Uncertainty relations for modified isotropic harmonic oscillator and Coulomb potentials S.H. Patil a , K.D. Sen b,∗ a Department of Physics, Indian Institute of Technology, Mumbai 400 076, India b School of Chemistry, University of Hyderabad, Hyderabad 500 046, India Communicated by V.M. Agranovich Abstract The dimensional analyses of the position and momentum variances which define the Heisenberg uncertainty product are carried out for two non-relativistic model central potentials generated by adding a/r 2 term to (i) the isotropic harmonic oscillator, and (ii) the Coulombic hydrogenlike potentials. The uncertainty products are shown to be independent of the scaling of the part (i) and (ii) but are dependent on the strength a of the additional term. The scaling properties are found to be reflected in the entropic uncertainty measure of the Shannon information entropy sum and the Fisher information product. Numerical results are presented in support of the analytic results derived. Keywords: Isotropic harmonic oscillator; H atom; Davidson potential; Kratzer potential; Central potentials in D dimensions; Heisenberg uncertainty relation; Shannon entropy; Fisher information measure 1. Introduction Uncertainty relations are the basic properties of quantum mechanics. In particular, Heisenberg uncertainty principle [1] for the product of the uncertainties in position and momentum, in terms of Planck’s constant, 1 σx σp h̄, 2 2 σx2 = x − x , 2 σp2 = px − px , (1) is an important element of quantum properties. In this case the Gaussian wave functions have the minimum uncertainty product of h̄/2. For extensive numerical tests of Eq. (1) for the central potentials we refer the reader to the published literature [2]. Here it may be observed that the uncertainty product for the bound states in homogeneous power-law potentials, is independent of the strength of the potential. This follows from the dimensionality argument that for this case h̄ is the only quantity which has the dimension of xp. Recently, an extension of this effect for the eigendensities of the homogeneous potentials has be proposed [3]. An interesting case would be the uncertainty relation for the bound states for a superposition of two power-law potentials, and its dependence on the strengths of the two terms. We consider two special cases of superpositions of potentials, the isotropic harmonic oscillator (h.o.) potential with an additional a/r 2 term, 1 a V1 (r) = kr 2 + 2 2 r (2) and the Coulomb potential with an additional a/r 2 term, Z a + 2. (3) r r We note here that V1 (r) represents the modified oscillator potential proposed by Davidson [4] which has been found useful in analyzing [5] the roto-vibrational states of diatomic molecules. A five-dimensional Davidson model potential has been employed to study the nuclear rotations [6] and vibrations. FurV2 (r) = − 110 ther, a generalization of the free isotropic harmonic oscillator potential given by Br 2 + A/r 2 , with A and B constants, expressed as the Gol’dman and Krivchenkov Hamiltonian [7] has been recently used to represent the unperturbed part of a class of anharmonic singular potentials [8]. Similarly, the potential V2 (r) can be identified with the Kratzer potential [9,10] which has been successfully