Next Article in Journal
State-Selective Double Photoionization of Atomic Carbon and Neon
Previous Article in Journal
Modified Ji-Huan He’s Frequency Formulation for Large-Amplitude Electron Plasma Oscillations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two-Electron Atomic Systems—A Simple Method for Calculating the Ground State near the Nucleus: Some Applications

by
Evgeny Z. Liverts
1,* and
Rajmund Krivec
2
1
Racah Institute of Physics, The Hebrew University, Jerusalem 91904, Israel
2
Department of Theoretical Physics, J. Stefan Institute, 1000 Ljubljana, Slovenia
*
Author to whom correspondence should be addressed.
Atoms 2024, 12(12), 69; https://doi.org/10.3390/atoms12120069
Submission received: 9 October 2024 / Revised: 3 December 2024 / Accepted: 11 December 2024 / Published: 14 December 2024

Abstract

A simple method of non-relativistic variational calculations of the electronic structure of a two-electron atom/ion, primarily near the nucleus, is proposed. The method as a whole consists of a standard solution of a generalized matrix eigenvalue equation, all matrix elements of which are reduced to a numerical calculation of one-dimensional integrals. The distinctive features of the method are as follows: The use of the hyperspherical coordinate system. The inclusion of logarithms of the hyperspherical radius R in the basis functions, similar to the Fock expansion. Using a special basis function including the leading angular Fock coefficients to provide the correct behavior of the wave function near the nucleus. The main numerical parameters characterizing the properties of the helium atom and a number of helium-like ions near the nucleus are calculated and presented in tables. Among others, the specific coefficients, a 21 , of the Fock expansion, which can only be calculated using a wave function with the correct behavior near the nucleus, are presented in table and graphs.

1. Introduction

The trivial statement is that all the main characteristics of a two-electron atomic system in the S state, having an infinitely massive nucleus with charge, Z, and non-relativistic energy, E, are determined by the corresponding wave function (WF) Φ ( r 1 , r 2 , r 12 ) , where r 1 and r 2 are the electron-nucleus distances and r 12 is the distance between the electrons. It is also well known that the behavior of the ground state WF in the vicinity of the nucleus located at the origin is determined by the Fock expansion (FE) [1,2],
Φ ( r 1 , r 2 , r 12 ) Ψ ( R , α , θ ) = k = 0 R k p = 0 [ k / 2 ] ψ k , p ( α , θ ) ln p R ,
where the hyperspherical coordinates (HSC) R , α and θ are defined by the relations
R = r 1 2 + r 2 2 , α = 2 arctan r 2 r 1 , θ = arccos r 1 2 + r 2 2 r 12 2 2 r 1 r 2 ,
whereas ψ k , p ( α , θ ) are the angular Fock coefficients (AFC), many of which have been calculated previously (see, e.g., [3,4,5] and references therein). It is worth noting that the exact representation of the AFCs ψ k , p with even k and maximum values p = k / 2 was recently presented in Ref. [6]. These AFCs will be used in what follows. Recall that the convergence of expansion (1) has been proven in Ref. [7].
There are a large amount of methods for calculating the electronic structure of the two-electron atomic systems. An excellent review on this topic can be found in Refs. [3,8,9,10,11]. However, we know only one technique that correctly represents the WF near the nucleus. It is the so-called correlation function hyperspherical harmonic method (CFHHM) [12,13,14]. This method uses an expansion similar to the FE to represent the WF near the nucleus and uses an expansion in hyperspherical harmonics (HHs) to represent the AFCs. However, the HH-expansion is known to converge very slowly. Although this method makes it possible to increase the convergence of the HH-expansion, a sufficiently good accuracy requires a large HH basis size (thousands of basis functions). Moreover, instead of the widely used variational method for solving the problem, an algebraic method for solving coupled systems of ordinary differential equations is used. All of the above creates great computational difficulties in implementing the method.
It follows that it would be extremely useful to develop a much simpler method for calculating the WF with correct behavior near the nucleus. The main objective of this article is to present such a method, as well as some of its non-trivial applications.

2. General Characteristics of the Method

The simplicity of the proposed method primarily implies the use of a standard variational method, which involves solving the generalized matrix eigenvalue equation
H ^ = E S ^ ,
where H ^ V ^ + T ^ represents the matrix of the non-relativistic Hamiltonian, whereas S ^ , V ^ and T ^ are the overlap, potential and kinetic energy matrices, respectively, the elements of which are defined as follows:
S j , j = f j f j d v , V j , j = f j V f j d v , T j , j = f j T f j d v .
The potential, V, and kinetic energy, T, operators will be defined later. The number of basis functions (BFs), f j , is limited by the basis size, N b ( 1 j N b ) . The resulting WF can then be constructed in the form
Ψ = j = 1 N b C j f j ,
where C j is the j-th component of the eigenvector corresponding to solution of Equation (3) with eigenvalue E. Note that all matrix elements (MEs) will be calculated using the HSC defined by Equation (2), and the atomic units system. This choice of coordinate system will help us to construct a WF, Ψ , with the behavior near the nucleus represented by the FE (1). Thus, for the volume element in the HSC we have [3]:
d v = R 5 d R d Ω , d Ω = π 2 sin 2 α d α sin θ d θ , R [ 0 , ) , α [ 0 , π ] , θ [ 0 , π ]
The kinetic energy operator, T, is equal to ( 1 / 2 ) Δ , where the Laplacian is:
Δ = 2 R 2 + 5 R R Λ 2 R 2 .
The hyperspherical angular momentum operator Λ 2 Λ 2 ( α , θ ) is defined as follows (see, e.g., [3,4]):
Λ 2 = 4 2 α 2 + 2 cot α α + 1 sin 2 α 2 θ 2 + cot θ θ .
It is reasonable to discuss the form of potential operator, V, for a two-electron atomic system later. At this stage, we will begin to introduce a set of BFs, which is the main feature of the proposed method. In particular, we propose to divide the entire set of BFs into two unequal parts. The first part will consist of only one, but rather complex, special basis function (SBF). The second part will be represented by a large set of simple BFs.
To proceed to introducing SBF, let us first recall the form of AFC for the linear in the R term of the FE:
ψ 1 , 0 ( α , θ ) = Z ( r 1 + r 2 ) R + r 12 2 R = Z η + 1 2 ξ ,
where
η = 1 + sin α , ξ = 1 sin α cos θ .
Secondly, it was already mentioned in the “Introduction” that the AFC ψ k , k / 2 ψ k , k / 2 ( α , θ ) with even k have recently been presented in the following closed analytical form:
ψ k , k / 2 = Z k 2 l = k / 2 , ( 2 ) 0 c k , l Y k , l ( α , θ ) ,
where Y k , l ( α , θ ) are HHs, and the closed form of the coefficients c k , l for k 10 can be found in Ref. [6]. It is seen that both the AFCs (11) and the linear AFC (9) depend on the nucleus charge, Z, and do not depend on the energy, E. However, more importantly, the right-hand side (RHS) of the representation (11) can be reduced to the polynomial in variables η and ξ defined by Equation (10). In particular, we obtain:
ψ k , k / 2 = Z π k 2 b ˜ k 2 n , m b k 2 , n , m η n ξ m ,
where the coefficients b ˜ k 2 and b k 2 , n , m are represented in Table 1.
Taking into account Equations (9) and (12), and also the fact that the products ( ln R ) k / 2 R k ψ k , k / 2 represent the leading terms of the logarithmic series [6] of the FE, it is reasonable to introduce the SBF in the form:
f 0 f 0 ( R , α , θ ) = exp R Z η 1 2 ξ k = 0 ( 2 ) K R k ( ln R ) k / 2 ψ k , k / 2 ( α , θ ) ,
where the AFCs are represented by Equation (12) and K is the parameter of the problem under consideration. It should be emphasized that the SBF (13) is designed to serve two purposes. The first one is to provide the correct representation of the FE terms with even k and maximum power p = k / 2 of ln R . And the second one is to guarantee the correct form of the AFC ψ 1 , 0 ( α , θ ) or (which is equivalent) to guarantee that the WF will satisfy the Kato’s cusp conditions [15,16].
The main part of the complete basis set can be taken as:
f k p , n m f k p , n m ( R , α , θ ) = exp R λ η γ ξ R k ( ln R ) p η n ξ m , ( k > 1 )
where, in general, λ and γ are variational parameters. It is clear that the basis set (14) is bounded by k 2 , since the constant ( k = 0 ) and linear ( k = 1 ) terms in R are described by the SBF (13).
The Coulomb potential operator in the HSC can now be introduced as follows:
V 1 r 12 Z 1 r 1 + 1 r 2 = 1 R 1 ξ 2 Z η sin α .
Note that, according to our building of the complete basis set, the MEs (4) of each of the three matrices T ^ , V ^ and S ^ can be divided into three groups. The first group includes MEs between BFs of the form (14), the second group between BFs (14) and (13) and the third group between BFs (13).
It is also important to note that the constant π 2 included in the angular space element d Ω (see definitions (6)) is a factor for all terms of the matrix Equation (3), and will therefore be omitted in what follows.

