SciELO - Scientific Electronic Library Online

 
vol.64 issue5Hamiltoniano efectivo de un circuito LC con carga discreta bajo la aproximación semiclásicaInvestigation of the electron paramagnetic resonance signal response in gamma-irradiated poly(lactic acid) for high-dose dosimetry author indexsubject indexsearch form
Home Pagealphabetic serial listing  

Services on Demand

Journal

Article

Indicators

Related links

  • Have no similar articlesSimilars in SciELO

Share


Revista mexicana de física

Print version ISSN 0035-001X

Rev. mex. fis. vol.64 n.5 México Sep./Oct. 2018  Epub Nov 30, 2019

 

Research

Bound states in the continuum and time evolution of the generalized eigenfunctions

D. Lohra 

E. Hernándeza 

A. Jáureguib 

A. Mondragóna 

aInstituto de Física, Universidad Nacional Autónoma de México, Apartado Postal 20-364, 01000 Ciudad de México, México.

bDepartamento de Física, Universidad de Sonora, Apartado Postal 106, Hermosillo, Sonora, México.


Abstract

We study the Jost solutions for the scattering problem of a von Neumann-Wigner type potential, constructed by means of a two times iterated and completely degenerated Darboux transformation. We show that for a particular energy the unnormalized Jost solutions coalesce to give rise to a Jordan cycle of rank two. Performing a pole decomposition of the normalized Jost solutions we find the generalized eigenfunctions: one is a normalizable function corresponding to the bound state in the continuum and the other is a bounded, non-normalizable function. We obtain the time evolution of these functions as pseudo-unitary, characteristic of a pseudo-Hermitian system. An explicit calculation of the cross section as a function of the wave number k reveals no sign of the bound state in the continuum.

Keywords: Bound states in the continuum; Darboux transformations; Jordan chain

PACS: 03.65.-w; 03.65.Ge; 03.65.Nk

1.Introduction

Bound states in the continuum (BICs) are a wave phenomenon that refers to states whose wave functions remain localized among a continuum of radiating waves with positive energies. This phenomenon of localized waves has been identified in atomic and molecular physics, optics, acoustics and other fields 1-5. One review concerning this type of bound states was recently published 6.

BICs were first proposed by von Neumann and Wigner in 1929 7, since then they have stimulated many experimental and theoretical research in various branches of physics. The first experimental evidence in this direction was reported by F. Capasso et al., they found an electronic bound state with energy greater than the barrier height in a semiconductor superlattice 8. More recently, BICs have been observed in optical waveguide arrays 9-11. M. Koirala et al. found a class of critical states embedded in the continuum in a one-dimensional arrangement of waveguides 12, and M. I. Molina et al. observed surface bound states in the continuum in a linear optical band of a discrete lattice 13.

In quantum physics, there exists a particular type of potentials that support BICs and are solely defined by their oscillatory asymptotic behaviour. Typically, a radial potential with an oscillatory asymptotic behaviour that falls to zero slowly,

V(r)=asinbrr+O1r2, (1)

supports a BIC with energy E=b2/4, if the parameters a and b satisfy the relation |a|>|b| 14. Potentials of the form (1) are known as von Neumann-Wigner potentials 7, and have been studied with methods of supersymmetric quantum mechanics (SUSY QM) 15, Darboux transformation 16 and the inverse scattering method (Gel’fand-Levitan equation) 17. Further development in the techniques for this purpose was made in the following years. A. A. Andrianov and A. V. Sokolov studied a non-Hermitian Hamiltonian in the whole axis with complex von Neumann-Wigner type potential obtained with SUSY methods, and found the normalized eigenfunction and associated eigenfunction and their orthogonal relations; as the Hamiltonian is non-Hermitian and self-orthogonality of the normalized eigenfunction occurs, they related the BIC to an exceptional point 18. A. Khelashvili and N. Kiknadze established a one to one correspondence between the decay law in von Neumann-Wigner type potentials and the asymptotic behaviour of the wave functions representing the bound states 19, while T. A. Weber and D. L. Pursey showed that by truncating a von Neumann-Wigner potential the BIC manifest itself as a resonance . E. Hernández et al. studied a particular spectral singularity produced by the coalescence of two BICs 21.