applied in the analysis of diatomic molecular spectra [11]. Recently there has been renewed interest [12] in finding accurate solutions of Kratzer potential for all angular momentum states. We begin with a dimensional analysis for σr and σp where 2 σr2 = r − r , 2 , σp2 = p − p (4) which defines the (square of ) Heisenberg uncertainty product, HP, as σr2 σp2 . The purpose of this Letter is to show that the HP corresponding to the potentials V1 (r) and V2 (r) is independent of the parameters k and Z, respectively. Equivalently, HP depends only on the parameter a of the potentials given by Eqs. (2)–(3). We deduce explicit expressions for these uncertainties using Feynman–Hellmann theorem and other relations. These expressions lead to the uncertainty relations which confirm the prediction of the dimensional analysis. A representative set of numerical calculations are presented to verify these results. Further, the arguments based on the inter-dimensional degeneracy of the Schrödinger equation has been used to discuss the HP in two-dimensional forms of the potentials as well. We also consider the well-known entropic formulation of the uncertainty relationship in terms of sum of the Shannon information entropy [13] of the probability density in the position and momentum spaces given by the lower bound due to Bialynicki-Birula and Mycielski [14], and test its validity and scaling behavior in case of V1 (r) and V2 (r) for the arbitrary values of the parameters Z, k and a. Finally, the application of our results in obtaining the Fisher information measure [15] corresponding to a single particle in a central potential is presented. 2. Some general properties Here some general properties of the uncertainties will be considered by analyzing the dimensional properties. For the modified isotropic h.o. potential in Eq. (2), the Schrödinger equation is − h̄2 2 1 a ∇ ψ + kr 2 ψ + 2 ψ = Eψ. 2M 2 r (5) Here, the basic parameters are h̄, M, k and a. Of these Ma/h̄2 is the only dimensionless quantity, so that 1 σr2 = f1 Ma/h̄2 , α 2 σp = α h̄2 f2 Ma/h̄2 , 1/2 α = Mk/h̄2 , (6) where 1/α has the dimension of (length)2 . This implies that the uncertainty product σr2 σp2 = h̄2 f1 Ma/h̄2 f2 Ma/h̄2 (7) depends on a but is independent of k. It is also interesting to note that the bound state energies are of the form E = h̄(k/M)1/2 f3 Ma/h̄2 . (8) For the modified Coulomb potential in Eq. (3), the Schrödinger equation is − h̄2 2 Z a ∇ ψ − ψ + 2 ψ = Eψ. 2M r r (9) Here also Ma/h̄2 is the only dimensionless quantity, so that 2 σr2 = h̄2 /MZ g1 Ma/h̄2 , 2 σp2 = h̄2 MZ/h̄2 g2 Ma/h̄2 , (10) where h̄2 /MZ is the Bohr radius. This implies that the uncertainty product σr2 σp2 = h̄2 g1 Ma/h̄2 g2 Ma/h̄2 (11) depends on a but is independent of Z. The bound state energies are of the form E = MZ 2 /h̄2 g3 Ma/h̄2 . (12) These forms are implied by just the simple dimensional properties of the parameters. Clearly, the product of the uncertainties is also of the same form as in Eqs. (7) and (11) for other power potentials with an additional a/r 2 term. 3. Uncertainty relation for the modified isotropic h.o. potential The radial equation for the modified isotropic h.o. potential in Eq. (2) can be written as h̄2 d 2 l(l + 1) 1 a − − ηl + kr 2 ηl + 2 ηl = Eηl , 2M dr 2 2 r2 r ηl = rRl (r). (13) Now the a/r 2 term can be combined with the angular momentum term leading to l (l + 1) 1 h̄2 d 2 − ηl + kr 2 ηl = Eηl , − 2M dr 2 2 r2 2Ma ⇒ l (l + 1) = l(l + 1) + 2 h̄ 1 1 2 2Ma 1/2 l =− + l+ (14) + 2 . 