3. Overlap and Potential Energy Matrices

According to definition (4), the overlap MEs calculated between BFs of the form (14) are:
S k p , n m ; k p , n m π 2 f k p , n m f k p , n m d v = 0 R 5 d R 0 π sin 2 α d α 0 π R k + k ( ln R ) p + p exp 2 R ( λ η γ ξ ) η n + n ξ m + m sin θ d θ = K 2 ( k + k , p + p , n + n , m + m ; 2 λ , 2 γ ) ,
where the specific calculations of the general 3D (three-dimensional) integral
K L ( k , p , n , m ; a , b ) = 0 π η n sin L α d α 0 π ξ m sin θ d θ 0 R k + 5 ( ln R ) p exp R ( a η b ξ ) d R
will be discussed in Appendix A.
The overlap MEs calculated between BFs (14) and SBF (13) can be represented in terms of the K -integrals (17) as follows:
S k p , n m ; 0 π 2 f k p , n m f 0 d v = = k = 0 , ( 2 ) K Z π k 2 b ˜ k 2 n , m b k 2 , n , m K 2 k 1 , p 2 , n 1 , m 1 ; λ + Z , γ + 1 2 ,
where we introduced notations:
k 1 = k + k , p 1 = p + p , p 2 = p + k / 2 , n 1 = n + n , m 1 = m + m ,
that remain relevant in what follows.
It is worth noting that the integer parameter K is determined by the number of the FE terms R k ( ln R ) p ψ k , p ( α , θ ) , with the maximum value p = k / 2 included into the SBF (13) for this method option.
The third group of MEs, consisting of only one overlap ME calculated between SBFs (13), can be determined as:
S 0 ; 0 π 2 f 0 2 d v = k = 0 , ( 2 ) K b ˜ k 2 n , m b k 2 , n , m k = 0 , ( 2 ) K Z π k + k 2 b ˜ k 2 n , m b k 2 , n , m K 2 k 1 , k 1 2 , n 1 , m 1 ; 2 Z , 1 .
Following the previously proposed scheme for calculating the overlap matrix, it is easy to derive the corresponding formulas for the elements of the potential energy matrix. Thus, using the notations (19), we obtain:
V k p , n m ; k p , n m π 2 f k p , n m 1 R 1 ξ 2 Z η sin α f k p , n m d v = K 2 ( k 1 1 , p 1 , n 1 , m 1 1 ; 2 λ , 2 γ ) 2 Z K 1 ( k 1 1 , p 1 , n 1 + 1 , m 1 ; 2 λ , 2 γ ) ,
V k p , n m ; 0 π 2 f k p , n m 1 R 1 ξ 2 Z η sin α f 0 d v = k = 0 , ( 2 ) K Z π k 2 b ˜ k 2 n , m b k 2 , n , m × K 2 k 1 1 , p 2 , n 1 , m 1 1 ; λ + Z , γ + 1 2 2 Z K 1 k 1 1 , p 2 , n 1 + 1 , m 1 ; λ + Z , γ + 1 2 ,
V 0 ; 0 π 2 f 0 2 1 R 1 ξ 2 Z η sin α d v = k = 0 , ( 2 ) K b ˜ k 2 n , m b k 2 , n , m k = 0 , ( 2 ) K Z π k + k 2 b ˜ k 2 n , m b k 2 , n , m × K 2 k 1 1 , k 1 2 , n 1 , m 1 1 ; 2 Z , 1 2 Z K 1 k 1 1 , k 1 2 , n 1 + 1 , m 1 ; 2 Z , 1 .

