The Laguerre functions constitute one of the fundamental basis sets for calculations in atomic and molecular electron-structure theory, with applications in hadronic and nuclear theory as well. While similar in form to the Coulomb bound-state eigenfunctions (from the Schrödinger eigenproblem) or the Coulomb-Sturmian functions (from a related Sturm-Liouville problem), the Laguerre functions, unlike these former functions, constitute a complete, discrete, orthonormal set for square-integrable functions in three dimensions. We construct the SU(1, 1) × SO(3) dynamical algebra for the Laguerre functions and apply the ideas of factorization (or supersymmetric quantum mechanics) to derive shift operators for these functions. We use the resulting algebraic framework to derive analytic expressions for matrix elements of several basic radial operators (involving powers of the radial coordinate and radial derivative) in the Laguerre function basis. We illustrate how matrix elements for more general spherical tensor operators in three dimensional space, such as the gradient, may then be constructed from these radial matrix elements.

The Laguerre functions1–3 constitute one of the fundamental basis sets for calculations in atomic and molecular electron-structure theory,1,4–6 with application in hadronic7–12 and nuclear structure13–15 theory as well. The Laguerre functions share the same functional form as the Coulomb bound-state wave functions (obtained as solutions to the Schrödinger equation central force eigenproblem) or the Coulomb-Sturmian functions (obtained as solutions to a related Sturm-Liouville problem). However, unlike these functions, the Laguerre functions constitute a complete, discrete, orthonormal set of square-integrable functions in three dimensions.

Numerical solution of quantum mechanical problems—for the present discussion, we can take this to mean one-body and many-body Schrödinger equation problems, although the applicability is more general—may be represented in terms of an expansion of the wave function on a set of basis functions. The Schrödinger equation eigenproblem, when expanded in a discrete basis, is recast as a Hamiltonian matrix eigenproblem.16 To proceed with this approach, it is necessary to be able to evaluate the matrix elements of the operators of the problem, taken with respect to the expansion basis. Relevant operators may include those contributing to the Hamiltonian, transition operators, and operators representing moments or other static observables.

For numerical solution of many-body problems, the many-body basis functions are obtained as products of single-particle basis functions (more precisely, either symmetrized or antisymmetrized products, according to the statistics of the problem17). When the many-body problem is rotationally invariant, it is natural to consider an expansion in terms of single-particle basis functions which are themselves obtained as solutions to a rotationally invariant central force problem, for instance, the three-dimensional harmonic oscillator functions.18 The solution sets to central force problems are well known to factorize into radial and angular functions, arising from the factorizability of the Hilbert space of square-integrable functions on ℝ3 as ℒ2(ℝ3) = ℒ2(ℝ+) × ℒ2(S2). The angular functions, on the sphere S2, are the well-known spherical harmonics.19 For each choice of angular momentum l, the associated radial functions form a complete set of square-integrable functions on ℝ+. Matrix elements of spherical tensor operators (including rotationally invariant scalar operators and vector operators) may be simplified through a corresponding factorization into radial and angular dependences, together with the Wigner-Eckart theorem19 for the angular dependence.

Most basic among the central force problems are those for the harmonic oscillator potential and the Coulomb (hydrogenic) potential. Both these problems have a rich group theoretical structure (reviewed in Ref. 20), which allows analytic expressions to be obtained for matrix elements of many operators of interest, built from the coordinate vector r and momentum vector p = − .

Underlying the convenience of these bases is the presence of an SU(1, 1) × SO(3) dynamical group,20 which reflects the factorization of the problem into radial and angular coordinates—with SU(1, 1) operating on the radial functions and SO(3) on the angular functions. The SU(1, 1) spectrum generating algebra relates radial wave functions with different radial quantum number (node number) n, at the same angular momentum l. This SU(1, 1) algebra has been widely studied for both the oscillator21–23 and Coulomb21,22,24–27 problems. The SO(3) algebra, in turn, is the standard angular momentum algebra, which relates angular wave functions (spherical harmonics) with different angular momentum projection quantum number m, at given l.19 

While the SU(1, 1) × SO(3) algebra thus provides the framework for relating states of the same angular momentum (and, in fact, for calculating matrix elements among these states, at least for certain suitable operators), an additional structure is required to connect basis states of different angular momentum. This connection is provided either by the factorization method (first applied by Schrödinger for the Coulomb problem28 and subsequently generalized by Infeld and Hull29) or, equivalently, through the ideas of supersymmetric quantum mechanics (SUSY QM).30 

More concretely, the relations among basis states provided by the algebraic structures are indicated in Fig. 1. To each angular momentum l is associated a set of radial wave functions. These constitute a positive-discrete, infinite-dimensional irreducible representation (irrep) of SU(1, 1). Radial wave functions differing in radial quantum number n by ±1, at fixed l, are related by the SU(1, 1) ladder operatorsT±. These change the eigenvalue of the weight operatorT0, i.e., the weight mt = n + l + 3/2, by ±1. Each SU(1, 1) irrep is labeled by its lowest weight t = l + 3/2. Thus, to each irrep (with label l) of the angular algebra SO(3) there is uniquely associated a dual irrep (with label t = l + 3/2) of the SU(1, 1) radial algebra.31 Then, shift operators, A and A,26 obtained by the factorization or SUSY QM approaches alluded to above, relate radial wave functions corresponding to different angular momenta. Specifically, the shift operators connect radial wave functions of the same mt but differing in t by ±1, i.e., radial wave functions associated with angular functions of angular momentum l differing by ±1. The algebraic structure (ladder operators) and shift operators in combination are what enable the evaluation of matrix elements of operators in the full three-dimensional space.20 

FIG. 1.

Algebraic classification of basis functions, with respect to the SU(1, 1)  ×  SO(3) dynamical algebra, and the pattern of laddering operators T± and shift operators, A and A, connecting these functions. The quantum number l labels the SO(3) irrep, and the dual quantum number t labels the SU(1, 1) irrep by its lowest weight, where mt is the SU(1, 1) weight (see text).

FIG. 1.

Algebraic classification of basis functions, with respect to the SU(1, 1)  ×  SO(3) dynamical algebra, and the pattern of laddering operators T± and shift operators, A and A, connecting these functions. The quantum number l labels the SO(3) irrep, and the dual quantum number t labels the SU(1, 1) irrep by its lowest weight, where mt is the SU(1, 1) weight (see text).

Close modal

Let us now re-examine the choice of basis functions. Since we shall restrict attention to functions with radial-angular factorization, these would also commonly be termed orbitals. In actual calculations, the basis must be truncated to finite size (either the set of single-particle basis functions may be truncated directly, or some more general limitation may be placed on the set of many-body antisymmetrized product states constructed from this basis32). Therefore, we must be concerned not only with mathematical completeness of the basis but also with how well matched the basis is to the physical problem at hand. That is, the basis functions should be chosen according to the following criteria (see Ref. 6): (i) The functions should allow for a systematic approach to completeness as the single-particle basis set is expanded. (ii) The functions should provide for rapid convergence of the calculated energies or eigenfunctions to the true results, requiring only a small number of basis functions for an accurate description. (iii) It is also practically desirable that the functions have a simple analytic form which permits convenient evaluation of relevant operator matrix elements. (Ideally, the functions should also be orthogonal and easily normalized, to simplify the formulation of the many-body problem, but this is not essential.)

The harmonic oscillator eigenfunctions are of the form (for simplicity, we omit normalization factors and set the radial length parameter to unity)3,18

(1)

where Lnα(x) is the associated Laguerre polynomial of degree n and order α, i.e., for parameter α in the associated Laguerre equation,33 and Ylm(rˆ) is a spherical harmonic. These functions constitute a complete, orthonormal set on ℒ2(ℝ3). However, while the Schrödinger equation solutions for a finite-range potential (or for an infinite-range potential of Coulomb 1/r type) exhibit an exponential asymptotic decay at large r (∝eβr), the oscillator functions have Gaussian asymptotics (∝eαr2), as an artifact of the infinite (quadratic) binding potential of the harmonic oscillator problem. The oscillator functions are thus poorly matched to the physical solutions in the asymptotic region and may therefore violate criterion (ii) above for a suitable basis, i.e., rapid convergence. In nuclear physics, for instance, while structure involving only deeply bound orbitals might be adequately represented using basis functions with Gaussian asymptotics, weakly bound orbitals with distinctly exponential asymptotics influence many aspects of nuclear structure, especially in light or neutron-rich nuclei (see, e.g., Fig. 4 of Ref. 34 and Fig. 4 of Ref. 35).