2 2 h̄ This equation is of the same form as the usual s.h.o. equation except that l is replaced by l . The solutions therefore are of the same form, E = (k/M)1/2 h̄(2n + l + 3/2), 1/2 , l + 1/2 = (l + 1/2)2 + 2Ma/h̄2 1 2 Rl = Ar l F −n, l + 3/2, αr 2 e− 2 αr , 1/2 , α = Mk/h̄2 (15) 111 where n is the radial quantum number, and F is the confluent hypergeometric function. Now, since the solutions have welldefined parity, expectation values r and p are zero, so that σp2 = p 2 . σr2 = r 2 , (16) These expectation values can be obtained in a very simple form by using the Feynman–Hellmann theorem, ∂H /∂λ = ∂E/∂λ, (17) of Eq. (20) in that the HP is independent of the parameter k but depends on a. It is interesting to note that the uncertainty product increases with n and a. In particular, for a → ∞, carrying out an expansion in inverse powers of a leads to 1/4 , σr σp → h̄(2n + 1)1/2 2Ma/h̄2 for a → ∞. (21) 4. Uncertainty relation for the modified Coulomb potential where λ is a parameter in the Hamiltonian. Taking k and M as the parameters and using Eqs. (5) and (15), we get 2 r = h̄(Mk)−1/2 (2n + l + 3/2), (18) 2 2 2Ma/h̄ . p = h̄(Mk)1/2 (2n + l + 3/2) − (19) l + 1/2 The radial equation for the modified Coulomb potential in Eq. (3) reduces to This leads to Combining the a/r 2 term with the angular momentum term, one gets σr σp = h̄(2n + l + 3/2) 1 − 1/2 2Ma/h̄2 (l + 1/2)(2n + l + 3/2) 1/2 l + 1/2 = (l + 1/2)2 + 2Ma/h̄2 . , (20) This is of the form in Eq. (7) implied by the dimensional properties, independent of parameter k but dependent on a. In Table 1 we have presented the results of our numerical calculations for the HP using the solution of the Schrödinger equation with the potential V1 (r) and the method described in our earlier work [16]. A representative set of our calculations for the three lowest states with l = 0, 1 shown in Table 1 confirm the predictions Table 1 Numerical results of energy (E), expectation values r 2 , p2 and the product r 2 p 2 corresponding to the potential V1 (r) = kr 2 /2 + a/r 2 with different values of k and a. The first three levels with the angular quantum number 0 and 1 are given. At a fixed value of a, the product is independent of k. The entries in italics display the dependence on the value of a. All values are in a.u. n+1 k/2 a E r 2 p 2 1 1 1 1 2 2 2 2 3 3 3 3 1 1 1 1 2 2 2 2 3 3 3 3 0 0 0 0 0 0 0 0 0 0 0 0 1 1 1 1 1 1 1 1 1 1 1 1 0.5 0.7 1 0.5 0.5 0.7 1 0.5 0.5 0.7 1 0.5 0.5 0.7 1 0.5 0.5 0.7 1 0.5 0.5 0.7 1 0.5 0.5 0.5 0.5 1.5 0.5 0.5 0.5 1.5 0.5 0.5 0.5 1.5 0.5 0.5 0.5 1.5 0.5 0.5 0.5 1.5 0.5 0.5 0.5 1.5 2.11803 2.50609 2.99535 2.80278 4.11803 4.87252 5.82378 4.80278 6.11803 7.23896 8.65221 6.80278 2.80278 3.31629 3.96372 3.29129 4.80278 5.68272 6.79215 5.29129 6.80278 8.04915 9.62058 7.29129 2.11803 1.79007 1.49768 2.80278 4.11803 3.48037 2.91189 4.80278 6.11803 5.17068 4.3261 6.80278 2.80278 2.36878 1.98186 3.29129 4.80278 4.05909 3.39608 5.29129 6.80278 5.74939 4.81029 7.29129 1.22361 1.44779 1.73044 1.13868 3.22361 3.81422 4.55887 