4. Kinetic Energy Matrix

To simplify the derivation of calculation formulas for the kinetic energy MEs defined by Equations (4) and (7), we divide them into two parts as follows:
T j , j = 1 2 T j , j ( 1 ) T j , j ( 2 ) ,
where
T j , j ( 1 ) = π 2 f j 2 R 2 + 5 R R f j d v , T j , j ( 2 ) = π 2 f j Λ 2 R 2 f j d v .
To proceed, we use the results of the action of the Laplace operator components on the constituents of our complete basis set defined by Equations (13) and (14). In particular, we obtain:
2 R 2 + 5 R R exp R a η b ξ R k ( ln R ) p = exp R a η b ξ R k ( ln R ) p × p ( p 1 ) R 2 ln 2 R + 2 p ( k + 2 ) R 2 ln R + k ( k + 4 ) R 2 + ( a η b ξ ) 2 2 p R ln R + 2 k + 5 R ( a η b ξ ) ,
Λ 2 exp R a η b ξ η n ξ m = exp R a η b ξ η n ξ m × ( m + n ) ( m + n + 4 ) 2 ( a 2 + b 2 ) R 2 2 n ( n 1 ) η 2 + 4 n a R η + a 2 η 2 R 2 2 m ( m + n + 1 ) ξ 2 2 ( 2 m + n + 2 ) b R ξ + b ( 2 m + 2 n + 5 ) ξ R + b 2 ξ 2 R 2 a ( 2 m + 2 n + 1 ) η R + 2 m a η R ξ 2 + 2 a b η 1 ξ ξ R 2 + 2 sin α n ( m + 2 ) + m n ξ 2 + n b 1 ξ ξ R + a ( m + 4 ) η R 2 a η 3 R m a η R ξ 2 a b η 1 ξ ξ R 2 .
It can be seen that the angular parts of both results represented by the RHSs of the last two equations are expressed in terms of the angular variables η and ξ , and the factor sin 1 α . This enables us to represent the kinetic energy MEs through the integrals of the form (17).
Thus, using Equation (25), as well as Equations (26) and (27) with a = λ and b = γ , we obtain for the kinetic energy MEs calculated between BFs of the form (14):
T k p , n m ; k p , n m ( 1 ) = p ( p 1 ) K 2 ( k 1 2 , p 1 2 , n 1 , m 1 ) + 2 p ( k + 2 ) K 2 ( k 1 2 , p 1 1 , n 1 , m 1 ) + k ( k + 4 ) K 2 ( k 1 2 , p 1 , n 1 , m 1 ) + λ 2 K 2 ( k 1 , p 1 , n 1 + 2 , m 1 ) 2 λ γ K 2 ( k 1 , p 1 , n 1 + 1 , m 1 + 1 ) + γ 2 K 2 ( k 1 , p 1 , n 1 , m 1 + 2 ) 2 p λ K 2 ( k 1 1 , p 1 1 , n 1 + 1 , m 1 ) γ K 2 ( k 1 1 , p 1 1 , n 1 , m 1 + 1 ) ( 2 k + 5 ) λ K 2 ( k 1 1 , p 1 , n 1 + 1 , m 1 ) γ K 2 ( k 1 1 , p 1 , n 1 , m 1 + 1 ) ,
T k p , n m ; k p , n m ( 2 ) = ( m + n ) ( m + n + 4 ) K 2 ( k 1 2 , p 1 , n 1 , m 1 ) 2 ( λ 2 + γ 2 ) K 2 ( k 1 , p 1 , n 1 , m 1 ) 2 n ( n 1 ) K 2 ( k 1 2 , p 1 , n 1 2 , m 1 ) + 4 n λ K 2 ( k 1 1 , p 1 , n 1 1 , m 1 ) 2 m ( m + n + 1 ) K 2 ( k 1 2 , p 1 , n 1 , m 1 2 ) 2 ( 2 m + n + 2 ) γ K 2 ( k 1 1 , p 1 , n 1 , m 1 1 ) + γ ( 2 m + 2 n + 5 ) K 2 ( k 1 1 , p 1 , n 1 , m 1 + 1 ) + γ 2 K 2 ( k 1 , p 1 , n 1 , m 1 + 2 ) λ ( 2 m + 2 n + 1 ) K 2 ( k 1 1 , p 1 , n 1 + 1 , m 1 ) + 2 m λ K 2 ( k 1 1 , p 1 , n 1 + 1 , m 1 2 ) + 2 λ γ K 2 ( k 1 , p 1 , n 1 + 1 , m 1 1 ) K 2 ( k 1 , p 1 , n 1 + 1 , m 1 + 1 ) + λ 2 K 2 ( k 1 , p 1 , n 1 + 2 , m 1 ) 2 n ( m + 2 ) K 1 ( k 1 2 , p 1 , n 1 , m 1 ) + 2 n γ K 1 ( k 1 1 , p 1 , n 1 , m 1 1 ) K 1 ( k 1 1 , p 1 , n 1 , m 1 + 1 ) + 2 m n K 1 ( k 1 2 , p 1 , n 1 , m 1 2 ) + 2 λ ( m + 4 ) K 1 ( k 1 1 , p 1 , n 1 + 1 , m 1 ) 4 λ K 1 ( k 1 1 , p 1 , n 1 + 3 , m 1 ) 2 m λ K 1 ( k 1 1 , p 1 , n 1 + 1 , m 1 2 ) 2 λ γ K 1 ( k 1 , p 1 , n 1 + 1 , m 1 1 ) K 1 ( k 1 , p 1 , n 1 + 1 , m 1 + 1 ) .
For simplicity, we replaced the designations K L ( k , p , n , m ; 2 λ , 2 γ ) with K L ( k , p , n , m ) for all K -integrals in Equations (28) and (29), that is, we simply omitted the last two common parameters. We also used the notations introduced in Equation (19).
Similarly, for the kinetic energy MEs calculated between BFs (13) and SBF (14), we obtain:
T 0 ; k p , n m ( j ) = k = 0 , ( 2 ) K Z π k 2 b ˜ k 2 n , m b k 2 , n , m T ^ k p , n m ; k p , n m ( j ) , ( j = 1 , 2 )
where T ^ k p , n m ; k p , n m ( 1 ) and T ^ k p , n m ; k p , n m ( 2 ) represent the RHSs of Equations (28) and (29), respectively, but in which p 1 should be replaced by p 2 (see Equation (19)), and the last two common parameters in the K -integrals should be taken successively as λ + Z and γ + 1 / 2 , similar to Equations (18) and (22).
Accordingly, the kinetic energy MEs calculated between SBFs can be represented as:
T 0 ; 0 ( j ) = k = 0 , ( 2 ) K b ˜ k 2 n , m b k 2 , n , m k = 0 , ( 2 ) K Z π k + k 2 b ˜ k 2 n , m b k 2 , n , m T ^ k p , n m ; k p , n m ( j ) , ( j = 1 , 2 )
where T ^ k p , n m ; k p , n m ( 1 ) and T ^ k p , n m ; k p , n m ( 2 ) represent the RHSs of Equations (28) and (29), respectively, but in which the parameters p 1 , λ and γ should be replaced by k 1 / 2 , Z and 1 / 2 , respectively, and the last two common parameters in the corresponding K -integrals should be taken successively as 2 Z and 1.

