2}. (64)

Using the first term of (64) in (59), it becomes

(1r2)χ¯¯1′′(r)+2rχ¯¯1(r)+(r224(1r2)+3mr(1r2)+m2+2(1r2)4ϵ2)χ¯¯1(r)=0.-(1-r^{2})\bar{\bar{\chi}}_{1}^{\prime\prime}(r)+2r\bar{\bar{\chi}}^{\prime}_{1}(r)+\left(\frac{r^{2}-2}{4(1-r^{2})}+3m\frac{r}{(1-r^{2})}+\frac{m^{2}+2}{(1-r^{2})}-4-\epsilon^{2}\right)\bar{\bar{\chi}}_{1}(r)=0. (65)

If we consider the functions of Sturm-Lioville type equation given in (59) and (65),

(1r2)χ¯¯1′′(r)+2rχ¯¯1(r)+Vi(r)χ¯¯1=0,i=1,2-(1-r^{2})\bar{\bar{\chi}}_{1}^{\prime\prime}(r)+2r\bar{\bar{\chi}}^{\prime}_{1}(r)+V_{i}(r)\bar{\bar{\chi}}_{1}=0,~~~~i=1,2 (66)

where

V1(r)\displaystyle V_{1}(r) =\displaystyle= r224(1r2)+3mr(1r2)+m2+2(1r2)4ϵ2(1r2)2\displaystyle\frac{r^{2}-2}{4(1-r^{2})}+3m\frac{r}{(1-r^{2})}+\frac{m^{2}+2}{(1-r^{2})}-4-\frac{\epsilon^{2}}{(1-r^{2})^{2}} (67)
V2(r)\displaystyle V_{2}(r) =\displaystyle= r224(1r2)+3mr(1r2)+m2+2(1r2)4ϵ2\displaystyle\frac{r^{2}-2}{4(1-r^{2})}+3m\frac{r}{(1-r^{2})}+\frac{m^{2}+2}{(1-r^{2})}-4-\epsilon^{2} (68)

In the low energies values, V1(r)V_{1}(r) and V2(r)V_{2}(r) have similar behaviour while they act different in the higher energy values as is shown in the figure below.

Refer to caption
Figure 2: Graph of (67) and (68). ϵ=0.5\epsilon=0.5, m=2m=2 are used for each curve. The red curve corresponds to V1(r)V_{1}(r), green one is the curve of V2(r)V_{2}(r).

.

Refer to caption
Figure 3: Graph of (67) and (68). ϵ=15\epsilon=15, m=2m=2 are used for each curve. The red curve corresponds to V1(r)V_{1}(r), blue one is the curve of V2(r)V_{2}(r).

.

On the other hand, (67) and (68) can be expanded in series when r0r\rightarrow 0 as

V1(r)\displaystyle V_{1}(r) =\displaystyle= 52+m2ϵ2+3mr+(74+m22ϵ2)r2+O(r3)\displaystyle-\frac{5}{2}+m^{2}-\epsilon^{2}+3mr+(\frac{7}{4}+m^{2}-2\epsilon^{2})r^{2}+O(r^{3}) (69)
V2(r)\displaystyle V_{2}(r) =\displaystyle= 52+m2ϵ2+3mr+(74+m2)r2+O(r3),\displaystyle-\frac{5}{2}+m^{2}-\epsilon^{2}+3mr+(\frac{7}{4}+m^{2})r^{2}+O(r^{3}), (70)

the system reduced into the linear potential plus harmonic oscillator system. Let us consider (66) with V1(r)V_{1}(r). A mapping χ¯¯1=Z(r)1r2\bar{\bar{\chi}}_{1}=\frac{Z(r)}{\sqrt{1-r^{2}}} can be applied to (66) and dividing the whole equation by 1r21-r^{2} leads to