In this paper, in an attempt to develop further results in this field, particularly concerning the nature of the eigenfunction corresponding to a BIC, we study the Jost solutions of a radial Hamiltonian with a real von Neumann-Wigner type potential and obtain the Jordan cycle of length two of generalized eigenfunctions, one of which is the BIC, and their respective time evolution. Although the exact model presented in this paper may not be readily applied to a real physical system, it may prove useful in determining further properties of more realistic Hamiltonians. The paper is organized as follows: In Sec. 2, by performing a two times iterated and completely degenerated Darboux transformation with initial Hamiltonian the free particle, a von Neumann-Wigner type potential is obtained supporting a BIC with energy E = q 2 . In Sec. 3, the set of Jost solutions normalized to unit flux are explicitly constructed and it is shown that the unnormalized Jost solutions coalesce at k = q to give rise to a Jordan chain of rank two. In Sec. 4, the generalized eigenfunctions are found from the pole decomposition of the normalized Jost solutions and the corresponding Jordan chain is established. The time evolution of the generalized eigenfunctions is described in Sec. 5. In Sec. 6, the scattering matrix is obtained as a function of k and the cross section is calculated. In Sec. 7, a summary of the results and conclusions is presented.

2.Von Neumann-Wigner potentials and Darboux transformations

A particular class of completely degenerated Darboux transformations, with initial Hamiltonian the free particle, generates new potentials of von Neumann-Wigner type supporting any number of bound states embedded in the continuum (16). These transformations are obtained from Crum’s theorem, a generalization of the Darboux transformation, which states that the function

ψE=W(ϕ1,ϕ2,,ϕn,ϕE)W(ϕ1,ϕ2,,ϕn) (2)

is an eigenfunction of the Hamiltonian

Hn=-d2dr2+Un, (3)

with eigenvalue E, and the potential U n is given by

Un=V0-2d2dr2lnW(ϕ1,ϕ2,,ϕn). (4)

The auxiliary or transformation functions ϕi are eigenfunctions of the initial Hamiltonian H 0 with eigenvalues E i , i=1,,n, and ϕE is also an eigenfunction of H 0 with eigenvalue E. In the above expressions W(ϕ1,ϕ2,,ϕn) is the Wronskian of the eigenfunctions ϕi 22,23.

The completely degenerated case occurs when we take all E i energies close to each other, i.e.EiE¯+ϵi with ϵi1, and later taking the limit ϵi0. After this procedure, all eigenfunctions ϕi coalesce in ϕ and the respective eigenvalues E 𝑖 in E¯. From Crum’s theorem it follows that the function

ψE=W(ϕ,ϕ,,n-1ϕ,ϕE)W(ϕ,ϕ,,n-1ϕ) (5)

is an eigenfunction of the Hamiltonian H n , with eigenvalue E, and the potential U n is substituted by V n given as

Vn=V0-2d2dr2lnW(ϕ,ϕ,,n-1ϕ). (6)

The partial derivative is with respect to the energy E¯.

From here on, we consider as the initial Hamiltonian the free particle in spherical coordinates with potential V 0 = 0, and auxiliary eigenfunction

ϕ=sin(qr+δ(q)) (7)

with eigenvalue E¯=q2 and δ(q) an arbitrary phase shift which sets the parameters of the system. In this case, differentiation with respect to the energy E¯ is equivalent to differentiation with respect to wave number q in expressions (5) and (6), and in what follows we derive with respect to q.

Let us consider now the simplest case that produces a von Neumann-Wigner potential: n = 2. The Hamiltonian H 2 is given in (3), with the potential V 2 obtained from (6) as

V2(r)=-2d2dr2lnW(ϕ,ϕ), (8)

and calculating the Wronskian W2W(ϕ,ϕ) with (7) we obtain:

W2(r)=12sin2θ-2qγ, (9)

where

θ=qr+δ(q), (10)

γ=r+δ'(q), (11)

and as a convenient notation the prime in δ'(q) indicates the first derivative of δ(q).

