Spectroscopy of the Dirac oscillator perturbed by a surface delta potential

Similar documents
arxiv: v1 [quant-ph] 18 Dec 2013

Completeness of the Dirac oscillator eigenfunctions

arxiv:hep-th/ v1 23 Apr 1994

3.3 Lagrangian and symmetries for a spin- 1 2 field

arxiv:cond-mat/ v1 5 Jul 1994

The 3 dimensional Schrödinger Equation

On the relativistic L S coupling

Mathematical Tripos Part IB Michaelmas Term Example Sheet 1. Values of some physical constants are given on the supplementary sheet

Columbia University Department of Physics QUALIFYING EXAMINATION

Qualifying Exam. Aug Part II. Please use blank paper for your work do not write on problems sheets!

charges q r p = q 2mc 2mc L (1.4) ptles µ e = g e

Particle Physics. Michaelmas Term 2011 Prof. Mark Thomson. Handout 2 : The Dirac Equation. Non-Relativistic QM (Revision)

Gated-controlled electron pumping in connected quantum rings

ψ s a ˆn a s b ˆn b ψ Hint: Because the state is spherically symmetric the answer can depend only on the angle between the two directions.

Particle Physics Dr. Alexander Mitov Handout 2 : The Dirac Equation

Lecture 4 Quantum mechanics in more than one-dimension

LECTURES ON QUANTUM MECHANICS

Chm 331 Fall 2015, Exercise Set 4 NMR Review Problems

Lecture 10. The Dirac equation. WS2010/11: Introduction to Nuclear and Particle Physics

Spin, Isospin and Strong Interaction Dynamics

Attempts at relativistic QM

The Yukawa Lagrangian Density is Inconsistent with the Hamiltonian

Basic quantum Hamiltonian s relativistic corrections. Abstract

Ket space as a vector space over the complex numbers

Transient Phenomena in Quantum Bound States Subjected to a Sudden Perturbation

Fine structure in hydrogen - relativistic effects

Quantum Physics 2006/07

Introduction to Modern Quantum Field Theory

Total Angular Momentum for Hydrogen

Binding energy of hydrogenic impurities in quantum dots under intense laser radiation

1 Commutators (10 pts)

1 r 2 sin 2 θ. This must be the case as we can see by the following argument + L2

DISCRETE SYMMETRIES IN NUCLEAR AND PARTICLE PHYSICS. Parity PHYS NUCLEAR AND PARTICLE PHYSICS

Columbia University Department of Physics QUALIFYING EXAMINATION

Space-Time Symmetries

Quantum Mechanics Solutions

Relativistic Corrections to the Neutron Electric. Dipole Moment in Valence Quark Models. Abstract

Angular momentum and spin

Semi-Classical Theory of Radiative Transitions

arxiv:cond-mat/ v1 1 Apr 1999

BOUND STATES OF THE DIRAC EQUATION FOR A CLASS OF EFFECTIVE QUADRATIC PLUS INVERSELY QUADRATIC POTENTIALS

Algebraic Approach to Radial Ladder Operators in the Hydrogen Atom

Nuclear size corrections to the energy levels of single-electron and -muon atoms

Columbia University Department of Physics QUALIFYING EXAMINATION

Solution of One-dimensional Dirac Equation via Poincaré Map

(a) Write down the total Hamiltonian of this system, including the spin degree of freedom of the electron, but neglecting spin-orbit interactions.

a = ( a σ )( b σ ) = a b + iσ ( a b) mω 2! x + i 1 2! x i 1 2m!ω p, a = mω 2m!ω p Physics 624, Quantum II -- Final Exam

PHY4604 Introduction to Quantum Mechanics Fall 2004 Final Exam SOLUTIONS December 17, 2004, 7:30 a.m.- 9:30 a.m.

Quantum Theory of Angular Momentum and Atomic Structure

Relativistic Approach to the Hydrogen Atom in a Minimal Length Scenario

Non-relativistic scattering

1. Electricity and Magnetism (Fall 1995, Part 1) A metal sphere has a radius R and a charge Q.

Physics 221A Fall 1996 Notes 21 Hyperfine Structure in Hydrogen and Alkali Atoms

20 The Hydrogen Atom. Ze2 r R (20.1) H( r, R) = h2 2m 2 r h2 2M 2 R

Phys 622 Problems Chapter 5

The general solution of Schrödinger equation in three dimensions (if V does not depend on time) are solutions of time-independent Schrödinger equation