Z′′(r)+U(r)Z(r)=0,-Z^{\prime\prime}(r)+U(r)Z(r)=0, (71)

where U(r)=144m212mr+(31+4m2)r2+12mr3+17r4+4ϵ24(1+r2)3U(r)=\frac{14-4m^{2}-12mr+(-31+4m^{2})r^{2}+12mr^{3}+17r^{4}+4\epsilon^{2}}{4(-1+r^{2})^{3}}. We can expand U(r)U(r) in series when r0r\rightarrow 0, then,

Z′′(r)+(3mr+(114+2m23ϵ2)r272+m2ϵ2)Z(r)=0.-Z^{\prime\prime}(r)+\left(3mr+(-\frac{11}{4}+2m^{2}-3\epsilon^{2})r^{2}-\frac{7}{2}+m^{2}-\epsilon^{2}\right)Z(r)=0. (72)

(72) can also be written as

Z′′(r)\displaystyle-Z^{\prime\prime}(r) +\displaystyle+ U1(r)Z(r)=ϵ2Z(r),\displaystyle U_{1}(r)Z(r)=\epsilon^{2}Z(r), (73)
U1\displaystyle U_{1} =\displaystyle= 3mr+(114+2m23ϵ2)r272+m2\displaystyle 3mr+\left(-\frac{11}{4}+2m^{2}-3\epsilon^{2}\right)r^{2}-\frac{7}{2}+m^{2} (74)

We note that (74) is known as energy dependent potentials in the literature becasue of it’s coefficient including energy parameter edp . Moreover, it is important that 2m2>114+3ϵ22m^{2}>\frac{11}{4}+3\epsilon^{2} which makes the coefficient of r2r^{2} term as positive due to the physical solutions. The solutions of (74) can be found as

Zn(r)=Dn(6m+8m2r(11+12ϵ2)r(11+8m212ϵ2)3/4)Z_{n}(r)=D_{n}\left(\frac{6m+8m^{2}r-(11+12\epsilon^{2})r}{(-11+8m^{2}-12\epsilon^{2})^{3/4}}\right) (75)

where Dn(ar)D_{n}(ar) are the parabolic cylinder functions and ϵ2\epsilon^{2} satisfies the relationship

ϵ2=12f(ϵ)3/4(16m4+(11+12ϵ2)(7+f(ϵ))8m2(123ϵ2+f(ϵ)))=n,\epsilon^{2}=\frac{1}{2f(\epsilon)^{3}/4}\left(-16m^{4}+(11+12\epsilon^{2})(-7+f(\epsilon))-8m^{2}(-12-3\epsilon^{2}+f(\epsilon))\right)=n, (76)

here, f(ϵ)=11+8m212ϵ2f(\epsilon)=\sqrt{-11+8m^{2}-12\epsilon^{2}}. We are not giving the solutions of (76) which are very huge, one can compute the numerical energy eigenvalues. In the recent paper, our interest is fixing the ϵ2\epsilon^{2} term in (59) without sacrificing more originality. Hence, let us make another transformation for the function χ¯¯1(r)\bar{\bar{\chi}}_{1}(r) which leads to a transformation of (65) into a Jacobi type differential equation

χ¯¯1(r)=(r+1)b(1r)aY(r)\bar{\bar{\chi}}_{1}(r)=(r+1)^{b}(1-r)^{a}Y(r) (77)

then, (65) turns into

(1r2)Y′′(r)2(ab+(1+a+b)r)Y(r)+[14(174a24b2)+ϵ2\displaystyle(1-r^{2})Y^{\prime\prime}(r)-2\left(a-b+(1+a+b)r\right)Y^{\prime}(r)+[\frac{1}{4}(17-4a^{2}-4b^{2})+\epsilon^{2} +\displaystyle+ (78)
716a2+12m+4m28(r1)+7+16b2+12m4m28(r+1)ba(1+2b)]Y(r)=0.\displaystyle\frac{7-16a^{2}+12m+4m^{2}}{8(r-1)}+\frac{-7+16b^{2}+12m-4m^{2}}{8(r+1)}-b-a(1+2b)]Y(r)=0.