Using (9) in (8) the potential obtained is:

V2(r)=32q2(sinθ-qγcosθ)sinθ(sin2θ-2qγ)2, (12)

with asymptotic behaviour given by:

V2(r)=-4qsin2θr+O1r2, (13)

and comparing it with (1) we see that it is a potential of von Neumann-Wigner type and, given that a=-4q and b=2q, supports a bound state in the continuum with energy E=q2.

A requirement for the validity of the Darboux transformation is the absence of singularities in the new potential not present in the initial potential. From (12) we see that the singularities of V 2 occur at the zeros of W 2 . The Wronskian W 2 as a function of r grows linearly with a negative slope, and has only one real zero.

As the Hamiltonian H 2 is defined in the positive semi-axis, we set the condition

W2(0)<0, (14)

which locates the real zero for negative r and therefore setting the potential V 2 an analytical function of r in the physical space. Condition (14) provides a differential relation for the phase shift δ(q) as a function of q. Evaluating (9) in r = 0 we get,

W2(0)=12sin2δ(q)-2qdδ(q)dq,

and defining t(q)=tanδ(q) we can write condition (14) as

qdt(q)dq-t(q)>0. (15)

This relation is readily solved by considering a positive constant β to eliminate the inequality, and solving the resulting differential equation yields the result

δ(q)=arctan(αq-β), (16)

with α and β real constants, and β > 0.

The potential V 2 in (12) is now determined by the parameters α and β. Figure 1 shows the potential V 2 as a function of r for a given choice of parameters.

FIGURE 1 Potential V2 (continuous line) and density probability of the eigenfunction of the bound state embedded in the continuum (dashed line) for parameters values α = 2, β = 3 and q = 1. 

3. Jost solutions of the Hamiltonian H2

The Schrödinger equation for the scattering problem of potential V 2 is

H2f±(k,r)=k2f±(k,r), (17)

for positive energies E=k2, where f±(k,r) are the two linearly independent unnormalized Jost solutions of Hamiltonian H2, which behave asymptotically as outgoing and incoming spherical waves 24. They are obtained from the Darboux transformation in (5) with n = 2 and ϕE are the free particle wave solutions e±ikr with eigenvalue E=k2:

f±(k,r)=W(ϕ,ϕ,e±ikr)W(ϕ,ϕ).

As the last column of the Wronskian in the numerator is proportional to e±ikr, the above expression can be written as

f±(k,r)=w±(k,r)W2(r)e±ikr, (18)

with w±(k,r) a complex function of real arguments k and r defined as

w±k,r=-12k2+q2sin2θ+k2-q2qγ±i2jq sin2θ (19)

Using the expressions (9) and (19) in (18), we can write the unnormalized Jost solutions in explicit form as

f±k,r=2k2-q2qγ-k2+q2sin2θ ±4ikq sin2θe±ikrsin2θ-2qγ, (20)

and from the linear behaviour of γ in r for large values of r we find the asymptotic behaviour:

f±(k,r)=-(k2-q2)+O1re±ikr. (21)

Hence, to obtain the two Jost solutions of H 2 normalized to unit flux at infinity, we must divide (20) by the factor -(k2-q2),

F±(k,r)=-f±(k,r)k2-q2, (22)

and the Jost solutions exhibit a simple pole at k = q.

In Appendix A, it is shown that the previous results can be obtained in an alternative way using the confluent case of the intertwining method of SUSY QM.

Hamiltonian H 2 has its spectrum defined for positive energies and, for each spectral point, there exist two linearly independent eigenfunctions. However, at the point k = q the two unnormalized Jost solutions coalesce. The Wronskian of the unnormalized Jost solutions can be obtained with the asymptotic behaviour given in (21):

W(f+,f-)=-2ik(k2-q2)2, (23)

which vanishes at the point k = q and therefore the unnormalized Jost solutions are linearly dependent at that spectral point. An eigenfunction is lost and the basis of linearly independent eigenfunctions of the Hamiltonian H 2 appears to be incomplete. In its place a Jordan chain of two generalized eigenfunctions is formed. The subspace spanned by the generalized eigenfunctions is in the domain of H 2 for E=q2.