LS coupling. 2 2 n + H s o + H h f + H B. (1) 2m

Relativistic quantum mechanics of a Dirac oscillator

Angular momentum. Quantum mechanics. Orbital angular momentum

PHYS 771, Quantum Mechanics, Final Exam, Fall 2011 Instructor: Dr. A. G. Petukhov. Solutions

Chapter 3: Relativistic Wave Equation

Control of spin-polarised currents in graphene nanorings

Optical Lattices. Chapter Polarization

PH 451/551 Quantum Mechanics Capstone Winter 201x

PHY 407 QUANTUM MECHANICS Fall 05 Problem set 1 Due Sep

arxiv: v2 [hep-th] 6 Jul 2009

J07M.1 - Ball on a Turntable

Plane wave solutions of the Dirac equation

Physics 221A Fall 1996 Notes 14 Coupling of Angular Momenta

Eigenfunctions of Spinless Particles in a One-dimensional Linear Potential Well

Physics 129, Fall 2010; Prof. D. Budker

New insights on pseudospin doublets in nuclei

Quantum Physics III (8.06) Spring 2007 FINAL EXAMINATION Monday May 21, 9:00 am You have 3 hours.

TPP entrance examination (2012)

Notes on Quantum Mechanics

Wigner s Little Groups

Two-photon Absorption Process in Semiconductor Quantum Dots

Forbidden Electric Dipole Transitions in the Hydrogen Molecular Ion First Estimates

β matrices defined above in terms of these Pauli matrices as

A generalization of the Three-Dimensional Harmonic Oscillator Basis for Wave Functions with Non-Gaussian Asymptotic Behavior

THE QUANTUM MECHANICAL CURRENT OF THE PAULI EQUATION

Quantum Mechanics: Fundamentals

Quantum Field Theory. and the Standard Model. !H Cambridge UNIVERSITY PRESS MATTHEW D. SCHWARTZ. Harvard University

We now turn to our first quantum mechanical problems that represent real, as

ELECTRON-PION SCATTERING II. Abstract

Planar QED in Magnetic or Electric Solitonic Backgrounds

Quantum Theory of Many-Particle Systems, Phys. 540

( ). Expanding the square and keeping in mind that

Symmetries for fun and profit

Classical Oscilators in General Relativity

Symmetries in Quantum Physics

CHAPTER 8 The Quantum Theory of Motion

Energy spectrum for a short-range 1/r singular potential with a nonorbital barrier using the asymptotic iteration method

Harmonic Oscillator Eigenvalues and Eigenfunctions

2.4 Parity transformation

Lecture 4 Quantum mechanics in more than one-dimension

6. QED. Particle and Nuclear Physics. Dr. Tina Potter. Dr. Tina Potter 6. QED 1

Representation theory and quantum mechanics tutorial Spin and the hydrogen atom

Second quantization: where quantization and particles come from?

QUALIFYING EXAMINATION, Part 1. 2:00 PM 5:00 PM, Thursday September 3, 2009

Transcription:

Spectroscopy of the Dirac oscillator perturbed by a surface delta potential J. Munárriz and F Domínguez-Adame GISC. Departamento de Física de Materiales, Universidad Complutense, E-28040 Madrid, Spain R. P. A. Lima Instituto de Física, Universidade Federal de Alagoas, Maceió, AL 57072-970, Brazil Abstract We study theoretically the level shift of the Dirac oscillator perturbed by any sharply peaked potential approaching a surface delta potential. A Green function method is used to obtain closed expressions for all partial waves and parities. Keywords: Dirac oscillator, singular potentials, relativistic wave equations PACS: 03.65.pm, 03.65. w, 03.65.Ge, Physics Letters A 376, 3475 (2012) According to their behavior under Lorentz transformations, the interaction potentials for the Dirac equation are either vector, scalar or pseudoscalar. Here the term vector means the time component of a Lorentz vector, while a scalar potential is equivalent to a dependence of the mass upon position. Relativistic wave equations with vector and scalar linearly rising potentials have been widely used to investigate the confinement of particles in nuclear and hadron physics. When the linear potential is vectorlike, there exist no bound states and only tunneling solutions arise [1, 2]. Therefore, vector linear potentials cannot confine particles, being another fine example of the Klein tunneling [3]. On the contrary, scalar linear potentials can bind relativistic particles and give rise to confinement since the mass of the particle increases Preprint submitted to Physics Letters A May 29, 2015