(78) is in the form of Jacobi differential equation which is given by

(1r2)y′′(r)+(βα(α+β+2)r)y(r)+n(n+α+β+1)y(r)=0.\displaystyle(1-r^{2})y^{\prime\prime}(r)+\left(\beta-\alpha-(\alpha+\beta+2)r\right)y^{\prime}(r)+n(n+\alpha+\beta+1)y(r)=0. (79)

Matching (78) and (79) gives

a\displaystyle a =\displaystyle= 147+12m+4m2,\displaystyle\frac{1}{4}\sqrt{7+12m+4m^{2}}, (80)
b\displaystyle b =\displaystyle= 14712m+4m2.\displaystyle\frac{1}{4}\sqrt{7-12m+4m^{2}}. (81)

The parametrs a,ba,b makes the rational terms in (78) equal to zero. and we also obtain,

α\displaystyle\alpha =\displaystyle= 2a\displaystyle 2a (82)
β\displaystyle\beta =\displaystyle= 2b.\displaystyle 2b. (83)

χ¯¯1(r)\bar{\bar{\chi}}_{1}(r) can be obtained as

χ¯¯1(r)=(r+1)b(1r)aPn(2a,2b)(r),\bar{\bar{\chi}}_{1}(r)=(r+1)^{b}(1-r)^{a}P_{n}^{(2a,2b)}(r), (84)

where Pn(2a,2b)(r)P_{n}^{(2a,2b)}(r) are the Jacobi polynomials. Hence, we can give the whole solutions of our system for n=νn=\nu (45)

χn(1)(u)=N1(1uu2+R2)147+12m+4m21(uu2+R2+1)14712m+4m21Pn(2a,2b)(uu2+R2),\displaystyle\chi^{(1)}_{n}(u)=N_{1}\left(1-\frac{u}{\sqrt{u^{2}+R^{2}}}\right)^{\frac{1}{4}\sqrt{7+12m+4m^{2}}-1}\left(\frac{u}{\sqrt{u^{2}+R^{2}}}+1\right)^{\frac{1}{4}\sqrt{7-12m+4m^{2}}-1}P^{(2a,2b)}_{n}\left(\frac{u}{\sqrt{u^{2}+R^{2}}}\right), (85)

N1N_{1} is the corresponding normalization constant and the energy eigenvalues can be found as

En(1)=±VF22R[27+4m2+2(M1+M2)+M1M2+8n2+4n(M1+M2)]1/2,E^{(1)}_{n}=\pm\frac{V_{F}}{2\sqrt{2}R}\left[-27+4m^{2}+2(M_{1}+M_{2})+M_{1}M_{2}+8n^{2}+4n(M_{1}+M_{2})\right]^{1/2}, (86)

where M1=7+4m(m3)M_{1}=\sqrt{7+4m(m-3)} and M2=7+4m(m+3)M_{2}=\sqrt{7+4m(m+3)}. The results agree with the solutions already obtained in the literature sukhatme . It is important to highlight here that since the energy must be real, it imposes a constraint condition between the principal quantum number nn and the orbital angular momentum mm, such that the inside of the square root should be positive for physical solutions. In fig.(4a) we can see the probability density and in fig.(4b) the probability ring for n=1n=1 and m=2m=-2. These values for nn and mm were chosen in order to obey the constraint imposed by the realness of the energy. From the plots we can see a region of confinement of the electron around the throat of the catenoid. Similarly, for n=3n=3 and m=2m=-2 we can see in fig.(5) that besides the confinement region at the throat of the catenoid, we have also two other regions of confinement, one in the upper sheet and another in the lower sheet.

We can also calculate the χ¯n(2)(u)\bar{\chi}^{(2)}_{n}(u) which is the solution of (48) using (41):

χn(2)\displaystyle\chi^{(2)}_{n} =\displaystyle= N21En+1(du+mR2+u2)χn+1(1)(u)\displaystyle N_{2}\frac{1}{\sqrt{E_{n+1}}}\left(\frac{d}{u}+\frac{m}{\sqrt{R^{2}+u^{2}}}\right)\chi^{(1)}_{n+1}(u)

Considering the partner Hamiltonian 2\mathcal{H}_{2}, let us introduce the corresponding transformed partner potential V2,eff(x)V_{2,eff}(x) corresponding to (48).