The Coulomb wave functions, as solutions to a problem with finite binding, display the appropriate exponential asymptotics. The bound-state (or negative-energy) Coulomb functions3,6,33 are of the form36 

(2)

The full set of solutions to the Schrödinger equation with the Coulomb potential, including both these bound-state solutions and the continuum (or positive-energy) solutions, constitutes a complete set on ℒ2(ℝ3). However, the bound-state solutions by themselves do not. This is not merely a formal deficiency: attempting to use them as an expansion basis is well-known to result in convergence to erroneous energies and wave functions.1 It is therefore necessary to supplement the bound-state solutions with the continuum functions. The inconvenience of including continuum functions limits the practical applicability of the Coulomb basis: both in terms of arranging a systematic approach to completeness with increasing basis size [criterion (i)] and due to the complicated nature of the continuum wave functions [criterion (iii)].3,6

One may instead seek a discrete complete set, retaining the same exponential asymptotics as the Coulomb wave functions, by modifying the Schrödinger equation into a Sturm-Liouville problem associated with the Coulomb potential (as detailed in Sec. II), thereby obtaining the Coulomb-Sturmian functions.3,37–40 These have been applied as expansion bases in atomic and molecular theory,37,38,41–43 nuclear scattering,44,45 and other few-body problems.46 The Coulomb-Sturmian functions are of the form

(3)

that is, of the same functional form as Coulomb eigenproblem solutions (2) but independently rescaled in the radial coordinate so that r/(n + l + 1) → r. These radial functions constitute a discrete set, complete on the Sobolev space W2(1)(R3) [i.e., the space of all functions in ℒ2(ℝ3), the generalized second derivatives of which are also in ℒ2(ℝ3)],3 which implies completeness on ℒ2(ℝ3).47 However, these functions are orthogonal with respect to the Coulomb potential 1/r as weighting function, rather than the usual Euclidean metric.

We therefore finally turn to the Laguerre functions48,49 which are of the form

(4)

They are identical in form to Coulomb-Sturmian functions (3), differing only in the parameter α of the associated Laguerre polynomial: odd α = 2l + 1 for the Coulomb-Sturmian functions vs. even α = 2l + 2 for the Laguerre functions.50 The Laguerre functions satisfy all the posited criteria: constituting a discrete set, complete on ℒ2(ℝ3), and orthogonal with respect to the usual Euclidean metric, but with the physically appropriate exponential asymptotics.

We demonstrate that an algebraic framework, analogous to that obtained for the usual central force eigenfunctions, can likewise be constructed for the Laguerre functions. These results permit analytic expressions to be obtained for matrix elements. The Laguerre functions are first introduced (Sec. II). After a brief review of the SO(3) and SU(1, 1) Lie algebras and the structure of their irreps (Sec. III A), we develop the SU(1, 1) algebra of the Laguerre radial functions (Sec. III B), and we use this algebra to obtain analytic expressions for the action of several basic operators (involving powers of the radial coordinate r and radial derivative operator d/dr) on the basis of Laguerre radial functions—or, equivalently, analytic expressions for matrix elements of these operators (Sec. III C). The results thus obtained apply to functions within a single SU(1, 1) irrep, i.e., to the set of radial functions associated with the same angular momentum. We then extend the ideas of factorization (or SUSY QM) to a quantum number dependent formulation (Sec. IV A), to construct shift operators relating Laguerre radial functions in different irreps (Sec. IV B), and thus to obtain analytic expressions for the matrix elements of radial operators between different irreps (Sec. IV C). Finally, returning to the full Laguerre functions in three dimensions, we demonstrate how the results for radial matrix elements and shift operators are combined to yield reduced matrix elements of spherical tensor operators, taking the Laplacian and gradient operators for illustration (Sec. V).

Coulomb functions (2) are the solutions to the Schrödinger equation

(5)

As outlined above, these do not form a complete set on ℒ2(ℝ3) without including the continuum,1,41 making them an impractical choice of basis.

There is, however, a set of functions, closely related to the Coulomb functions, called Coulomb-Sturmian functions (3), which combine the discrete character of the harmonic oscillator functions with the desired large r behavior. These functions are related to the Coulomb functions by making the substitution

(6)

which recasts differential equation (5) into the form of a Sturm-Liouville equation

(7)

and yields the Coulomb-Sturmian solutions

(8)

Observe that, while in Schrödinger equation eigenproblem (5) one varies the energy Enl to satisfy the boundary conditions, in Sturm-Liouville eigenproblem (7) one now holds the energy constant and varies the coefficient βnl, which governs the strength or depth of the potential, to satisfy the boundary conditions.38 From the basic theory of the Sturm-Liouville equation,51 it follows that the solutions Ψnlm are orthogonal with respect to the Coulomb potential 1/r as weighting function.38 These functions Ψnlm are also complete in the Sobolev space W2(1)(R3).3 

However, for calculations in physics it is useful to work with functions that are orthonormal with respect to the usual integration metric d3r and complete on ℒ2(ℝ3). To obtain functions with this property, the integration weight and norm are absorbed into the function by multiplying Ψnlm by [b(n + l + 1)/r]1/2, and the shift ll + 1/2 is introduced.3 The resulting functions are the Laguerre functions,

(9)

which satisfy the radial differential equation

(10)

The Laguerre functions Λnlm(r) form a complete, discrete basis for L2(ℝ3),47 have the correct asymptotic behavior and are orthogonal with respect to the ℝ3 metric. That is,

(11)

It is now the matrix elements calculated in this Laguerre basis that are of interest. Certain matrix elements can be calculated via direct integration.8,10 However, in this paper we use the factorability of the Hilbert space to develop a simpler and more elegant algebraic method for calculating matrix elements.

The Laguerre function Λnlm(r) can be factorized into radial and angular functions as

(12)

The integration weight has been absorbed into the radial function Snl(r), given by

(13)

which satisfies the orthogonality condition

(14)

The set of functions {Snl|n = 0, 1, 2, …} forms a complete and orthogonal set on L2(ℝ+), for any l = 0, 1, 2, …, independent of b,39 so we make the simplifying substitution b → 1 in the following discussion. The resulting functions,

(15)

satisfy the differential equation

(16)

with αnl as defined in (10). Although (16) is not explicitly written in the form of a Sturm-Liouville equation (see Ref. 51), it nonetheless may readily be put in this form by left multiplying by r. For convenience of notation, we will sometimes denote Λnlm as |nlm〉, Snl as |nl〉, and Ylm as |lm〉.

Although we want the matrix elements in the Λnlm basis, the actual calculations will be carried out more simply using the functions

(17)

since they are orthogonal with respect to the measure drdΩ rather than the measure r2drdΩ. We start with the matrix element of a component of some irreducible, rank a tensor operator

(18)

where R is a scalar radial operator and Θa is a rank a tensor with no radial dependence. Although we are only considering operators that can be factored into radial and angular parts, many operators of interest can be rewritten as sums of such separable operators using various factorization lemmas, and so this method can also be used to calculate matrix elements of these operators. As an example, matrix elements of the gradient operator ∇1 are calculated in Sec. V C.

The matrix element of the operator Oaα separates into radial and angular factors as

(19)

That is,

(20)

where

(21)

If the operator is purely radial (Θ100), then we have that

(22)

The similarity transformation given by (21) will allow us to calculate matrix elements in the Λnlm basis by factoring the matrix element into linear combinations of products of a matrix element in the Snl basis and a matrix element in the well known angular momentum basis, as shown (for the case of positive weights) in Sec. V.

Factorization (12) of the Λnlm basis for ℒ(ℝ3) has the underlying symmetry SU(1,1) × SO(3), where SU(1,1) is the symmetry of the radial functions and SO(3) is the symmetry of the Ylm angular functions. It is well known that the SO(3) symmetry can be used to calculate matrix elements of operators on the Ylm basis. In this section we discuss how SU(1,1) symmetry can be used to calculate the matrix elements on the Snl basis.