4. Poles of the Jost solutions and Jordan chain

To obtain the generalized eigenfunctions, we rewrite the normalized Jost solution as a decomposition of its pole in a sum of singular and regular parts. From (22) and using (20) we can write

F±(k,r)=1+4q(qcosθiksinθ)sinθ(k2-q2)(sin2θ-2qγ)e±ikr, (24)

which can also be written as

F±(k,r)=ψB(q,r)k2-q2+ψR±(k,r), (25)

with the following defined functions:

ψBq,r=limkqk2-q2F±k,r=4q2sinθsin2θ-2qγeiδ(q) (26)

and

ψR±(k,r)=F±(k,r)-ψB(q,r)k2-q2, (27)

where ψB(q,r) is a square integrable function of r, while ψR±(k,r) are analytic functions of k = q, and behave asymptotically in 𝑟 as outgoing and incoming spherical waves. The explicit form of ψR±(k,r) obtained form (27) is

ψR±k,r=e±ikr+e±ik,rr-2qk+qk-q-e±i(k+qr+2δq)k+qψB(q,r)2q (28)

and its expression in k = q is given by

ψR±(q,r)=e±iqr+[1-e±2iθ±i2qr]ψB(q,r)4q2. (29)

After some algebraic manipulation the previous expression can be written as

ψR±(q,r)=1iqδ'ψB(q,r)2q2+ψA(q,r), (30)

and ψA(q,r), the associated eigenfunction to ψB(q,r), is defined as

ψA(q,r)=-2qγcosθsin2θ-2qγeiδ(q), (31)

which is a bounded, non-normalizable function.

As eigenfunctions of the Hamiltonian H 2 , the normalized Jost solutions satisfy the time independent Schrödinger equation

H2F±(k,r)=k2F±(k,r) (32)

for all k, except k = q where F±(k,r) is not defined. To explore the limit kq we substitute (25) in (32):

H2ψBq,rk2-q2+ψR±(k,r)=k2ψBq,rk2-q2+ψR±k,r. (33)

Multiplying (33) by (k2-q2) and taking the limit kq we obtain:

H2ψB(q,r)=q2ψB(q,r), (34)

the square integrable solution ψB(q,r) representing the bound state embedded in the continuum is an eigenfunction of H2 with energy E=q2. Figure 1 shows |ψB(q,r)|2 as a function of 𝑟.

Using the result (34) in (33) and taking the limit kq we get

H2ψR±(q,r)=q2ψR±(q,r)+ψB(q,r). (35)

However, as the additive term proportional to ψB(q,r) in (30) satisfies (34) it can be omitted and we can write

H2ψA(q,r)=q2ψA(q,r)+ψB(q,r), (36)

with ψA(q,r) the associated eigenfunction. Thus, ψB(q,r) and ψA(q,r) are generalized eigenfunctions of the Hamiltonian H 2 , and they form a Jordan chain of rank two for E=q2. The Jordan chain is the result of the coalescence of two energy levels (25), for a direct proof see Appendix B.

5. Pseudo-unitary time evolution of the generalized eigenfunctions

The two generalized eigenfunctions ψB(q,r) and ψA(q,r) belong to the same spectral point, E=q2; in consequence, they evolve in time together. Hence, it should be convenient to introduce a matrix notation to deal with the two together. From (34) and (36) we can write them as

H2Ψq,r=HqΨq,r, (37)

where

Ψ(q,r)=ψB(q,r)ψA(q,r) (38)

is the two component vector of the doublet, and

Hq=q201q2 (39)

is the 2 x 2 energy matrix.

The time dependent generalized eigenfunctions are

Ψr,t=Uq,tΨq,r, (40)

where Uq,t is the 2 x 2 matrix of time dependent coefficients and gives the time evolution of the wave function Ψ(q,r).