V2,eff(x)=m2sec2x+mtanxsecxV_{2,eff}(x)=m^{2}\sec^{2}x+m\tan x\sec x (87)

which is sharing the same energy spectrum with H1\emph{H}_{1}.

Refer to caption
(a)
Refer to caption
(b)
Figure 4: (a) Probability density for χn(1)(u)\chi_{n}^{(1)}(u) when n=1n=1 and m=2m=-2. (b) Probability ring for χn(1)(u)\chi_{n}^{(1)}(u) when n=1n=1 and m=2m=-2
Refer to caption
(a)
Refer to caption
(b)
Figure 5: (a) Probability density for χn(1)(u)\chi_{n}^{(1)}(u) when n=3n=3 and m=2m=-2. (b) Probability rings for χn(1)(u)\chi_{n}^{(1)}(u) when n=3n=3 and m=2m=-2.

III.2 position-dependent Fermi velocity case

We consider now the transformed equation for the position-dependent Fermi velocity case presented in the previous Section

ω′′(u)+[m2R2+u2+mu(R2+u2)3/2+((vF(u)))24vF(u)212vF(u)(vF′′(u)+2mvF(u)R2+u2)]ω(u)=E2vF(u)2ω(u).-\omega^{\prime\prime}(u)+\left[\frac{m^{2}}{R^{2}+u^{2}}+\frac{mu}{(R^{2}+u^{2})^{3/2}}+\frac{((v_{F}(u))^{\prime})^{2}}{4v_{F}(u)^{2}}-\frac{1}{2v_{F}(u)}\left(v^{\prime\prime}_{F}(u)+\frac{2mv^{\prime}_{F}(u)}{\sqrt{R^{2}+u^{2}}}\right)\right]\omega(u)=\frac{E^{2}}{v_{F}(u)^{2}}\omega(u). (88)

Using a point transformation which is u=Rtanxu=R\tan x in (88) leads to

ω′′(x)+2tanxω(x)+Δ(x)ω(x)=E2R2vF(x)2ω(x),-\omega^{\prime\prime}(x)+2\tan x\omega^{\prime}(x)+\Delta(x)\omega(x)=\frac{E^{2}R^{2}}{v_{F}(x)^{2}}\omega(x), (89)

where

Δ(x)=m2sec2x+mtanxsecx+Rsecx4vF(x)vF(x)vF′′(x)2vF(x)+vF(x)vF(x)tanxmRsecxvF(x)vF(x)\Delta(x)=m^{2}\sec^{2}x+m\tan x\sec x+\frac{R\sec x}{4}\frac{v^{\prime}_{F}(x)}{v_{F}(x)}-\frac{v^{\prime\prime}_{F}(x)}{2v_{F}(x)}+\frac{v^{\prime}_{F}(x)}{v_{F}(x)}\tan x-\frac{mR\sec x}{v_{F}(x)}v^{\prime}_{F}(x) (90)

Using ω(x)=secxω1(x)\omega(x)=\sec x\omega_{1}(x) and a suggestion on the Fermi velocity which is

vF(x)=λsec2xv_{F}(x)=\lambda\sec^{2}x (91)

makes (88) as

ω1′′(x)+[1+(m21)sec2x+12(2m+R4mR)secxtanx]ω1(x)=E2R2λ2ω1(x),-\omega^{\prime\prime}_{1}(x)+\left[-1+(m^{2}-1)\sec^{2}x+\frac{1}{2}(2m+R-4mR)\sec x\tan x\right]\omega_{1}(x)=\frac{E^{2}R^{2}}{\lambda^{2}}\omega_{1}(x), (92)

where λ\lambda is a real parameter. The trigonometric ansatz for the Fermi velocity could be related to periodic inhomogeneities in the graphene lattice, such as the presence of heteroatoms or topological defects. The potential function in (92) is known as trigonometric Scarf-I potential in the literature whose solutions are very well-known sukhatme . Then, the solutions of (92) can be written as