The Lie algebra SU(1,1) has a very similar structure to that of the Lie algebra SO(3), so we begin by summarizing the properties of the more familiar SO(3) algebra for comparison with the SU(1,1) algebra.

The group SO(3) is the compact Lie group of rotations in three dimensions. Its algebra is generated by the operators L3 and L±, which satisfy the commutation relations

(23)

and

(24)

where L+=L. The eigenfunctions of L3 are the spherical harmonics |lm〉, where

(25)

The laddering operators L± have the well known action on the spherical harmonics

(26)

Because SO(3) is a compact group, the weight m is bounded above and below. The bound is determined by the SO(3) Casimir operator

(27)

which acts on the eigenfunctions as

(28)

The bounds above and below on m are given by ±l, that is,

(29)

as shown in Fig. 2(a).

FIG. 2.

Weights contained within (a) SO(3) and (b) SU(1, 1) irreps of the types considered in the present work (see text). Weights (m or mt, respectively, for these two algebras) within an irrep (labeled by extremal weight l or t, respectively) are connected by solid lines. The bounding weights, as functions of l or t, are indicated by dashed lines.

FIG. 2.

Weights contained within (a) SO(3) and (b) SU(1, 1) irreps of the types considered in the present work (see text). Weights (m or mt, respectively, for these two algebras) within an irrep (labeled by extremal weight l or t, respectively) are connected by solid lines. The bounding weights, as functions of l or t, are indicated by dashed lines.

Close modal

Similarly, the generators of SU(1,1) are T3 and T±, which satisfy the commutation relations

(30)

and

(31)

where T+=T. The actions of the generators on the eigenfunctions |tmt〉 of T3 are given by

(32)

and

(33)

The commutator in (31) differs from the commutator in (24) by a negative sign. The group SU(1,1) is a non-compact group, and the weight mt is either bounded above or bounded below but takes on infinitely many values. As with SO(3), the bound on mt is determined by the Casimir operator

(34)

which acts on the eigenfunction as

(35)

If t is an integer or half-integer, then the irrep is discrete. If t > 0 then the irrep is bounded below by t, and if t < 0 then the irrep is bounded above by t. In other words,

(36)

as shown in Fig. 2(b) for t > 0.

In this section, we construct a representation of the Lie algebra of the radial group, SU(1,1), using the Snl(r) radial functions as the basis of the vector space. Recall from (16) that the Snl basis is generated by varying αnl. We rearrange (16) to rewrite it as an eigenvalue problem for αnl,

(37)

Following a similar approach to the method used in Refs. 21 and 26 to construct representations of SU(1,1) for the Coulomb functions, we define

(38)

and

(39)

which satisfy the commutation relations in (30) and (31) and the relation T+=T. The Casimir invariant is given by

(40)

Thus, the representation is discrete, with t = l + 3/2 and weights mt = n + l + 3/2 (n = 0, 1, ….). Note that mt = αnl, with αnl as defined in (10). From (32) and (33), the actions of the generators are expressed in terms of the labels n and l as

(41)
(42)
(43)

We can use the SU(1,1) algebra to easily determine the actions of r, 1/r, and d/dr on the functions Snl(r) following a method similar to that used by Rowe in Ref. 23 to calculate the matrix elements for the generalized harmonic oscillator functions. We begin with the elements of the algebra

(44)

and

(45)

Applying the operators in (44) and (45) to Snl, we immediately have that

(46)

and

(47)

We can obtain the action of the operator 1/r, which is not an element of the SU(1,1) algebra, using a recurrence relationship that we deduce using the orthogonality of the Snl(r) functions. Starting with (46), we divide through by r, multiply through by Sml(r), and integrate to obtain

(48)

where

(49)

Starting with f00 = 1/(l + 1), obtained via direct integration, recurrence relationship (48) can be solved to find that

(50)

A similar result, obtained by simply interchanging m and n, holds for n < m. From these results, we have that

(51)

Finally, the action of the radial derivative can be determined by noting that d/dr = r−1(rd/dr). Combining (47) and (51),

(52)

The actions of other operators, built from these basic operators, can be calculated by combining (46), (51), and (52).

Once we have the action of a radial operator R(r) on Snl, we can extract the matrix element 〈nl|R(r)|nl〉 by inspection as the coefficient of Snl in the sum, using the orthogonality of the radial functions Snl. For example, if n′ ≥ n + 1, then

(53)

Expressions for the matrix elements for r, r2, 1/r, 1/r2, rd/dr, d/dr, and d2/dr2 are given in the  Appendix (Table I).

TABLE I.

Matrix elements between functions in the Snl basis.

nl|r|nl〉 n′ ≥ n + 2 
 n′ = n + 1 12((n+1)(n+2l+3))1/2 
 n′ = n n+l+32 
 n′ = n − 1 12(n(n+2l+2))1/2 
 n′ ≤ n − 2 
nl|r2|nl〉 n′ ≥ n + 3 
 n′ = n + 2 14((n+1)(n+2)(n+2l+3)(n+2l+4))1/2 
 n′ = n + 1 −(n + l + 2)((n + 1)(n + 2l + 3))1/2 
 n′ = n (n+l+32)(n+l+2)+12n(n+2l+2) 
 n′ = n − 1 −(n + l + 1)(n(n + 2l + 2))1/2 
 n′ = n − 2 14(n(n1)(n+2l+2)(n+2l+1))1/2 
 n′ ≤ n − 3 
nl|1r|nl n′ ≥ n 1l+1n!(n+2l+2)!n!(n+2l+2)!1/2 
 n′ ≤ n 1l+1n!(n+2l+2)!n!(n+2l+2)!1/2 
nl|1r2|nl n′ ≥ n 2n(2l+3)2n(2l+1)+4l+6(l+1)(2l+1)(2l+3)n!(n+2l+2)!n!(n+2l+2)!1/2 
 n′ ≤ n 2n(2l+3)2n(2l+1)+4l+6(l+1)(2l+1)(2l+3)n!(n+2l+2)!n!(n+2l+2)!1/2 
nl|rddr|nl n′ ≥ n + 2 
 n′ = n + 1 12((n+1)(n+2l+3))1/2 
 n′ = n 12 
 n′ = n − 1 12(n(n+2l+2))1/2 
 n′ ≤ n − 2 
nl|ddr|nl n′ ≥ n + 1 n!(n+2l+2)!n!(n+2l+2)!1/2 
 n′ = n 
 n′ ≤ n − 1 n!(n+2l+2)!n!(n+2l+2)!1/2 
nl|d2dr2|nl n′ ≥ n + 1 (2l+4)(2l+1)n2l(2l+3)n+(2l+3)(2l+2)(2l+3)(2l+1)n!(n+2l+2)!n!(n+2l+2)!1/2 
 n′ = n 4n(l+1)+2l+3(2l+3)(2l+1) 
 n′ ≤ n − 1 (2l+4)(2l+1)n2l(2l+3)n+(2l+3)(2l+2)(2l+3)(2l+1)n!(n+2l+2)!n!(n+2l+2)!1/2 
nl|r|nl〉 n′ ≥ n + 2 
 n′ = n + 1 12((n+1)(n+2l+3))1/2 
 n′ = n n+l+32 
 n′ = n − 1 12(n(n+2l+2))1/2 
 n′ ≤ n − 2 
nl|r2|nl〉 n′ ≥ n + 3 
 n′ = n + 2 14((n+1)(n+2)(n+2l+3)(n+2l+4))1/2 
 n′ = n + 1 −(n + l + 2)((n + 1)(n + 2l + 3))1/2 
 n′ = n (n+l+32)(n+l+2)+12n(n+2l+2) 
 n′ = n − 1 −(n + l + 1)(n(n + 2l + 2))1/2 
 n′ = n − 2 14(n(n1)(n+2l+2)(n+2l+1))1/2 
 n′ ≤ n − 3 
nl|1r|nl n′ ≥ n 1l+1n!(n+2l+2)!n!(n+2l+2)!1/2 
 n′ ≤ n 1l+1n!(n+2l+2)!n!(n+2l+2)!1/2 
nl|1r2|nl n′ ≥ n 2n(2l+3)2n(2l+1)+4l+6(l+1)(2l+1)(2l+3)n!(n+2l+2)!n!(n+2l+2)!1/2 
 n′ ≤ n 2n(2l+3)2n(2l+1)+4l+6(l+1)(2l+1)(2l+3)n!(n+2l+2)!n!(n+2l+2)!1/2 
nl|rddr|nl n′ ≥ n + 2 
 n′ = n + 1 12((n+1)(n+2l+3))1/2 
 n′ = n 12 
 n′ = n − 1 12(n(n+2l+2))1/2 
 n′ ≤ n − 2 
nl|ddr|nl n′ ≥ n + 1 n!(n+2l+2)!n!(n+2l+2)!1/2 
 n′ = n 
 n′ ≤ n − 1 n!(n+2l+2)!n!(n+2l+2)!1/2 
nl|d2dr2|nl n′ ≥ n + 1 (2l+4)(2l+1)n2l(2l+3)n+(2l+3)(2l+2)(2l+3)(2l+1)n!(n+2l+2)!n!(n+2l+2)!1/2 
 n′ = n 4n(l+1)+2l+3(2l+3)(2l+1) 
 n′ ≤ n − 1 (2l+4)(2l+1)n2l(2l+3)n+(2l+3)(2l+2)(2l+3)(2l+1)n!(n+2l+2)!n!(n+2l+2)!1/2 