Substitution of Ψ(r,t) in the time dependent Schrödinger equation gives the following set of coupled equations written in matrix form

iUq,ttΨq,r=Uq,tH2Ψq,r=Uq,tHqΨq,r. (41)

Making abstraction of Ψ(q,r), we obtain

iU(q,t)t=U(q,t)H(q) (42)

Integrating equation (42) we get

Uq,t=e-iH(q)t (43)

writing Hq in explicit form in (43), and computing the exponential, we obtain

Uq,t=e-iq2t10-it1 (44)

Substitution of the expression (44) in (40) gives the evolution in time of the two generalized components of the doublet Ψ(r,t):

ψB(r,t)=ψB(q,r)e-iq2t (45)

and

ψA(r,t)=ψA(q,r)e-iq2t-itψB(q,r)e-iq2t. (46)

The component ψB(r,t), describing the time evolution of the bound state eigenfunction embedded in the continuum, exhibits a unitary evolution in time, while the component ψA(r,t) has a linear growth with time. Therefore, the wave function Ψ(r,t) grows linearly with time t. This type of behaviour has been found by Longhi et al. in a non-Hermitian Hamiltonian.

The above result is a direct consequence of the pseudo-Hermiticiy of Hamiltonian H2 at E=q2, represented as the matrix Hq in (39). An operator is pseudo-Hermitian if there exists a linear, invertible, Hermitian operator η such that 27

H=ηHη-1 (47)

The general form of η satisfying (47) for Hq in (39) is

η=abb0, (48)

with a and b any real parameters. In turn, U(q,t) is pseudo-unitary because Hamiltonian H2 at E=q2 is a pseudo-Hermitian operator. An operator is pseudo-unitary if its inverse and its adjoint satisfy the transformation 27

U=ηU-1η-1 (49)

as it may be verified by substitution of (44) and (48).

Therefore, the generalized eigenfunctions ψB(q,r) and ψA(q,r) in the Jordan cycle have a pseudo-unitary time evolution.

6. Cross section and scattering matrix

In this section, we will show that the scattering matrix has no singularities at the corresponding spectral point k = q, contrary to the case of conventional bound states with negative energies.

The scattering solution is defined as 24:

ψs(k,r)=i2[F-(k,r)-S(k)F+(k,r)], (50)

where F±(k,r) are the Jost solutions in (22) and S(k) is the scattering matrix defined as

S(k)=F-(k,0)F+(k,0), (51)

and F+(k,0) is the Jost function.

From (20) and (22) evaluated at r = 0 we obtain F±(k,0) and a direct substitution in (51) gives:

S(k)=2(k2-q2)qδ'-(k2+q2)sin2δ-4ikqsin2δ2(k2-q2)qδ'-(k2+q2)sin2δ+4ikqsin2δ (52)

which can be written as

S(k)=e2iΔ(k), (53)

where Δ(k) is the phase shift and is given by

Δ(k)=-arctan4kqsin2δ2(k2-q2)qδ'-(k2+q2)sin2δ. (54)

Taking the limit kq we get Δ(q)=δ. Hence, the scattering matrix evaluated at k = q is finite and equal to

limkqS(k)=e2iδ(q). (55)

However, there can be singularities of S(k) for different values of k. The singularities of S(k) are the zeros of the Jost function F+(k0,0)=0:

(2qδ´-sin2δ)k02+4iqsin2δk0-2qδ´+sin2δq2=0 (56)

which is a quadratic equation for k0. In terms of α and β we have

βk02+2iq(αq-β)2k0-2αq+β=0. (57)

The zeros in the fourth quadrant of the complex k-plane near the real axis may be resonances, while zeros on the imaginary positive axis correspond to bound states with negative energy.

The cross section is defined as

σ(k)=4πk2sin2Δ(k), (58)

and with the explicit form of Δ(k) in (54) we are able to observe its behaviour and dependence on k. Figure 2 shows σ(k) as a function of k and for the chosen values of parameters α and β a resonance shape is found, belonging to the value k0=2/3-i/3, far from the value k = q = 1 of the BIC. The BIC has no effect in the cross section, only when the system is perturbed the BIC may manifest itself as a resonance 20.