ω1(x)=(1sinx)AB2(1+sinx)A+B2Pn(AB1/2,A+B1/2)(sinx),\omega_{1}(x)=(1-\sin x)^{\frac{A-B}{2}}(1+\sin x)^{\frac{A+B}{2}}P^{(A-B-1/2,A+B-1/2)}_{n}(\sin x), (93)

where

A\displaystyle A =\displaystyle= 18(R+m(4R2))×\displaystyle\frac{1}{8(-R+m(4R-2))}\times
×[4R+8m(2R1)+42m2c2(3+c3+4m2(1+m2R)(32R+4m(1+m+2R)))],\displaystyle\times[-4R+8m(2R-1)+4\sqrt{2}m^{2}c-\sqrt{2}(3+c\sqrt{-3+4m^{2}(1+m-2R)(-3-2R+4m(-1+m+2R))})],
B\displaystyle B =\displaystyle= c22,\displaystyle\frac{c}{2\sqrt{2}}, (95)
c\displaystyle c =\displaystyle= 3+4m2+(3+4m(1+m2R)+2R)(32R+4m(1+m+2R)),\displaystyle\sqrt{-3+4m^{{}^{2}}+\sqrt{(-3+4m(1+m-2R)+2R)(-3-2R+4m(-1+m+2R))}}, (96)

and energy spectrum is given as

En(1)=±λR(A+n)21,n=0,1,2,E^{(1)}_{n}=\pm\frac{\lambda}{R}\sqrt{(A+n)^{2}-1},~~n=0,1,2,... (97)

Also in this case, the quantity (A+n)21(A+n)^{2}-1 must be positive definite in order to assure that the energy is real. Such imposition leads to a constraint condition between the nn and mm. The solutions of ω\omega can be written in terms of uu as

ωn(u)=ωn(1)(u)\displaystyle\omega_{n}(u)=\omega^{(1)}_{n}(u) =\displaystyle= N11+u2R2(1uR2+u2)AB2(1+uR2+u2)A+B2×\displaystyle N_{1}\sqrt{1+\frac{u^{2}}{R^{2}}}\left(1-\frac{u}{\sqrt{R^{2}+u^{2}}}\right)^{\frac{A-B}{2}}\left(1+\frac{u}{\sqrt{R^{2}+u^{2}}}\right)^{\frac{A+B}{2}}\times (98)
×Pn(AB1/2,A+B1/2)(uR2+u2).\displaystyle\times P^{(A-B-1/2,A+B-1/2)}_{n}\left(\frac{u}{\sqrt{R^{2}+u^{2}}}\right).

We also remind that the corresponding superpotential of the system in (92) is:

W(x)=Atanx+Bsecx.W(x)=-A\tan x+B\sec x. (99)

We can introduce the partner Hamiltonian of the system given in (87) as

ω2′′(u)+[m2R2+u2mu(R2+u2)3/2((vF(u)))24vF(u)2+12vF(u)(vF′′(u)+2mvF(u)R2+u2)]ω2(u)=E2vF(u)2ω2(u).-\omega_{2}^{\prime\prime}(u)+\left[\frac{m^{2}}{R^{2}+u^{2}}-\frac{mu}{(R^{2}+u^{2})^{3/2}}-\frac{((v_{F}(u))^{\prime})^{2}}{4v_{F}(u)^{2}}+\frac{1}{2v_{F}(u)}\left(v^{\prime\prime}_{F}(u)+\frac{2mv^{\prime}_{F}(u)}{\sqrt{R^{2}+u^{2}}}\right)\right]\omega_{2}(u)=\frac{E^{2}}{v_{F}(u)^{2}}\omega_{2}(u). (100)

Using (99) and (93), we can calculate the solutions of (100) which shares the same energy states (97) with (87) except the ground state:

ω2(x)\displaystyle\omega_{2}(x) \displaystyle\sim secx(cosx2sinx2)(1sinx)A12(cosx2+sinx2){(1+2A+n)cos2xP(n,12+A+B,12+AB,sinx)\displaystyle\sec x\frac{(\cos\frac{x}{2}-\sin\frac{x}{2})(1-\sin x)^{A-1}}{2(\cos\frac{x}{2}+\sin\frac{x}{2})}\left\{(1+2A+n)\cos^{2}xP\left(n,\frac{1}{2}+A+B,\frac{1}{2}+A-B,\sin x\right)\right. (101)
2P(n+1,12+A+B,12+AB,sinx)(AB+2Asinx)},\displaystyle\left.-2P\left(n+1,-\frac{1}{2}+A+B,-\frac{1}{2}+A-B,\sin x\right)(A-B+2A\sin x)\right\},

And the whole solutions read as

ω2(u)\displaystyle\omega_{2}(u) =\displaystyle= ωn(2)(u)=N2En+1(1)1+u2R21+RR2+u2(1uR2+u2)A121RR2+u2{(1+2A+n)RR2+u2×\displaystyle\omega^{(2)}_{n}(u)=\frac{N_{2}}{E^{(1)}_{n+1}}\sqrt{1+\frac{u^{2}}{R^{2}}}\frac{\sqrt{1+\frac{R}{\sqrt{R^{2}+u^{2}}}}\left(1-\frac{u}{\sqrt{R^{2}+u^{2}}}\right)^{A-1}}{2\sqrt{1-\frac{R}{\sqrt{R^{2}+u^{2}}}}}\left\{(1+2A+n)\frac{R}{\sqrt{R^{2}+u^{2}}}\right.\times (102)
×P(n,12+A+B,12+AB,uR2+u2)2P(n+1,12+A+B,12+AB,uR2+u2)×\displaystyle\times P\left(n,\frac{1}{2}+A+B,\frac{1}{2}+A-B,\frac{u}{\sqrt{R^{2}+u^{2}}}\right)-2P\left(n+1,-\frac{1}{2}+A+B,-\frac{1}{2}+A-B,\frac{u}{\sqrt{R^{2}+u^{2}}}\right)\times
×(AB+2AuR2+u2)}\displaystyle\left.\times(A-B+2A\frac{u}{\sqrt{R^{2}+u^{2}}})\right\}

IV final remarks

We studied the Dirac equation for an electron constrained to move on a catenoid surface realized as a connection bridge between two layers of graphene. We decoupled the two components of the spinor and in order to obtain two Klein-Gordon-like equations. Analytical solutions were obtained using supersymmetric quantum mechanics for two cases, namely, the constant Fermi velocity case and the position dependent Fermi velocity case. In case of constant Fermi velocity, both zero energy modes and approximated bound state energies are obtained. The bound state solutions for the constant Fermi velocity are obtained in the limit of r0r\rightarrow 0.

For the constant Fermi velocity case, the supersymmetric quantum mechanical approach allows us to identify the potential as a very well known trigonometric Scarf-I potential. The eigenvalues of the Hamiltonian operator and the eigenfunctions were found exactly. It is important to highlight that the solutions of the system imposes a relation between the principal quantum number nn and the orbital angular momentum mm in order to guarantee the positively of the energy.

For the position-dependent Fermi velocity case, similarly, the supersymmetric quantum mechanical approach allows us to identify the potential as a very well known trigonometric Scarf-I potential. In this case exact solutions were also found under a trigonometric ansatz for the Fermi velocity.

The paper not only presents important properties about the dynamics of an electron constrained to move on a catenoid bridge but also opens up new possibilities of investigation. The electron-phonon interaction as well as the thermodynamic properties will be addressed in a future work.

Data Availability Statement
The authors confirm that the data supporting the findings of this study are available within the article and its supplementary materials.

Acknowledgments

J.E.G.Silva thanks the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), grants no 312356/2017-0 for financial support.