In Sec. III, we derived expressions for matrix elements of operators between Laguerre radial functions within the same SU(1,1) irrep, i.e., matrix elements between the three-dimensional basis functions with the same angular quantum number l. However, many operators of interest in physics are not strictly radial, scalar operators (e.g., r and are vector operators or, equivalently, spherical tensors of rank 1) and may therefore have nonvanishing matrix elements between states with different l. If we consider a spherical tensor operator Oaα which factorizes into radial and angular parts, as discussed in Sec. II A, then recall from (20) that the problem reduces to evaluating the radial matrix element 〈nl′|γ(R)|nl〉, with l′ ≠ l, that is, connecting Laguerre radial functions lying in different SU(1,1) irreps.

The matrix elements of radial operators between radial functions in different irreps (l′ ≠ l) can be written in terms of the matrix elements we have already calculated, i.e., between members of the same irrep (l′ = l), provided we can obtain analytic relations allowing us to expand the members of one of the irreps (namely, l′) as linear combinations of those in the other irrep (namely, l). This relation may be obtained through shift operators, A and A, introduced in Sec. I.

To see how we may obtain such shift operators, recall from (16) that the Laguerre functions are eigenfunctions of a differential operator. In particular, the operator

(54)

may be taken as a Hamiltonian operator defining the Laguerre functions (of a single irrep l) as its eigenfunctions, except that this Hamiltonian involves an n-dependent coefficient αnl and is thus actually a quantum number dependent Hamiltonian (the eigenfunctions of interest are also degenerate, with eigenvalue −1).

The problem of relating Laguerre functions from different irreps (of different l) thus amounts to the problem of relating the eigenfunctions of a sequence of Hamiltonians, which are defined through discrete variation of a parameter l. This is the problem addressed by the factorization method of Infeld and Hull29 or, equivalently, for present purposes, the factorization method used in SUSY QM30 for shape invariant potentials.52 

The essential idea of factorization is that if a Hamiltonian can be written in a factorized form H = AA, this immediately gives rise to a partner Hamiltonian H′ = AA, with (at least some) eigenvalues which are degenerate to those of H, and with eigenfunctions which are related to those of H—in fact, related by the action of the operators A and A. Factorization is a particularly powerful technique in problems where we have an extended (perhaps infinite) sequence or “tower” of related Hamiltonians Hm=AmAm (m = 1, 2, …), where m is a discrete parameter in the Hamiltonian. Then we likewise have extended sets of degenerate states ψnm (n = 0, 1, … and m = 1, 2, …), related to each other by the action of the operators Am and Am. SUSY QM provides a general formalism for finding such a factorization, through the superpotential.30 However, for many familiar exactly solvable Schrödinger equation problems, the factorizations have been cataloged by Infeld and Hull,29 and the problem of finding a factorization reduces to that of transforming the problem into one of their standard forms.

It may be helpful to note both the analogy to the familiar factorization of the one-dimensional harmonic oscillator Hamiltonian, H = aa + 1/2, explored by Dirac,53 as well as the differences from this factorization. In the oscillator problem, the operators a and a are simply laddering operators, between eigenfunctions of the same Hamiltonian, serving to change the number of nodes. However, the operators A and A in the more general factorization are shift operators, between eigenfunctions of different Hamiltonians, though in fact their action also changes the number of nodes.

While factorization is a powerful technique for solving eigenproblems, we are interested here primarily in the shift operation which it provides, towards our goal of relating Laguerre radial functions in different irreps. We therefore summarize the necessary ideas behind the shift operation, then show how the relevant relations can readily be extended to accommodate quantum number dependent Hamiltonians, such as encountered for the Laguerre functions in (54). We refer the reader to the more extensive discussions in Refs. 29, 30, and 54 for a more comprehensive discussion of the families of potentials which lead to factorizable Hamiltonians, the solution process, and the nature of the eigenspectra which result. Our notational conventions for factorization follow those of, e.g., Refs. 23 and 30, in the use of A and A for the shift operators.

To start with this simplest case, of two partner Hamiltonians, suppose the Hamiltonian for some one-dimensional eigenproblem can be written as

(55)

where, in addition to the factorized portion of the Hamiltonian, it is convenient to permit a constant term K1. (Since the primary concern in the present work is with radial wave functions on ℝ+, we have chosen to denote the independent variable as r rather than as x, but the discussion here is relevant without modification to one-dimensional problems on ℝ.) Then the eigenspectrum of H1 is related in a simple way to that of the partner Hamiltonian

(56)

Specifically, suppose H1 has eigenfunctions ψn1 (n = 0, 1, …), with associated eigenvalues En1, and H2 has eigenfunctions ψn2 (n = 0, 1, …), with associated eigenvalues En2, that is,

(57)
(58)

Then, simply from the form of (55) and (56), one can readily verify (as we shall illustrate below in a more general context) that acting with A1 on any eigenfunction ψn1 of H1 gives an eigenfunction of H2, still with eigenvalue En1. Similarly, acting with A1 on any eigenfunction ψn2 of H2 gives an eigenfunction of H1, still with eigenvalue En2. Thus, eigenfunctions of H1 and H2 must occur as degenerate partners. The exception arises if A1 (or A1) annihilates the wave function on which it acts, that is, if it yields the null function instead of a bona fide eigenfunction of the partner Hamiltonian. We shall restrict our attention to the typical case (see Refs. 29, 30, and 54) in which A1ψ01 = 0. That is, the ground state eigenfunction of H1 has no partner. Then the remaining eigenfunctions and eigenvalues follow the correspondence (for n = 1, 2, …)

(59)
(60)
(61)

The factorization process may, however, continue. Suppose H2, in turn, can be refactored in some way, distinct from (56), as

(62)

Then the same pattern can be followed, as above, to relate the eigenspectrum of H2 to that of a partner Hamiltonian

(63)

and relations on the eigenfunctions and eigenvalues analogous to (59)(61) can be obtained.

More generally, we may consider an infinite “tower” of related Hamiltonians Hm, parametrized by a discrete index m (m = 1, 2, …), where successive Hamiltonians Hm and Hm+1 have related factorizations

(64)
(65)

We now have a sequence of related eigenproblems

(66)

Their eigenfunctions and eigenvalues are related by

(67)
(68)
(69)

etc. The identification of n quantum numbers (n = 0, 1, …) across the different eigenproblems here again rests on the further conventional condition that the ground state for each Hm be annihilated by the A operator, i.e., Amψ0m = 0 or, equivalently, as may be verified, E0m = Km.

