Shannon information entropies for position-dependent mass Schrödinger problem with a hyperbolic well
Guo-Hua Sun†a), Popov Dušan‡b), Camacho-Nieto Oscar§c), Shi-Hai Dong¶c)
Cátedra CONACyT, Centro de Investigación en Computación, Instituto Politécnico Nacional, UPALM, Mexico D. F. 07738, Mexico
Politehnica University Timisoara, Department of Physical Foundations of Engineering, Bd. V. Parvan No. 2, 300223 Timisoara, Romania
CIDETEC, Instituto Politécnico Nacional, UPALM, Mexico D. F. 07700, Mexico

E-mail: sunghdb@yahoo.com

E-mail: dusan_popov@yahoo.com

§E-mail: ocamacho@ipn.mx

Corresponding author. E-mail: dongsh2@yahoo.com

Abstract

The Shannon information entropy for the Schrödinger equation with a nonuniform solitonic mass is evaluated for a hyperbolic-type potential. The number of nodes of the wave functions in the transformed space z are broken when recovered to original space x. The position S x and momentum S p information entropies for six low-lying states are calculated. We notice that the S x decreases with the increasing mass barrier width a and becomes negative beyond a particular width a, while the S p first increases with a and then decreases with it. The negative S x exists for the probability densities that are highly localized. We find that the probability density ρ( x) for n = 1, 3, 5 are greater than 1 at position x = 0. Some interesting features of the information entropy densities ρ s( x) and ρ s( p) are demonstrated. The Bialynicki–Birula–Mycielski (BBM) inequality is also tested for these states and found to hold.

PACS: 03.65.–w; 03.65.Ge; 03.67.–a
Keyword: position-dependent mass; Shannon information entropy; hyperbolic potential; Fourier transform
1.Introduction

In the late 1940s Shannon proposed the idea of assigning probabilities to the outcome of uncertain events and introduced the information entropy to measure the uncertainty.[1] This concept has crossed a lot of barriers between traditionally separated disciplines and became a universal concept of statistical physics.[26] In 1975, Beckner, Bialynicki– Birula, and Mycieslki obtained an entropic uncertainty relation[7, 8] as Sx + SpD(1 + ln π ), where D represents the spatial dimension. In the one-dimensional case, the position– space (Sx) and momentum– space (Sp) information entropies are defined, respectively, by

where ψ (x) is a normalized eigenfunction in spatial coordinates and ϕ (p) is its normalized Fourier transform. It should be pointed out that the interval of the integration with respect to the variable x depends on the concrete quantum system.

Apart from their own significance, the entropic uncertainty relations have been utilized widely in both physics and chemistry.[913] Up to now, we have recognized that it is difficult to obtain explicit and analytical expressions for the formulas (1) due to the logarithmic factors in the integrals. At most they can be obtained analytically only for a few low-lying states of the harmonic oscillator, [1, 14] the Pö schl– Teller potential, [15, 16] the Morse potential, [15, 17] the Coulomb potential in high excited states, [18] as well as position– space entropy for the potential isospectral to the Pö schl– Teller potential, [19] several classical orthogonal polynomials, [2022] and other studies.[2325] The main reason why the high lying states cannot be treated is from the difficulty in their Fourier transforms. We have to say that the above studies are done for those well-known physical potentials. Except for them, we have studied the quantum information entropies for some interesting and particular quantum systems, which are not familiar to general readers, such as the symmetrically and asymmetrically trigonometric Rosen– Morse potential, [26, 27] the Pö schl– Teller-like potential, [28] a squared tangent potential, [29] hyperbolic-type potential, [30] and hyperbolic double-well potential.[31] It should be pointed out that almost all contributions to this topic have been done only in the constant mass Schrö dinger equation case[131] except for Ref. [32], in which we have studied the position-dependent mass Schrö dinger equation with a null potential. In this work, we attempt to study the Shannon entropy for the position-dependent mass Schrö dinger equation for a particle with a nonuniform solitonic mass density in the case of a hyperbolic-type potential.

The present paper is organized as follows. In Section 2 we present exact solutions of the system. In Section 3 we illustrate position and momentum information entropy densities, ρ s(x) and ρ s(p), respectively, and probability density ρ (x). We study Sx and Sp numerically for those six lowest-lying states. We find that the Sx entropy becomes negative beyond a particular mass width a due to the highly localized probability densities. Notice that the momentum information entropy Sp first increases with the parameter a and then decreases with it. Some concluding remarks are given in Section 4.