3.13868 5.22361 6.18065 7.3873 5.13868 2.24808 2.65996 3.17926 1.98198 4.24808 5.02639 6.00769 3.98198 6.24808 7.39282 8.83611 5.98198 HP 2.59164 2.59164 2.59164 3.19145 13.27492 13.27492 13.27492 15.07435 31.9582 31.9582 31.9582 34.95725 6.30085 6.30085 6.30085 6.52327 20.40255 20.40255 20.40255 21.06981 42.50426 42.50426 42.50426 43.61634 l(l + 1) Z a h̄2 d 2 − ηl − ηl + 2 ηl = Eηl , − 2M dr 2 r r2 r ηl = rRl (r). h̄2 d 2 l (l + 1) Z − ηl − ηl = Eηl , 2M dr 2 r r2 2 1/2 1 2Ma 1 + 2 . l = − + l + 2 2 h̄ (22) − (23) This equation is of the same form as the usual equation for the Coulombic potential except that l is replaced by l . The solutions for this case are MZ 2 1 E=− , 2 + 1)2 (n + l 2h̄ 1/2 , l + 1/2 = (l + 1/2)2 + 2Ma/h̄2 Rl = Ar l F (−n, 2l + 2, 2αr)e−αr , 1/2 , α = −2ME/h̄2 (24) with n being the radial quantum number. In these cases, mean squared deviations are σr2 = r 2 , σp2 = p 2 . (25) Now the expectation value of r 2 can be obtained by using integrals involving the confluent hypergeometric functions [17], or by using Pasternak–Kramers recursion relations [18], leading to 2 2 1 2 2 h̄ 2 N 5N − 3l (l + 1) + 1 , r = MZ 2 N = n + l + 1. (26) Note that l and N are generally not integers. The expectation value p2 can be obtained by using the Feynman–Hellmann theorem in Eq. (17). For the Hamiltonian in Eq. (9) and the energy in Eq. (24), with M as the parameter one obtains p2 = MZ h̄2 2 h̄2 1 2Ma/h̄2 . 1 − (l + 1/2)N N2 (27) 112 Table 2 Numerical results of energy (E), expectation values r 2 , p2 and the product r 2 p 2 corresponding to the potential V2 (r) = −Z/r + a/r 2 with different values of Z and a. The first three levels with the angular quantum number 0 and 1 are given. At a fixed value of a, the product is independent of Z. The entries in italics display the dependence on the value of a. All values are in a.u. −Z a E r 2 p 2 HP 1 1 1 2 2 2 3 3 3 1 1 1 2 2 2 3 3 3 0 0 0 0 0 0 0 0 0 1 1 1 1 1 1 1 1 1 −1 −2 −1 −1 −2 −1 −1 −2 −1 −1 −2 −2 −1 −2 −2 −1 −2 −2 0.5 0.5 1.5 0.5 0.5 1.5 0.5 0.5 1.5 0.5 0.5 1.5 0.5 0.5 1.5 0.5 0.5 1.5 −0.19098 −0.76393 −0.09429 −0.07295 −0.2918 −0.04584 −0.0382 −0.15279 −0.02701 −0.09429 −0.37716 −0.2567 −0.04584 −0.18335 −0.13914 −0.02701 −0.10803 −0.08712 14.51722 3.62931 49.08747 110.5927 27.64817 253.8457 415.2912 103.8228 782.8537 49.08747 12.27187 24.30535 253.8457 63.46142 103.9757 782.8537 195.7134 289.2001 0.17082 0.68328 0.0523 0.09605 0.38421 0.04548 0.05751 0.23003 0.03312 0.14315 0.57262 0.27258 0.07628 0.30511 0.18218 0.04705 0.1882 0.12663 2.47984 2.47984 2.56741 10.62279 10.62279 11.54584 23.88247 23.88247 25.93107 7.0271 7.0271 6.62507 19.36249 19.36249 18.94218 36.83349 36.83349 36.62081 These expressions together lead to 1 2 2 2Ma/h̄2 = h̄ 5N − 3l (l + 1) + 1 1 − , 2 (l + 1/2)N 1/2 , l + 1/2 = (l + 1/2)2 + 2Ma/h̄2 σr2 σp2 (28) This again is independent of the strength Z of the Coulombic potential, but depends on Ma/h̄2 , consistent with the result in Eq. (11) implied by the dimensional properties. In Table 2 we have listed the results of our numerical calculations for the HP using the potential V2 (r). Our results for the three lowest states with l = 0, 1 given in Table 2 present the numerical evidence in support of the predictions of Eq. (28), i.e. the HP is independent of the parameter Z but depends on a. In particular, for a → ∞, one can carry out an expansion in inverse powers of a to get 1/4 , for a → ∞. σr σp → h̄(n + 1/2)1/2 2Ma/h̄2 (29) It is interesting to observe that this is similar to the corresponding expression for the modified isotropic h.o. potential, in Eq. (21) except for the factor of 2. 