5. Numerical Results and Discussion

For convenience, we will call the method described above as the variational method of near the nucleus calculations (VMNN). We applied VMNN to calculate the ground state energies, E, and the corresponding WFs, Ψ , for the helium atom and a number of two-electron ions.
The results corresponding to the simplest version of the method with λ = Z and γ = 1 / 2 are presented in Table 2. Note that we call this variant “simplest” because the mentioned choice of parameters implies that both the SBF and the set of BFs (14) contain the same exponential. In the sixth column we present one of the most important characteristics of the obtained state, namely, the corresponding deviation from the virial theorem for the Coulomb interactions. The 10 M factor has been replaced with ( M ) . To confirm the main purpose of VMNN, in the last two columns we present two matrix elements of the relevant operators, which represent the main properties of WF near the nucleus and can be calculated in the following simple way:
δ ( r 1 ) = 4 π N 0 Ψ 2 ( R , 0 , 0 ) R 2 d R , δ ( r 1 ) δ ( r 2 ) = 1 N Ψ 2 ( 0 , 0 , 0 ) ,
where δ ( r ) denotes the Dirac delta function, and the normalization coefficient is:
N = Ψ 2 ( R , α , θ ) d v .
The last significant digit in energy and matrix elements (32) presented in Table 2 is the first one that differs from the more accurate calculations (see, e.g., [11,17,18,19]) for 2 Z 28 , or from the Pekeris-like method [20,21] for non-integer values of Z. The case Z = 28 is the maximum nucleus charge for which we found numerical results in the relevant scientific literature. Note that fairly accurate values of the ground state energy, especially for large Z, can be calculated using the so called 1 / Z expansion (see, e.g., [22] and references therein). In particular, all 14 digits after the decimal point for the energy values presented in Table 2 for Z 40 are in complete agreement with the corresponding results obtained using the 1 / Z expansion. In addition, the energies for any Z obtained using this approximation can be used to solve the eigenvalue Equation (3) by a very efficient “Arnoldi” method, which is also known as the “Lanczos” method.
It is well-known that the negative ion of hydrogen is the only ion in the two-electron atomic sequence that has only one bound (ground) state. In other words, there are no excited bound states in this ion. At the same time, its WF is rather diffuse. These features cause special computational difficulties when using not only variational methods, including the one presented here, or the Pekeris-like method [20,21], but also CFHHM [12,13,14]. To solve this problem, both a significantly higher calculation accuracy and a significantly longer basis size are required. It should be emphasized that the given problem concerns only a two-electron negative ion with a nuclear charge, Z, equal to exactly 1. Note that there is no mathematical (computational) method of calculating the electronic structure for a two-electron system with non-integer Z. In particular, the unrealistic case Z = 1.5 , presented in Table 2, was calculated without any specific problems. These results will be useful in solving the problem of calculating the function a 21 ( Z ) , which will be discussed below.
It is seen from representation (5) that the WF is defined, first of all, by the set of BFs, which in turn are determined by the sets of integers k , p , n , m for the BFs (14), and by integer number K for the SBF (13). To construct the large set (14) we used two characteristic numbers, k m a x and ω m a x , such that
1 < k k m a x , 0 n ω m a x , 0 m ω m a x ,
but under the additional condition
k + p + n + m max ( k m a x , ω m a x ) .
The power, p, of logarithm can vary as follows: 0 p [ k / 2 ] , where [ x ] denotes (as in the FE) the integer part of x. It is clear that all cases of p = k / 2 for even k must be excluded from the set (14), since these cases are represented by the SBF (13) with upper limit K = 2 [ k m a x / 2 ] . To reduce the length of the set (14), one can use the well-known closed form of the AFC [3,4]:
ψ 3 , 1 ( α , θ ) = Z ( π 2 ) 36 π 6 Z η ( 1 ξ 2 ) + ξ ( 5 ξ 2 6 ) .
The representation (36), together with the fact that the series expansion of the SBF (13) gives the correct coefficient Z ( 2 π ) / ( 6 π ) for ξ in the angular part for R 3 ln R , enables us to restrict BFs (14) with k , p = 3 , 1 to angular parts defined by only three sets of integers: n , m = 1 , 0 , 1 , 2 , 0 , 3 . Thus, using all the above conditions, we obtain basis sets, in particular those presented in Table 2, of length 516 and 441 for k m a x , ω m a x equal to 7 , 13 and 8 , 12 , respectively.
The following comments are needed regarding the FE parameter a 21 , calculated by VMNN and presented in Table 2.
The AFCs ψ k , p satisfy the Fock recurrence relation,
Λ 2 k ( k + 4 ) ψ k , p ( α , θ ) = h k , p ( α , θ ) ,
where the operator Λ 2 is defined by Equation (8). The specific form of the RHS h k , p ( α , θ ) , which is not important for our consideration, can be found, for example, in Ref. [3]. What really matters is that the general solution of the inhomogeneous differential Equation (37) can be expressed as the sum of a particular solution, ψ ˜ k , p , of that equation and the general solution ϕ k , p of the associated homogeneous equation
Λ 2 k ( k + 4 ) ϕ k , p ( α , θ ) = 0 ,
which can be represented as (see, e.g., [3,4])
ϕ k , p ( α , θ ) = l = 0 [ k / 2 ] a k l ( p ) Y k l ( α , θ ) ,
where Y k l are the unnormalized HHs. A particular solution, ψ ˜ k , p , can be obtained by directly solving Equation (37) under suitable boundary conditions. The contribution of ϕ k , p is determined by the coefficients a k , l ( p ) , which can only be found using the WF, Ψ ( R , α , θ ) , calculated, e.g., by VMNN.
We are interested in the case k = 2 , p = 0 which has been discussed, e.g., in Ref. [16]. For this case, Equation (39) becomes
ϕ 2 , 0 ( α , θ ) = a 20 ( 0 ) Y 20 ( α , θ ) + a 21 ( 0 ) Y 21 ( α , θ ) .
For singlet S states (including the ground state) the coefficient a 20 ( 0 ) is identically zero, since the unnormalized HH, Y 20 ( α , θ ) 2 cos α does not preserve the sign under the transformation α π α , which essentially corresponds to a permutation of electrons. Recall that the WF of a two-electron atomic system must preserve its parity with such a permutation. To calculate the only remaining non-zero coefficient, a 21 a 21 ( 0 ) , we must first obtain the angular function ϕ ˜ 20 ( α , θ ) ϕ 20 ( α , θ ) , which is a factor for R 2 in the series expansion of the WF, Ψ ( R , α , θ ) near the the nucleus ( R 0 ) , calculated by VMNN. These functions can then be used to calculate the required coefficient a 21 as follows:
a 21 = π 2 N 21 2 0 π 0 π ϕ ˜ 20 ( α , θ ) Y 21 ( α , θ ) sin 2 α sin θ d α d θ ,
where N 21 = 2 π 3 / 2 is the normalization coefficient for the HH, Y 21 ( α , θ ) = sin α cos θ . Note that the last relation represents a particular case for calculating the expansion coefficients in HHs for the general function of angles α and θ (see, e.g., [3]). It is the coefficients a 21 , computed according to Equation (41), that are presented in Table 2.
It is worth noting the following. The term ϕ 20 ( α , θ ) = a 21 sin α cos θ represents the total contribution of the HH, Y 21 ( α , θ ) into the AFC ψ 2 , 0 ( α , θ ) = ψ ^ 2 , 0 ( α , θ ) + ϕ 20 ( α , θ ) . On the other hand, in practice, all obtained particular solutions ψ ^ 2 , 0 ( α , θ ) of the corresponding Fock recurrence relation also contain an admixture of Y 21 ( α , θ ) . Therefore, before adding the components ϕ 20 ( α , θ ) and ψ ^ 2 , 0 ( α , θ ) , it is necessary to get rid of the admixture C 21 Y 21 ( α , θ ) in the latter component to obtain the correct AFC, ψ 20 ( α , θ ) . In particular, the authors of Ref. [23] were the first to obtain a particular solution ψ ^ 2 , 0 ( α , θ ) in a closed analytic form (see also, [4,10]) for the helium-like sequence. The corresponding admixture coefficient (at least for singlet S states) was obtained [24] in the form:
C 21 = Z 62 + 17 π 48 G 72 π ,
where G is Catalan’s constant.
The results for a 21 as a function of the nucleus charge, Z, are graphically presented in Figure 1. The curves corresponding to the full range, Z [ 1.5 , 100 ] , and the partial range, Z [ 1.5 , 22 ] , are shown in the left and right columns, respectively. It is seen that the function a 21 ( Z ) has two characteristic points. These are the maximum point, Z m a x 8.30613237 , and the zero point, Z 0 21.645108 , of the function. Note that, in order to obtain more accurate results, we calculated a 21 for a set of additional half-integer points, in particular for Z = 1.5 , 2.5 , 3.5 , 7.5 , 8.5 , 9.5 . It was found that a simple analytic function of the form
g ( Z ) = c 1 Z + c 0 + c 1 Z ( c 2 + ln Z ) ,
with properly chosen parameters c 1 , c 0 , c 1 and c 2 provides an excellent interpolation of the curve calculated using VMNN. Direct numerical interpolation gives a good hint for the following choice of two of the four parameters:
c 0 = 1 9 , c 1 = 2 π 3 π .
It is easy to verify that the second parameter in the last equation represents the factor for the AFC ψ 21 . The two remaining parameters can be found by solving the system of two equations, g ( Z m a x ) = 0 and g ( Z 0 ) = 0 , which, given the results (44), yields
c 1 = 0.0012024376 , c 2 = 3.117138 .
The accuracy of the resulting function, g ( Z ) , was estimated using the function log 10 1 a 21 / g ( Z ) , which characterizes the relative difference between the functions under consideration. This evaluation function is shown in the bottom row of Figure 1. It can be seen that the difference is less than a hundredth of a percent. It is important to note that all the coefficients a 21 that were used to construct Figure 1, most of which are presented in Table 2, coincide (at least to 5 significant digits) with those calculated using the AFCs computed by CFHHM.
It should be emphasized that although the above results confirm the correct behavior of the WF near the nucleus, this does not mean that the WF behaves incorrectly at sufficiently large R. On the contrary, the sufficiently high accuracy of the calculations of the energy and parameters characterizing the virial theorem tells us the opposite.