Let us take a moment to interpret the energy and operator relations indicated in (67)(69), which are illustrated schematically in Fig. 3. From (67), it is seen that the operator Am acts on the wave function ψn−1,m+1 to give the wave function ψnm. It thus raises the quantum number n, which counts the number of nodes in the wave function, and is therefore reminiscent of Dirac’s a operator. However, the initial and final wave functions are eigenfunctions of different Hamiltonians, as the Hamiltonian index has simultaneously been lowered. Thus, Am is not merely a raising operator, but rather a shift operator. Similarly, from (68), it is seen that Am is likewise a shift operator, but with the inverse action on the node number and Hamiltonian indices. Finally, from (69), the eigenvalues form degenerate multiplets across the different Hamiltonians, with En0 = En−1,1 = En−2,2 = ⋯. In fact, by repeated application of (67)(69), the eigenfunctions and eigenvalues of any two Hamiltonians in the sequence can be related. Comparing the shift relations among eigenspectra arising from factorization (Fig. 3) with those anticipated for shift operators between SU(1, 1) irreps (Fig. 1), we may already see the essential resemblance.

FIG. 3.

Structure of the relations among eigenfunctions and eigenvalues for a sequence of factorized Hamiltonians Hm, defined in (67)(69).

FIG. 3.

Structure of the relations among eigenfunctions and eigenvalues for a sequence of factorized Hamiltonians Hm, defined in (67)(69).

Close modal

We may now generalize these relations to the case in which the Hamiltonians are n-dependent, that is, to eigenproblems of the form

(70)

Factorizations (64) and (65) then generalize to

(71)
(72)

again with the assumption that the ground state eigenfunctions are annihilated, i.e., A0m(r) ψ0m(r) = 0. We label the latter Hamiltonian with the index n − 1, rather than n, in recognition of the anticipated correspondence of eigenfunctions across the Hamiltonians.

We may now obtain n-dependent shift operator relations

(73)
(74)

while the eigenvalue relation En−1,m+1 = Enm from (69) still holds. These relations follow by first establishing that Anm acts on the eigenfunction ψn−1,m+1 of Hn−1,m+1 to give an eigenfunction of Hnm, with the same eigenvalue En−1,m+1,

(75)

where we have successively applied the definition of Hnm from (71), reassociated, recognized the definition of Hn−1,m+1 from (72), and applied the eigenproblem definition from (70). It then may be verified, by a similar argument, that Anm acts on the eigenfunction ψnm of Hnm to give an eigenfunction of Hn−1,m+1, with the same eigenvalue Enm, i.e.,

(76)

To apply the shift operators in practice, it is desirable not only to know that a proportionality exists, as in (73) and (74), but also to know the constant of proportionality, so that we may construct normalized shift operatorsAnm and Anm. We require that these operators, when acting on normalized eigenfunctions, then yield normalized eigenfunctions, so that proportionalities (73) and (74) become equalities

(77)
(78)

The constants of proportionality in (73) and (74) follow uniquely, within phase, from the assumed adjoint relationship between the operators Anm and Anm and from the requirement that the wave functions ψnm be normalized. In particular, the norm of Anmψn1,m+1 is readily evaluated, since (we switch to bracket notation to facilitate representing the inner product)

(79)

where we have recognized the relationship between AnmAnm and Hn−1,m+1 by (72) and then applied the eigenproblem definition from (70). The result may simply be rewritten as |Anmψn1,m+1|2=EnmKnm by eigenvalue relation (69). Similarly, considering the norm of Anmψnm gives an identical value |Anmψnm|2 = EnmKnm.

Specifically, to construct the normalized shift operators, let

(80)
(81)

where σnm and σnm represent the phases, and cnm and cnm represent the normalization factors. Thus, insisting that both ψnm=Anmψn1,m+1 and ψn1,m+1=Anmψnm be normalized gives cnm=cnm=(EnmKnm)1/2.

The phases (signs) of the solutions to an eigenproblem (66) or (70) are not determined by the eigenproblem itself, but rather are a matter of convention. For instance, it may be chosen that the wave functions are positive as r → 0 or positive as r → ∞. Thus, the phases σnm and σnm appearing in the normalized shift operators cannot be deduced purely from the factorization formalism but must rather be chosen so that the shift operators yield wave functions with phases consistent with the adopted convention. The self-consistency requirement that ψnm=Anmψn1,m+1=AnmAnmψnm imposes the constraint that σnm=σnm.

If the normalized shift operators are known for all m, then we have a way of expressing eigenfunctions of any Hamiltonian Hm+k in terms of those of Hm, via (77) or (78). In particular, if k is a positive integer, then we repeatedly apply A to obtain

(82)

We now deduce shift operators on the l quantum number of the Laguerre radial functions Snl(r). We deduce these shift operators by recasting differential equation (16) into the form of a factorizable n-dependent Hamiltonian eigenproblem, so that we may use the formalism laid out in Sec. IV A. Specifically, (16) can be rewritten in terms of Hamiltonian operator (54) as

(83)

where Enl = − 1 for all n and l. However, to put this Hamiltonian into one of the standard factorizable forms cataloged by Infeld and Hull,29 we must first eliminate the first-derivative term, through a similarity transformation r1/2Hnlr−1/2. Hamiltonian equation (83) becomes

(84)

where

(85)

Comparing (84) with the standard form for Coulomb-type problems, given by (8.0.1) of Ref. 29, yields the factorization55 

(86)

and constant term

(87)

The tildes on à and à indicate that these are the shift operators corresponding to similarity transformed Hamiltonian operator (84). We also note that K0l = − 1 = E0l, thus satisfying the condition to obtain relationships (67)(69) between partner Hamiltonian eigenfunctions and eigenvalues, equivalent to Ã0lR0l=0 and thus A0lS0l = 0 (see Sec. IV A).

From (86) and (87), we can write (84) in the n-dependent factorized form of (71), as

(88)

Normalized shift operators (80) and (81) for the functions Rnl(r) are given by

(89)
(90)

Here we adopt phases σnl = − 1, to enforce the convention that Rnl(r) [and thus Snl(r) below] is always positive at the origin.

To obtain the shift operators for the functions Snl(r), we must apply the inverse similarity transformation. Thus, Anl(r)=r1/2Ãnl(r)r1/2 gives

(91)

and, similarly, Anl(r)=r1/2Ãnl(r)r1/2 gives

(92)

Now that we have shift operators relating Snl(r) functions with different l,

(93)

we can derive expressions for the matrix elements between functions Snl and Snl for l′ = l ± k, for k = 1, 2, …, by writing Sn′,l+k in terms of Snl using relationships (93). Note that the matrix elements for l′ = lk can be obtained from the matrix elements of l′ = l + k by making the appropriate substitutions ( Appendix), and thus it is sufficient to consider only l′ = l + k.

For l′ = l ± 1, we simply apply the raising shift operator to Sn+1,l,

(94)

Applying (51) and (52) to (94),

(95)

For l′ = l + 2, the raising shift operator must be applied twice,

(96)

From this, we have that

where

(97)

Similarly, we could obtain the explicit relationship between Sn′,l+k and Snl for any k by applying the shift operator to Snk,lk times. However, we only explicitly calculate examples of matrix elements for l′ = l ± 1 and l′ = l ± 2 in this paper.

Once we have an expression for Sn,l+k(r) in terms of Snl(r), we can then use the matrix elements already calculated for l = l′ to derive the matrix elements for l′≠l. For example,

(98)

From (46), we then deduce that

(99)

Using the orthonormality of the Snl(r), we can read off the matrix elements 〈nl|r|n, l + 1〉 from this expression, i.e.,

(100)
(101)
(102)

and all other matrix elements are zero. Matrix elements of select operators between functions with l′ = l ± 1 or l′ = l ± 2 are given in the  Appendix (Tables II and III).

TABLE II.