5. Extension to the potentials in 2D The results we have obtained can be extended to the potentials in 2D by using inter-dimensional degeneracy [19]. The radial equation for the Schrödinger equation in 2D can be written in the form m2 − 1/4 1/2 (2) h̄2 d 2 (2) − r Rm + V (r)r 1/2 Rm − 2M dr 2 r2 (30) where m is the angular momentum quantum number in 2D. Comparing it with the corresponding equation in 3D, (31) one obtains the relations (3) (2) = Er 1/2 Rm , (3) = ErRl , (2) = r 1/2 Rm−1/2 , Rm n+1 N = n + l + 1. (l + 1/2)2 − 1/4 h̄2 d 2 − rRl(3) + V (r)rRl(3) − 2M dr 2 r2 (3) (2) Em = Em−1/2 (32) for the wavefunctions and the energies in 2D and 3D. Therefore all the energies and expectation values in 2D can be obtained from the 3D results by replacing l with m − 1/2. In particular, for the modified 2D isotropic h.o. potential in Eq. (2) one gets (2) = (k/M)1/2 h̄(2n + m + 1), Em 1/2 m = m2 + 2Ma/h̄2 , 1/2 2Ma/h̄2 , σr σp = h̄(2n + m + 1) 1 − m (2n + m + 1) (33) where n is the radial quantum number. For the modified 2D Coulomb potential in Eq. (3), one obtains MZ 2 1 (2) , Em = − 2 (n + m + 1/2)2 2h̄ 1/2 , m = m2 + 2Ma/h̄2 1 2Ma/h̄2 σr2 σp2 = h̄2 5N 2 − 3 m 2 − 1/4 + 1 1 − , 2 m N N = n + m + 1/2. (34) 6. Heisenberg product, Shannon entropy, Fisher information In this section we discuss the relevance of the results obtained in Eqs. (18)–(20) and (26)–(28) to the other popular information theoretical measures, namely, the Shannon information entropy [13] and the Fisher information [15]. The Shannon information entropy of the electron density ρ(r) in coordinate space is defined as Sr = − ρ(r) ln ρ(r) dr (35) and the corresponding momentum space entropy Sp is given by Sp = − ρ(p) ln ρ(p) dp (36) where ρ(p) denotes the momentum density. The densities ρ(r) and ρ(p) are each normalized to unity and all quantities are given in atomic units. The Shannon entropy sum ST = Sr + Sp contains the net information and obeys the well-known lower bound derived by Bialynicki-Birula and Mycielski BBM [14]. This lower bound provides a stronger version of the uncertainty measure than the Heisenberg uncertainty product. According to it, the entropy sum in D dimensions satisfies the inequality [14,20–22] ST = Sr + Sp D(1 + ln π). (37) The lower bound in Eq. (37) is saturated by a Gaussian distribution. We have generated the momentum space wave function 113 through Fourier transformation of the position space wave function [16] and the corresponding densities were used to compute the Shannon entropies. The relevance of Eqs. (18)–(20) and (26)–(28) on the above information measures will now be considered with the example of the potential V1 (r) in an equivalent formulation as follows. This can be similarly extended to V2 (r). We begin by defining = [h̄2 /(Mk)]1/4 . Since has dimensions of length, a dimensionless coordinate x can be defined via r = x. As already noted, β = Ma/h̄2 is a pure (i.e., dimensionless) number. Expressing the Laplacian in terms of x as ∇ 2 = ∇˜ 2 /2 and after dividing by h̄ω, where ω = [k/M]1/2 , the Schrödinger equation becomes 1 β 1 − ∇˜ 2 φ + x 2 φ + 2 φ = φ. 2 2 x (38) Here, = E/(h̄ω) and φ(x) depends only on the pure number β which in turn depends on a. Therefore r 2 = 2 x 2 and, similarly, p 2 is proportional to 1/2 . It follows that the uncertainty product depends on β but not on k. Since the position space wavefunction φ(x) depends on β but not on k, all the uncertainty measures that depend of the position and momentum space density will be independent of k but depend on β. In addition, all products of the expectation values such as r n r −n and p n p −n will exhibit similar dependence on the parameters of V1 (r). Analogously, for V2 (r) the information measures considered above can be shown to be independent of Z but depend on a. We shall now present the numerical data to support these observations. In Table 3 we have collected numerical results corresponding to Sr , Sp and ST derived from the potentials V1 (r) and V2 (r) for a set of parameters Z, k and a. The variation of ST under scaling of the parameters Z, k and a in the potentials is found to be the same as in the case of HP. The results also confirm that the BBM bound is valid for these sums of power potentials for the arbitrary values of the parameters. Very recently Dehesa and coworkers [23,24] have reported novel results connecting the HP with the Fisher information measure [15,25]. The Fisher information measure or intrinsic Table 3 Numerical results of the Shannon information entropy in position (= Sr ) and momentum space (= Sp ) and their sum (ST ) for the ground state corresponding to the Davidson potential, V1 (r) = kr 2 /2 + a/r 2 , and the Kratzer potential, V2 (r) = −Z/r +a/r 2 , with different values of (k, a) and (Z, a). The parameter C1 represents k/2 and −Z in V1 (r) and V2 (r). The entries in italics display the dependence on the value of a. All values are in a.u. Potential n+1 C1 a Sr Sp ST Davidson 1 1 1 1 1 1 1 1 0 0 0 0 0 0 0 0 0.5 0.7 1 0.5 1 1.5 2 1 0.5 0.5 0.5 0.6 0.5 0.5 0.5 0.6 3.70352 3.45116 3.18366 3.75631 6.59449 5.37809 4.51505 6.86299 2.82717 3.07952 3.34703 2.7941 −0.1191 1.09729 1.96033 −0.3911 6.53068 6.53068 6.53068 6.55041 6.47538 6.47538 6.47538 6.47189 Kratzer accuracy in position space is defined as |∇ρ(r)|2 dr Ir = ρ(r) (39) and the corresponding momentum space measure is given by |∇ρ(p)|2 dp. Ip = (40) ρ(p) The individual Fisher measures are bounded through the Cramer–Rao inequality [26] according to Ir 12 and Ip 12 . σr σp The equality corresponds to a Gaussian distribution. A sharp and strongly localized probability density gives rise to a larger value of Fisher information in the position space. For a variety of applications of the Fisher information measure we refer to the recent book [25] and for applications to the electronic structure of atoms, to the pioneering work of Dehesa et al. [23,24]. These authors have shown that for a single particle in a central potential in the quantum state defined by the set of quantum numbers (n, , m) is given by Ir = 4 p 2 − 2(2l + 1)|m| r −2 , (41) −2 2 Ip = 4 r − 2(2l + 1)|m| p . (42) Using the dimensional analysis arguments