2.Smooth mass profile and its exact solutions

The non-ambiguous position-dependent mass Schrö dinger equation can be written as[3337]

In the present study, we choose the following smooth effective mass distribution,

which is a suitable representative of a solitonic profile[36, 37] found in several effective models of condensed matter and low-energy nuclear physics. It is shown in Fig. 1 that the position-dependent mass is symmetric with respect to the variable x. Clearly, the mass barrier width is inversely proportional to the parameter a. In this case, if we take y = ax then the effective differential equation reads

Fig. 1. Position-dependent mass as a function of the position x.

Defining ψ (y) = coshv (y)ϕ (y) and substituting it into the above equation leads to

The transformation

maps y ∈ (− ∞ , ∞ ) → z ∈ (− π /2, π /2). The corresponding equation becomes

In the derivation of Eq. (6), we have used the following useful relations

If choosing v = − 1/2, we obtain

Thus, equation (6) is equivalent to a regular constant-mass stationary Schrö dinger equation (8) for a particle of mass m0 in a confining potential

In order to make Eq. (8) soluble, we may take V(z) = − (3a2ħ 2/(8m0)) sinh2(x) − a2ħ 2/(4m0) in order to cancel the second term in if we use the relation sinh(x) = tan(z). In order to study the Shannon information entropy, let us write down the symmetric and antisymmetric solutions in the z space[35]

which, in turn, yield the solutions in the x space

with the energy levels ε n2 (n = 0, 1, 2, 3, … .).

For convenience, we list a few lowest-lying normalized states to calculate the Shannon information entropy

We give a useful remark about the solutions listed above. The analytical and explicit solutions have been obtained by Cunha and Christiansen in Ref. [35], but we have to point out that the solution ψ 1(x) for the special case n = 0 was not discussed. A possible reason is that this case was not considered from the energy level, i.e., ε = 0. In fact, it is known that such a quantum state with zero energy is very special and is also important in studying the relation between the number of bound states and the scattering phase shifts. This kind of phenomenon is called the Levinson theorem.[38, 39] If the wave function with the zero-energy is square integral and it tends to zero at infinity, this solution is physically acceptable. Otherwise, it is called a half bound state, which is not a bound state. On the other hand, we notice that the wave functions in the x space, the number of the nodes of the wave functions are broken. This is unlike that in the usual quantum systems. We show the wave functions in Figs. 2 and 3. Notice that the basic variations of the wave functions are very similar since the change of the parameter a directly modifies the range of the variable x.

Fig. 2. Wave functions of the first (a) and second (b) states with definite parity depending on the parameter a as a function of the position x.

Fig. 3. Wave functions (a = 1) with definite parity as a function of the position x.

3.Information entropy

The position- and momentum-space information entropies for the one-dimensional potential can be calculated by using Eqs. (1). In general, explicit derivations of the information entropy are quite difficult. We have to study them numerically. In order to calculate the momentum information entropy, one has to transform the wave functions obtained in position space to momentum space using the Fourier transform. We also illustrate the position and momentum information entropy densities ρ s(x) = | ψ (x)| 2 ln | ψ (x)| 2 and ρ s(p) = | ϕ (p)| 2 ln | ϕ (p)| 2, since they play a similar role as the probability density ρ = | ψ (x)| 2 in quantum mechanics.

Due to the difficulty in calculating wave functions in momentum space, we only consider a few lowest states. For example, using the wave functions of the ground state, we can derive the corresponding momentum representation through the Fourier transform

where p1 = p/a (also below) and F1(a; b, c; d; − 1; − i, i) represents the Appell function.

For the excited state n = 2, after some complicated algebraic manipulations we have

where

To obtain the analytical Fourier transforms of the high lying states n = 3− 6, we have to choose different approaches to treat the symmetrical and antisymmetrical solutions. For the symmetrical solutions , considering their parity property, we use the following formula[40] for the sech(x) to study the Fourier transform

In this case, we are able to study this kind of integral (k = 2m + 1) as follows:

However, for the wave functions with odd parity, says , their Fourier transforms become very complicated. For simplicity we do not list them here. The wave functions in the momentum space are displayed in Figs. 4 and 5.