without bound when the separation from the center gets larger [2, 4, 5]. Similar conclusions can be drawn for vector and scalar parabolic potentials [6]. In spite of the fact that vector linear and quadratic potentials cannot bind particles, it is possible to extend the quantum harmonic oscillator to the relativistic domain [6]. In this context, Moshinsky and Szczepaniak introduced a new type of interaction in an attempt to describe a relativistic oscillator by means of a Dirac equation linear in both momenta and coordinates [7]. The authors gave the name of Dirac oscillator to this model, whose nonrelativistic limit leads to a standard harmonic oscillator equation with spin-orbit term. The complete energy spectrum and the electromagnetic potential associated to the Dirac oscillator were found by Benítez et al. [8]. The spectrum presents degeneracies which are explained by a symmetry Lie algebra [9]. Regardless the intensity of the coupling, this interaction leaves the vacuum unchanged and the Klein tunneling is avoided [8, 10]. Therefore, in contrast to the vector linear potential, the resulting states are truly bound and the Dirac oscillator is a good candidate to explain the observed confinement of quarks [11]. This interaction has also been considered in the two-body Dirac equation, presenting interesting features which are not shared with the onebody oscillator [12]. Recently, a photonic realization of the Dirac oscillator based on light propagation in engineered fiber Bragg gratings has been discussed by Longhi [13]. Therefore, the interest of the Dirac oscillator is rather general and well beyond high energy physics. The aim of this paper is to study the spectroscopy of the Dirac oscillator in a (3 + 1)-dimensional space perturbed by a surface delta potential, solving the associated Lippmann-Schwinger equation. Therefore, we present a generalization of the results obtained in Ref. [14], in the sense that we do not restrict ourselves to (1 + 1)-dimensional space. In addition, in Ref. [14] the perturbation was described by a nonlocal separable potential. The δ-function limit of a nonlocal potential in the Dirac equation is mathematically well defined [15] and the corresponding Lippmann-Schwinger equation is valid even in this limit. However, as discussed in Ref. [16], there exist some ambiguities in defining the surface delta potential (local) that require a careful analysis of the Lippmann-Schwinger equation, as we show below. We arrive at the Dirac oscillator equation by the non-minimal substitution p p imωβr, where m is the mass of the particle, ω is the oscillator frequency and β is the usual Dirac matrix defined below. From the above considerations, the Dirac Hamiltonian may be written in the standard nota- 2

tion as (in units with = 1 and c = 1) H = α (p imωβr) + βm + V R (r) H 0 + V R (r). (1) We choose the representation ( 0 σ α = σ 0 ) ( I2 0, β = 0 I 2 ), where σ = (σ x, σ y, σ z ) contains the Pauli matrices and I 2 is the 2 2 unity matrix. We are interested in the energy spectrum of the Dirac oscillator, whose Hamiltonian is given by H 0 in (1), perturbed by a potential V R (r) approaching the δ-shell limit with radius R. However, as mentioned above, the resulting equation is ambiguous if one takes the limit V R (r) λδ(r R) from the outset. The origin of the ambiguity is the following. Since the Dirac equation is linear in momentum, the wavefunction itself must be discontinuous at r = R to account for the singular term V R (r) λδ(r R). However, the product of a discontinuous function and the δ-function is mathematically ill defined. This ambiguity can be avoided by solving the corresponding Dirac equation for any arbitrary sharply peaked at r = R function, R being the radius of the shell, and then take the δ-function limit with the constraint R+ R R R V R (r) dr = λ, R 0, (2) where λ is the dimensionless coupling constant. For the moment, we only assume that the potential V R (r) is spherically symmetric. The eigenfunctions of definite parity and total angular momentum (J 2, J z ) are written in the form Ψ(r) = 1 r ( if(r) g(r)σ r/r ) Φ l jm j, (3) where Φ l jm j are the normalized two-components eigenfunctions of J 2, J z, L 2 and S 2 [17]. Using (1) and (3), the Dirac equation leads to d [ )] dr φ(r) = σ z γ(r) σ x m + iσ y (E V R (r) φ(r), (4) where the upper and lower components of the radial spinor φ(r) are f(r) and g(r), respectively. Here κ = (j + 1/2) for l = j ± 1/2 and we have 3