FIGURE 2 Cross section σ(k) for parameters values α=2,β=3 and q = 1. The Breit-Wigner peak in σ(k) corresponds to a resonance, far from the value k = q = 1 of the BIC, located at k0=23-i/3

7. Summary and conclusions

In this work, we presented a von Neumann-Wigner type potential V 2 constructed by means of a two times iterated and completely degenerated Darboux transformation. The Hamiltonian H 2 and the free particle Hamiltonian H0 are isospectral, and to each point in their continuous spectrum corresponds two linearly independent Jost solutions, which behave at infinity as incoming and outgoing waves. However, we have shown that in the continuous spectrum of H 2 there is a singular point, with energy E=q2, such that the two unnormalized Jost solutions are linearly dependent and coalesce to give rise to a Jordan chain of rank two of generalized eigenfunctions and a Jordan block representation of the Hamiltonian H 2 . The normalized Jost solutions have a simple pole at wave number k = q and after a pole decomposition the Jordan chain and respective generalized eigenfunctions are obtained. One of the generalized eigenfunctions is normalizable and corresponds to the BIC, the other is a bounded, non-normalizable function associated with the BIC. Finally, we obtained the time evolution of the generalized eigenfunctions: the BIC has a unitary time evolution, while the associated eigenfunction has a linear growth in time. Together, they exhibit a pseudo-unitary behaviour characteristic of a pseudo-Hermitian system. Finally, we have shown that the BIC is not associated with a singularity of the scattering matrix S(k) and, as a result, the BIC is not observed in the cross section σ(k).

Acknowledgements

We would like to thank C. Calcáneo for his interest in this paper. E. Hernández would like to thank Professor A. Mondragón for his advice and suggestions on this kind of problems and for his great human quality.

