1. Introduction
With the discovery of molecular magnetism in the early 1990s [
1,
2] and the subsequent synthesis of a variety of exchange-coupled clusters, theoretical modeling based on multi-spin Hamiltonians has come to the forefront of analyzing the intriguing properties of multinuclear transition-metal complexes [
3,
4,
5]. The leading term is usually of Heisenberg type,
[
3], and this model allows the description of thermodynamic properties like magnetic susceptibilities or heat capacities over a wide temperature range; anisotropic contributions become significant only at low temperatures for systems with a magnetic ground state (total spin
). Additionally, for an individual spin multiplet (comprising states with
eigenvalues
), anisotropic terms that lift the
-fold degeneracy (zero-field splitting) can often be treated perturbatively [
3,
6,
7]. Thus, techniques for the efficient solution of isotropic spin Hamiltonians, which may also include other interaction terms, like biquadratic exchange, are of particular importance.
In its most basic form, exact diagonalization (ED) quickly encounters limits due to the rapid growth of the Hilbert-space dimension,
, with the number
N of sites, where
denotes a local spin quantum number. Therefore, it becomes necessary to factorize the Hamiltonian according to different irreducible representations of the symmetry group [
8]. This not only reduces computation times but also decreases memory requirements, which are usually the limiting factor in practice. The symmetries that can be utilized are spin-rotational symmetry, which assigns a quantum number
S to each level, and the point-group symmetry (PG), which manifests as symmetry under permutations of sites [
8]. Here, we will focus on spin symmetry.
It is straightforward to use only the
z-component
of the total spin by working in a basis of uncoupled states
defined by a set of
eigenvalues
of the individual sites with a selected value
.
and PG symmetry can be combined easily [
8,
9], which is essential for calculating the thermodynamic properties of large lattice fragments, such as systems with
spin-1/2 particles, using the finite-temperature Lanczos method [
10].
On the other hand, adapting the basis to have definite spin
S leads to smaller matrices than working with
only, but the conventional procedure is more complex, as it relies on irreducible tensor-operator (ITO) techniques, where successively coupled states are decoupled in the calculation of Hamiltonian matrix elements using Wigner-9
j symbols. This scheme is explained in textbooks [
3,
11], and it is implemented in the MAGPACK [
12] and PHI [
13] programs. For a detailed recent account with numerous examples, see Ref. [
14].
Our present strategy of spin-projecting uncoupled states
is simpler, as it does not require spin-coupling or ITO techniques. To the best of our knowledge, such an approach has, in only two instances, been applied to Heisenberg clusters: Bernu et al. [
15] used Löwdin’s projector (see Theory section) in conjunction with spatial-symmetry adaptation for Lanczos ED of triangular-lattice sections containing up to 36 spin-1/2 sites, and, by a similar procedure, we recently derived eigenvalues in symbolic form for particularly small systems [
16]. Here, we are concerned with spin symmetry only, where two aspects come to the forefront: the selection of configurations and the practical realization of projection. In the following Theory section, we initially address the former question, which for
has already been answered by Löwdin’s theorem [
17,
18,
19,
20]. We extend this theorem to arbitrary
s and then discuss additional aspects of spin projection, emphasizing that a group-theoretical projector can be evaluated directly [
19] without the need to discretize the relevant integral. This offers numerical advantages over Löwdin’s projector. In the subsequent Results and Discussion sections, we illustrate the selection of uncoupled basis functions according to our extension of Löwdin’s theorem and investigate the numerical accuracy of different projection methods through examples.
We do not claim that the projection method for spin adaptation offers computational advantages over the traditional vector-coupling method, nor that it enables the exact diagonalization of larger systems. However, such advantages could be realized by combining total spin symmetry with PG symmetry for full symmetry factorization of the Hamiltonian, which cannot generally be achieved in a practically useful way with the vector-coupling method [
8,
21,
22,
23].
2. Theory
The essence of the present scheme for the spin factorization of the Hamiltonian consists of applying a projector
to configurations
with a definite
z-projection
. As mentioned in the Introduction, Löwdin’s operator [
24] in Equation (1) has, in a few cases [
15,
16], been used in the diagonalization of Heisenberg models:
Applying to a state with definite M affords a pure-spin state , or the state is annihilated if it has no contributions of spin S.
Before discussing practical aspects and presenting an alternative formulation of
, we want to address the question of how to select functions
such that, upon application of
, a complete and linearly independent set spanning the respective
S sector is generated. The total number of these multiplets is denoted
. If one chooses
configurations randomly, the projected functions will generally exhibit linear dependencies. This problem has long been solved for
[
17]. Proofs of the underlying so-called Löwdin theorem [
20] were presented by Gershgorn [
18] and Pauncz [
19]. This theorem states that, for the application of
in the
sector of an
system, only those uncoupled states should be selected where the cumulative sum of the
is never negative, i.e.,
for
. For a system of twelve spin-1/2 centers, we illustrate this in
Figure 1 for
.
To our knowledge, the selection of configurations in
systems to ensure the linear independence of spin-projected states has not been addressed before. In the following, we explain how Löwdin’s theorem can be intuitively extended from
to any
s, including systems with different local spins on different sites. The idea leading to this generalization is to embed
in a larger space where the site is decomposed into
auxiliary spin-1/2 degrees of freedom
; see Equation (2).
A state with maximum total spin is symmetric under all permutations of its constituents. For instance,
can be represented as a symmetric linear combination of six spin-1/2 configurations related by all possible permutations; see Equation (3).
The number of linearly independent states for a given S in the spin-1/2 representation is greater than (for ) or equal to [for , ] for the corresponding number in the actual system (in the latter, are coupled to yield ). Thus, by first selecting spin-1/2 configurations according to Löwdin’s theorem, then applying and finally eliminating contributions with local spin quantum numbers below their respective maximum value , an overcomplete basis would be spanned, which we like to avoid.
For the sake of the argument, a local projector is defined in Equation (4), although this need not be constructed in practical calculations:
A state in the spin-1/2 space is mapped onto the smaller space with local spins
through application of the product
of all separate projectors, as in Equation (5):
Because
is a linear combination of scalar products
,
and
commute with the total spin,
; thus,
. Now, consider a spin-1/2 configuration
that was selected for
using Löwdin’s theorem. As mentioned, since
is symmetric under all permutations of the spin-1/2 centers within their respective sets of size
we can locally shift the
in
all the way to the left. Consider, for example, four
centers, each of which is split into two spin-1/2 components. A configuration selected for
according to Löwdin’s theorem would be
. With respect to
,
is equivalent to
, with all
moved to the left in their respective sets; see Equation (6).
However,
does not fulfill the cumulative sum criterion, so
can be expressed as a linear combination of projected states that do satisfy
for all
n:
We state without proof that
. Thus, when rewriting Equation (7) as Equation (8), both sides are non-zero:
Now, we apply
, using
and
:
Equation (9) shows that is a linear combination (with coefficients , ) of , where the unprimed spin-1/2 configurations obey Löwdin’s theorem and does not.
These considerations suggest a simple procedure to select the states of the actual system such that they span a complete and linearly independent set upon spin projection : replace each with a configuration of spin-1/2 centers, arranging a number of sites with on the left and the remaining with on the right. The thus obtained configurations of all N centers are concatenated into a single configuration of length , and the cumulative sum criterion, , , as per Löwdin’s theorem, is applied. A few examples are provided in the Results section. For systems with mixed local spin (i.e., not all are the same), contrary to the case of uniform s, the selected sets of configurations for different site numberings are, in general, not related by site permutations. However, each projected set completely spans the S space, irrespective of the numbering. Although we have not presented a strict proof of the described procedure, numerous checks have confirmed its correctness.
One can straightforwardly check that a correct number
of basis states has been found by computing the difference between the dimensions of the
and
spaces, which are obtained by counting states with a respective
. (A general formula for the dimensions of the
M-spaces was derived in Ref. [
25], and for numerous systems with uniform
s, the dimensions
are collected in Table 1 of Ref. [
14].)
Despite being linearly independent, the basis selected according to the (extended) Löwdin theorem is not necessarily optimally conditioned from the perspective of numerical stability. Particularly for larger systems, small eigenvalues of the overlap matrix can compromise the accuracy. Thus, as an alternative to selecting uncoupled basis functions following Löwdin’s theorem, we use the pivoted Cholesky decomposition [
26] (PCD), which is occasionally employed in quantum chemistry (see, e.g., Ref. [
27], and references cited therein) to address the problem of over-complete basis sets by pruning them to yield optimal low-rank approximations, which enhances the numerical stability and efficiency of electronic-structure calculations. A PCD of the full overlap matrix
(see below) between spin-projected uncoupled states in a constant-
M space provides an optimal basis in terms of numerical stability but is generally not practical. Therefore, we suggest the following iterative procedure: (i) initial selection: spin-project a randomly selected set of
configurations and form the overlap matrix; (ii) rank determination: calculate the rank
r of the overlap matrix within a tolerance well above the numerical accuracy threshold; (iii) optimal subset: select
r uncoupled states through a PCD of the overlap matrix, (iv) supplementary selection: if
, add additional configurations to the selected
r states so that the total slightly exceeds
, but avoid significantly exceeding
; (v) iteration: repeat the rank determination and state selection until
r has reached
; (vi) adjust numerical tolerance: if necessary, lower the tolerance for rank determination if
r remains too low even after considering significantly more than
states. The details of this procedure, or any similar approach, may be subject to optimization.
The selected uncoupled states form the columns of a matrix
R, i.e., each column of
R contains a single entry 1. The construction of
in the uncoupled constant-
M basis is a standard task. For completeness, we shall briefly sketch it for
(these considerations can be easily extended to
, see Ref. [
28]). Each uncoupled state, e.g.,
, is represented by a bit string of length
N, where 0 at the
i-th position denotes
, and 1 denotes
. For
between nearest neighbors
, the Hamiltonian is formulated in terms of raising and lowering operators in Equation (10),
where
contributes to the diagonal, and
flips the respective bits’ values, thus accounting for off-diagonal elements. The number of non-zero entries in each row or column of
approximately corresponds to the number of interacting spin pairs, and
can be efficiently stored in the sparse-matrix format [
28]. As outlined below, we need to form the matrix product
(the superscript
T denotes transposition) and, therefore, limit ourselves to the rows corresponding to the selected configurations by directly constructing the rectangular matrix
instead of
H.
We now turn to the practical aspects of spin-adapting the basis. It is simplest to work directly with the Löwdin projector, Equation (1). Since the spin square
is the sum of the scalar products of all pairs up to a constant (first sum in Equation (11)),
its representation
can be calculated just as easily as
, or the relation
may be used [
15]. Spin adaptation is formally accomplished by multiplying
R with the projector,
, although in practice, one would successively apply the factors in the product of Equation (1) to
R instead of explicitly computing the dense matrix
. Since the projector is Hermitian
and idempotent
and commutes with the Hamiltonian,
, it only appears once in the matrix products for the Hamiltonian and the overlap matrix (not to be confused with the spin vector), Equations (12) and (13), respectively.
and define a generalized eigenvalue problem (GEP) whose solution yields the complete spectrum in the respective S sector. This requires storing in addition to . We solve the GEP with Matlab. Note that numerical rounding errors generally cause minor asymmetries in and , which may lead to small imaginary components in the eigenvalues. We, therefore, perform a symmetrization, and , which also significantly improves the computational efficiency of solving the GEP.
As indicated, one can alternatively write
in terms of an integration with respect to Euler angles
,
and
,
where
is a diagonal element of the Wigner rotation matrix, and the asterisk * denotes complex conjugation. In a space with definite
M, the group-theoretical projector, Equation (14), is fundamentally identical with Löwdin’s operator, Equation (1), but the former is preferred in methods like Projected Hartree–Fock [
29,
30] (PHF), where it simplifies the optimization of the reference state using self-consistent field or gradient-based procedures. PHF has also been applied to Heisenberg spin clusters [
31,
32,
33]. In this approach, a symmetry-broken mean-field reference—which can be a product state of either individual spins [
31] (such as a non-collinear spin configuration) or spin centers grouped into subclusters [
32,
33]—is optimized for spin- or PG-projection.
Here, we would like to make use of the fact that the group-theoretical projector can be straightforwardly evaluated in closed form for
, as explained by Pauncz (Chapter 4.9 in Ref. [
19]), and provide a compact derivation for completeness. Since we work with
eigenstates, the integrations over the Euler angles
and
, both associated with rotations about the
z-axis, can be directly evaluated [
19,
34] and yield factors that are irrelevant for our purpose. The non-trivial part of the projector thus reduces to an integration over
. Specifically, the wave function of
in the uncoupled basis
is formulated as an integral over products of elements of Wigner’s small
d-matrices in Equation (15).
The
d-matrix for
is given in Equation (16).
Noting that
, Equation (15) represents a standard integral, Equation (17), where
B is the Euler Beta function, which can be expressed in terms of the Gamma function
.
This leads to the result of Equation (18),
where
is a Sanibel coefficient for the special case of
[
19],
is essentially a sign factor divided by a binomial coefficient,
k counts the number of sites with
, and
is the number of
sites, i.e.,
,
. As explained, the proportionality factor in Equation (18) is not relevant here, and we omit it to avoid clutter.
For
, the closed-form solution of the integral in Equation, in general, represents a linear combination of standard integrals, Equation (17), because the
with
and
are linear combinations of
with different sets of (real integer) exponents (
a,
b), see Equations (20) and (21) for
and
, respectively, where
and
.
By taking these linear combinations into account, we can still evaluate the integral in closed form, which we have implemented up to . Note that, in general, the combinatorial growth of the length of these linear combinations with the number of spin centers may be mitigated by selecting the uncoupled basis through PCD. For example, for , , we found that PCD, based on the full overlap matrix in the sector, yields only Ising-type configurations with local projections . Since is thereby excluded, the integral for the group-theoretical projection is just a single term as for systems. Thus, in a limited PCD procedure, as explained above, one may preferentially select configurations with only a few local z-projections that do not have their maximal magnitude to keep linear combinations in closed-form integral solutions short. However, optimizing such a procedure is beyond the scope of this work.
Alternatively, we can discretize the integration using a Gauss–Legendre grid [
34]. This scales approximately linearly with the number of grid points (larger systems generally require larger grids), and for
it is significantly less efficient than using the closed-form solution of Equation (18). In the Results section, we illustrate some examples using integration on a grid. In our MATLAB code, we have embedded several tasks, including the integration of products of Wigner-
d-matrix elements (Equation (15)), with mex-C functions for efficiency. For
, projection using either Löwdin’s operator or discretized or exact SU(2) integration all require similar amounts of computation time. However, we cannot rule out that in an optimal implementation, one of these options might be significantly advantageous.
3. Results and Discussion
Here, we first present examples for selecting basis states according to the extended Löwdin theorem and then compare the various presented options for spin projection in terms of their numerical accuracy by calculating spectra of antiferromagnetic rings (
between nearest neighbors). Note that our intent is not to provide new insights into any specific magnetic molecule; see, e.g., Ref. [
14] for how the properties of various existing exchange-coupled clusters can be calculated based on ED.
Figure 2 illustrates the cumulative sum of the local
z-projections as a function of the number
n of auxiliary spin-1/2 sites for a system with eight
centers. Each path in the diagram, moving from left to right, represents a configuration selected by the extension of Löwdin’s theorem. When
is applied to these configurations, they form a complete, linearly independent basis for the given
S. In each panel, an arbitrarily chosen configuration is highlighted for illustration, which is expressed as a product of states from three spin-1/2 centers, each representing a single
. This spin-1/2 configuration is translated back into a state of the actual system by summing the
z-projections of the respective spin-1/2 components. As explained, within each center, if
sites are present, they are locally positioned to the left; thus, a configuration is only compatible if
.
Figure 3 shows the energy levels of the antiferromagnetic
spin-1/2 ring as a function of
S. Our reference for assessing the numerical accuracy of the different projection schemes is the exact energy levels
from the full ED of
H. By comparing the spectra across all different
M-spaces, each level is assigned a spin
S. Since
H is constructed numerically exactly, i.e., without (or with negligible) rounding errors, these eigenvalues are accurate within double precision in Matlab, corresponding to approximately 15 or 16 decimal places. In
Figure 4, we plot the logarithmic difference between the numerically exact levels and the levels obtained from the GEP that was set up through either the Löwdin-projector, Gauss–Legendre integration with
grid points, or Sanibel coefficients. One point is plotted for each spin multiplet (the number of points thus corresponds to the dimension of the
space), and within each
S sector, the energies are ordered in ascending fashion from left to right. The grid integration is employed here mainly for comparison with the preferable use of Sanibel coefficients, which, as explained in the Theory section, result from a closed-form solution of the SU(2) integral.
All three methods are accurate to within
, with the maximum error being greatest for Löwdin’s projector. The
grid size is sufficient. However, for
, a peculiar pattern emerges for
: some states are described accurately, while others exhibit significant errors, resulting in a large gap between the two groups; see
Figure 5. Interestingly, all states with
or
belong to the latter group, whereas for
or
, some levels are accurate and others are not.
Figure 6 illustrates the corresponding numerical results for an
ring. Again, Löwdin’s projector is less accurate compared to using Sanibel coefficients. Notably, the maximum error for eigenvalues obtained based on Löwdin’s projector is almost four orders of magnitude larger than for the
ring. Projecting seven contaminating spin contributions (
in Equation (1)) from each configuration, as opposed to five for
, requires more matrix-vector products and thus leads to a greater accumulation of rounding errors. Even the projection using Sanibel coefficients now shows larger errors, although these remain
Finally, as an example of an
system, we choose a ring with
and
. For example, existing Cr
8 rings are described by such a spin Hamiltonian, although anisotropic interactions may need to be added to describe spectroscopic properties [
35]. However, as stated, our aim here remains to examine the accuracy of the spin projection method rather than to study any specific molecule. The full spectrum of the Heisenberg model is shown in
Figure 7.
Figure 8 shows that grid integration yields significantly more accurate spectra than Löwdin’s projector. Using the latter, the maximal error is
(in units of the coupling constant
). Such large deviations could become noticeable when fitting inelastic neutron scattering (INS) spectra or thermodynamic data such as magnetic susceptibilities. Bernu et al. [
15] have already pointed out that rounding errors can quickly accumulate with Löwdin’s projector, necessitating additional mitigation measures. In this regard, an advantage of the group-theoretical formulation becomes evident. However, errors remain comparatively large even when the integrals of Equation (15) are evaluated exactly (based on a closed-form solution of Equation (15)); see
Figure 8c. The limited accuracy indeed derives from near linear dependencies on the spin-projected basis.
This is demonstrated by using PCD to select the uncoupled states. The left panel of
Figure 9 depicts results based on the impractical PCD of the full overlap matrix, while the right panel was obtained from basis selection through the described iterative PCD procedure that considers only a randomly selected fraction of states. It is evident that Löwdin’s projector still affords larger errors than the group-theoretical formulation, but all results are significantly more accurate than when selecting states according to the extended Löwdin theorem.