defined γ(r) = κ/r + mωr for brevity. Equation (4) is solved by a Newmann solution as follows [18] { r φ(r) = P ( exp dr [σ z γ(r ) σ x m + iσ y E V R (r ))] } φ(r 0 ), r 0 where P is the ordering operator. Setting r = R + R and r 0 = R R, taking the limit R 0 and using the constraint (2) we finally obtain the following boundary condition φ(r + R) = exp ( iλσ y ) φ(r R), (5) which becomes independent of how the δ-function limit is taken and thus we avoid any ambiguity defining the relativistic surface δ potential. Once the correct boundary condition at the singularity has been obtained, we proceed to find the energy spectrum of the perturbed Dirac oscillator. To this end, we write the Lippmann-Schwinger solution of the radial Dirac equation (4) R+ R φ(r) = G(r, r ; E)V R (r )φ(r )dr, (6) R R where the Green function for the unperturbed problem is a 2 2 matrix satisfying the inhomogeneous differential equation [ ] iσ y r + σ xγ(r) + σ z m E G(r, r ; E) = I 2 δ(r r ). (7) The Green function exhibits a jump discontinuity at the line r = r. The value of the jump can be obtained by integration of (7) in the vicinity of this line. The result is G(r + R, r; E) G(r R, r; E) = iσ y. (8) The product V R (r)φ(r) in the integral (6) is not well defined if one takes the limit V R (r) λδ(r R), as we already discussed. Thus, we consider the same limiting procedure discussed previously and solve (6) for any arbitrary sharply peaked at r = R function and then take the δ-function limit. Using 4

the radial Dirac equation (4) one finds that the integral equation (6) leads to (see Ref. [19] for details) ] φ(r) = ig(r, R; E)σ y [φ(r + R) φ(r R). (9) Hence we have obtained a closed expression for the perturbed eigenfunctions. The energy levels can be obtained by setting r = R + R in (9) and using the boundary condition (5) [ ( ) ] det I 2 ig(r + R, R; E)σ y e iσ yλ I 2 = 0. (10) Therefore, the energy spectrum of the perturbed Dirac oscillator can be obtained provided the Green function for the unperturbed problem is known. The Green function of the Dirac oscillator can be cast in the form ( ) G(r, r G++ (r, r ; E) = ; E) G + (r, r ; E) G + (r, r ; E) G (r, r ; E) and it was calculated in Ref. [20]. For the sake of brevity we only quote here the final expressions of the diagonal terms G ±± = E ± m Γ ( ) ν ± µ ± + 1 2 mω Γ (2ν ± + 1) rr M ( ) ( ) µ ±,ν ± mωr 2 < Wµ±,ν ± mωr 2 >, (11a) where M µ,ν and W µ,ν are the Whittaker functions, r < = min(r, r ) and r > = max(r, r ). The parameters µ ± and ν ± are defined as µ ± = 1 ( ) E 2 m 2 2κ ± 1, ν ± = 1 4 mω 2 κ ± 1 2. (11b) The off-diagonal terms are calculated from the following expressions 1 G ± = [ r E m + κr ] + mωr G. (11c) Real solutions of (10) with the Green function given by (11) yield the energy levels of the perturbed Dirac oscillator. Due to the boundary condition (5), the energy levels of the perturbed Dirac oscillator satisfy the property E(m, κ, R, λ) = E(m, κ, R, λ + lπ), l being an integer. Since the energy 5

levels are π-periodic functions of the coupling constant, we can restrict ourselves to the range π/2 < λ π/2 hereafter. Figure 1 shows these levels as a function of the coupling constant for the case ω = m. The energy levels of the perturbed Dirac oscillator are shifted upwards on increasing the coupling constant from π/2 to π/2. Notice that the levels of the perturbed Dirac oscillator cross those of the unperturbed oscillator only when λ = nπ, n being an integer. According to Eq. (5), in this case φ(r+ R) = ( 1) n φ(r R) and the surface δ potential is actually transparent. Figure 1: Energy levels of a Dirac oscillator perturbed by a surface delta potential as a function of the coupling constant λ for ω = m. Left and right panels correspond to κ = 1 and κ = 1, respectively. Solid (dashed) lines correspond to the results for R = 0.3/m (R = 1/m). The solutions of the unperturbed Dirac oscillator are marked with red points. In the limit λ 0, the eigenvalues of the unperturbed system are recovered [21]. They correspond to the poles of G, namely the set of values E 0 (m, κ) for which (E 0 ± m)γ(ν ± µ ± + 1/2) E 2 0(m, κ) m 2 = 4mω [n + θ(κ) (2κ + 1)], E 0 m, (12) where θ is the Heaviside step function and n is a nonnegative integer. Due to the confining properties of the Dirac oscillator, the spatial extent of the eigenstates increases with the absolute value of the energy, E. If this spatial extent is much smaller than the radius of the surface δ potential, the net effect of the perturbation on the eigenfunction is small, and therefore the energy does not change noticeably. This trend is clearly observed in Fig. 2, where the energy levels are plotted as a function of the radius R for ω = m, κ = 1 and λ = π/4. When the dimensionless coupling constant is large, the energy levels display anticrossings, as seen in the figure. Moreover, each energy level approaches the same level of the unperturbed Dirac oscillator in the two limiting situations, mr 1 and mr 1. 6