Fig. 4. Wave functions in momentum space for the first (a) and second (b) states with definite parity depending on the parameter a as a function of the momentum p.

Fig. 5. Wave functions (a = 1) with definite parity as a function of the momentum p.

Characteristic features of the position and momentum probability densities are displayed in Figs. 6 and 7. We find that ρ (x) for the states n = 1, 3, 5 is greater than 1 at positions x = 0. This is not surprising since the most of the density is less than 1, the curve has to rise higher than 1 in order to have a total area of 1 as required for all probability distributions. Let us explain this in some detail. As pointed out by Catalá n et al., [41] this situation is mathematically possible, but it is unfound from a physical point of view when we consider the features of extension of a statistical measures of complexity introduced by LMC[42] is extended to continuous systems.

Fig. 6. Position– space densities (a = 1) for the n = 1– 6 eigenstates as a function of the position x.

Fig. 7. Momentum– space densities (a = 1) for the n = 1– 6 eigenstates as a function of the momentum p.

We find that the maximum amplitudes of ρ (x) increase with the high excited states. On the other hand, we show the position and momentum entropy densities ρ s(x) and ρ s(p) in Figs. 8 and 9 with the parameter a = 1. It is found that the behavior of the maximum amplitude of the ρ s(p) is contrary to that of ρ s(x). Specially, we find that ρ s(x) and ρ s(p) for n = 2, 4, 6 are equal to zero at x = p = 0 since all the wave functions both in the position x and both in the momentum p are zero. We notice that all of them are symmetric with respect to the variable x. This arises from the symmetric mass barrier as well as the chosen potential sinh2(x) which we study here. In short, narrower mass barriers yield entropy densities that are more (less) localized in position (momentum) space. It should be addressed that the behaviors of the ρ s and ρ both in the position and momentum in the case of the various parameter values a are very similar to those in the case a = 1. We do not show them for simplicity.

Fig. 8. Position– space entropy probability densities (a = 1) for the n = 1– 6 eigenstates as a function of the position x.

Fig. 9. Momentum– space entropy probability densities (a = 1) for the n = 1– 6 eigenstates as a function of the momentum p.

In Fig. 10, we show the position information entropy Sx as a function of the parameter a and find that it decreases with this parameter. With increasing a, the mass barrier becomes narrower, thus the time that the moving particle passes it becomes shorter, the system becomes more stable. Beyond some value a, the Sx becomes negative. This means that the physical “ object” , with nonuniform distributed solitonic mass, condenses or rather falls into order in a greater degree. For a fixed, the Sx decreases with the energy levels. That is to say, when the energy of the system increases, the quantum system becomes more localized in the position space. In particular, we find that the momentum space Shannon entropy Sp first increases with the parameter a and then decreases with it.

Fig. 10. Position information entropy Sx as a function of the parameter a for the n = 1− 6 eigenstates.

In Table 1 we present the numerical results of the information entropies Sx, Sp, and their sum for the low-lying six states and various parameter values a. The behavior coincides with that illustrated in Figs. 10 and 11. Particularly, we find that the system entropy could become negative for some given parameter values a. How to understand the negative entropy? It might be understood that the moving particle becomes condensed so that it does not move at all. The system becomes more stable at this moment. Note that with the increasing Sp, the Sx decreases but only to the extent that their sum stays above the stipulated lower bound of the value (1 + lnπ ).

Fig. 11. Momentum information entropy Sp as a function of the parameter a for the low-lying states.

Table 1. Information entropies of the n = 1− 6 eigenstates for different a values.
4.Concluding remarks

