Next Article in Journal
Spatially Resolved Precision Measurement of Magnetic Field Using Ultracold Cesium Atoms as Sensors
Previous Article in Journal
Properties of a Static Dipolar Impurity in a 2D Dipolar BEC
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Statistics of Quantum Numbers for Non-Equivalent Fermions in Single-j Shells

by
Jean-Christophe Pain
1,2
1
CEA, DAM, DIF, F-91297 Arpajon, France
2
Université Paris-Saclay, CEA, Laboratoire Matière en Conditions Extrêmes, F-91680 Bruyères-le-Châtel, France
Atoms 2025, 13(4), 25; https://doi.org/10.3390/atoms13040025
Submission received: 27 February 2025 / Revised: 20 March 2025 / Accepted: 24 March 2025 / Published: 25 March 2025
(This article belongs to the Section Atomic, Molecular and Nuclear Spectroscopy and Collisions)

Abstract

:
This work addresses closed-form expressions for the distributions P ( M ) of the magnetic quantum numbers M and Q ( J ) of total angular momentum J for non-equivalent fermions in single-j orbits. Such quantities play an important role in both nuclear and atomic physics, through the shell models. Using irreducible representations of the rotation group, different kinds of formulas are presented, involving multinomial coefficients, generalized Pascal triangle coefficients, or hypergeometric functions. Special cases are discussed, and the connections between P ( M ) (and therefore Q ( J ) ) and mathematical functions such as elementary symmetric, cyclotomic, and Jacobi polynomials are outlined.

1. Introduction

The problem of determining the number of states corresponding to a given ( J , M ) pair, where J represents the magnitude of the total angular momentum operator and M is the eigenvalue (in units of ) of its projection along the z-axis ( J M J ) in the case of indistinguishable particles, was first addressed by Bethe in 1936 for nuclear systems [1]. The total angular momentum of N fermions
J = i = 1 N j [ i ]
is the addition of the intrinsic angular momenta j [ i ] of each fermion. The one-electron states are defined by | j [ i ] m [ i ] , where j [ i ] m [ i ] j [ i ] . Multielectron states of the ion are represented by | J M . The notations Q ( J ) and P ( M ) are introduced to represent the number of levels and states, respectively, with the same value of J (or M) in common. The study of the distribution of quantum number J in an atomic system is somewhat cumbersome, in particular because J is not the eigenvalue of any simple operator. Furthermore, there is no simple rule on J to account for the Pauli exclusion principle in configurations with more than two equivalent fermions. It is usually more convenient to study the distribution of the projection M of J, because M is the eigenvalue of the one-fermion operator J z and is thus additive with respect to the magnetic quantum numbers of all the electrons:
M = i = 1 N m [ i ] .
The Pauli exclusion principle can be handled through the combinatorics, by searching for all the possibilities to populate the one-electron states (Figure 1 shows an example of a state of a single-j shell ( 9 / 2 ) 4 corresponding to M = 3 ). For a general configuration made of r shells
j 1 N 1 j 2 N 2 j r N r ,
the total degeneracy is
g = k = 1 r 2 j k + 1 N k .
The single-j shell (corresponding to r = 1 and j 1 = j ) is a model system and a cornerstone of nuclear structure studies [2]. Many aspects of this paradigm have been investigated, such as symmetries [3], two-body random Hamiltonians [4,5], isospin relations, and the J-pairing interaction (see, for instance, Refs. [6,7]). The enumeration of M-states and J-levels has been the subject of numerous studies over the years within the context of nuclear physics. For example, in a study on the quantum Hall effect [8], Ginocchio and Haxton derived a simple formula for Q ( 0 ) in the case of j 4 , which is also equal to Q ( j ) for j 3 . Zhao and Arima found empirical formulas for Q ( J ) for three, four, and five particles [9] for specific J values. Zamick and Escuderos interpreted the Ginocchio–Haxton formula using a combinatorial approach for J = j with n = 3 [10], and Talmi derived a recursion relation for Q ( J ) of n fermions in a j orbit in terms of k n fermions in a ( j 1 ) orbit [11]. In Refs. [12,13,14], the authors extended the studies for n = 3 and n = 4 to determine the number of states with a given spin J and isospin T. The number of states for a given spin was found to be closely related to the sum rules of many six-j and nine-j symbols, as well as coefficients of fractional parentage [15,16,17,18,19,20]. In particular, Zamick and Escuderos identified a relationship between coefficients of fractional parentage obtained from the principal-parent method and from a seniority classification. They applied this relationship to Redmond’s recursion relation formula [21,22], transformed it to the seniority scheme, and used it to determine the number of spin-J states for J = j (a result previously obtained by Zhao and Arima [9]). Bao et al. derived recursive formulas by induction with respect to n and j and applied them to systems of two, three, and five identical particles [23]. A couple of years ago, we proposed new recursion relations for P ( M ) obtained using generating functions [24,25]. We also studied the statistical properties of the distributions using the cumulants [26,27], as well as the odd–even staggering (the number of odd-J states is larger than the number of even-J states) [25]. In addition, it is worth mentioning that we obtained analytical expressions of the total number of J levels for identical particles in a single-j shell using coefficients of fractional parentage [28] for j 3 and j 4 .
The enumeration of M-states and J-levels is also of fundamental and practical interest for atomic physics and the statistical modeling of hot-plasma atomic spectral properties, such as emission and absorption spectra (see for instance [29,30] and references therein) or the Auger effect [31]. It is not so common for the same model to apply to both nuclear and atomic physics.
Obtaining general analytical expressions for P ( M ) and Q ( J ) is a very difficult task. In Ref. [32], we derived analytical expressions in the case of three, four, and five fermions (i.e., configurations j 3 , j 4 and j 5 ) using recursion relations. Very recently, we developed a general method yielding closed-form expressions of P ( M ) , whatever the number of fermions (electrons in the case of atomic physics, neutrons and protons in the framework of nuclear physics), as linear combinations of piecewise polynomials and congruences, and presented illustrations of up to six fermions [33]. The distributions P ( M ) and Q ( J ) for four fermions in a shell j = 9 / 2 are displayed in Figure 2 and Figure 3, respectively. The corresponding analytical expressions for j 4 , obtained in Ref. [32], are recalled in Appendix A.
In the present work, we focus on a particular case, a configuration made of n single-j shells, with each shell being populated with a single fermion and each having its own value of j. In such a situation, we could derive two analytical expressions, the first one involving multinomial coefficients and the second one, more compact and more satisfactory from a practical point of view, based on generalized Pascal triangles, i.e., products of simple binomial coefficients.
In Section 2, the distributions of angular momentum J and its projection M in a general configuration made of single-j shells are introduced, the Sunko algorithm is explained and illustrated [34,35], and useful expressions in terms of Gaussian polynomials (or q−binomial coefficients) [36], symmetric functions and cyclotomic polynomials [37] are outlined. In Section 3, integral expressions of P ( M ) and Q ( J ) in the case of a configuration made of non-equivalent fermions in single-j orbits (corresponding to j k = j and N k = 1 , k [ 1 , r ] in Equation (2)) [38,39,40] are derived in the framework of group theory, using irreducible representations of SO ( 3 ) (or of the Lie group SU ( 2 ) ; see Appendix B), and exact expressions involving multinomial coefficients are deduced. In Section 4, exact expressions in terms of generalized Pascal triangle coefficients are given. Special cases for j = 1 / 2 are presented in Section 5 in terms of hypergeometric functions, and sum rules over angular momentum J are given in Section 6. Expressions for the total number of levels in a configuration and possibilities offered by the Molien functions are explained in Appendices Appendix C and Appendix D, respectively.

2. Characterization of the Magnetic Quantum Number Distribution

2.1. Generalities

Since the quantum number J is the eigenvalue of no simple operator, its mathematical study is tedious. Therefore, it is more appropriate to study the distribution of the magnetic quantum number M. The number Q ( J ) of levels with angular momentum J in a configuration can be obtained from the M values by means of
Q ( J ) = M = J J + 1 ( 1 ) J M P M = P M = J P M = J + 1 ,
where P represents the distribution of the angular momentum projection M. Let us consider a system of N identical fermions in a configuration consisting of a single orbital j (which is a half-integer) of degeneracy g = 2 j + 1 , with m i being the angular momentum projection of electron state i ( m 1 = j ,   m 2 = j + 1 ,   m 3 = j + 2 , ,   m g 1 = j 1 ,   m g = j ). The following constraint must be satisfied:
N = n 1 + + n g = i = 1 g n i ,
where n i is the number of fermions in state i ( n i = 0 or 1 i ). P ( M ) is the number of N-electron states such as
M = n 1 m 1 + + n g m g = i = 1 g n i m i .
The latter condition can be formulated in an alternative way: if we have N fermions of spin j, the number of possible states with total projection M is equal to the number of all arrangements of length 2 j + 1 for which
m 1 + m 2 + m N = M ,
where m i ( m i < m i + 1 ) are the single-particle projections ranging from j to j. The maximum total angular momentum is
J max = m = j N + 1 j m = ( 2 j + 1 N ) N / 2
and the minimum angular momentum J min is 0 if N is even and 1 / 2 if N is odd.

2.2. Binomial Coefficients and the Sunko Algorithm

Let us consider a single-j shell populated by N fermions (in other words, a configuration j N ). In this case, P ( M ) is in fact the coefficient of q M in
q N 2 ( 2 j + 1 N ) 2 j + 1 N q = q J max 2 j + 1 N q ,
where
2 j + 1 N q = q 2 j + 1 1 q 2 j + 1 q q 2 j + 1 q 2 q 2 j + 1 q N 1 q N 1 q N q q N q 2 q N q N 1 .
One also has (see Ref. [26], Equations (3.1)–(3.4))
P ( M ) = 1 ( J max + M ) ! d J max + M d q J max + M q 2 j + 1 1 q 2 j 1 q 2 j + 2 N 1 q N 1 q N 1 1 q 1 q = 0 ,
which amounts to the derivative of a rational fraction. Since Equation (6) is not very convenient in practice, Sunko and Srvtan suggested writing (omitting the factor q J max and replacing P ( M ) by P ( q ) ) [34,35]
P ( q ) = m 0 c m q m = exp k 1 1 k p k q k ,
where the c m and p k are connected by the determinantal identity [41]
c m = 1 m ! p 1 1 0 0 0 0 p 2 p 1 2 0 0 p 3 p 2 p 1 3 0 0 p 1 ( m + 2 ) 0 p m 1 p m 2 p m 3 p 2 p 1 ( m + 1 ) p m p m 1 p m 2 p 3 p 2 p 1 ,
or equivalently
m c m = p 1 c m 1 + p 1 c m 1 + + p m 1 c 1 + p m .
The point is that it is easy to obtain a closed-form expression for the p k s of a Gaussian polynomial because its logarithm is a simple expression. Using the expansion of ln ( 1 x ) , one obtains (setting p = 2 j + 1 and r = N )
p m = s = 1 , s | m min ( r , p r ) s s = max ( r , p r ) + 1 , s | m p s ,
where s | m means that s is divided by m, and we have
c m = 1 m p m + q = 1 m 1 c q p m q
with c 0 = 1 and c k = 0 for k < 0 . If we consider the example ( 7 / 2 ) 3 : j = 7 / 2 and N = 3 , corresponding to p = 2 j + 1 = 8 and r = N = 3 , the maximum value of J is J max = 15 / 2 and therefore
p m = s = 1 , s | m 3 s s = 6 , s | m 8 s .
The parameters p k and c k are given in Table 1 and the values of Q ( J ) are in Table 2. For instance, if k = 4 ,
p 4 = s = 1 , s | 4 3 s s = 6 , s | 4 8 s .
As concerns the first sum in Equation (9), the only divisors of 4 between 1 and 3 are 1 and 2. This sum is equal to 1 + 2 = 3. The second sum in Equation (9) is equal to 0, because there are no divisors of 4 between 6 and 8. Thus, p 4 = 3 0 = 3 . In the same way, for k = 7 , one has to calculate
p 7 = s = 1 , s | 7 3 s s = 6 , s | 7 8 s .
As concerns the first sum in Equation (10), the only divisor of 7 between 1 and 3 is 1. This sum is equal to 1. The second sum in Equation (10) is equal to 7, because the only divisor of 7 between 6 and 8 is 7. Thus, p 7 = 1 7 = 6 . Once the p k is known, the c k can be obtained by the recurrence relation (8).
Considering the example ( 7 / 2 ) 3 ( 5 / 2 ) 2 ( 1 / 2 ) 1 ( 9 / 2 ) 2 , one has J max = 20 and
p m = s = 1 , s | m 3 s s = 6 , s | m 8 s + s = 1 , s | m 2 s s = 5 , s | m 6 s + s = 1 , s | m 1 s s = 2 , s | m 2 s + s = 1 , s | m 2 s s = 9 , s | m 10 s .
The parameters p k and c k are listed in Table 3 and Table 4. The corresponding values of Q ( J ) are displayed in Table 5 and Table 6.
The last example ( 7 / 2 ) 3 ( 3 / 2 ) 2 ( 5 / 2 ) 1 ( 9 / 2 ) 2 corresponds to J max = 43 / 2 and
p m = s = 1 , s | m 3 s s = 6 , s | m 8 s + s = 1 , s | m 2 s s = 3 , s | m 4 s + s = 1 , s | m 2 s s = 5 , s | m 6 s + s = 1 , s | m 2 s s = 9 , s | m 10 s ,
yielding the p k and c k parameters of Table 7, Table 8 and Table 9. The Q ( J ) values are in Table 10 and Table 11.
This algorithm does not yield an analytic expression of P ( M ) or Q ( J ) , but is used as our reference calculation.

2.3. Connection with Elementary Symmetric Functions

Since we have
i = 1 2 j + 1 1 + q i t = k = 0 2 j + 1 t k e k q , q 2 , q 3 , , q 2 j + 1 ,
where e k is the kth elementary symmetric function, e.g.,
e 2 ( x 1 , x 2 , x 3 ) = x 1 x 2 + x 2 x 3 + x 1 x 3
and
i = 1 2 j + 1 1 + q i t = k = 0 2 j + 1 t k q k ( k + 1 ) / 2 2 j + 1 k q ,
we obtain
q k ( k + 1 ) / 2 2 j + 1 k q = e k q , q 2 , q 3 , , q 2 j + 1 ,
i.e., P ( M ) is the coefficient of q M in
q N ( j + 1 ) e N q , q 2 , q 3 , , q 2 j + 1 .

2.4. Link with Cyclotomic Polynomials

The binomial coefficient can be expressed in terms of cyclotomic polynomials Φ n as
n k q = d = 2 n Φ d ( q ) n / d k / d ( n k ) / d ,
where x denotes the integer part of x and
Φ n ( q ) = 1 k < n , ( k , n ) = 1 q ζ k with ζ = e 2 i π n ,
where ( k , n ) = 1 means that k and n are relatively prime to each other. One also has
Φ n ( x ) = d | n x d 1 μ n d ,
where μ represents the Möbius function:
μ ( n ) = 1 if n = 1 , ( 1 ) k if n square free and k number of prime divisors of n , 0 otherwise .
Using the formula for the multiple derivative of a product, we have
n x n i = 1 k f i = j 1 + j 2 + + j k = n n j 1 , j 2 , , j k i = 1 k j i x j i f i ,
where
n j 1 , j 2 , , j k = n ! j 1 ! j 2 ! j k !
is the usual multinomial coefficient. We can write
p q p 1 k < n , ( k , n ) = 1 q ζ k = 1 + 2 + + n 1 = p p 1 , 2 , , n 1 i = 1 n 1 i q i q ζ k ,
with
i q i q ζ k = i + 1 i q ζ k θ 1 i .
One has in particular
x n 1 = d | n Φ d ( x ) ,
yielding for instance x 16 1 = Φ 1 ( x ) Φ 2 ( x ) Φ 4 ( x ) Φ 8 ( x ) Φ 16 ( x ) .

3. Expressions Involving Multinomial Coefficients

3.1. Integral Representation from Group Theory

Let us consider D ( j ) a representation of dimension 2 j + 1 of the group SU ( 2 ) , and U ( θ ) an element of SU ( 2 ) corresponding to the rotation of angle θ around axis (Oz). The quantity
χ j ( U ) = Tr D ( j ) = m = j j D m m j ( U ) = m = j j j m | D ( j ) | j m ,
where D is the Wigner function, is called the characteristic function or the character of the irreducible representation of rank j [42,43,44,45]. In contrast to D m m j ( U ) , the function χ j ( U ) is invariant under rotations of the coordinate system. χ j ( U ) is entirely determined by the rotation angle θ and is independent of the rotation axis: χ j ( U ) χ j ( θ ) . θ is related to the Euler angles by
cos θ 2 = cos β 2 cos α + γ 2 .
There are many possible expressions for the latter quantity (by an integral representation, etc.); in particular, it can be formulated in terms of F 1 2 , the Gauss hypergeometric function, and Jacobi P 2 j ( 1 / 2 , 1 / 2 ) [46,47], Gegenbauger C 2 j 1 [48,49] or Chebyshev polynomials of the second kind U 2 j [50]:
χ j ( θ ) = ( 2 j + 1 )   F 1 2 j , j + 1 3 / 2 ; sin 2 θ 2 = ( 2 j + 1 )   F 1 2 2 j , 2 ( j + 1 ) 3 / 2 ; sin 2 θ 2 = U 2 j cos θ 2 = C 2 j 1 cos θ 2 = ( 4 j + 2 ) ! ! 2 ( 4 j + 1 ) ! ! P 2 j ( 1 / 2 , 1 / 2 ) cos θ 2 ,
where m ! ! = m ( m 2 ) ( m 3 ) . One has
χ j ( θ ) = m = j j e i m θ = sin ( j + 1 / 2 ) θ sin ( θ / 2 ) ,
and the characters obey the orthogonality relation
0 2 π d θ sin 2 θ 2 χ j 1 ( θ ) χ j 2 ( θ ) = π δ j 1 , j 2 ,
where δ j 1 , j 2 represents the Kroneckers symbol and
0 2 π d θ sin 2 θ 2 χ j 1 ( θ ) χ j 2 ( θ ) χ j 3 ( θ ) = π j 1 , j 2 , j 3 ,
where j 1 , j 2 , j 3 = 1 if j 1 + j 2 + j 3 is an integer and | j 1 j 2 | j 3 j 1 + j 2 , and 0 otherwise. Among particular values, one has χ j ( 0 ) = 2 j + 1 , χ j ( 2 π ) = ( 1 ) 2 j ( 2 j + 1 ) and χ j ( π ) = 0 if J is a half-integer and ( 1 ) j if J is an integer. It is worth noting that the three coordinates of one vector span the irreducible representation (1) of SO ( 3 ) and the 3 N coordinates of the N vectors span the reducible representation ( 1 ) ( 1 )  N times. In fact, the group SO ( 3 ) acts diagonally on the direct sum of the various symmetric tensor powers of this representation or, equivalently, on polynomials in the coordinates. The Molien generating function associated with a given irreducible representation j is the power series in a variable λ where the coefficient of λ n is the multiplicity of j in the SO ( 3 ) -module of n-degree polynomials in the coordinates (see Appendix D and references therein).
In contrast to the D m m j , the characters χ j ( U ) are real. The characters which depend on the combined rotations U 1 U 2 U n do not change under cyclic permutations of the rotations χ j ( U 1 U 2 U n ) = χ j ( U 2 U n U 1 ), and therefore χ j ( U 1 U 2 ) = χ j ( U 2 U 1 ) in spite of the non-commutativity of U 1 and U 2 .
For n non-equivalent fermions in shells j (i.e., a configuration P n j as defined in the introduction), let us consider the representation
D ( j ) D ( j ) D ( j ) n times .
According to a property of the tensorial product,
Tr D ( j ) D ( j ) D ( j ) n times = Tr D ( j ) Tr D ( j ) Tr D ( j ) n times
and the character χ ( n ) of such a representation is therefore
χ ( n ) ( θ ) = χ j ( θ ) n = sin j + 1 2 θ sin θ 2 n .
All rotations through the same angle θ about different axes through the same center have the same character. From a group theory point of view, they belong to the same class. The characters of different irreducible representations (called the irreducible characters) are orthogonal over the range ( 0 , π ) with the weight function ( 1 cos θ ) . Indeed,
2 π 0 π χ j ( θ ) χ j ( θ ) sin 2 θ 2 d θ = δ j , j .
The multiplicity Q n j ( J ) which we wish to determine appears in the decomposition
χ ( n ) ( θ ) = J = J 0 n j Q n j ( J ) χ J ( θ ) ,
where J 0 = 0 or 1/2 depending upon whether n j is an integer or half-integer and J moves in steps of 1 from J 0 to n j . Q n j ( J ) denotes the distribution Q ( J ) for n non-equivalent fermions in single-j shells, i.e., for a configuration
j 1 j 1 j 1 n times .
The orthogonality (Equation (13)) of the irreducible characters then results in an integral representation for the multiplicity. Mikhailov has given an analytic expression for the multiplicity Q n j ( J ) of occurrence of any angular momentum J in the decomposition of the direct product of n identical angular momenta j [42]. He guessed the expression by examining special cases for low values of n. However, a proof based on combinatorics has been published, to our knowledge, only for J s . In the following, we shall present another method of arriving at this expression which does not distinguish between the two cases J j and J < j .
The Haar measure enables one to integrate over the elements U of the group SU ( 2 ) , which enables one to extend to a compact Lie group the notion, usual for finite groups, of summation over elements of that group. Such a measure reads
d μ ( θ ) = 2 ( 1 cos θ ) d θ d 2 n ,
where d 2 n is the integration measure over unit sphere S 2 . θ varies between 0 and 4 π . One finds then that the volume of the group SU ( 2 ) is 16 π 2 . One can show that
d μ ( θ ) 16 π 2 χ j ( θ ) χ j ( θ ) = δ j , j
or
0 2 π ( 1 cos θ ) χ j ( θ ) χ j ( θ ) d θ = 2 π δ j , j .
In order to obtain Q n j ( J ) , one has to project such a character on the character χ J of D ( J ) , which yields
Q n j ( J ) = d μ ( θ ) 16 π 2 χ ( n ) ( θ ) χ J ( θ ) ,
or, setting ϕ = θ / 2 ,
Q n j ( J ) = 2 π 0 π sin 2 j + 1 ϕ sin ϕ n sin ( 2 J + 1 ) ϕ sin ϕ d ϕ ,
which is also equal to
Q n j ( J ) = 1 π 0 2 π sin 2 j + 1 ϕ sin ϕ n sin ( 2 J + 1 ) ϕ sin ϕ d ϕ ,
and the integral representation of the distribution of angular momentum projection M reads
P n j ( M ) = 1 π 0 π sin ( 2 j + 1 ) ϕ sin ϕ n cos ( 2 M ϕ ) d ϕ ,
where P n j ( J ) denotes the distribution P ( M ) for n non-equivalent electrons in single-j shells, i.e., for a configuration of Equation (14). Thus, for the simple case j = 1 / 2 , one has
Q n 1 2 J = 2 n + 1 π 0 π cos n ϕ sin ( 2 J + 1 ) ϕ sin ϕ d ϕ ,
which is in fact equal to
Q n 1 2 J = ( 2 J + 1 ) n ! n 2 + J + 1 ! n 2 J ! .
We also have a connection to the Vilenkin function P m n j [51,52,53]:
m = j j P m m j cos ( θ ) = sin j + 1 2 θ sin θ 2 = U 2 j sin θ 2
related to the Jacobi polynomials P n ( α , β ) through
P m m j cos ( θ ) = 1 2 m 1 + cos θ m P j m ( 0 , 2 m ) .

3.2. Expression Using Multinomial Coefficients: A New Sum Rule

Using the Moivre formula, it can be easily proven that
sin ( 2 j + 1 ) ϕ = 0 2 p + 1 2 j + 1 ( 1 ) p 2 j + 1 2 p + 1 cos 2 j 2 p ϕ sin 2 p + 1 ϕ .
Following the above-mentioned procedure, it is possible to derive a general formula for Q n j ( J ) , involving multinomial coefficients. One also has
sin ( 2 j + 1 ) ϕ sin ϕ n = p = 0 j 1 / 2 a p ( j ) n ,
i.e.,
sin ( 2 j + 1 ) ϕ sin ϕ n = k 0 + k 1 + k 2 + + k j 1 / 2 = n n k 0 , k 1 , k 2 , , k j 1 / 2 a 0 k 0 a 1 k 1 a j 1 / 2 k j 1 / 2 ,
with
a p ( j ) = ( 1 ) p 2 j + 1 2 p + 1 cos 2 j 2 p ϕ sin 2 p ϕ .
Since j is a half-integer, we can replace Equation (17) by
sin ( 2 j + 1 ) ϕ = p = 0 j 1 / 2 ( 1 ) p 2 j + 1 2 p + 1 cos 2 j 2 p ϕ sin 2 p + 1 ϕ .
We also have
sin ( 2 J + 1 ) ϕ = r = 0 B ( 1 ) r 2 J + 1 2 r + 1 cos 2 J 2 r ϕ sin 2 r + 1 ϕ ,
where B = J 1 / 2 if J is a half-integer and B = J if J is an integer, i.e.,
B = J + ( 1 ) 2 J 1 4 .
We can therefore write
Q n j ( J ) = 2 π k 0 + k 1 + k 2 + + k j 1 / 2 = n r = 0 B n k 0 , k 1 , k 2 , , k j 1 / 2 ( 1 ) r 2 J + 1 2 r + 1 × 0 π p = 0 j 1 / 2 ( 1 ) p k p 2 j + 1 2 p + 1 cos 2 j 2 p ϕ sin 2 p ϕ k p cos 2 J 2 r ϕ sin 2 r + 2 ϕ d ϕ ,
i.e.,
Q n j ( J ) = k 0 + k 1 + k 2 + + k j 1 / 2 = n r = 0 B n k 0 , k 1 , k 2 , , k j 1 / 2 ( 1 ) r 2 J + 1 2 r + 1 × ( 1 ) s = 0 j 1 / 2 s k s p = 0 j 1 / 2 2 j + 1 2 p + 1 k p × 0 π cos ϕ 2 J 2 r + s = 0 j 1 / 2 ( 2 j 2 s ) k s sin ϕ 2 r + 2 + 2 s = 0 j 1 / 2 s k s d ϕ .
We then use
0 π cos a ϕ sin b ϕ d ϕ = 1 + ( 1 ) a 2 Γ a + 1 2 Γ b + 1 2 Γ a + b + 2 2 ,
where Γ is the usual Gamma function, with
a = 2 J 2 r + s = 0 j 1 / 2 ( 2 j 2 s ) k s = 2 J r + s = 0 j 1 / 2 ( j s ) k s
and
b = 2 r + 2 + 2 s = 0 j 1 / 2 s k s = 2 r + 1 + s = 0 j 1 / 2 s k s .
We have
a + 1 2 = J r + s = 0 j 1 / 2 ( j s ) k s + 1 2 = J r + j s = 0 j 1 / 2 k s s = 0 j 1 / 2 s k s + 1 2 = J r + j n s = 0 j 1 / 2 s k s + 1 2 ,
because s = 0 j 1 / 2 k s = n . In the same way,
b + 1 2 = r + 1 + s = 0 j 1 / 2 s k s + 1 2 and a + b + 2 2 = J + j n + 2 .
We obtain therefore finally
Q n j ( J ) = s = 0 j 1 / 2 k s = n r = 0 B n k 0 , k 1 , k 2 , , k j 1 / 2 ( 1 ) r 2 J + 1 2 r + 1 × ( 1 ) s = 0 j 1 / 2 s k s p = 0 j 1 / 2 2 j + 1 2 p + 1 k p 1 + ( 1 ) 2 J r + s = 0 j 1 / 2 ( j s ) k s 2 × Γ J r + j n s = 0 j 1 / 2 s k s + 1 2 Γ r + 1 + s = 0 j 1 / 2 s k s + 1 2 Γ J + j n + 2 .
and a similar formula can of course be derived for P n j ( M ) .

4. Expression in Terms of Pascal Triangles

For the simple configuration j 1 , we have
M P j ( M ) z M = z j + + z j .
In a general way, the so-called generalized Pascal triangle T m , n , q coefficients are given by [54]:
( 1 + t + t 2 + + t m ) n = q T m , n , q t q
and since
( 1 + t + t 2 + + t m ) n = 1 t m + 1 1 t n = 1 t m + 1 n q q + n 1 q t q = k = 0 n n k ( t m + 1 ) k q q + n 1 q t q = k = 0 n q ( 1 ) k n k q + n 1 q t m k + k + q = q k = 0 n ( 1 ) k n k q m k k + n 1 q k m k t q ,
we obtain
T m , n , q = k = 0 n ( 1 ) k n k q m k k + n 1 q k m k .
In the framework of the one-dimensional isotropic Heisenberg magnet, the quantum inverse scattering method has led to an algebraic version of the classical substitution method of Bethe (the so-called Bethe ansatz) for the determination of the trace of the monodromy matrix. These eigenvectors were referred to as Bethe vectors [55]. The completeness of the multiplet system constructed from the Bethe vectors was proven by Kirrilov for the Heisenberg model of arbitrary spin. Kirrilov [56] found that the number of Bethe vectors with fixed for the Heisenberg model of spin S reads (N being the total number of possible states):
Z 1 ( N , S | ) = j ( 1 ) j N j N + 2 + ( 2 S + 1 ) j N 2 ,
which is also a generalized Pascal triangle coefficient. Therefore, in the case of a configuration made of n shells j 1 (i.e., n non-equivalent j fermions), then
M = j n j n P n j ( M ) z M = z j + + z j n ,
and in that case, we obtain
P n j ( M ) = k = 0 min n , j n + M 2 j + 1 ( 1 ) k n k ( j + 1 ) n + M ( 2 j + 1 ) k 1 n 1 .
Using
s t = s 1 t + s 1 t 1 ,
one finds, for Q ( J ) ,
Q n j ( J ) = k = 0 min n , j n + J + 1 2 j + 1 ( 1 ) k + 1 n k ( j + 1 ) n + J ( 2 j + 1 ) k 1 n 2 .
Such coefficients were also encountered in multiphoton processes, for the proportion of neutral atoms in a statistical description of multiple ionization [57], or in the determination of the number of configurations in atomic physics [58].

Case of Different Angular Momenta

Let us consider a system of n non-equivalent fermions [59,60], each in a single- j i shell (of degeneracy g i = 2 j i + 1 , i = 1 , , n ). In the following, such a configuration
j 1 1 j 2 1 j n 1
will be denoted: j 1 j 2 j n 1 j n (abbreviated j 1 j n in the following). The number of J levels of such a configuration can be obtained as [61]
Q j 1 j n ( J ) = M = J J + 1 ( 1 ) J M P j 1 j n ( M ) = P j 1 j n ( J ) P j 1 j n ( J + 1 ) ,
where P j 1 j n ( M ) represents the distribution of the angular momentum projection M, i.e., the number of states of a given value of M. The number of such states for a configuration j 1 j 2 j n 1 j n is equal to the number of times that a given representation U appears in the product U U 1 U 2 U n 1 U n [51,52]. Such a quantity reads
P j 1 j n ( M ) = χ ( U ) χ ( U 1 ) χ ( U n ) d μ
where μ is a relevant measure (such as the Haar measure; see Equation (15)). The latter expression can be put in the form
P j 1 j n ( M ) = 1 π 0 2 π χ M ( 2 θ ) χ j 1 ( 2 θ ) χ j n ( 2 θ ) sin 2 ( θ ) d θ .
For a configuration made of n non-equivalent fermions with respective angular momenta j 1 , , j n , one has
M P j 1 j n ( M ) z M = i = 1 n z j i + + z j i ,
i.e.,
P j 1 j n ( M ) = i = 1 n ( 2 j i + 1 ) k i j i + M 0 k i 1 ( 1 ) k n + i = 1 n j i + M i = 1 n ( 2 j i + 1 ) k i 1 n 1 i = 1 n 1 k i
and
Q j 1 j n ( J ) = i = 1 n ( 2 j i + 1 ) k i j i + J 0 k i 1 ( 1 ) k + 1 n + i = 1 n j i + J i = 1 n ( 2 j i + 1 ) k i 1 n 2 i = 1 n 1 k i .
In a more general case of a configuration made of n 1 non-equivalent j 1 fermions, n 2 non-equivalent j 2 fermions, …, n σ non-equivalent j σ fermions, we have, with n = α = 1 σ n α ,
P j α n α ( M ) = s α / α = 1 σ ( 2 j α + 1 ) s α α = 1 σ n α j α + M ; 0 s α n α ( 1 ) s 1 + s 2 + + s σ × n + α = 1 σ n α j α + M 1 α = 1 σ ( 2 j α + 1 ) s α n 1 n 1 s 1 n σ s σ
and
Q j α n α ( J ) = s α / α = 1 σ ( 2 j α + 1 ) s α α = 1 σ n α j α + J ; 0 s α n α ( 1 ) s 1 + s 2 + + s σ + 1 × n + α = 1 σ n α j α + J 1 α = 1 σ ( 2 j α + 1 ) s α n 2 n 1 s 1 n σ s σ ,
where the notation j α n α means j 1 j 1 n 1 times j 2 j 2 n 2 times j σ j σ n σ times .

5. Special Cases

The formulas for Q ( J ) obtained above can be simplified in particular cases. Making the change of variables t = cos 2 θ , one has
Q n j ( J ) = 2 π 0 1 U 2 J ( t ) U 2 j ( t ) n 1 t t d t ,
which enables one to express Q n j ( J ) in terms of roots of Chebyshev polynomials [62]. For a half-integer j,
U 2 j ( cos θ ) = 4 j t l = 1 j 1 / 2 t r l ( j )
with
r q ( p ) = cos 2 q π 2 p + 1 .
Similarly, for a half-integer J (corresponding to an odd number n of j shells),
U 2 J ( cos θ ) = 4 J t l = 1 J 1 / 2 t r l ( J ) ,
and for integer J (corresponding to an even number n of j shells):
U 2 J ( cos θ ) = 4 J l = 1 J t r l ( J ) .
Therefore, one has, in the case of an even number of fermions n = 2 m ,
Q n j ( J ) = 2 π 4 J + 2 m j 0 1 k = 1 J t r k ( J ) l = 1 j 1 / 2 t r l ( j ) 2 m t m 1 t t d t ,
with the usual convention that the empty product is 1. It is possible, using these formulas, to obtain analytical expressions of Q ( J ) for j = 1 / 2 and j = 3 / 2 , n = 2 m , and where m is an integer. For j = 1 / 2 , the distribution of J values reads
Q 2 m 1 2 ( J ) = 2 π 4 J + m 0 1 k = 1 J t cos 2 k π 2 J + 1 t m 1 t t d t ,
which is equal to (see Equation (16))
Q 2 m 1 2 ( J ) = ( 2 J + 1 ) ( 2 m ) ! ( m J ) ! ( m + J + 1 ) ! .
In the case j = 3 / 2 , one has
Q 2 m 3 2 ( J ) = 2 π 4 3 m + 1 0 1 k = 1 J t cos 2 k π 2 J + 1 t 1 2 2 m t m 1 t t d t
and specifically for J = 1 ,
Q 2 m 3 2 ( 1 ) = 2 π 4 3 m + 1 0 1 t 1 4 t 1 2 2 m t m 1 t t d t
or also
Q 2 m 3 2 ( 1 ) = 2 π 4 3 m + 1 0 1 t + 1 2 t 1 2 2 m + 1 t m 1 t t d t
which turns out to be equal to, using a computer algebra system,
Q 2 m 3 2 ( 1 ) = 2 3 + 6 m π ( 1 ) 2 m 2 5 / 2 3 m Γ ( 1 / 2 + m ) Γ ( 1 + 2 m ) 3 Γ ( 3 / 2 + 3 m ) × 3   F 1 2 1 / 2 , 1 / 2 + m 3 / 2 + 3 m ; 1 2   F 1 2 1 / 2 , 3 / 2 + m 5 / 2 + 3 m ; 1 / 2 + 2 5 / 2 3 m Γ ( 3 / 2 3 m ) Γ ( 1 + 2 m ) Γ ( 1 / 2 m ) × 3   F 1 2 1 / 2 , 1 / 2 + m 3 / 2 + 3 m ; 1 / 2 + 2   F 1 2 1 / 2 , 3 / 2 + m 5 / 2 + 3 m ; 1 / 2 + π Γ ( 1 / 2 + 3 m ) 8 Γ ( 3 + 3 m ) 2 ( 1 + 6 m )   F 1 2 2 3 m , 2 m 1 / 2 3 m ; 1 / 2 ( 2 + 3 m )   F 1 2 1 3 m , 2 m 1 / 2 3 m ; 1 / 2 .
As an example, for m = 3 , we obtain
Q 6 3 2 ( 1 ) = 90
For higher values of J, the Computer Algebra System Mathematica [63] can give expressions involving hypergeometric functions F 1 2 , but the latter become more and more cumbersome as J increases.

6. Sum Rules

The total number of states yields the following trivial sum rule:
J = 0 n j ( 2 J + 1 ) Q n j ( J ) = ( 2 j + 1 ) n .
It is instructive to study the so-called odd–even staggering [27], i.e., the difference between the number of odd values of J minus the number of even values of J. Such a study can be easily performed using the sum rule
J = 0 n j F ( 2 J + 1 , x ) Q n j ( J ) = F ( 2 j + 1 , x ) n ,
where
F ( k , x ) = x ( k 1 ) 2 + x ( k 3 ) 2 + + x ( k 3 ) 2 + x ( k 1 ) 2 = x k / 2 x k / 2 x 1 / 2 x 1 / 2
yielding
J = 0 n j ( 1 ) J Q n j ( J ) = ( 1 ) n j
if it is an integer, and 0 otherwise. Polychronakos and Sfetsos derived the following sum [64]:
J = 0 n j J ( J + 1 ) ( 2 J + 1 ) Q n j ( J ) = j ( j + 1 ) n ( 2 j + 1 ) n .
Other sum rules can be obtained by differentiating Relation (22) with respect to x and taking the limit x 1 . The odd derivative of the order k + 1 gives the same result as the even derivative of order k. The first sum rules are
J = 0 n j J ( J + 1 ) ( 2 J + 1 ) ( J 2 + J + 18 ) Q n j ( J ) = j ( j + 1 ) ( j 2 + j + 18 ) n ( 2 j + 1 ) n ,
J = 0 n j J ( J + 1 ) ( 2 J + 1 ) 600 + J ( J + 1 ) ( 118 + J ( J + 1 ) ) Q n j ( J ) = j ( j + 1 ) 600 + j ( j + 1 ) ( 118 + j ( j + 1 ) ) n ( 2 j + 1 ) n
and
J = 0 n j J ( J + 1 ) ( 2 J + 1 ) 35280 + J ( J + 1 ) 11772 + J ( J + 1 ) ( 412 + J ( J + 1 ) ) Q n j ( J ) = j ( j + 1 ) 35280 + j ( j + 1 ) 11772 + j ( j + 1 ) ( 412 + j ( j + 1 ) ) n ( 2 j + 1 ) n .
In addition, knowing the distribution P ( M ) and evaluating it at the minimum value of M gives the total number of J levels (see Appendix C).

7. Conclusions

We have presented analytical formulas for the number of levels in the case of an arbitrary number of non-equivalent fermions in single-j orbits. We obtained expressions in terms of Gaussian polynomials, symmetric functions, and cyclotomic polynomials. From integral representations, derived within the framework of group theory and using irreducible representations of S O ( 3 ) , we deduced exact expressions of P ( M ) and Q ( J ) involving multinomial coefficients. We also obtained expressions in terms of generalized Pascal triangle coefficients, and presented special cases for j = 3 / 2 in terms of hypergeometric functions, as well as sum rules for Q ( J ) . Formulas for the total number of levels in a configuration were also given, and the possibilities offered by the Molien functions were evoked. The formulas presented here apply to both atomic and nuclear physics. This also illustrates that the two disciplines can benefit greatly from each other.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

I would like to thank P. Mandelbaum for helpful discussions.

Conflicts of Interest

The author declares no conflicts of interest.

Appendix A. Distributions P(M) and Q(J) for a j Shell with N = 4 Fermions

We found in Ref. [32] that P ( M ) for j 4 can be expressed as
P ( 2 j n ; j , 4 ) = 1 18 j + n 1 2 3 1 6 π ( n ) 8 j + n 1 2 H ( n ) f 1 n 2 + ξ ( n )
+ ω ( 2 j + n 1 ) ,
where H ( n ) is the usual Heaviside function, π ( n ) = n mod 2 ,
ξ ( n ) = 1 9 , 1 72 , 0 , 17 72 , 1 9 , 1 8
if n mod 6 = ( 0 , 1 , 2 , 3 , 4 , 5 ) , respectively,
f 1 ( n ) = 2 9 n 3 n 2 6 n 6 + 1 9 ,
and
ω ( 2 j + n 1 ) = 0 , 1 72 , 1 9 , 1 8 , 1 9 , 17 72 , 0 , 17 72 , 1 9 , 1 8 , 1 9 , 1 72
if 2 j + n 1 mod 12 = ( 0 , 1 , 2 , 3 , 4 , 5 , 6 , 7 , 8 , 9 , 10 , 11 ) , respectively. In this formula, 2 j n must be non-negative, and n may be negative. Explicitly, n must be such that 2 j + 6 n 2 j . The fundamental relation (3), together with the expression (A1) of the M distribution for a four-fermion system, gives the distribution Q ( J ) of the total momentum J for j 4 . One has to evaluate Q ( 2 j n ; j , 4 ) = P ( 2 j n ; j , 4 ) P ( 2 j ( n 1 ) ; j , 4 ) which can be expressed as Q 1 + Q 2 + Q 3 . The quantity Q 1 consists of the contribution of the first two terms of (A1), which is obtained noticing that π ( n 1 ) = 1 π ( n ) :
Q 1 = 1 12 j + n 3 2 2 7 72 + π ( n ) 8 ( 2 j + n 3 / 2 ) .
The quantity Q 2 reads
Q 2 = H ( n ) f 1 ( n / 2 ) + ξ ( n ) f 1 ( ( n 1 ) / 2 ) ξ ( n 1 ) = H ( n ) n 2 12 n 6 1 72 + ξ ¯ ( n ) ,
with
ξ ¯ ( n ) = ξ ( n ) ξ ( n 1 ) = 1 72 , 7 72 , 1 72 , 17 72 , 25 72 , 17 72
for n mod 6 = 0 , 1 , 2 , 3 , 4 , 5 , respectively. Finally the ω -dependent term is
Q 3 = ω ¯ ( 2 j + n 1 ) = ω ( 2 j + n 1 ) ω ( 2 j + n 2 ) = 1 72 , 1 72 , 7 72 , 17 72 , 1 72 , 25 72 , 17 72 , 17 72 , 25 72 , 1 72 , 17 72 , 7 72
for 2 j + n 1 mod 12 = 0 11 , respectively. The complete formula therefore reads
Q ( 2 j n ; j , 4 ) = 1 12 j + n 3 2 2 7 72 + π ( n ) 8 ( 2 j + n 3 / 2 ) H ( n ) ( n 1 ) 2 12 7 72 + ξ ¯ ( n ) + ω ¯ ( 2 j + n 1 ) .

Appendix B. SO(3) and SU(2)

In this article, we used the notations SO ( 3 ) and SU ( 2 ) indifferently. This is due to the fact that there is a connection between SO ( 3 ) and SU ( 2 ) [65]. SU ( 2 ) is the set of all 2 × 2 complex matrices of determinant equal to ± 1 and satisfying A A = I :
A = α β β α ,
where | α | 2 + | β | 2 = 1 . There are three independent parameters required to specify an element of SU ( 2 ) . One of them is
A = cos ( θ / 2 ) exp i ( ψ + ϕ ) 2 sin ( θ / 2 ) exp i ( ψ ϕ ) 2 sin ( θ / 2 ) exp i ( ψ ϕ ) 2 cos ( θ / 2 ) exp i ( ψ + ϕ ) 2 ,
where 0 θ π , 0 ψ 4 π and 0 ϕ 2 π . The objects that transform under SU ( 2 ) elements like this are called “spinors”. They are analogous to vectors in the three-dimensional rotation group SO ( 3 ) . Actually, there is a two-to-one homomorphic mapping of the group SU ( 2 ) onto the group SO ( 3 ) . If A SU ( 2 ) maps onto a representation R ( A ) SO ( 3 ) , then R ( A ) = R ( A ) and the mapping is chosen so that
R ( A ) j k = 1 2 Tr σ j A σ k A 1 , j , k = 1 , 2 , 3 ,
where
σ 1 = 0 1 1 0 , σ 2 = 0 i i 0 , σ 3 = 1 0 0 1 ,
are the Pauli spin matrices. This implies that the representations of SO ( 3 ) are also representations of SU ( 2 ) . In addition, there are spinor representations of SU ( 2 ) that have no analog in SO ( 3 ) .

Appendix C. Total Number of J Levels, Restricted Partitions and q−Binomial Coefficients

We are interested in the following quantity:
S = J = J min J max Q ( J ) = P J min ,
which plays an important role in the determination of the sizes of the blocks of the Hamiltonian matrix in atomic and nuclear physics. For instance, as concerns the example of j 4 of Appendix A, we found in Ref. [32] that the total number of levels for four fermions of spin j reads
Q tot j 4 = P ( 0 ; j , 4 ) = 2 9 j 3 j 2 6 + j 6 5 / 72 if j 1 / 2 mod 3 = 0 , + 3 / 8 if j 1 / 2 mod 3 = 1 , + 11 / 72 if j 1 / 2 mod 3 = 2 .
Setting
a i = m i + j + 1 ,
the determination, of P ( M ) , which can be inferred through the constraint (4), consists also of determining the number of N−vectors a 1 , , a N satisfying
a 1 + a 2 + + a N = M + N ( j + 1 ) ,
where
1 a 1 < a 2 < < a N 2 j + 1 .
One has therefore
S j N = P [ 2 j + 1 N , N ] ,
where P [ n , k ] is the number of partitions of n k / 2 in at most n positive integers, each one being lower than or equal to k,   In the Computer Algebra System Mathematica [63], this reads
  • Pt[n_,k_]:=Length[IntegerPartitions[Floor[n*k/2],n,Range[k]]];
  • j = …;
  • Flatten[Table[Pt[2j+1-k,k],{k,0,2j+1}]]
Thus, S j N represents the number of partitions of N ( 2 j + 1 N ) / 2 in at most ( 2 j + 1 N ) positive integers, each one being lower than or equal to N. In fact, P [ t , k ] is the coefficient of q t k / 2 in the expansion of
t + k t q
with
t + k t q = q t + k 1 q t + k q q t + k q 2 q t + k q t 1 q t 1 q t q q t q 2 q t q t 1 .
The generating function of the number of partitions of m in at most r positive integers, each one being lower than or equal to s, can be found in Ref. [66], p. 127. Thus, S j N is the coefficient of q N ( 2 j + 1 N ) / 2 in the expansion of
2 j + 1 2 j + 1 N q = 2 j + 1 N q
with 2 j + 1 N q given by Equation (5).
Table A1 and Table A2 represent the total number of levels for configurations j N with j = 9 / 2 and 11 / 2 , respectively. The total number of levels for configurations n j , with different values of j and n, is given in Table A3. It is worth mentioning that, for j = 9 / 2 , the values in the row correspond to the number of integers in [ 0 , 10 n 1 ] whose sums of digits are equal to the most common value, which is 9 n / 2 for even n and ( 9 n ± 1 ) / 2 for odd n > 1 (e.g., the most common value of sums of digits of numbers from 0 to 9999 is 9 × 4 / 2 = 18 , so there are 670 numbers in this range whose sums of digits are 18). In the case of j = 5 / 2 , the values in the row of Table A3 correspond to the greatest multiplicity of one- or two-dimensional standard representation of Lie algebras sl ( 2 ) in the decomposition of tensor power F 6 k , where F 6 is the standard 6−dimensional irreducible representation of sl ( 2 ) .
Table A1. Maximum number of J levels in configuration 9 2 N .
Table A1. Maximum number of J levels in configuration 9 2 N .
N J = J min J max Q ( J ) = P ( J min )
11
25
310
418
520
618
710
85
91
101
Table A2. Maximum number of J levels in configuration 11 2 N .
Table A2. Maximum number of J levels in configuration 11 2 N .
N J = J min J max Q ( J ) = P ( J min )
11
26
315
433
549
658
749
833
915
106
111
121
Table A3. Values of the total number of spin-J states ( P ( J min ) = P ( n j ) ) for a configuration n j for different values of j and n.
Table A3. Values of the total number of spin-J states ( P ( J min ) = P ( n j ) ) for a configuration n j for different values of j and n.
j / n 123456789
1/2123610203570126
3/21412441555802128809230,276
5/21627146780433224,017135,954767,394
7/21848344246018,152134,5121,012,6647,635,987
9/211075670600055,252512,3654,816,03045,433,800
11/2112108115612,435137,2921,528,68817,232,084195,170,310

Appendix D. Deriving Multiplicities from the Knowledge of Molien Functions

It is worth mentioning that Molien functions can help in finding the multiplicities. Let γ be a representation (irreducible or reducible) of a group G and γ [ s ] the s t h anti-symmetrized Kronecker product ( A ^ represents the anti-symmetrization operator).
γ [ s ] = A ^ γ × γ × γ s times .
γ [ s ] is normally reducible and can be expressed as a direct sum of representations Γ of G, i.e.,
γ [ s ] = Γ n ( Γ , γ , s ) Γ ,
where n ( Γ , γ , s ) is the multiplicity of Γ , i.e., the number of times it appears in the decomposition of γ [ s ] as a direct sum of Γ . The Molien function is the generating function of n ( Γ , γ , s ) , which reads
ϕ ( Γ , γ , s ) = s = 0 d n ( Γ , γ , s ) λ s ,
where d is the dimension of Γ . If G = S O ( 3 ) , then γ = D ( j ) , an irreducible representation of the three-dimensional rotation group S O ( 3 ) with the highest weight j. Its character reads
χ J ( ω ) = z J + 1 z J z 1
with z = e i ω . It can be shown, in particular using
J = 0 t J χ J ( z ) = ( 1 + t ) z ( z t ) ( 1 z t ) ,
that the Molien function reads [67,68,69], for half-integer j:
ϕ ( J , j , λ ) = i 4 π d z 1 z 2 2 z 3 χ J z 2 m = j j 1 + λ z 2 m ,
which can be obtained using the method of residues. The associated generating function is
F ( j , λ , t ) = J = 0 ϕ ( J , j , λ ) t J = J = 0 s = 0 d n ( J , j , s ) λ s t J ,
where d is the number of anti-symmetrized Kronecker products. The quantity n ( J , j , s ) represents the number of tensors of rank J which can be found in the anti-symmetrized s t h Kronecker product of the representation D ( j ) of SU ( 2 ) . Since
J = 0 t J χ J ( z ) = ( 1 + t ) z ( z t ) ( 1 z t ) ,
we obtain
F ( j , λ , t ) = 1 2 Res 1 z 2 2 z 2 ( z t ) ( 1 z t ) m = j j 1 + λ z 2 m ,
i.e.,
F ( j , λ , t ) = J , s n ( J , j , s ) λ s t 2 J ,
from which Q n j ( J ) can be deduced.

References

  1. Bethe, H.A. An attempt to calculate the number of energy levels of a heavy nucleus. Phys. Rev. 1936, 50, 332–341. [Google Scholar] [CrossRef]
  2. Talmi, I. Simple Models of Complex Nuclei; Harwood: Amsterdam, The Netherlands, 1993. [Google Scholar]
  3. Zamick, L.; Van Isacker, P. Partial dynamical symmetries in the shell: Progress and puzzles. Phys. Rev. C 2008, 78, 044327. [Google Scholar]
  4. Zhao, Y.M.; Arima, A.; Yoshinaga, N. Many-body systems interacting via a two-body random ensemble. I. Angular momentum distribution in the ground states. Phys. Rev. C 2002, 66, 064322. [Google Scholar]
  5. Mulhall, D.; Volya, A.; Zelzvinsky, V. Geometric chaoticity leads to ordered spectra for randomly interacting fermions. Phys. Rev. Lett. 2000, 85, 4016–4019. [Google Scholar] [CrossRef]
  6. Fu, G.J.; Zhao, Y.M.; Arima, A. Nucleon-pair approximation of the shell model with isospin symmetry. Phys. Rev. C 2013, 88, 054303. [Google Scholar] [CrossRef]
  7. Zamick, L.; Escuderos, A. Odd-J pairing in nuclei. Phys. Rev. C 2013, 87, 044302. [Google Scholar] [CrossRef]
  8. Ginocchio, J.N.; Haxton, W.C. Symmetries in Science VI: From the Rotation Group to Quantum Algebras; Gruber, B., Ed.; Plenum: New York, NY, USA, 1993; p. 263. [Google Scholar]
  9. Zhao, Y.M.; Arima, A. Number of states with a given angular momentum for identical fermions and bosons. Phys. Rev. C 2003, 68, 044310. [Google Scholar] [CrossRef]
  10. Zamick, L.; Escuderos, A. Alternate derivation of Ginocchio-Haxton relation [(2j + 3)/6]. Phys. Rev. C 2005, 71, 054308. [Google Scholar]
  11. Talmi, I. Number of states with given spin J of n fermions in a j orbit. Phys. Rev. C 2005, 72, 037302. [Google Scholar] [CrossRef]
  12. Zamick, L.; Escuderos, A. Companion problems in quasispin and isospin. Phys. Rev. C 2005, 71, 014315. [Google Scholar] [CrossRef]
  13. Zamick, L.; Escuderos, A. Isospin relations for four nucleons in a single j shell. Phys. Rev. C 2005, 72, 044317. [Google Scholar]
  14. Zhao, Y.M.; Arima, A. Number of states for nucleons in a single-j shell. Phys. Rev. C 2005, 72, 064333. [Google Scholar]
  15. Zamick, L.; Robinson, S.J.Q. Zeros of 6-j symbols: Atoms, nuclei, and bosons. Phys. Rev. C 2011, 84, 044325. [Google Scholar]
  16. Qi, C.; Wang, X.B.; Xu, Z.X.; Liotta, R.J.; Wyss, R.; Xu, F.R. Alternate proof of the Rowe-Rosensteel proposition and seniority conservation. Phys. Rev. C 2010, 82, 014304. [Google Scholar]
  17. Wang, X.B.; Xu, F.R. Special relations of coefficients of fractional parentage and partial dynamical symmetries in j = 9/2 shells. Phys. Rev. C 2012, 85, 034304. [Google Scholar]
  18. Zamick, L.; Escuderos, A. Large-j limit for certain 9-j symbols: Power law behavior. Phys. Rev. C 2013, 88, 014326. [Google Scholar] [CrossRef]
  19. Kleszyk, B.; Zamick, L. Analytical and numerical calculations for the asymptotic behavior of unitary 9j coefficients. Phys. Rev. C 2014, 89, 044322. [Google Scholar] [CrossRef]
  20. Hertz-Kintish, D.; Zamick, L.; Kleszyk, B. Asymptotics of the 3j and 9j coefficients. Phys. Rev. C 2014, 90, 027302. [Google Scholar]
  21. Redmond, P.J. An explicit formula for the calculation of fractional parentage coefficients. Proc. Roy. Soc. A 1954, 222, 84. [Google Scholar]
  22. Rosensteel, G.; Rowe, D.J. Seniority-conserving forces and Usp(2j + 1) partial dynamical symmetry. Phys. Rev. C 2003, 67, 014303. [Google Scholar]
  23. Bao, M.; Zhao, Y.M.; Arima, A. Number of states for identical particles. Phys. Rev. C 2016, 93, 014307. [Google Scholar]
  24. Pain, J.-C. Special six-j and nine-j symbols for a single-j shell. Phys. Rev. C 2011, 84, 047303. [Google Scholar]
  25. Pain, J.-C. Number of spin-J states and odd-even staggering for identical particles in a single-j shell. Phys. Rev. C 2018, 97, 064311. [Google Scholar]
  26. Poirier, M.; Pain, J.-C. Angular momentum distribution in a relativistic configuration: Magnetic quantum number analysis. J. Phys. B At. Mol. Opt. Phys. 2021, 54, 145002. [Google Scholar]
  27. Poirier, M.; Pain, J.-C. Distribution of the total angular momentum in relativistic configurations. J. Phys. B At. Mol. Opt. Phys. 2021, 54, 145006. [Google Scholar]
  28. Pain, J.-C. Total number of J levels for identical particles in a single-j shell using coefficients of fractional parentage. Phys. Rev. C 2019, 99, 054321. [Google Scholar]
  29. Karazija, R. Sums of Atomic Quantities and Mean Characteristics of Spectra; Mokslas: Vilnius, Lithuania, 1992. (In Russian) [Google Scholar]
  30. Bauche, J.; Bauche-Arnoult, C.; Peyrusse, O. Atomic Properties in Hot Plasmas; Springer: Cham, Switzerland, 2015. [Google Scholar]
  31. Kucˇas, S.; Karazija, R.; Jonauskas, V.; Aksela, S. Global characteristics of atomic spectra and their use for the analysis of spectra. III. Auger spectra. Phys. Scr. 1995, 52, 639–648. [Google Scholar]
  32. Poirier, M.; Pain, J.-C. Exact expressions for the number of levels in single-j orbits for three, four, and five fermions. Phys. Rev. C 2021, 104, 064324. [Google Scholar]
  33. Poirier, M.; Pain, J.-C. Exact expressions of the distributions of total magnetic quantum number and angular momentum in single-j orbits: A general technique for any number of fermions. Phys. Rev. C 2024, 109, 024306. [Google Scholar]
  34. Sunko, D.K. Rapid calculation of exact spin distributions of multilevel configurations. Phys. Rev. C 1986, 33, 1811–1813. [Google Scholar]
  35. Sunko, D.K.; Svrtan, D. Generating function for angular momentum multiplicities. Phys. Rev. C 1985, 31, 1929–1933. [Google Scholar] [CrossRef] [PubMed]
  36. Andrews, G.E. The Theory of Partitions, Encyclopedia of Mathematics and Its Applications; Cambridge University Press, Cambridge, UK, 1984.
  37. Schroeder, M. Cyclotomic polynomials. In Number Theory in Science and Communication: With Applications in Cryptography, Physics, Digital Information, Computing, and Self-Similarity; Springer: Berlin/Heidelberg, Germany, 2009; pp. 289–304. [Google Scholar]
  38. Mendonça, J.R.G. Exact eigenspectrum of the symmetric simple exclusion process on the complete, complete bipartite and related graphs. J. Phys. A Math. Theor. 2013, 46, 295001. [Google Scholar]
  39. Gyamfi, J.A.; Barone, V. On the composition of an arbitrary collection of SU(2) spins: An enumerative combinatoric approach. J. Phys. A Math. Theor. 2018, 51, 105202. [Google Scholar]
  40. Klapisch, M.; Krasnitz, A.; Mandelbaum, P.; Bauche-Arnoult, C.; Bauche, J. New results of the unresolved transition arrays method, Eight International Colloquium on Ultraviolet and X-ray Spectroscopy of Astrophysical and Laboratory Plasmas. In Proceedings of the IAU Colloq. 86, Washington, DC, USA, 27–29 August 1984; Naval Research Laboratory: Washington, DC, USA, 1984; p. 114. [Google Scholar]
  41. Faà di Bruno, F. Einleitung in Die Theorie der Binären Formen; B. G. Teubner: Leipzig, Germany, 1881; p. 126. [Google Scholar]
  42. Hamermesh, M. Group Theory and Its Applications to Physical Problems; Addison-Wesley: Reading, MA, USA, 1962. [Google Scholar]
  43. Murnaghan, F.D. The Theory of Group Representations; Dover Publications: Dover, NY, USA, 1963. [Google Scholar]
  44. Rashid, M.A. Addition of arbitrary number of identical angular momenta. J. Phys. A Math. Gen. 1977, 10, L135–L137. [Google Scholar]
  45. Ramadevi, P.; Dubey, V. Group Theory for Physicists: With Applications; Cambridge University Press: Cambridge, UK, 2019. [Google Scholar]
  46. Szegö, G. Ch. 4 in Orthogonal Polynomials. In Jacobi Polynomials, 4th ed.; American Mathematical Society: Providence, RI, USA, 1975. [Google Scholar]
  47. Iyanaga, S.; Kawada, Y. (Eds.) Jacobi Polynomials. Appendix A, Table 20.V in Encyclopedic Dictionary of Mathematics; MIT Press: Cambridge, MA, USA, 1980; p. 1480. [Google Scholar]
  48. Talman, J.D. Special Functions: A Group Theoretic Approach; Based on Lectures by Eugene P. Wigner; Mathematical Physics Monograph Series; Benjamin: New York, NY, USA; Amsterdam, The Netherlands, 1968. [Google Scholar]
  49. Nikiforov, A.F.; Uvarov, V.B. Special Functions of Mathematical Physics: A Unified Introduction with Applications; Birkhäuser: Boston, MA, USA, 2013. [Google Scholar]
  50. Mason, J.C.; Handscomb, D.C. Chebyshev Polynomials, 1st ed.; Chapman and Hall/CRC: Boca Raton, FL, USA, 2002. [Google Scholar]
  51. Vilenkin, N.J. Special Functions and Theory of Group Representations; Izdat. “Nauka”, Moscow, 1965, Transl. Math. Monographs; American Mathematical Society: Providence, RI, USA, 1968; Volume 22. [Google Scholar]
  52. Vilenkin, N.J.; Klimyk, A.U. Representation of Lie Groups and Special Functions, Recent Advances; Mathematics and its Applications 316; Kluwer Academic Publishers: Dordrecht, The Netherlands, 1994. [Google Scholar]
  53. Haettel, T.; Was, M. Fonctions Spéciales et Théorie des Représentations de Groupes. Available online: https://imag.umontpellier.fr/~haettel/Memoire%20Maitrise.pdf (accessed on 24 March 2025). (In French).
  54. Bollinger, R.C. Extended Pascal Triangles. Math. Mag. 1993, 66, 87–94. [Google Scholar]
  55. Takhtadzhyan, L.A.; Fadeev, L.D. Spectrum and scattering of excitations in the one-dimensional isotropic Heisenberg model. J. Sov. Math. 1984, 24, 241–267. [Google Scholar]
  56. Kirrilov, A.N. Combinatorial identities, and completeness of eigenstates of the Heisenberg magnet. J. Math. Sci. 1985, 30, 2298–2310. [Google Scholar]
  57. Crance, M. A statistical description for multiphoton stripping of atoms. J. Phys. B At. Mol. Opt. Phys. 1984, 17, 4333–4341. [Google Scholar]
  58. Poirier, M.; Pain, J.-C. Analytical and numerical expressions for the number of atomic configurations contained in a supershell. J. Phys. B At. Mol. Opt. Phys. 2020, 53, 115002. [Google Scholar]
  59. Mikhailov, V.V. Addition of a large number of identical angular momenta. Statistical distributions. J. Phys. A Math. Gen. 1979, 12, 2329–2335. [Google Scholar]
  60. Mikhailov, V.V. Addition of an arbitrary number of different spins. Phys. Lett. 1980, 6A, 229–230. [Google Scholar] [CrossRef]
  61. Condon, E.U.; Shortley, G.H. The Theory of Atomic Spectra; Cambridge University Press: Cambridge, UK, 1935. [Google Scholar]
  62. Curtright, T.L.; Van Kortryk, T.S.; Zachos, C.K. Spin multiplicities. Phys. Lett. A 2017, 381, 422–427. [Google Scholar] [CrossRef]
  63. Wolfram Research, Inc. Mathematica, Version 11.3; Wolfram Research: Champaign, IL, USA, 2018.
  64. Polychronakos, A.P.; Sfetsos, K. Composition of many spins, random walks and statistics. Nucl. Phys. B 2016, 913, 664–693. [Google Scholar] [CrossRef]
  65. Cornwell, J.F. The double crystallographic point groups. In Group Theory in Physics; Academic Press: Cambridge, MA, USA, 1984; Volume I, pp. 342–355. [Google Scholar]
  66. Comtet, L. Advanced Combinatorics: The Art of Finite and Infinite Expansions (Reidel, Dordrecht, 1974); Translation of Analyse Combinatoire; Presses Universitaires de France: Paris, France, 1970. [Google Scholar]
  67. Asherova, R.M.; Zhilinsky, B.I.; Pavlov-Verevkin, V.B.; Smirnov, Y.F. Preprint IFT-88-70P; Institute for Theoretical Physics: Kiev, Ukraine, 1988. [Google Scholar]
  68. Raychev, P.P.; Smirnov, Y.F. On the generalization of Molien functions to supergroups. J. Phys. A Math. Gen. 1990, 23, 4417–4425. [Google Scholar] [CrossRef]
  69. Dhont, G.; Cassam-Chenaï, P.; Patras, F. Molien generating functions and integrity bases for the action of the SO(3) and O(3) groups on a set of vectors. J. Math. Chem. 2021, 10, 2294–2326. [Google Scholar] [CrossRef]
Figure 1. Example of a state of a single-j shell ( 9 / 2 ) 4 corresponding to M = m [ 1 ] + m [ 2 ] + m [ 3 ] + m [ 4 ] = 9 / 2 5 / 2 + 3 / 2 + 5 / 2 = 3 .
Figure 1. Example of a state of a single-j shell ( 9 / 2 ) 4 corresponding to M = m [ 1 ] + m [ 2 ] + m [ 3 ] + m [ 4 ] = 9 / 2 5 / 2 + 3 / 2 + 5 / 2 = 3 .
Atoms 13 00025 g001
Figure 2. Distribution P ( M ) of magnetic quantum number M (angular momentum projection) for j = 9 / 2 and N = 4 .
Figure 2. Distribution P ( M ) of magnetic quantum number M (angular momentum projection) for j = 9 / 2 and N = 4 .
Atoms 13 00025 g002
Figure 3. Distribution Q ( J ) of total angular momentum J for j = 9 / 2 and N = 4 .
Figure 3. Distribution Q ( J ) of total angular momentum J for j = 9 / 2 and N = 4 .
Atoms 13 00025 g003
Table 1. Parameters p k and c k required by Sunko’s method in the ( 7 / 2 ) 3 case.
Table 1. Parameters p k and c k required by Sunko’s method in the ( 7 / 2 ) 3 case.
k01234567
p k /134310−6
c k 11234566
Table 2. Distribution of Q ( J ) in the ( 7 / 2 ) 3 case.
Table 2. Distribution of Q ( J ) in the ( 7 / 2 ) 3 case.
J1/23/25/27/29/211/213/215/2
Q ( J ) c 7 c 6 c 6 c 5 c 5 c 4 c 4 c 3 c 3 c 2 c 2 c 1 c 1 c 0 c 0
Q ( J ) 01111101
Table 3. Parameters p k and c k required by Sunko’s method in the ( 7 / 2 ) 3 ( 5 / 2 ) 2 ( 1 / 2 ) 1 ( 9 / 2 ) 2 case for k = 0 to 10.
Table 3. Parameters p k and c k required by Sunko’s method in the ( 7 / 2 ) 3 ( 5 / 2 ) 2 ( 1 / 2 ) 1 ( 9 / 2 ) 2 case for k = 0 to 10.
k012345678910
p k /4878−1−1−30−2−7
c k 141229621192113495458081144
Table 4. Parameters p k and c k required by Sunko’s method in the ( 7 / 2 ) 3 ( 5 / 2 ) 2 ( 1 / 2 ) 1 ( 9 / 2 ) 2 case for k = 11 to 20.
Table 4. Parameters p k and c k required by Sunko’s method in the ( 7 / 2 ) 3 ( 5 / 2 ) 2 ( 1 / 2 ) 1 ( 9 / 2 ) 2 case for k = 11 to 20.
k11121314151617181920
p k 4−141204−104−7
c k 1553202825543108366141794628497551955270
Table 5. Distribution Q ( J ) in the ( 7 / 2 ) 3 ( 5 / 2 ) 2 ( 1 / 2 ) 1 ( 9 / 2 ) 2 case for J = 0 to 10.
Table 5. Distribution Q ( J ) in the ( 7 / 2 ) 3 ( 5 / 2 ) 2 ( 1 / 2 ) 1 ( 9 / 2 ) 2 case for J = 0 to 10.
J012345678910
Q ( J ) c 20 c 19 c 18 c 17 c 16 c 15 c 14 c 13 c 12 c 11 c 10
c 19 c 18 c 17 c 16 c 15 c 14 c 13 c 12 c 11 c 10 c 9
Q ( J ) 75220347449518553554526475409336
Table 6. Distribution Q ( J ) in the ( 7 / 2 ) 3 ( 5 / 2 ) 2 ( 1 / 2 ) 1 ( 9 / 2 ) 2 case for k = 11 to 20.
Table 6. Distribution Q ( J ) in the ( 7 / 2 ) 3 ( 5 / 2 ) 2 ( 1 / 2 ) 1 ( 9 / 2 ) 2 case for k = 11 to 20.
J11121314151617181920
Q ( J ) c 9 c 8 c 7 c 6 c 5 c 4 c 3 c 2 c 1 c 0
c 8 c 7 c 6 c 5 c 4 c 3 c 2 c 1 c 0
Q ( J ) 26319613892573317831
Table 7. Parameters p k and c k required by Sunko’s method in the ( 7 / 2 ) 3 ( 3 / 2 ) 2 ( 5 / 2 ) 1 ( 9 / 2 ) 2 case for k = 0 to 10.
Table 7. Parameters p k and c k required by Sunko’s method in the ( 7 / 2 ) 3 ( 3 / 2 ) 2 ( 5 / 2 ) 1 ( 9 / 2 ) 2 case for k = 0 to 10.
k012345678910
p k /41248−10−30−5−3
c k 1414368417132456593214492153
Table 8. Parameters p k and c k required by Sunko’s method in the ( 7 / 2 ) 3 ( 3 / 2 ) 2 ( 5 / 2 ) 1 ( 9 / 2 ) 2 case for k = 11 to 15.
Table 8. Parameters p k and c k required by Sunko’s method in the ( 7 / 2 ) 3 ( 3 / 2 ) 2 ( 5 / 2 ) 1 ( 9 / 2 ) 2 case for k = 11 to 15.
k1112131415
p k 4−445−1
c k 30544168547269438515
Table 9. Parameters p k and c k required by Sunko’s method in the ( 7 / 2 ) 3 ( 3 / 2 ) 2 ( 5 / 2 ) 1 ( 9 / 2 ) 2 case for k = 16 to 21.
Table 9. Parameters p k and c k required by Sunko’s method in the ( 7 / 2 ) 3 ( 3 / 2 ) 2 ( 5 / 2 ) 1 ( 9 / 2 ) 2 case for k = 16 to 21.
k161718192021
p k 04−94−7−3
c k 10,12311,66813,06214,20415,01815,440
Table 10. Distribution Q ( J ) in the ( 7 / 2 ) 3 ( 3 / 2 ) 2 ( 5 / 2 ) 1 ( 9 / 2 ) 2 case for J = 1 / 2 to 21/2.
Table 10. Distribution Q ( J ) in the ( 7 / 2 ) 3 ( 3 / 2 ) 2 ( 5 / 2 ) 1 ( 9 / 2 ) 2 case for J = 1 / 2 to 21/2.
J1/23/25/27/29/211/213/215/217/219/221/2
Q ( J ) c 21 c 20 c 19 c 18 c 17 c 16 c 15 c 14 c 13 c 12 c 11
c 20 c 19 c 18 c 17 c 16 c 15 c 14 c 13 c 12 c 11 c 10
Q ( J ) 42281411421394154516081572147113041114901
Table 11. Distribution Q ( J ) in the ( 7 / 2 ) 3 ( 3 / 2 ) 2 ( 5 / 2 ) 1 ( 9 / 2 ) 2 case for J = 23 / 2 to 43/2.
Table 11. Distribution Q ( J ) in the ( 7 / 2 ) 3 ( 3 / 2 ) 2 ( 5 / 2 ) 1 ( 9 / 2 ) 2 case for J = 23 / 2 to 43/2.
J23/225/227/229/231/233/235/237/239/241/243/2
Q ( J ) c 10 c 9 c 8 c 7 c 6 c 5 c 4 c 3 c 2 c 1 c 0
c 9 c 8 c 7 c 6 c 5 c 4 c 3 c 2 c 1 c 0
Q ( J ) 7045173672411538748221031
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

Pain, J.-C. Statistics of Quantum Numbers for Non-Equivalent Fermions in Single-j Shells. Atoms 2025, 13, 25. https://doi.org/10.3390/atoms13040025

AMA Style

Pain J-C. Statistics of Quantum Numbers for Non-Equivalent Fermions in Single-j Shells. Atoms. 2025; 13(4):25. https://doi.org/10.3390/atoms13040025

Chicago/Turabian Style

Pain, Jean-Christophe. 2025. "Statistics of Quantum Numbers for Non-Equivalent Fermions in Single-j Shells" Atoms 13, no. 4: 25. https://doi.org/10.3390/atoms13040025

APA Style

Pain, J.-C. (2025). Statistics of Quantum Numbers for Non-Equivalent Fermions in Single-j Shells. Atoms, 13(4), 25. https://doi.org/10.3390/atoms13040025

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