Figure 2: Shift of the energy levels as a function of the radius R for ω = m, κ = 1 and λ = π/4. The levels of the unperturbed Dirac oscillator are plotted with dashed gray lines. The wave functions corresponding to the four red points are shown in Fig. 3 The wave functions at points marked a d are shown in figure 3 below. Finally, the change of the probability density φ(r) 2 due to the surface δ potential is shown in Fig. 3. The set of parameters chosen correspond to the four red circles in Fig. 2, i.e., all of them are taken from the same original unperturbed eigenstate. As R is increased, the eigenstate transits between two loosely perturbed states upper left and lower right panels. In between, states with probability density strongly peaked around r = R are found. Figure 3: Plots of φ(r) 2 as a function of the dimensionless radial coordinate z = r m for the four different values of the delta-shell radius R indicated in figure 2 by red circles, when ω = m, κ = 1 and λ = π/4. The corresponding unperturbed eigenstate is plotted using gray solid regions. In summary, we have calculated the shift of the energy levels of the Dirac 7

oscillator perturbed by a surface δ potential using a Green function technique. The method is valid for any sharply peaked potential approaching the δ-function and consequently it is free of the ambiguities appearing in defining relativistic δ-interactions [16]. Remarkably, the energy spectrum is a π-periodic function of the coupling constant λ, a situation not found in the (1 + 1)-dimensional Dirac oscillator perturbed by a nonlocal δ potential [14]. Work in Madrid was supported by MICINN (project MAT2010-17180). R. P. A. Lima would like to thank CAPES via project PPCP-Mercosul, CNPq, and FINEP (Brazilian Research Agencies) as well as FAPEAL (Alagoas State Research Agency) for partial financial support. [1] M. L. Glasser and N. Shawagfeh, J. Math. Phys. 25 (1984) 2533. [2] F. Domínguez-Adame and M. A. González, Europhys. Lett. 13 (1990) 193. [3] H. Galic, Am. J. Phys. 56 (1988) 312. [4] C. L. Critchfield, J. Math. Phys. 17 (1976) 261. [5] B. Ram, J. Phys. A: Math. Gen. 20 (1987) 5023. [6] F. M. Toyama and Y. Nogami, Phys. Rev. A 59 (1999) 1056. [7] M. Moshinsky and A. Szczepaniak, J. Phys. A: Math. Gen. 22 (1989) L817. [8] J. Benítez, R. P. Martínez y Romero, H. N. Núñez-Yepes and A. L Salas-Brito, Phys. Rev. Lett. 64 (1990) 1643. [9] C. Quesne and M. Moshinsky, J. Phys. A: Math. Gen. 23 (1990) 2262. [10] F. Domínguez-Adame, Phys. Lett. A 162 (1992) 18. [11] M. Moreno and A. Zentella, J. Phys. A: Math. Gen. 22 (1989) L821. [12] F. Domínguez-Adame and B. Méndez, Can. J. Phys. 69 (1991) 780. [13] S. Longhi, Opt. Lett. 35 (2010) 1302. [14] F. Domínguez-Adame, Europhys. Lett. 15 (1991) 569. 8

[15] M. G. Calkin, D. Kiang and Y. Nogami, Phys. Rev. C 38 (1988) 1076. [16] F. Domínguez-Adame, J. Phys. A: Math. Gen. 23 (1990) 1993. [17] D. Bjorken and S. D. Drell, Relativistic Quantum Mechanics (Mc-Graw Hill, New York, 1964) p. 53. [18] B. H. J. McKellar and G. J. Stephenson, Phys. Rev. A 36 (1987) 2566. [19] B. Méndez and F. Domínguez-Adame, J. Phys. A: Math. Gen. 25 (1992) 2065. [20] A. D. Alhaidari, Int. J. Theor. Phys. 43 (2004) 939. [21] R.. Szmytkowski and M. Gruchowski, J. Phys. A: Math. Gen. 34 (2001) 4991. 9