In this paper we have presented the Shannon entropy for the position and momentum eigenstates of the position-dependent Schrö dinger equation for a particle with a nonuniform solitonic mass density in the case of a hyperbolic-type potential. The position Sx and momentum Sp information entropies for n = 1− 6 are calculated. The fact that the Sx might be negative arises from the probability densities that are highly localized. We find that the position information entropy Sx decreases for narrower mass widths (larger a values), but the behavior of Sp first increases with the parameter a and then decreases with it. The mass barrier being inversely proportional to the a parameter determines the stability of the system. Some interesting features of the information entropy densities, ρ s(x) and ρ s(p), are demonstrated. The Bialynicki– Birula– Mycielski (BBM) inequality is also verified for a number of states. Before ending this work, we give a few useful remarks on the Shannon information entropy for the Schrö dinger equation with a constant mass or with a position-dependent mass case. Generally speaking, the quantum systems with a constant mass could be solved rather easily in comparison with those with a position-dependent mass. As we know, most of quantum systems with well-known physical potentials have been solved (see, e.g., Refs. [1]– [31]), but the quantum systems with a position-dependent mass become rather complicated because of the non-commutative property between the operator ∇ = d/dx and m(x) as shown in Eq. (2). Therefore, it is not easy to obtain the exact solutions of those quantum systems with a position-dependent mass even for those simplest cases as studied in Ref. [32] and in the present work. Nevertheless, the good points for position-dependent mass case lie in providing us for some new features of the Shannon information entropy related to the mass barrier width. On the other hand, let us give a brief comparison of the studies performed in Ref. [32] and in the present work. When we consider the effective potential (9) which is obtained by some suitable canonical transforms, the different choices of the V(z) determine different properties of studied quantum system, i.e., the solutions will be totally different. For instance, we choose V(z) = 0 in Ref. [32] and call it as the so-called null potential. However, in this work we choose V(z) as the present form (see below Eq. (9)) in order to make this system soluble. The different solutions given both in Ref. [32] and in this work result in the different properties of the Shannon entropy.

Acknowledgments

We would like to thank the kind referee for invaluable and positive suggestions, which have improved the manuscript greatly. This work is supported partially by project 20150964-SIP-IPN, COFAA-IPN, Mexico.