Matrix elements between functions in the Sn,l+1 and Snl bases.

nl|n, l + 1〉 n′ ≥ n + 2 
 n′ = n + 1 n+1n+2l+41/2 
 n′ ≤ n (2l+3)n!(n+2l+2)!n!(n+2l+4)!1/2 
nl|r|n, l + 1〉 n′ ≥ n + 3 
 n′ = n + 2 12((n+1)(n+2))1/2 
 n′ = n + 1 −((n + 1)(n + 2l + 4))1/2 
 n′ = n 12((n+2l+4)(n+2l+3))1/2 
 n′ ≤ n − 1 
nl|1r|n,l+1 n′ ≥ n + 1 
 n′ ≤ n (2(n+1)2n)n!(n+2l+2)!n!(n+2l+4)!1/2 
nl|ddr|n,l+1 n′ ≥ n + 2 
 n′ = n + 1 n+1n+2l+41/2 
 n′ ≤ n (n(2l+4)n(2l+2)+1)n!(n+2l+2)!n!(n+2l+4)!1/2 
nl|n, l + 1〉 n′ ≥ n + 2 
 n′ = n + 1 n+1n+2l+41/2 
 n′ ≤ n (2l+3)n!(n+2l+2)!n!(n+2l+4)!1/2 
nl|r|n, l + 1〉 n′ ≥ n + 3 
 n′ = n + 2 12((n+1)(n+2))1/2 
 n′ = n + 1 −((n + 1)(n + 2l + 4))1/2 
 n′ = n 12((n+2l+4)(n+2l+3))1/2 
 n′ ≤ n − 1 
nl|1r|n,l+1 n′ ≥ n + 1 
 n′ ≤ n (2(n+1)2n)n!(n+2l+2)!n!(n+2l+4)!1/2 
nl|ddr|n,l+1 n′ ≥ n + 2 
 n′ = n + 1 n+1n+2l+41/2 
 n′ ≤ n (n(2l+4)n(2l+2)+1)n!(n+2l+2)!n!(n+2l+4)!1/2 
TABLE III.

Matrix elements between functions in the Sn,l+2 and Snl bases.

nl|n, l + 2〉a n′ ≥ n + 3 
 n′ = n + 2 (n+1)(n+2)(n+2l+6)(n+2l+5)1/2 
 n′ ≤ n + 1 cnnn!(n+2l+2)!n!(n+2l+6)!1/2 
nl|r|n, l + 2〉 n′ ≥ n + 4 
 n′ = n + 3 12(n+3)!n!(n+2l+6)1/2 
 n′ = n + 2 (n+3l+152)(n+2)!(n+2l+4)!n!(n+2l+6)!1/2 
 n′ = n + 1 12(n(n+14)+6l(n+2l+9)+60)(n+1)(n+2l+3)!(n+2l+6)!1/2 
 n′ ≤ n (l+2)(2l+3)(2l+5)n!(n+2l+2)!n!(n+2l+6)!1/2 
nl|r2|n, l + 2〉 n′ ≥ n + 5 
 n′ = n + 4 14(n+4)!n!1/2 
 n′ = n + 3 (n+2l+6)(n+3)!n!1/2 
 n′ = n + 2 32(n+2l+6)!(n+2)!n!(n+2l+4)!1/2 
 n′ = n + 1 (n+2l+6)!(n+1)(n+2l+3)!1/2 
 n′ = n 14(n+2l+6)!(n+2l+2)!1/2 
 n′ ≤ n − 1 
nl|1r2|n,l+2 n′ ≥ n + 1 
 n′ ≤ n 23(nn+1)(nn+2)(nn+3)n!(n+2l+2)!n!(n+2l+6)!1/2 
nl|ddr|n,l+2b n′ ≥ n + 3 
 n′ = n + 2 (n+2)!(n+2l+4)!n!(n+2l+6)!1/2 
 n′ ≤ n + 1 dnnn!(n+2l+2)!n!(n+2l+6)!1/2 
nl|d2dr2|n,l+2c n′ ≥ n + 3 
 n′ = n + 2 (n+2)!(n+2l+4)!n!(n+2l+6)!1/2 
 n′ ≤ n + 2 fnnn!(n+2l+2)!n!(n+2l+6)!1/2 
nl|n, l + 2〉a n′ ≥ n + 3 
 n′ = n + 2 (n+1)(n+2)(n+2l+6)(n+2l+5)1/2 
 n′ ≤ n + 1 cnnn!(n+2l+2)!n!(n+2l+6)!1/2 
nl|r|n, l + 2〉 n′ ≥ n + 4 
 n′ = n + 3 12(n+3)!n!(n+2l+6)1/2 
 n′ = n + 2 (n+3l+152)(n+2)!(n+2l+4)!n!(n+2l+6)!1/2 
 n′ = n + 1 12(n(n+14)+6l(n+2l+9)+60)(n+1)(n+2l+3)!(n+2l+6)!1/2 
 n′ ≤ n (l+2)(2l+3)(2l+5)n!(n+2l+2)!n!(n+2l+6)!1/2 
nl|r2|n, l + 2〉 n′ ≥ n + 5 
 n′ = n + 4 14(n+4)!n!1/2 
 n′ = n + 3 (n+2l+6)(n+3)!n!1/2 
 n′ = n + 2 32(n+2l+6)!(n+2)!n!(n+2l+4)!1/2 
 n′ = n + 1 (n+2l+6)!(n+1)(n+2l+3)!1/2 
 n′ = n 14(n+2l+6)!(n+2l+2)!1/2 
 n′ ≤ n − 1 
nl|1r2|n,l+2 n′ ≥ n + 1 
 n′ ≤ n 23(nn+1)(nn+2)(nn+3)n!(n+2l+2)!n!(n+2l+6)!1/2 
nl|ddr|n,l+2b n′ ≥ n + 3 
 n′ = n + 2 (n+2)!(n+2l+4)!n!(n+2l+6)!1/2 
 n′ ≤ n + 1 dnnn!(n+2l+2)!n!(n+2l+6)!1/2 
nl|d2dr2|n,l+2c n′ ≥ n + 3 
 n′ = n + 2 (n+2)!(n+2l+4)!n!(n+2l+6)!1/2 
 n′ ≤ n + 2 fnnn!(n+2l+2)!n!(n+2l+6)!1/2 
a

cnn = (2l + 4)(2l + 3) n − (2l + 5)(2l + 4) n′ + (2l + 4)(2l + 3).

b

dnn = n′[4n(l + 2)2 + l(2l + 9) + 11 − n′(2l + 5)(l + 3)] − (n + 1)(2l + 3)(n(l + 1) − 2).

c

fnn=23{n[(l+3)n((l+4)n3((l+2)n+1))+(l+2)n(3n(l+1)6)l(l+7)9]n[(l+1)n(nl9)l(l+13)9]+3l}.

Now that we have a way to calculate matrix elements of radial operators with respect to the Snl basis on ℝ+, we can calculate the matrix elements for irreducible tensor operators in the Λnlm basis on ℝ3. In Sec. II A we saw that the matrix element for Oaα=RΘaα can be written (20) as

(103)

where γ(R) is given by (21). From standard angular momentum theory for spherical tensor operators,19 matrix elements involving different angular momentum projection quantum numbers m and m′ are all related via the Wigner-Eckart theorem, so we need only consider the reduced matrix elements. The reduced matrix element of Oaα is given by the product of the SO(3) reduced matrix element of the angular operator and the matrix element of the radial part in the Snl basis,

(104)

Note that we follow the normalization and phase convention of, e.g., Refs. 19 and 56 for the reduced matrix elements in the Wigner-Eckart theorem, i.e.,

(105)

where the quantity delimited by parentheses is a Clebsch-Gordan coefficient, and lˆ=2l+1.

To illustrate this method of calculating the matrix elements in the Λnlm(r) basis, we will consider several examples, specifically, rk for k ∈ ℤ, ∇2 and ∇1.

Matrix elements of polynomial operators that depend only on r are particularly simple to calculate. In spherical tensor form, rkrk100. Applying the factorization in (104), we can write

(106)

The reduced matrix element of the identity operator is l100l=lˆδll, and, under the transformation (21), γ(rk) = rk for any k. Therefore, (106) reduces to

(107)

The radial matrix elements are already known from Sec. III C and can be obtained from Table I for select values of k.