following Eq. (38), it is clear that all the four products of the radial and momentum expectation values of the form r n p n defining the product Ir Ip via Eqs. (41) and (42), follow similar dependence on k, Z, and a for the potentials given by Eqs. (2)–(3). This leads to an interesting result that for any central potential including those given by Eqs. (2)–(3), the Fisher information product is independent of the scaling parameters k and Z but depends only upon a. The parameters can vary arbitrarily. Following Dehesa et al. [23], our results on r 2 p 2 , ST and Ir Ip can be extended to the D-dimensional hyperspherical systems. It is to be noted that a universal lower bound on the Fisher product for an arbitrary potential is not as yet available. Very recently [24] the lower bounds for the single particle in any central potential has been obtained as 2 2 r p (l + D/2)2 . (43) We have numerically tested the above bound for a very large number of states corresponding to the potentials given by Eqs. (2)–(3) with several (l) values and found Eq. (43) to be valid for the arbitrary values of the scaling parameter k, Z, and a. 7. Summary We have analyzed the uncertainties in the vectorial position r and momentum p for the isotropic h.o. potential and the Coulomb potential, with an additional a/r 2 term which are identified with the empirical potentials generally known as the Davidson and Kratzer potentials, respectively. The dimensional analysis indicates that the product of the associated uncertainties is independent of the strength of isotropic h.o. and Coulomb potentials, but depends on the dimensionless parameter Ma/h̄2 . 114 Explicit analytical expressions have been obtained for the uncertainties and their products, which confirm the predictions of the dimensional analysis. The results obtained on the variances based Heisenberg uncertainty have been shown to hold good for (a) the sum of the Shannon entropy, and (b) the product of the Fisher information measure in the position and momentum spaces. The results obtained here can be generalized to D dimensions (D 2). We are currently carrying out numerical tests of the recently proposed entropic formulation [27] based on the Rényi entropies for the potentials considered in this work. Acknowledgements S.H.P. acknowledges support from AICTE, New Delhi for emeritus fellowship. K.D.S. thanks Professor Jacob Katriel for a comment leading to Eq. (38), and gratefully acknowledges the constant encouragement received from Professor H.E. Montgomery Jr. References [1] W. Heisenberg, Z. Phys. 43 (1927) 172; E.H. Kennard, Z. Phys. 44 (1927) 326. [2] B. Tsapline, Chem. Phys. Lett. 6 (1970) 596; V. Majernik, L. Richterek, J. Phys. A: Math. Gen. 30 (1997) L49; M.E. Grypeoos, C.G. Koutroulos, K.J. Oyewumi, Th. Petridou, J. Phys. A: Math. Gen. 37 (2004) 7895; C. Kuo, Ann. Phys. 316 (2005) 431. [3] K.D. Sen, J. Katriel, J. Chem. Phys. 125 (2006) 074117. [4] P.M. Davidson, Proc. R. Soc. London A 135 (1932) 459. [5] D.J. Rowe, C. Bahri, J. Phys. A 38 (2005) 10181. [6] D.J. Rowe, C. Bahri, J. Phys. A 31 (1998) 4947. [7] I.I. Gol’dman, D.V. Krivchenkov, Problems in Quantum Mechanics, Pergamon, London, 1961. [8] N. Saad, R.L. Hall, Q.D. Katatbeh, J. Math. Phys. 46 (2005) 022104. [9] A. Kratzer, Z. Phys. 