6. Conclusions

The results presented here are devoted to the development of a simple method of non-relativistic variational calculations of the electronic structure of a two-electron atom/ion in the ground state. Although we show numerical results (see Table 2), including for positive ions with a nuclear charge Z 30 , which suggest the presence of a significant relativistic component, it should be emphasized that our method is intended only for calculating energies and wave functions corresponding to the non-relativistic Hamiltonian. The results for large Z are intended only to demonstrate the accuracy of the presented method, which is fully consistent with correlation function hyperspherical harmonic [12,13,14] and Pekeris-like [20,21] methods, as well as with the 1 / Z expansion [22], which is intended solely for calculating energies. Unlike the 1 / Z expansion and some other techniques, the method described in Section 2, Section 3 and Section 4 is designed to calculate both the non-relativistic energy and the wave function, representing with high accuracy the ground state of a two-electron atom/ion, especially near the nucleus.
The method as a whole is variational, consisting of a standard solution of a generalized matrix eigenvalue equation, all matrix elements of which are reduced to the numerical calculation of one-dimensional integrals (see Appendix A).
High accuracy of the wave function near the nucleus is achieved primarily due to the following features of the calculation method.
The complete basis set is divided into two unequal parts, each of which is constructed using hyperspherical coordinates α , θ and R.
Unlike other methods that use hyperspherical harmonics Y k , l ( α , θ ) , our method instead uses powers of specific functions η 1 + sin α = ( r 1 + r 2 ) / R and ξ 1 sin α cos θ = r 12 / R (see Equation (2)) as constructive elements of the angular parts of the basis functions. This enables us to significantly reduce the basis size, which in turn simplifies and speeds up computations.
The main (in quantity, but not in importance) part of the complete basis set is represented by the basis functions of the form (14), including logarithms of the hyperspherical radius R, similar to the Fock expansion (1).
It should be emphasized that the most characteristic feature of our method is the additional inclusion of a special basis function of the form (13) in the complete basis set. It is this basis function, involving the leading angular Fock coefficients, that ensures the correct behavior of the wave function near the nucleus, including the guarantee of satisfying the Kato’s cusp conditions [15].
To test the method under consideration, we calculated the four most characteristic parameters presented in Table 2 for various charges Z of the nuclei. These are the non-relativistic energy, E, and the deviation from the virial theorem, V / T + 2 , demonstrating the high accuracy of the method, as well as the matrix elements of the operators δ ( r 1 ) and δ ( r 1 ) δ ( r 2 ) , demonstrating the high accuracy of the wave function near the nucleus in particular.
The correct behavior of the wave function at all points of the configuration space enabled us to calculate the Fock expansion coefficient a 21 (see Equation (41)) as a function of the nucleus charge Z. The corresponding results presented in Table 2 are in full agreement with the correlation function hyperspherical harmonic method [12,13,14], giving the most accurate results related to the behavior of the wave function near the nucleus. A plot of the function a 21 ( Z ) , together with a very simple interpolating function, g ( Z ) (see Equations (43)–(45)), are shown in Figure 1.

Author Contributions

Conceptualization, E.Z.L.; methodology, E.Z.L.; software, E.Z.L. and R.K.; validation, E.Z.L.; writing—original draft, E.Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Acknowledgments

We would like to thank Lorenzo Lodi for a number of helpful comments.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
FEFock expansion
WFwave function
HSChyperspherical coordinates
AFCangular Fock coefficient
HHhyperspherical harmonic
MEmatrix element
BFbasis function
SBFspecial basis function
RHSright hand side
CFHHMcorrelation function hyperspherical harmonic method
VMNNvariational method of near the nucleus (calculations)