The Laplacian ∇2 has the well known form19 

(108)

We factor each of the terms contributing to the matrix element of this operator into radial and angular parts, according to (104), and recall L2|lm〉 = l(l + 1)|lm〉, yielding

(109)

The matrix elements of d2/dr2 and r−2 can be read from Table I, and we find that

(110)

We now want to calculate the reduced matrix elements of the gradient operator , which can be expressed as a rank 1 tensor ∇1 with components19 

(111)

where Y1 is the irreducible tensor whose components are the spherical harmonics Y1μ(rˆ), and L1 is the orbital angular momentum operator, likewise considered as a spherical tensor. The reduced matrix element of ∇1 is thus given by

(112)

d by applying (104). The angular part of the second term can be factored as

(113)

by the standard reduction formula for reduced matrix elements of product operators on a single space.19 Using the reduced matrix elements of Y1 and L1 given in Ref. 19, the expression for the reduced matrix element of ∇, becomes

(114)

The terms contributing to the radial matrix elements can then be read from Table II, and we have

(115)

and

(116)

It has long been recognized that an SU(1, 1) × SO(3) algebraic structure, together with shift operators between SU(1, 1) irreps, forms a successful combination in the treatment of the classic sets of central force eigenfunctions—those of the harmonic oscillator and Coulomb problems. We have now demonstrated that a similar framework can be developed and effectively applied to the Laguerre functions, a basis set of practical value in quantum mechanical one-body, few-body, and many-body problems.

The essential results thus obtained provide for the evaluation of matrix elements of spherical tensor operators. This is the immediate application of interest in formulating quantum mechanical problems for numerical solution. The evaluation of matrix elements is accomplished by decomposing the full operator, in the three-dimensional coordinate space, into parts which act only radial or angular coordinates. Then, recall the structure outlined in Fig. 1: (1) the SU(1, 1) algebra of the radial functions is used to deduce radial matrix elements, while (2) the shift operators connect the different families of radial functions, or SU(1, 1) irreps, associated with different angular functions, and (3) the SO(3) angular momentum algebra is used to address the action of the operator on the angular functions themselves. The SU(1, 1) algebra follows directly from the eigenvalue equation satisfied by the radial functions. The ladder operators are obtained by factorization (or, equivalently, SUSY QM) methods, reformulated to allow for quantum number dependence of the shift operators. And the SO(3) algebra provides the familiar tools of the Wigner-Eckart theorem and Racah’s reduction formulas for the angular operators.

Although the results, as presented, have been specialized to the Laguerre functions on ℝ3, their applicability is broader. In particular, the results for the radial functions as a basis on ℝ+ stand alone or in conjunction with any decomposition of N-dimensional coordinates on ℝN into a radial coordinate on ℝ+ and angular coordinates on the (N − 1)-sphere SN−1. For instance, functions of the Laguerre type considered here are applied as hyperradial basis functions in few-body calculations in Jacobi coordinates (e.g., Ref. 57). Alternatively, much as the SU(1, 1) treatment of the harmonic oscillator radial functions may be extended to a broader family of modified harmonic oscillator functions, by transitioning from a discrete SU(1, 1) irrep label t = l + 3/2 to a continuous label23 (physically corresponding from the transition from a pure oscillator potential to a Davidson potential,58 containing an additional 1/r2 dependence), a similar generalization may be carried out for the present results, yielding an algebraic description of families of modified Laguerre functions. The general framework may also be carried over to the Coulomb-Sturmian functions, for which the SU(1, 1) radial algebra is considered in Ref. 59.

Finally, it is worth noting that, so far, we have focused exclusively on the SU(1, 1) character of the basis functions, rather than on the SU(1, 1) tensorial properties of operators on the radial space. Through these properties, it is possible to benefit from the Wigner-Eckart theorem, not only of the SO(3) angular algebra as we have done so far, but also of the SU(1, 1) radial algebra. Matrix elements involving different members of the same irrep (i.e., different radial functions for the same angular momentum) may then be related to each other through SU(1, 1) Clebsch-Gordan coefficients. The SU(1, 1) Wigner-Eckart theorem, applied in this fashion, has been a powerful tool in the construction of closed-form expressions for matrix elements in the classic harmonic oscillator and Coulomb eigenproblems.20,60

We would like to thank C. Kolda and J. de Blas for valuable discussions, M. Wolf for comments on the manuscript, and A. I. Rakoski for verification of expressions for matrix elements. This work was supported by the Research Corporation for Science Advancement under a Cottrell Scholar Award and by the US Department of Energy under Grant No. DE-FG02-95ER-40934.

In the following tables, we have summarized the matrix elements of some of the most basic radial operators that can be constructed from r, r−1, and d/dr, taken in the Laguerre function radial basis. Matrix elements within a single SU(1, 1) irrep, i.e., between radial functions Snl of the same l, are given in Table I. Matrix elements between radial functions belonging to different SU(1, 1) irreps, corresponding to angular momenta differing by Δl = 1 and 2, are given in Tables II and III, respectively.

Note that the matrix elements in Tables IIII are of the form 〈nl|R|nl〉, 〈nl|R|n, l + 1〉, or 〈nl|R|n, l + 2〉, respectively, as this is the form that is most naturally derived from applying the shift operators (see Sec. IV). However, one may express the l-changing matrix elements in the form 〈n′, l − 1|R|nl〉 or 〈n′, l − 2|R|nl〉 by making the appropriate substitution. For instance, to obtain 〈n′, l − 1|r|nl〉 from the matrix elements in Table II, one may simply make the substitution ll − 1. Alternatively, matrix elements of the form 〈n′, l − 1|R|nl〉 or 〈n′, l − 2|R|nl〉 may be obtained by Hermitian conjugation. For instance, to obtain 〈n′, l − 1|r|nl〉 from the matrix elements in Table II, one may simply note that 〈n′, l + 1|r|nl〉 = 〈nl|r|n′, l + 1〉, since the matrix elements are real-valued, and make the interchange nn′. When applying Hermitian conjugation to differential operators appearing in Tables IIII, recall that the operator d/dr is anti-Hermitian (e.g., Ref. 16).

1.
H.
Shull
and
P.-O.
Löwdin
,
J. Chem. Phys.
23
(
7
),
1362
(
1955
).
2.
E.
Filter
and
E. O.
Steinborn
,
J. Math. Phys.
21
,
2725
(
1980
).
3.
E. J.
Weniger
,
J. Math. Phys.
26
,
276
(
1985
).
4.
H.
Shull
and
P.-O.
Löwdin
,
J. Chem. Phys.
23
,
1565
(
1955
).
5.
P.
Lowdin
and
H.
Shull
,
Phys. Rev.
101
(
6
),
1730
(
1956
).
6.
T.
Helgaker
,
P.
Jørgensen
, and
J.
Olsen
,
Molecular Electron-Structure Theory
(
Wiley
,
Chichester
,
2000
).
7.
S.
Jacobs
,
M. G.
Olsson
, and
C.
Suchyta
III
,
Phys. Rev. D
33
,
3338
(
1986
).
8.
L. P.
Fulcher
,
Z.
Chen
, and
K. C.
Yeong
,
Phys. Rev. D
47
,
4122
(
1993
).
9.
C.
Olson
and
M. G.
Olsson
,
Phys. Rev. D
49
(
9
),
4675
(
1994
).
10.
M. G.
Olsson
and
S.
Veseli
,
Phys. Rev. D
51
(
9
),
5079
(
1995
).
11.
B. D.
Keister
and
W. N.
Polyzou
,
J. Comput. Phys.
134
,
231
(
1997
).
12.
M.
Pervin
, “
Semileptonic decay of heavy baryons in a constituent quark model
,” Ph.D. thesis,
Florida State University
,
2005
.
13.
S.
Veerasamy
and
W. N.
Polyzou
,
Phys. Rev. C
84
,
034003
(
2011
).
14.
M. A.
Caprio
,
P.
Maris
, and
J. P.
Vary
,
Phys. Rev. C
86
,
034312
(
2012
).
15.
M. A.
Caprio
,
P.
Maris
, and
J. P.
Vary
,
Phys. Rev. C
90
,
034305
(
2014
).
16.
J. J.
Sakurai
,
Modern Quantum Mechanics
, Rev. ed. (
Addison-Wesley
,
Reading, Massachusetts
,
1994
).
17.
J. W.
Negele
and
H.
Orland
,
Quantum Many-Particle Systems
(
Addison-Wesley
,
Redwood City, CA
,
1988
).
18.
M.
Moshinsky
and
Y. F.
Smirnov
,
The Harmonic Oscillator in Modern Physics
(
Harwood Academic Publishers
,
Amsterdam
,
1996
).
19.
D. A.
Varshalovich
,
A. N.
Moskalev
, and
V. K.
Khersonskii
,
Quantum Theory of Angular Momentum
(
World Scientific
,
Singapore
,
1988
).
20.
B. G.
Wybourne
,
Classical Groups for Physicists
(
Wiley
,
New York
,
1974
).
21.
J.
Čížek
and
J.
Paldus
,
Int. J. Quantum Chem.
12
(
5
),
875
(
1977
).
22.
T. H.
Cooke
and
J. L.
Wood
,
Am. J. Phys.
70
(
9
),
945
(
2002
).
24.
L.
Armstrong
, Jr.
,
Phys. Rev. A
3
,
1546
(
1971
).
25.
B. G.
Adams
,
J.
Cizek
, and
J.
Paldus
,
Adv. Quantum Chem.
19
,
1
(
1988
).
26.
27.
E.
Kelbert
,
A.
Hyder
,
F.
Demir
,
Z. T.
Hlousek
, and
Z.
Papp
,
J. Phys. A
40
,
7721
(
2007
).
28.
E.
Schrödinger
,
Proc. R. Ir. Acad.
46
,
9
(
1940
).
29.
L.
Infeld
and
T. E.
Hull
,
Rev. Mod. Phys.
23
,
21
(
1951
).
30.
F.
Cooper
,
A.
Khare
, and
U.
Sukhatme
,
Phys. Rep.
251
,
267
(
1995
).
31.
D. J.
Rowe
,
M. J.
Carvalho
, and
J.
Repka
,
Rev. Mod. Phys.
84
,
711
(
2012
).
32.
B. R.
Barrett
,
P.
Navrátil
, and
J. P.
Vary
,
Prog. Part. Nucl. Phys.
69
,
131
(
2013
).
33.
G. B.
Arfken
and
H. J.
Weber
,
Mathematical Methods for Physicists
, 6th ed. (
Academic Press
,
San Diego
,
2005
).
34.
K. T. R.
Davies
,
S. J.
Krieger
, and
M.
Baranger
,
Nucl. Phys.
84
,
545
(
1966
).
36.

In comparing expressions in this work, such as (2), with standard treatments of the Coulomb problem (and with the electron-structure theory literature in general, e.g., Ref. 6) one must be alert to the dichotomous use of the symbol n. We retain the generic notation for central force problems (e.g., Ref. 16), in which n is the radial quantum number (n = 0, 1, …), representing the number of nodes in the radial wave function. This use of n is common in nuclear physics and treatments of the harmonic oscillator.18 Then, the principal quantum number, which determines the energy degeneracies, is indicated by N. In particular, N = 2n + l (N = 0, 1, …) for the harmonic oscillator problem, with energy E(N) ∝ N + 3/2, or N = n + l + 1 (N = 1, 2, …) for the Coulomb problem, with energy E(N) ∝ − 1/N2. However, in treatments of the hydrogen atom, it is the principal quantum number which is traditionally denoted by n. Thus, an alternative symbol, such as ν, must be used for the radial quantum number, and we may write n = ν + l + 1, where now E(n) ∝ − 1/n2. In electron-structure theory, this hydrogenic definition of n as ν + l + 1 in terms of the other quantum numbers ν and l is conventionally carried over to other basis sets,6 such as harmonic oscillator functions and Laguerre functions, despite having no interpretation as a principal quantum number for those bases. To translate expressions from the former convention (n as radial quantum number) to the latter convention (n as hydrogenic principal quantum number), one must make replacements nnl − 1.

37.
E. A.
Hylleraas
,
Z. Phys.
48
,
469
(
1928
).
38.
M.
Rotenberg
,
Ann. Phys. (N. Y.)
19
,
262
(
1962
).
39.
B.
Klahn
and
W. A.
Bingel
,
Theor. Chim. Acta
44
,
27
(
1977
).
40.

Although the term Sturmian by itself has been used to refer to the Coulomb-Sturmian functions,38 this term is more general and may be used to refer to Sturm-Liouville expansion basis functions associated with other choices of potential V(r).

41.
M.
Rotenberg
,
Adv. At. Mol. Phys.
6
,
233
(
1970
).
42.
E. J.
Heller
and
H. A.
Yamani
,
Phys. Rev. A
9
,
1209
(
1974
).
43.
C.
Coletti
,
D.
Calderini
, and
V.
Aquilanti
,
Adv. Quantum Chem.
67
,
73
(
2013
).
45.
Z.
Papp
and
S.
Moszkowski
,
Mod. Phys. Lett. B
22
,
2201
(
2008
).
47.
B.
Klahn
and
W. A.
Bingel
,
Theor. Chim. Acta
44
,
9
(
1977
).
48.

Although we use the term Laguerre function to refer to the full function Λnlm(r) of (4) in three dimensions, as in Ref. 6, it should be noted that this term is also used in a more generic sense to refer to the corresponding unnormalized radial wave function. In particular, Ref. 33 defines the Laguerre function as ψnk(r)=rk/2Lnk(r)er/2. With the identification k = 2l + 2, and except for the omission of the normalization factor, this may be recognized as the auxiliary radial function (that is, the radial function after absorbing the radial integration metric) Snl(r) associated with (4).

49.

Alternatively, in an entirely different meaning, the term Laguerre function may be used to refer to the generalization of the associated Laguerre polynomial Lnα(x) to nonintegral parameters n and α.33 

50.

Due to this similarity in functional form, the Laguerre functions have in a few instances been referred to as “Coulomb-Sturmian functions” in the hadronic and nuclear physics literature.11,14,15 However, it should be emphasized that Coulomb-Sturmian functions (3) and Laguerre functions (4) are distinct, and only the former are actually obtained as solutions to the Coulomb potential’s associated Sturm-Liouville problem.

51.

More precisely, the associated radial equation is of the Sturm-Liouville form −(d/dx)[p(x)(d/dx) yn(x)] + q(x) yn(x) = λV(x) yn(x), with p(x) > 0 and q(x) > 0. If p(x) and q(x) are continuous and the derivative of p(x) is continuous on the closed interval [a, b], Sturm-Liouville theory then tells us that the normalized eigenfunctions form an orthonormal basis abyn(x)ym(x)V(x)dx=δmn in the Hilbert Space L2([a,b],V(x)dx), i.e., under an inner product with weighting function V(x).33 

52.
L. E.
Gendenshtein
,
Pis’ma Zh. Eksp. Teor. Fiz.
38
(
6
),
299
(
1983
).
53.
P. A. M.
Dirac
,
The Principles of Quantum Mechanics
, 4th (Revised) ed.
The International Series of Monographs on Physics
Vol.
27
(
Clarendon Press
,
Oxford
,
1967
).
54.
K. T.
Hecht
,
Quantum Mechanics
(
Springer
,
Berlin
,
2000
).
55.

The parameter m of Infeld and Hull29 may be identified with either l + 1/2 or −(l + 3/2) in the present problem (we adopt the latter choice to obtain a more convenient form for our subsequent results), and the parameter q of Infeld and Hull is identified with −αnl.

56.
A. R.
Edmonds
,
Angular Momentum in Quantum Mechanics
, 2nd ed.
Investigations in Physics
Vol.
4
(
Princeton University Press
,
Princeton, New Jersey
,
1960
).
57.
S.
Bacca
,
N.
Barnea
, and
A.
Schwenk
,
Phys. Rev. C
86
,
034321
(
2012
).
58.
P. M.
Davidson
,
Proc. R. Soc. London, Ser. A
135
,
459
(
1932
).
59.
G.
Lévai
,
B.
Kónya
, and
Z.
Papp
,
J. Math. Phys.
39
,
5811
(
1998
).
60.
C.
Quesne
and
M.
Moshinsky
,
J. Math. Phys.
12
,
1780
(
1971
).