3 (1920) 289. [10] R.L. Hall, N. Saad, J. Chem. Phys. 109 (1998) 2983. [11] G. Van Hooydonk, Eur. J. Inorg. Chem. 10 (1999) 1617. [12] C. Berkdemir, A. Berkdemir, J. Han, Chem. Phys. Lett. 417 (2006) 326; O. Moustafa, S.H. Mazharimousavi, J. Phys. A 39 (2006) 10537; J. Vigo-Aguiar, T.E. Simos, Int. J. Quantum Chem. 103 (2005) 278; O. Bayrak, O. Bostusun, C. Ciftci, Int. J. Quantum Chem. (2006), in press. [13] C.E. Shannon, Bell Syst. Tech. 27 (1948) 379; C.E. Shannon, Bell Syst. Tech. 27 (1948) 623. [14] I. Bialynicki-Birula, J. Mycielski, Commun. Math. Phys. 44 (1975) 129. [15] R.A. Fisher, Proc. Cambridge Philos. Soc. 22 (1925) 700. [16] K.D. Sen, J. Chem. Phys. 122 (2005) 1943241, and references therein. [17] L.D. Landau, E.M. Lifshitz, Quantum Mechanics, Pergamon, Oxford, 1977. [18] S. Pasternak, Proc. Natl. Acad. Soc. 23 (1937) 91; H.A. Kramers, Quantum Mechanics, North-Holland, Amsterdam, 1957. [19] D.D. Frantz, D.R. Herschbach, J. Chem. Phys. 92 (1990) 6668. [20] S.B. Sears, Applications of information theory in chemical physics, PhD thesis, University of North Carolina at Chapel Hill, 1980. [21] S.R. Gadre, S.B. Sears, S.J. Chakravorty, R.D. Bendale, Phys. Rev. A 32 (1985) 2602; S.R. Gadre, Phys. Rev. A 30 (1984) 620; S.R. Gadre, R.D. Bendale, Int. J. Quantum Chem. 28 (1985) 311; S.R. Gadre, S.A. Kulkarni, I.H. Shrivastava, Chem. Phys. Lett. 16 (1990) 445; S.R. Gadre, R.D. Bendale, S.P. Gejji, Chem. Phys. Lett. 117 (1985) 138; S.R. Gadre, R.D. Bendale, Curr. Sci. (India) 54 (1985) 970; S.R. Gadre, in: K.D. Sen (Ed.), Reviews of Modern Quantum Chemistry, World Scientific, Singapore, 2002, p. 108. [22] A.N. Tripathi, V.H. Smith Jr., R.P. Sagar, R.O. Esquivel, Phys. Rev. A 54 (1996) 1877; M. Ho, D.F. Weaver, V.H. Smith Jr., R.P. Sagar, R.O. Esquivel, Phys. Rev. A 57 (1998) 4512; M. Ho, V.H. Smith Jr., D.F. Weaver, C. Gatti, R.P. Sagar, R.O. Esquivel, J. Chem. Phys. 108 (1998) 5469; J.C. Ramirez, J.M.H. Perez, R.P. Sagar, R.O. Esquivel, M. Ho, V.H. Smith Jr., Phys. Rev. A 58 (1998) 3507; M. Ho, D.F. Weaver, V.H. Smith Jr., R.P. Sagar, R.O. Esquivel, S. Yamamoto, J. Chem. Phys. 109 (1998) 10620; R.P. Sagar, J.C. Ramirez, R.O. Esquivel, M. Ho, V.H. Smith Jr., Phys. Rev. A 63 (2001) 022509; N.L. Guevara, R.P. Sagar, R.O. Esquivel, J. Chem. Phys. 119 (2003) 7030; N.L. Guevara, R.P. Sagar, R.O. Esquivel, J. Chem. Phys. 122 (2005) 084101; Q. Shi, S. Kais, J. Chem. Phys. 121 (2004) 5611; Q. Shi, S. Kais, J. Chem. Phys. 309 (2005) 127; K.Ch. Chatzisavvas, Ch.C. Moustakidis, C.P. Panos, J. Chem. Phys. 123 (2005) 174111. [23] E. Romera, P. Sanchez-Moreno, J.S. Dehesa, Chem. Phys. Lett. 414 (2005) 468; J.S. Dehesa, S. Lopez-Rosa, B. Olmos, R.J. Yanez, J. Math. Phys. 47 (2006) 052104. [24] E. Romera, P. Sanchez-Moreno, J.S. Dehesa, J. Math. Phys., in press; P. Sanchez-Moreno, R. Gonzales-Ferez, J.S. Dehesa, New J. Phys., submitted for publication. We thank the authors for sharing this work in advance of publication. [25] B.R. Frieden, Science from Fisher Information, Cambridge Univ. Press, Cambridge, 2004. [26] C.R. Rao, Linear Statistical Interference and its Applications, Wiley, New York, 1965; A. Stam, Inf. Control 2 (1959) 101. [27] I. Bialynicki-Birula, Phys. Rev. A 74 (2006) 052101.
© Copyright 2025 Paperzz