References

1. F. H. Stillinger and T. A. Weber, Phys. Rev. A 10 (1974) 1122. [ Links ]

2. F. H. Stillinger and D. R. Herrick, Phys. Rev. A 11 (1975) 446. [ Links ]

3. H. Friedrich and D. Wintgen, Phys. Rev. A 31 (1985) 3964. [ Links ]

4. S. Longhi, Opt. Lett. 39 (2014) 1697. [ Links ]

5. A. A. Lyapina, D. N. Maksimov, A. S. Pilipchuk, and A. F. Sadreev, J. Fluid Mech. 780 (2015) 370. [ Links ]

6. C. W. Hsu, B. Zhen, A. D. Stone, J. D. Joannopoulos, and M. Soljačić, Nat. Rev. Mater. 1 (2016) 16048. [ Links ]

7. J. von Neumann and E. P. Wigner, Phys. Z. 30 (1929) 465. [ Links ]

8. F. Capasso, C. Sirtori, J. Faist, D. L. Sivco, S.-N. G. Chu, and A. Y. Cho, Nature 358 (1992) 565. [ Links ]

9. Y. Plotnik et al., Phys. Rev. Lett. 107 (2011) 183901. [ Links ]

10. G. Corrielli, G. Della Valle, A. Crespi, R. Osellame, and S. Longhi , Phys. Rev. Lett. 111 (2013) 220403. [ Links ]

11. S. Weimann et al., Phys. Rev. Lett. 111 (2013) 240403. [ Links ]

12. M. Koirala, A. Yamilov, A. Basiri, Y. Bromberg, H. Cao, and T. Kottos, New J. Phys. 17 (2015) 013003. [ Links ]

13. M. I. Molina, A. E. Miroshnichenko, and Y. S. Kivshar, Phys. Rev. Lett. 108 (2012) 070401. [ Links ]

14. M. Klaus, J. Math. Phys. 32 (1991) 163. [ Links ]

15. J. Pappademos, U. Sukhatme, and A. Pagnamenta, Phys. Rev. A 48 (1993) 3525. [ Links ]

16. A. A. Stahlhofen, Phys. Rev. A 51 (1995) 934. [ Links ]

17. T. A. Weber and D. L. Pursey, Phys. Rev. A 50 (1994) 4478. [ Links ]

18. A. A. Andrianov and A. V. Sokolov, SIGMA 7 (2011) 111. [ Links ]

19. A. Khelashvili and N. Kiknadze, J. Phys. A: Math. Gen. 29 (1996) 3209. [ Links ]

20. T. A. Weber and D. L. Pursey , Phys. Rev. A 57 (1998) 3534. [ Links ]

21. E. Hernández, A. Jáuregui, D. Lohr, and A. Mondragón, J. Phys. A: Math. Theor. 50 (2017) 265301. [ Links ]

22. M. M. Crum, Q. J. Math. 6 (1955) 121. [ Links ]

23. V. B. Matveev and M. A. Salle, Darboux Transformations and Solitons (Springer, Berlin, 1991). [ Links ]

24. R. G. Newton, Scattering Theory of Waves and Particles (Springer, Berlin, 2014). [ Links ]

25. A. A. Andrianov , F. Cannata, and A. V. Sokolov , Nucl. Phys. B 773 (2007) 107. [ Links ]

26. S. Longhi and G. Della Valle , Phys. Rev. A 89 (2014) 052132. [ Links ]

27. A. Mostafazadeh, J. Math. Phys. 45 (2004) 932. [ Links ]

28. D. Bermudez, D. J. Fernández, and N. Fernández-García, Phys. Lett. A 376 (2012) 692. [ Links ]

Appendix

A. Equivalence with the confluent case of SUSY QM

The completely degenerated case of the Darboux transformation and the confluent case of SUSY QM are equivalent methods for obtaining new, completely solvable, quantum systems from previously solved ones.

In SUSY QM of second order the relation between the initial Hamiltonian H0 and the transformed Hamiltonian H 2 is given by 28

H2B2=B2H0 (A.1)

where

H2=-d2dr2+V2 (A.2)

H0=-d2dr2+V0 (A.3)

B2=d2dr2+g(r)ddr+h(r) (A.4)

Operator B2 is known as the intertwining operator, and is a differential operator of second order, with g(r) and h(r) functions to be determined.

Therefore, by applying (A.1) and using V2, given in (12), we can find an explicit expression for the corresponding intertwining operator for the present problem. With V0 = 0 we obtain the following equations:

-2dg(r)dr+V2=0 (A.5)

-d2g(r)dr2-2dh(r)dr+V2g(r)=0 (A.6)

-d2h(r)dr2+V2h(r)=0. (A.7)

We notice that h(r) in (A.7) satisfies the equivalent equation H2h(r)=0; thus, h(r) is readily obtained from (20) by taking k = 0 :

h(r)=-q2sin2θ+2qγsin2θ-2qγ. (A.8)

Using (A.5) in (A.6) we write:

ddr-12V2+g2(r)-2h(r)=0,

and as we know both V2 and h(r) we get g(r) as:

g(r)=±12V2+2h(r)+c,

with c an arbitrary integration constant. Using V2 and h(r) given in (12) and (A.8), respectively, we get

g(r)=±16q2sin4θ(sin2θ-2qγ)2+2q2+c,

and choosing c=-2q2 the function g(r) is simplified to

g(r)=±4qsin2θsin2θ-2qγ. (A.9)

Differentiating (A.9) once with respect to r and substituting in (A.5), we conclude that we must take the positive root. Hence, with (A.8) and (A.9) the intertwining operator B2 in (A.4) is given by

B2=d2dr2+4qsin2θsin2θ-2qγ dr ddr-q2sin2θ+2qγsin2θ-2qγ. (A.10)

If ϕk is an eigenfunction of H0 with energy eigenvalue E=k2, satisfying the eigenvalue equation H0ϕk=k2ϕk, then B2ϕk is an eigenfunction of H2 for the same eigenvalue

H2B2ϕk=k2B2ϕk. (A.11)

Using the same eigenfunctions for the free particle of outgoing and incoming waves ϕk±=e±ikr we get the following eigenfunctions for H2:

B2ϕk±=2k2-q2qγ-k2+q2sin2θ±4ikq sin2θ×e±ikrsin2θ-2qγ, (A.12)

which are exactly the same Jost solutions in (20) obtained with the method of Darboux transformation, thus proving the equivalence of both methods.

B. Explicit coalescence of two energy levels and Jordan chain

In Sec. 2 the completely degenerated case of the Darboux transformation generalized in Crum’s theorem was presented, and the case n = 2 was studied. An equivalent way of approaching the problem consists in performing the Darboux transformation with two different transformation functions

ϕ1=sin(q1r+δ(q1)) (B.1)

ϕ2=sin(q2r+δ(q2)) (B.2)

with respective energy eigenvalues E1=q12 and E2=q22, and then consider the limit when both energy eigenvalues coalesce.