Appendix A

It was shown in Section 3 and Section 4 that all matrix elements needed to solve the matrix Equation (3) can be expressed in term of 3D integrals of the form (17). The purpose of this section is to show how the 3D integrals under consideration can be reduced to 1D integrals, which leads to increased accuracy and a reduction in computation time by hundreds of times.
The first step is to integrate over R using the relation:
0 R k + 5 ( ln R ) p exp ( β R ) d R = β k 6 t = 0 p p t C p t ( k ) ( ln β ) t ,
where the coefficients
C s ( k ) = d s Γ ( k + 6 ) d k s
are expressed in terms of the derivatives of the gamma function.
Substituting Equation (A1) with β = a η b ξ into definition (17), we obtain:
K L ( k , p , n , m ; a , b ) = 2 0 π / 2 η n ( sin α ) L t = 0 p ( 1 ) t p t C p t ( k ) J t ( α ) d α ,
where
J t ( α ) = 0 π ξ m ln t ( a η b ξ ) ( a η b ξ ) k + 6 sin θ d θ = 2 sin α ζ η x m + 1 ln t ( a η b x ) ( a η b x ) k + 6 d x .
Function η η ( α ) is defined by Equation (10) and
ζ ζ ( α ) = 1 sin α .
When deriving (A3), it was taken into account that the α -dependence of the integrand on the RHS of this equation is represented only by sin α , which determines the symmetry with respect to the point π / 2 .
The next step is to represent the integral (A4) as:
J t ( α ) = 2 sin α P k + 6 , m + 1 , t ( a , b , α , η ) P k + 6 , m + 1 , t ( a , b , α , ζ ) ,
where
P k , m , t ( a , b , α , ρ ) = 0 ρ x m ln t ( a η b x ) ( a η b x ) k d x .
The representation (A6)–(A7) enables us to apply the general relation (see, e.g., [25])
0 a x μ 1 ( a x ) ν 1 ln n ( c x + d ) ( c x + d ) r d x = a μ + ν 1 B ( μ , ν ) d r × q = 0 n ( 1 ) q + n n q ( ln d ) q n q r n q F 1 2 r , μ ; μ + ν ; a c d , [ a , Re μ , Re ν > 0 ]
to obtain the closed analytic form for the integral (A7). B ( μ , ν ) and F 1 2 ( ) denote the Euler beta function and the Gauss hypergeometric function, respectively. Setting μ = m + 1 , ν = 1 , n = t , r = k , a = ρ , c = b , d = a η in (A8) and simplifying, we obtain the following explicit representation for the integral (A7):
P k , m , t ( a , b , α , ρ ) = ρ m + 1 ( m + 1 ) ( a η ) k q = 0 t ( 1 ) q + t t q ln q ( a η ) F k , m , t q ; b ρ a η ,
where we introduced the notation
F ( k , m , n ; x ) = lim r k n r n F 1 2 r , m + 1 ; m + 2 ; x .
It is necessary to note that problems arise when it is necessary to calculate (directly) the Gauss hypergeometric function (and/or its derivatives in respect of the first parameter) from Equation (A10) with an integer first parameter. Thus, the next step is to obtain the trouble-free explicit representation for the function (A10). Using representation
F 1 2 r , m + 1 ; m + 2 ; x = ( m + 1 ) x ( m + 1 ) B x ( m + 1 , 1 r )
for the hypergeometric function, and the integral representation
B x ( a , b ) = 0 x t a 1 ( 1 t ) b 1 d t
for the incomplete beta function, we can give the following sequential derivation of the integral representation for the function (A10):
F ( k , m , n ; x ) = ( m + 1 ) x ( m + 1 ) lim r k 0 x t m n r n ( 1 t ) r d t = ( m + 1 ) x m + 1 0 x t m [ ln ( 1 t ) ] n ( 1 t ) k d t = ( m + 1 ) x m + 1 1 x 1 y k ( 1 y ) m ( ln y ) n d y = ( m + 1 ) x m + 1 q = 0 m ( 1 ) q m q 1 x 1 y q k ( ln y ) n d y .
Using the explicit result of the last integration, we finally obtain:
F ( k , m , n ; x ) = ( m + 1 ) x m + 1 q = 0 m ( 1 ) q m q w k , n ( x ; q ) ,
where
w k , n ( x ; q = k 1 ) = [ ln ( 1 x ) ] n + 1 n + 1 ,
w k , n ( x ; q k 1 ) = n ! ( q + 1 k ) n + 1 1 ( 1 x ) q + 1 k s = 0 n ( k q 1 ) s ln s ( 1 x ) s ! .
Results (A13)–(A16) hold for non-negative integers m and n, real x and integers k.
Thus, we have reduced the 3D integral (17) to a 1D integral defined by Equations (A3), (A6), (A9) and (A14)–(A16). It can be seen (see Section 3 and Section 4) that only the K L ( ) integrals (A3) with L = 1 and L = 2 are used in the VMNN. It is easy to show that these integrals are related as follows:
K 2 ( k , p , n , m ; a , b ) = K 1 ( k , p , n + 2 , m ; a , b ) K 1 ( k , p , n , m ; a , b ) .
It is also useful to note that both the accuracy and the calculation time depend significantly on the choice of the method of integration. We used Wolfram Mathematica for all our calculations. In particular, to calculate the integrals (A3) we used the “GaussBerntsenEspelidRule” method, which significantly improves the accuracy and reduces the calculation time. Gaussian quadrature uses optimal sampling points (through polynomial interpolation) to form a weighted sum of the integrand values over these points. On a subset of these sampling points a lower order quadrature rule can be made. The difference between the two rules can be used to estimate the error. Berntsen and Espelid derived error estimation rules by removing the central point of Gaussian rules with an odd number of sampling points.
An additional remark concerns the behavior of the integrand on the RHS of Equation (A3) as hyperspherical angle α approaches zero. It is clear that η = ξ = 1 for α = 0 . Therefore, it follows from representation (A6) that the integrand on the RHS of Equation (A3) approaches zero as α 0 . The last statement is true, at least for the cases L = 1 , 2 (see Section 3 and Section 4) we are interested in. The problem is that, according to (A6), such an integrand represents the difference between two very close quantities. There are two solutions to the problem. The first is to apply high-precision calculations, at least in the vicinity of α = 0 . The second is to use a series expansion for the RHS of Equation (A6) near α = 0 .