Reference
1 Shannon C E 1948 Bell Syst. Tech. J. 27 623 Reprinted in Shannon C E 1993 >Collected Papers New York IEEE Press [Cited within:4]
2 Yáñez R J, Van Assche W and Dehesa J S 1994 Phys. Rev. A 50 3065 DOI:10.1103/PhysRevA.50.3065 [Cited within:1]
3 Van Assche W, Yáñez R J and Dehesa J S 1995 J. Math. Phys. 36 4106 DOI:10.1063/1.530949 [Cited within:1]
4 Aptekarev A I, Dehesa J S and Yáñez R J 1994 J. Math. Phys. 35 4423 DOI:10.1063/1.530861 [Cited within:1]
5 Everett H 1973 The Many World Interpretation of Quantum Mechanics New Jersey Princeton University Press [Cited within:1]
6 Hirschmann I IJr 1957 Am. J. Math. 79 152 DOI:10.2307/2372390 [Cited within:1]
7 Beckner W 1975 Ann. Math. 102 159 DOI:10.2307/1970980 [Cited within:1]
8 Bialynicki-Birula I and Mycielski J 1975 Commun. Math. Phys. 44 129 DOI:10.1007/BF01608825 [Cited within:1]
9 Orlowski A 1997 Phys. Rev. A 56 2545 DOI:10.1103/PhysRevA.56.2545 [Cited within:1]
10 Atre R, Kumar A, Kumar C N and Panigrahi P 2004 Phys. Rev. A 69 052107 DOI:10.1103/PhysRevA.69.052107 [Cited within:1]
11 Romera E and de los Santos F 2007 Phys. Rev. Lett. 99 263601 DOI:10.1103/PhysRevLett.99.263601 [Cited within:1]
12 Galindo A and Pascual P 1978 Quantum Mechanics Berlin Springer [Cited within:1]
13 Angulo J C, Antolin J, Zarzo A and Cuchi J C 1999 Eur. Phys. J. D 7 479 DOI:10.1007/s100530050375 [Cited within:1]
14 Majernik V and Opatrný T 1996 J. Phys. A: Math. Gen. 29 2187 DOI:10.1088/0305-4470/29/9/029 [Cited within:1]
15 Dehesa J S, Martínez-Finkelshtein A and Sorokin V N 2006 Mol. Phys. 104 613 DOI:10.1080/00268970500493243 [Cited within:2]
16 Coffey M W 2007 Can. J. Phys. 85 733 DOI:10.1139/P07-062 [Cited within:1]
17 Aydiner E, Orta C and Sever R 2008 Int. J. Mod. Phys. B 22 231 DOI:10.1142/S021797920803848X [Cited within:1]
18 Dehesa J S, Van Assche W and Yáñez R J 1997 Methods Appl. Math. 4 91 [Cited within:1]
19 Kumar A 2005 Ind. J. Pure. Appl. Phys. 43 958 [Cited within:1]
20 Dehesa J S, Yáñez R J, Aptekarev A I and Buyarov V 1998 J. Math. Phys. 39 3050 DOI:10.1063/1.532238 [Cited within:1]
21 Buyarov V S, Dehesa J S and Martinez-Finkelshtein A 1999 J. Approx. Theory 99 153 DOI:10.1006/jath.1998.3315 [Cited within:1]
22 Buyarov V S, López-Artés P, Martínez-Finkelshtein A and Van Assche W 2000 J. Phys. A: Math. Gen. 33 6549 DOI:10.1088/0305-4470/33/37/307 [Cited within:1]
23 Katriel J and Sen K D 2010 J. Comp. Appl. Math. 233 1399 DOI:10.1016/j.cam.2008.04.039 [Cited within:1]
24 Patil S H and Sen K D 2007 Int. J. Quantum Chem. 107 1864 DOI:10.1002/(ISSN)1097-461X [Cited within:1]
25 Sen K D 2005 J. Chem. Phys. 123 074110 DOI:10.1063/1.2008212 [Cited within:1]
26 Sun G H and Dong S H 2013 Phys. Scr. 87 045003 DOI:10.1088/0031-8949/87/04/045003 [Cited within:1]
27 Sun G H, Dong S H and Naad S 2013 Ann. Phys. (Berlin) 525 934 DOI:10.1002/andp.v525.12 [Cited within:1]
28 Sun G H, Avila Aoki M and Dong S H 2013 Chin. Phys. B 22 050302 DOI:10.1088/1674-1056/22/5/050302 [Cited within:1]
29 Dong S, Sun G H, Dong S H and Draayer J P 2014 Phys. Lett. A 378 124 [Cited within:1]
30 Valencia-Torres R, Sun G H and Dong S H 2015 Phys. Scr. 90 035205 DOI:10.1088/0031-8949/90/3/035205 [Cited within:1]
31 Sun G H, Dong S H, Launey K D, Dytrich T and Draayer J P 2015 Int. J. Quantum Chem. DOI: 101002/qua. 24928 [Cited within:3]
32 Yańez-Navarro G, Sun G H, Dytrich T, Launey K D, Dong S H and Draayer J P 2014 Ann. Phys. 348 153 DOI:10.1016/j.aop.2014.05.018 [Cited within:5]
33 Yu J and Dong S H 2004 Phys. Lett. A 325 194 (references therein) [Cited within:1]
34 Yu J, Dong S H and Sun G H 2004 Phys. Lett. A 322 290 DOI:10.1016/j.physleta.2004.01.039 [Cited within:1]
35 Cunha M S and Christiansen H R 2013 Commun. Theor. Phys. 60 642 DOI:10.1088/0253-6102/60/6/02 [Cited within:2]
36 Panahiy H and Bakhshi Z 2010 Acta Phys. Pol. B 41 11 [Cited within:1]
37 Bagchi B, Gorain P, Quesne C and Roychoudhury R 2004 Mod. Phys. Lett. A 19 2765 DOI:10.1142/S0217732304016123 [Cited within:2]
38 Levinson N K 1949 Dan. Vidensk. Selsk. Mat. Fys. Medd. 25 9 [Cited within:1]
39 Dong S H, Hou X W and Ma Z Q 1998 Phys. Rev. A 58 2790 DOI:10.1103/PhysRevA.58.2790 [Cited within:1]
40 Gradshteyn I S and Ryzhik I M 1994 Tables of Integrals, Series and Products(5th edn. New York Academic Press [Cited within:1]
41 Catalán R G, Garay J and López-Ruiz R 2002 Phys. Rev. E 66 011102 DOI:10.1103/PhysRevE.66.011102 [Cited within:1]
42 López-Ruiz R, Mancini H L and Calbet X 1995 Phys. Lett. A 209 321 DOI:10.1016/0375-9601(95)00867-5 [Cited within:1]