The potential U 2 is given by (4) with n = 2 and 𝑉 0 =0. A calculation of the Wronskian of ϕ1 and ϕ2 gives

W(ϕ1,ϕ2)=q2sinθ1cosθ2-q1cosθ1sinθ2, (B.3)

with θi=qir+δ(qi). And the potential is

U2(r)=-2(q12-q22)(q22sin2θ1-q12sin2θ2)(q2sinθ1cosθ2-q1cosθ1sinθ2)2. (B.4)

Now we calculate the eigenfunctions of the Hamiltonian with potential (B.4) and study the limit when q2q1. From (2) with n = 2 and the free particle solutions ϕE=e±ikr, the eigenfunctions are obtained as

φ±k,r=q1k2-q22cosθ1sinθ2-q2k2-q12sinθ1cosθ2ikq12-q22sinθ1sinθ2q2sinθ1cosθ2-q1cosθ1sinθ2e±ikr (B.5)

The Wronskian of the eigenfunctions is directly calculated and has the form

Wφ+,φ-=2ik(k2-q12)(k2-q22) (B.6)

and we see that for eigenvalues E=q12 and E=q22 the Wronskian vanishes. In the limit q2q1=q expression (23) is recovered.

The respective eigenfunctions for the energies mentioned above are calculated by direct substitution k=q1 and k=q2 in (B.5) and we obtain

φ±q1,r=q1q12-q22sinθ2 eiδ(q1)q2sinθ1cosθ2-q1cosθ1sinθ2 (B.7)

and

φ±q2,r=q2q12-q22sinθ1 eiδ(q2)q2sinθ1cosθ2-q1cosθ1sinθ2 (B.8)

For q1q2, the eigenfunctions φ±(q1,r) and φ±(q2,r) are linearly independent.

In order to study the coalescence of the two energy levels we denote q1=q and q2=q+ϵ, and take the limit ϵ0. Because ϵ1 we consider the following series expansions

sinθ2=sinθ+ϵγcosθ+ϵ22δcosθ-γ2sinθ+cosθ2=cosθ-ϵγsinθ-ϵ22δsinθ+γ2cosθ+,

with γ=r+δ'(q)

Substituting both series in (B.7) and (B.8), we get the following expressions

φ)±q,r=-4q2sinθ+2qϵ2qγcosθ+sinθ+O(ϵ2)sin2θ-2qγ-ϵqδ´´+2γsin2θ+O+(ϵ2)eiδ(q) (B.9)

and

φ±q+ϵ,r=-4q2sinθ+6qϵsinθ+O(ϵ2)sin2θ-2qγ-ϵqδ´´+2γsin2θ+O(e2)eiδ(q+ϵ) (B.10)

and taking the limit ϵ0 we get, respectively,

limϵ0φ±q,r=-4q2sinθ eiδqsin2θ-2qγ=-ψB(q,r) (B.11)

and

limϵ0φ±q+ϵ,r=-4q2sinθ eiδqsin2θ-2γ=-ψBq,r, (B.12)

which means that both eigenfunctions coalesce to the same square integrable function ψB(q,r), defined in (26) and representing the bound stated embedded in the continuum.

When two eigenfunctions and their respective eigenvalues coalesce, a Jordan chain of rank two is formed and the associated eigenfunction, completing the Jordan chain, is given by 25

ψG±q,r=φ±(q,r)ϵ-φ±(q+ϵ,r)ϵϵq2-ϵ(q+ϵ)2|ϵ=0 (B.13)

Differentiating (B.9) and (B.10) with respect to ϵ and plugging the results in (B.13) we obtain the following expression for the generalized eigenfunction

ψG±q,r=-1±iqδ´ψBq,r2q2-ψAq,r, (B.14)

with ψA(q,r) defined in (31). In expressions (B.11), (B.12) and (B.14), obtained from the explicit coalescence of two energy levels, we notice a global sign difference to their counterparts in (26) and (30). This comes from the normalization factor -(k2-q2) used to normalize the Jost solutions (22) to unit flux at infinity.

Received: March 14, 2018; Accepted: May 16, 2018

Creative Commons License This is an open-access article distributed under the terms of the Creative Commons Attribution License