References

  1. Fock, V.A. On the Schrödinger Equation of the Helium Atom. Izv. Akad. Nauk SSSR Ser. Fiz. 1954, 18, 161–174. [Google Scholar]
  2. Fock, V.A. Selected Works: Quantum Mechanics and Quantum Field Theory; Fadeev, L.D., Khalfin, L.A., Komarov, I.V., Eds.; Chapman & Hall/CRC: Boca Raton, FL, USA, 2004; p. 525. [Google Scholar]
  3. Abbott, P.C.; Maslen, E.N. Coordinate systems and analytic expansions for three-body atomic wavefunctions: I. Partial summation for the Fock expansion in hyperspherical coordinates. J. Phys. A Math. Gen. 1987, 20, 2043–2075. [Google Scholar] [CrossRef]
  4. Liverts, E.Z.; Barnea, N. Angular Fock coefficients. Refinement and further development. Phys. Rev. A 2015, 92, 042512. [Google Scholar] [CrossRef]
  5. Liverts, E.Z. Analytic calculation of the edge components of the angular Fock coefficients. Phys. Rev. A 2016, 94, 022504. [Google Scholar] [CrossRef]
  6. Liverts, E.Z.; Krivec, R. Fock Expansion for Two-Electron Atoms: High-Order Angular Coefficients. Atoms 2022, 10, 135. [Google Scholar] [CrossRef]
  7. Morgan, J.D., III. Convergence properties of Fock’s expansion for S-state eigenfunctions of the helium atom. Theor. Chim. Acta 1986, 69, 181–223. [Google Scholar] [CrossRef]
  8. Nakashima, H.; Nakatsuji, H. Solving the Schrödinger equation for helium atom and its isoelectronic ions with the free iterative complement interaction (ICI) method. J. Chem. Phys. 2007, 127, 224104. [Google Scholar] [CrossRef]
  9. Rodriguez, K.V.; Gasaneo, G.; Mitnik, D.M. Accurate and simple wavefunctions for the helium isoelectronic sequence with correct cusp conditions. J. Phys. B 2007, 40, 3923–3939. [Google Scholar] [CrossRef]
  10. Forrey, R.C. Compact representation of helium wave functions in perimetric and hyperspherical coordinates. Phys. Rev. A 2004, 69, 022504. [Google Scholar] [CrossRef]
  11. Drake, G.W.F. High Precision Calculations for Helium. In Atomic, Molecular, and Optical Physics Handbook; Drake, G.W.F., Ed.; AIP Press: New York, NY, USA, 1996. [Google Scholar]
  12. Haftel, M.I.; Mandelzweig, V.B. Exact Solution of Coupled Equations and the Hyperspherical Formalism: Calculation of Expectation Values and Wavefunctions of Three Coulomb-Bound Particles. Ann. Phys. 1983, 150, 48–91. [Google Scholar] [CrossRef]
  13. Haftel, M.I.; Mandelzweig, V.B. Fast Convergent Hyperspherical Harmonic Expansion for Three-Body Systems. Ann. Phys. 1989, 189, 29–52. [Google Scholar] [CrossRef]
  14. Haftel, M.I.; Krivec, R.; Mandelzweig, V.B. Power Series Solution of Coupled Differential Equations in One Variable. J. Comp. Phys. 1996, 123, 149–161. [Google Scholar] [CrossRef][Green Version]
  15. Kato, T. On the eigenfunctions of many-particle systems in quantum mechanics. Commun. Pure Appl. Math. 1957, 10, 151–177. [Google Scholar] [CrossRef]
  16. Myers, C.R. Fock’s expansion, Kato’s cusp conditions, and the exponential ansatz. Phys. Rev. 1991, A9, 5537–5546. [Google Scholar] [CrossRef]
  17. Frolov, A.M. Exponential representation in the Coulomb threebody problem. J. Phys. B 2004, 37, 2917–2932. [Google Scholar] [CrossRef]
  18. Frolov, A.M. Field shifts and lowest order QED corrections for the ground 11S and 23S states of the helium atoms. J. Chem. Phys. 2007, 126, 104302. [Google Scholar] [CrossRef] [PubMed]
  19. Frolov, A.M. On the Q-dependence of the lowest-order QED corrections and other properties of the ground 11S-states in the two-electron ions. Chem. Phys. Let. 2015, 638, 108–115. [Google Scholar] [CrossRef]
  20. Liverts, E.Z.; Barnea, N. S-states of helium-like ions. Comput. Phys. Comm. 2011, 182, 1790–1795. [Google Scholar] [CrossRef]
  21. Liverts, E.Z.; Barnea, N. Three-body systems with Coulomb interaction. Bound and quasi-bound S-states. Comput. Phys. Comm. 2013, 184, 2596–2603. [Google Scholar] [CrossRef]
  22. Lopez Vieyra, J.C.; Turbiner, A.V. On 1/Z expansion for two-electron systems. arXiv 2013, arXiv:1309.2707. [Google Scholar]
  23. Gottschalk, J.E.; Maslen, E.N. Coordinate systems and analytic expansions for three-body atomic wavefunctions: III. Derivative continuity via solution to Laplace’s equation. J. Phys. A Math. Gen. 1987, 20, 2781–2803. [Google Scholar] [CrossRef]
  24. Liverts, E.Z. Two-particle atomic coalescences: Boundary conditions for the Fock coefficient components. Phys. Rev. A 2016, 94, 022506. [Google Scholar] [CrossRef]
  25. Prudnikov, A.P.; Brychkov, Y.A.; Marichev, O.I. Integrals and Series; Volume 1: Elementary Functions; Gordon and Breach Science Publishers: New York, NY, USA, 1986. [Google Scholar]
Figure 1. The Fock expansion coefficient, a 21 , as a function of the nucleus charge, Z, is presented in the top row, and the accuracy estimate of the analytic interpolation function, g ( Z ) , defined by Equations (43)–(45) is shown in the bottom row of the graph.
Figure 1. The Fock expansion coefficient, a 21 , as a function of the nucleus charge, Z, is presented in the top row, and the accuracy estimate of the analytic interpolation function, g ( Z ) , defined by Equations (43)–(45) is shown in the bottom row of the graph.
Atoms 12 00069 g001
Table 1. Factor b ˜ k and components of coefficients b k , n , m = q ˜ k , n , m j = 0 J k q k , n , m ( j ) π j ( 0 k k / 2 5 ), representing the polynomials in η and ξ according to Equation (12). The upper limit of summation is defined as J k = k 2 for k 2 and J k = 0 for k < 2 .
Table 1. Factor b ˜ k and components of coefficients b k , n , m = q ˜ k , n , m j = 0 J k q k , n , m ( j ) π j ( 0 k k / 2 5 ), representing the polynomials in η and ξ according to Equation (12). The upper limit of summation is defined as J k = k 2 for k 2 and J k = 0 for k < 2 .
k b ˜ k N o ̲ nm q ˜ k , n , m q k , n , m ( 0 ) q k , n , m ( 1 ) q k , n , m ( 2 ) q k , n , m ( 3 )
5 ( 2 π ) ( 5 π 14 ) / 100 1 1 089 387 008 240 1 047 586 705 720 335 763 343 914 35 868 493 935
5012817232500000202 15 4 286 509 281 040 4 129 722 511 880 1 326 134 522 774 141 939 797 505
304 60 5 335 564 976 800 5 141 563 833 740 1 651 435 721 371 176 799 081 450
406 40 12 056 179 167 440 11 618 550 815 440 3 732 041 258 007 399 570 465 060