References

  • (1) A. K. Geim, K. S. Novoselov, Nature Materials 6, 183 (2007).
  • (2) A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, A. K. Geim, Rev. Mod. Phys. 81, 109 (2009).
  • (3) M. Katsnelson, Graphene: Carbon in two dimensions, Cambridge University Press, Cambridge, (2012).
  • (4) J. E. Gomes Silva, J. Furtado and A. C. A. Ramos, Eur. Phys. J. B 93, no.12, 225 (2020)
  • (5) Dai-Nam Le et al, Eur. Phys. Lett. 10005 (2019).
  • (6) Ö. Yeşiltaş, Adv. High Energy Phys. 2018, 6891402 (2018).
  • (7) Dai-Nam Le et al, Physica E 107 60 (2019).
  • (8) F. Guinea, A. K. Geim, M. I. Katsnelson, and K. S. Novoselov, Phys. Rev. B 81, 035408 (2010).
  • (9) F. de Juan, A. Cortijo, M. A. H. Vozmediano, Phys. Rev. B 76, 165409, (2007).
  • (10) V. Atanasov, A. Saxena, Phys. Rev. B 81, 205409 (2010).
  • (11) C. Furtado , F. Moraes , A.M. de M. Carvalho, Phys. Lett. A 372, 5368, (2008).
  • (12) R. Dandoloff, T. T. Truong, Phys. Lett. A 325, 233 (2004).
  • (13) V. Atanasov, R. Dandoloff and A. Saxena Phys. Rev. B 79, 033404 (2009).
  • (14) V. Atanasov, A. Saxena, Phys. Rev. B 92, 035440 (2015).
  • (15) R. Dandoloff, A. Saxena and B. Jensen, Phys. Rev. A 81 014102 (2010).
  • (16) J. González, J. Herrero, Nucl. Phys. B 825, 426 (2010).
  • (17) R. Pincak, J. Smotlacha, Quantum Matter 5, 114 (2016).
  • (18) R. Dandoloff, Phys. Lett. A, 373 (2009).
  • (19) R. Dandoloff, A. Saxena, B. Jensen, Phys. Rev. A 81, 014102 (2010).
  • (20) M. Spivak, A comprehensive introduction to differential geometry, Publish or Perish, Houston, (1999).
  • (21) J.E.G. Silva et al., Phys. Lett. A 384, 126458 (2020).
  • (22) J.E.G. Silva et al., Eur. Phys. J. B 94, 127 (2021).
  • (23) J. Wess and J. Bagger, Supersymmetry and Supergravity, Princeton University Press, 1992 ( ISBN:0691085560 ,0691025304).
  • (24) A. Sinha and P. Roy, Int. J. Mod. Phys. A 21, 5807–5822 (2006).
  • (25) P.A.M. Dirac, The Principles of Quantum Mechanics, 2nd edn, Clarendon, Oxford (1935)
  • (26) E. Schrödinger, Proc. R. Irish Acad. 46 A 9-16 (1940)
  • (27) F. Cooper, A. Khare and U. Sukhatme, Supersymmetry in Quantum Mechanics, World Scientific, 2001( ISBN:9810246129, 9789810246129).
  • (28) R. L. Hall and O. Yesiltas, Int. J. of Mod. Phys. E, 1910 1923 (2010).
  • (29) F. Correa and V. Jakubsky, Phys. Rev. A, 95(3) 033807 (2017).
  • (30) D. Bermudez and D. J. Fernandez, Symm. Int. and Geo.-Meth. and Appl., 7 025 (2011).
  • (31) G. Junker, Symmetry-Basel, 12(10) 1590 (2020).
  • (32) G. Levai, Czech. J. of Phys. 54(10) 1121 (2004).
  • (33) M. Znojil, Symmetry-Basel, 12(6) 892 (2020).
  • (34) G. Junker, Supersymmetric Methods in Quantum, Statistical and Solid State Physics, IOP Publishing Ltd; Enlarged and Revised edition (April 20, 2019), ISBN-13: 978-0750320245.
  • (35) J. Formanek, R. J. Lombard, J. Mares, Czech. J. of Physics, 54 289 (2004).