508 15 18 776 793 358 080 18 095 537 797 140 5 812 646 794 643 622 341 848 670
6010 3 18 776 793 358 080 18 095 537 797 140 5 812 646 794 643 622 341 848 670
780 5 31 900 530 538 240 30 758 808 723 020 9 885 507 900 549 1 058 977 460 610
882 5 31 900 530 538 240 30 758 808 723 020 9 885 507 900 549 1 058 977 460 610
960 20 31 900 530 538 240 30 758 808 723 020 9 885 507 900 549 1 058 977 460 610
1062 20 31 900 530 538 240 30 758 808 723 020 9 885 507 900 549 1 058 977 460 610
1120 160 967 277 327 240 932 324 864 590 299 527 103 049 32 074 335 000
1222 120 9 531 787 759 840 9 188 929 121 180 2 952 640 515 759 316 236 217 230
1324 60 24 726 253 970 560 23 837 487 905 180 7 659 813 135 081 820 411 311 690
1426 20 24 726 253 970 560 23 837 487 905 180 7 659 813 135 081 820 411 311 690
1540 20 28 031 421 229 280 2 702 950 926 466 8 687 399 488 353 930 680 120 610
1642 160 413 145 907 340 399 002 669 935 128 448 294 159 13 783 601 115
1744 30 24 726 253 970 560 23 837 487 905 180 7 659 813 135 081 820 411 311 690
1846 10 24 726 253 970 560 23 837 487 905 180 7 659 813 135 081 820 411 311 690
4 ( π 2 ) ( 5 π 14 ) / 100 12 483 896 308 420 49 095
15431472000220 288 1 201 904 775 505 125 073
340 4 48 803 648 31 601 320 5 117 187
460 4 92 072 192 59 519 500 9 619 815
580 1 92 072 192 59 519 500 9 619 815
602 96 1 685 800 1 083 925 174 168
704 12 37 545 376 24 164 696 3 886 989
806 12 30 802 176 19 828 996 3 190 317
908 3 30 802 176 19 828 996 3 190 317
1022 24 46 119 680 29 773 780 4 804 833
1124 12 46 119 680 29 773 780 4 804 833
1242 12 46 119 680 29 773 780 4 804 833
1344 6 46 119 680 29 773 780 4 804 833
3 ( 2 π ) ( 5 π 14 ) / 100 2 106 33
170100220 2 2064 675
340 1 2064 675
402 4 609 195
504 3 1112 357
606 1 1112 357
722 2 2064 675
842 1 2064 675
2 ( π 2 ) ( 5 π 14 ) / 100 1 1
180220 1 4
340 1 2
402 1 4
504 1 2
1 ( π 2 ) / 3 100 1 1
202 1 1
01100 1 1
Table 2. Absolute values of the ground state energies, E, and the corresponding FE coefficients a 21 calculated by VMNN with basis size N b . The sixth column presents an estimate of the deviation from the virial theorem. The two relevant matrix elements are presented in the last two columns.
Table 2. Absolute values of the ground state energies, E, and the corresponding FE coefficients a 21 calculated by VMNN with basis size N b . The sixth column presents an estimate of the deviation from the virial theorem. The two relevant matrix elements are presented in the last two columns.
Atom/IonZ a 21 N b | E | V / T + 2 δ ( r 1 ) δ ( r 1 ) δ ( r 2 )
1.5 0.382385161.465 279 051 825 7404.6 (−13)0.6821157220.20291
He20.476755162.903 724 377 034 1198.7 (−15)1.81042931841.86874
Li+30.622855167.279 913 412 669 3064.1 (−17)6.85200943733.32118
Be2+40.7278751613.655 566 238 423 5871.4 (−16)17.198172547231.0389
B3+50.8023251622.030 971 580 242 781−1.4 (−16)34.75874366996.0099
C4+60.8524051632.406 246 601 898 530−1.3 (−16)61.443578053221.850
N5+70.8822144144.781 445 148 772 7035.3 (−16)99.16253458596.074
O6+80.8946644159.156 595 122 757 9243.6 (−16)149.825472619,974.31
F7+90.8919344175.531 712 363 959 4902.6 (−16)215.342252041,827.47
Ne8+100.8757244193.906 806 515 037 5482.0 (−16)297.622730880,761.89
Na9+110.84738441114.281 883 776 072 7211.5 (−16)398.5767736146,112.4
Mg10+120.80801441136.656 948 312 646 9291.2 (−16)520.1142308250,608.2
Al11+130.75854441161.032 003 026 058 3591.0 (−16)664.1449752411,112.2
Si12+140.69974441187.407 049 998 662 9258.5 (−17)832.5788588649,432.6
P13+150.63229441215.782 090 763 537 1597.2 (−17)1027.325721993,207.9
S14+160.55675441246.157 126 474 254 7386.2 (−17)1250.295442147,686.5 (+1)
Cl15+170.47364441278.532 158 015 400 0945.3 (−17)1503.397895214,264.8 (+1)
Ar16+180.38340441312.907 186 076 611 1484.7 (−17)1788.543432304,172.8 (+1)
K17+190.28642441349.282 211 203 453 1664.1 (−17)2107.640384423,536.9 (+1)
Ca18+200.18307441387.657 233 833 158 5553.7 (−17)2462.600113579,618.9 (+1)
Sc19+210.07366441428.032 254 320 234 6903.3 (−17)2855.332029780,947.6 (+1)
Ti20+22−0.04152441470.407 272 955 138 3832.9 (−17)3287.745943103,745.9 (+2)
V21+23−0.16221441514.782 289 978 111 7732.7 (−17)3761.751772136,064.1 (+2)
Cr22+24−0.28816441561.157 305 589 581 2712.4 (−17)4279.259269176,369.1 (+2)
Mn23+25−0.41916441609.532 319 958 075 7452.2 (−17)4842.178389226,167.5 (+2)
Fe24+26−0.55501441659.907 333 226 327 8042.0 (−17)5452.418955287,169.7 (+2)
Co25+27−0.69552441712.282 345 516 026 5501.8 (−17)6111.890859361,307.4 (+2)
Ni26+28−0.84052441766.657 356 931 557 0991.7 (−17)6822.503872450,752.1 (+2)
Zn28+30−1.14334441881.407 377 488 360 6051.5 (−17)8404.792902685,562.3 (+2)
Zr38+40−2.881334411575.157 449 525 559 447.8 (−18)20,034.060392,478.5 (+3)
Sn48+50−4.925374412468.907 492 812 711 764.9 (−18)39,260.262151,406.5 (+4)
Nd58+60−7.213314413562.657 521 697 319 283.3 (−18)67,993.257455,483.1 (+4)
Yb68+70−9.704064414856.407 542 342 006 262.4 (−18)108,142.90115,468.8 (+5)
Hg78+80−12.36854416350.157 557 832 474 811.8 (−18)161,619.06258,314.5 (+5)
Th88+90−15.18464418043.907 569 884 711 271.4 (−18)230,331.59525,303.7 (+5)
Fm98+100−18.13574419937.657 579 529 067 141.1 (−18)316,190.36990,905.1 (+5)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liverts, E.Z.; Krivec, R. Two-Electron Atomic Systems—A Simple Method for Calculating the Ground State near the Nucleus: Some Applications. Atoms 2024, 12, 69. https://doi.org/10.3390/atoms12120069

AMA Style

Liverts EZ, Krivec R. Two-Electron Atomic Systems—A Simple Method for Calculating the Ground State near the Nucleus: Some Applications. Atoms. 2024; 12(12):69. https://doi.org/10.3390/atoms12120069

Chicago/Turabian Style

Liverts, Evgeny Z., and Rajmund Krivec. 2024. "Two-Electron Atomic Systems—A Simple Method for Calculating the Ground State near the Nucleus: Some Applications" Atoms 12, no. 12: 69. https://doi.org/10.3390/atoms12120069

APA Style

Liverts, E. Z., & Krivec, R. (2024). Two-Electron Atomic Systems—A Simple Method for Calculating the Ground State near the Nucleus: Some Applications. Atoms, 12(12), 69. https://doi.org/10.3390/atoms12120069

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop