Interaction Renormalization
A unique feature of Sunny is its support for building classical models where quantum spin is represented as an $N$-level system, rather than just an expected dipole. This generalization can be important when modeling quantum spin Hamiltonians that include, e.g., a single-ion anisotropy, or a biquadratic coupling between sites. Sunny also supports constraining quantum spin to the space of pure dipoles; in this case, Sunny will automatically perform an interaction strength renormalization that enhances accuracy.
Local operators
A quantum spin-$s$ state has $N = 2s + 1$ levels. Each local spin operator $\hat{S}^{\{x,y,z\}}$ is faithfully represented as an $N×N$ matrix. These matrices can be accessed using spin_matrices
for a given label $s$. For example, the Pauli matrices are associated with $s = 1/2$.
When $s > 1/2$, it is possible to construct multipole moments beyond the spin-dipole. For example,
S = spin_matrices(3/2)
@assert S[3] ≈ diagm([3/2, 1/2, -1/2, -3/2])
@assert S[3]^2 ≈ diagm([9/4, 1/4, 1/4, 9/4])
If the operator -S[3]^2
is passed to set_onsite_coupling!
, it would set an easy-axis anisotropy in the $\hat{z}$ direction.
Any Hermitian operator can be expanded in the basis of Stevens operators $\hat{\mathcal{O}}_{k,q}$ up to a constant shift. To see this expansion, use print_stevens_expansion
:
print_stevens_expansion((S[1]^2 + S[2]^2)) # Prints -(1/3)𝒪₂₀ + 5/2
Alternatively, the same operator could have been constructed directly from stevens_matrices
:
O = stevens_matrices(3/2)
@assert S[1]^2 + S[2]^2 ≈ -O[2, 0]/3 + (5/2)*I
The Stevens operators $\mathcal{O}_{k, q}$ are defined below as $k$th order polynomials of the spin operators.
Renormalization procedure for :dipole
mode
Sunny will typically operate in one of two modes: :SUN
or :dipole
. The former faithfully represents quantum spin as an SU(N) coherent-state which, for our purposes, is an $N$-component complex vector. In contrast, :dipole
mode constrains the coherent-state to the space of pure dipoles. Here, Sunny will automatically renormalize the magnitude of each Stevens expectation value to achieve maximal consistency with :SUN
mode. This procedure was derived in D. Dahlbom et al., [arXiv:2304.03874].
By way of illustration, consider a quantum operator $\hat{\mathcal{H}}_{\mathrm{local}}$ giving a single-ion anisotropy for one site. It can be expanded in Stevens operators,
\[\hat{\mathcal H}_{\mathrm{local}} = \sum_{k, q} A_{k,q} \hat{\mathcal{O}}_{k,q},\]
for some coefficients $A_{k,q}$.
In :SUN
mode, Sunny will faithfully represent $\hat{\mathcal H}_{\mathrm{local}}$ as an $N×N$ matrix. In :dipole
mode, the expected energy $\langle \hat{\mathcal H}_{\mathrm{local}} \rangle$ must somehow be approximated using the expected dipole data.
One approach is to formally take $s \to \infty$, and this yields the traditional classical limit of a spin system. In this limit spin operators commute and expectation values of polynomials become polynomials of expectation values. For example, $\langle \hat{S}^\alpha \hat{S}^\beta\rangle \to \langle \hat{S}^\alpha \rangle \langle \hat{S}^\beta\rangle$, because any corrections are damped by the factor $s^{-1} \to 0$. The expectation of a Stevens operator $\langle \hat{\mathcal{O}}_{k,q} \rangle$ would then become a classical Stevens function $\mathcal{O}_{k,q}(\langle\hat{\mathbf{S}}\rangle)$, i.e., a polynomial of the same form, but now applied to the expected dipole. Classical Stevens functions are constructed as homogeneous polynomials of order $k$, because lower-order terms would vanish in the limit $s \to \infty$.
For real compounds with finite quantum spin-$s$, one can obtain a better approximation by avoiding the formal $s \to \infty$ limit. Corrections can be derived by starting from the full dynamics of SU(N) coherent states and then constraining to the space of pure dipole states $|\boldsymbol{\Omega}\rangle$. The latter are defined as any states where the expected dipole 3-vector,
\[\boldsymbol{\Omega} ≡ \langle \boldsymbol{\Omega}| \hat{\mathbf{S}} | \boldsymbol{\Omega}\rangle,\]
has maximal magnitude $|\boldsymbol{\Omega}| = s$ and arbitrary direction.
For a pure dipole state, group theory dictates that expectations of the Stevens operators can be expressed as a renormalization of the classical Stevens functions,
\[\langle \boldsymbol{\Omega}| \hat{\mathcal{O}}_{k,q} | \boldsymbol{\Omega}\rangle = c_k \mathcal{O}_{k,q}(\boldsymbol{\Omega}).\]
At fixed $k$, the two sides must be proportional because they are both spin-$k$ irreducible representations of SO(3). The renormalization factors can be calculated explicitly:
\[\begin{align*} c_1 &= 1 \\ c_2 &= 1-\frac{1}{2}s^{-1} \\ c_3 &= 1-\frac{3}{2}s^{-1}+\frac{1}{2}s^{-2} \\ c_4 &= 1-3s^{-1}+\frac{11}{4}s^{-2}-\frac{3}{4}s^{-3} \\ c_5 &= 1-5s^{-1}+\frac{35}{4}s^{-2}-\frac{25}{4}s^{-3}+\frac{3}{2}s^{-4} \\ c_6 &= 1-\frac{15}{2}s^{-1}+\frac{85}{4}s^{-2}-\frac{225}{8}s^{-3}+\frac{137}{8}s^{-4}-\frac{15}{4}s^{-5} \\ &\vdots \end{align*}\]
Constrained to the space of dipoles, the expected local energy becomes
\[E_{\mathrm{local}}(\boldsymbol{\Omega}) = \langle \boldsymbol{\Omega}| \hat{\mathcal H}_{\mathrm{local}} | \boldsymbol{\Omega}\rangle = \sum_{k, q} c_k A_{k,q} \mathcal{O}_{k,q}(\boldsymbol{\Omega}).\]
It can be shown that SU(N) dynamics reduces to the usual Landau-Lifshitz dynamics of dipoles, but involving $E_{\mathrm{local}}(\boldsymbol{\Omega})$ as the classical Hamiltonian. The renormalization factors $c_k$ can therefore be interpreted as a correction to the traditional large-$s$ classical limit.
Renormalization also applies to the coupling between different sites. In Sunny, couplings will often be expressed as a polynomial of spin operators using set_pair_coupling!
, but any such coupling can be decomposed as a sum of tensor products of Stevens operators. Without loss of generality, consider a single coupling between two Stevens operators $\hat{\mathcal{H}}_\mathrm{coupling} = \hat{\mathcal{O}}_{k,q} \otimes \hat{\mathcal{O}}_{k',q'}$ along a bond connecting sites $i$ and $j$. Upon constraining to pure dipole states $|\boldsymbol{\Omega}_i\rangle$ and $|\boldsymbol{\Omega}_j\rangle$, the expected energy takes the form $E_\mathrm{coupling} = c_k c_k' \mathcal{O}_{k,q}(\boldsymbol{\Omega}_i) \mathcal{O}_{k',q'}(\boldsymbol{\Omega}_j)$, which now involves a product of renormalized Stevens functions.
Use :dipole_uncorrected
mode to disable renormalization
The above renormalization procedure is valid under the assumption that the starting model is the true microscopic quantum Hamiltonian. Sometimes, however, the starting model is instead an effective classical Hamiltonian that has been fitted to experimental data. In this case, the model parameters will already incorporate any appropriate renormalizations, and no further renormalization should be applied. Similarly, renormalization is unneeded for effective models of micromagnets that are far from the quantum regime.
To specify an effective spin-dipole Hamiltonian with renormalization disabled, construct a System
using the mode :dipole_uncorrected
instead of :dipole
. Formally, :dipole_uncorrected
takes the $s → ∞$ limit, such that all local operators become infinite dimensional and commute. A symbolic representation of these operators can be obtained by passing Inf
to either spin_matrices
or stevens_matrices
. Polynomials of such spin operators can be used, e.g., in set_onsite_coupling!
.
Definition of Stevens operators
The Stevens operators $\hat{\mathcal{O}}_{k,q}$ are defined as polynomials of angular momentum operators $\hat{S}_{\{x,y,z\}}$ in some spin-$s$ representation.
Using
\[\begin{align*} X &= \mathbf{\hat{S}} \cdot \mathbf{\hat{S}} = s (s+1) \\ \hat{S}_\pm &= \hat{S}_x \pm i \hat{S}_y \\ \phi_+ &= \frac{1}{4},\quad \phi_- = \frac{1}{4 i}, \end{align*}\]
the relevant Stevens operators are defined as,
\[\begin{align*} \hat{\mathcal{O}}_{0,0} & =1\\ \\ \hat{\mathcal{O}}_{1,\pm1} & =\phi_{\pm}(\hat{S}_{+}\pm\hat{S}_{-})+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{1,0} & =\hat{S}_{z}\\ \\ \hat{\mathcal{O}}_{2,\pm2} & =\phi_{\pm}(\hat{S}_{+}^{2}\pm\hat{S}_{-}^{2})+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{2,\pm1} & =\phi_{\pm}(\hat{S}_{+}\pm\hat{S}_{-})\hat{S}_{z}+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{2,0} & =3\hat{S}_{z}^{2}-X\\ \\ \hat{\mathcal{O}}_{3,\pm3} & =\phi_{\pm}(\hat{S}_{+}^{3}\pm\hat{S}_{-}^{3})+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{3,\pm2} & =\phi_{\pm}(\hat{S}_{+}^{2}\pm\hat{S}_{-}^{2})\hat{S}_{z}+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{3,\pm1} & =\phi_{\pm}(\hat{S}_{+}\pm\hat{S}_{-})(5\hat{S}_{z}^{2}-X-1/2)+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{3,0} & =5\hat{S}_{z}^{3}-(3X-1)\hat{S}_{z}\\ \\ \hat{\mathcal{O}}_{4,\pm4} & =\phi_{\pm}(\hat{S}_{+}^{4}\pm\hat{S}_{-}^{4})+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{4,\pm3} & =\phi_{\pm}(\hat{S}_{+}^{3}\pm\hat{S}_{-}^{3})\hat{S}_{z}+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{4,\pm2} & =\phi_{\pm}(\hat{S}_{+}^{2}\pm\hat{S}_{-}^{2})(7\hat{S}_{z}^{2}-(X+5))+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{4,\pm1} & =\phi_{\pm}(\hat{S}_{+}\pm\hat{S}_{-})(7\hat{S}_{z}^{3}-(3X+1)\hat{S}_{z})+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{4,0} & =35\hat{S}_{z}^{4}-(30X-25)\hat{S}_{z}^{2}+(3X^{2}-6X)\\ \\ \hat{\mathcal{O}}_{5,\pm5} & =\phi_{\pm}(\hat{S}_{+}^{5}\pm\hat{S}_{-}^{5})+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{5,\pm4} & =\phi_{\pm}(\hat{S}_{+}^{4}\pm\hat{S}_{-}^{4})\hat{S}_{z}+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{5,\pm3} & =\phi_{\pm}(\hat{S}_{+}^{3}\pm\hat{S}_{-}^{3})(9\hat{S}_{z}^{2}-(X+33/2))+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{5,\pm2} & =\phi_{\pm}(\hat{S}_{+}^{2}\pm\hat{S}_{-}^{2})(3\hat{S}_{z}^{3}-(X+6)\hat{S}_{z})+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{5,\pm1} & =\phi_{\pm}(\hat{S}_{+}\pm\hat{S}_{-})(21\hat{S}_{z}^{4}-14X\hat{S}_{z}^{2}+(X^{2}-X+3/2))+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{5,0} & =63\hat{S}_{z}^{5}-(70X-105)\hat{S}_{z}^{3}+(15X^{2}-50X+12)\hat{S}_{z}\\ \\ \hat{\mathcal{O}}_{6,\pm6} & =\phi_{\pm}(\hat{S}_{+}^{6}\pm\hat{S}_{-}^{6})+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{6,\pm5} & =\phi_{\pm}(\hat{S}_{+}^{5}\pm\hat{S}_{-}^{5})\hat{S}_{z}+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{6,\pm4} & =\phi_{\pm}(\hat{S}_{+}^{4}\pm\hat{S}_{-}^{4})(11\hat{S}_{z}^{2}-X-38)+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{6,\pm3} & =\phi_{\pm}(\hat{S}_{+}^{3}\pm\hat{S}_{-}^{3})(11\hat{S}_{z}^{3}-(3X+59)\hat{S}_{z})+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{6,\pm2} & =\phi_{\pm}(\hat{S}_{+}^{2}\pm\hat{S}_{-}^{2})(33\hat{S}_{z}^{4}-(18X+123)\hat{S}_{z}^{2}+X^{2}+10X+102)+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{6,\pm1} & =\phi_{\pm}(\hat{S}_{+}\pm\hat{S}_{-})(33\hat{S}_{z}^{5}-(30X-15)\hat{S}_{z}^{3}+(5X^{2}-10X+12)\hat{S}_{z})+\mathrm{h.c.}\\ \hat{\mathcal{O}}_{6,0} & =231\hat{S}_{z}^{6}-(315X-735)\hat{S}_{z}^{4}+(105X^{2}-525X+294)\hat{S}_{z}^{2}-5X^{3}+40X^{2}-60X \end{align*}\]
Computer-generated tables of Stevens operators with $k > 6$ are available from C. Rudowicz and C. Y. Chung, J. Phys.: Condens. Matter 16, 5825 (2004), but these typically do not appear in magnetic simulations.
The case $k=1$ gives the dipole operators,
\[(\hat{\mathcal{O}}_{1,1}, \hat{\mathcal{O}}_{1,0}, \hat{\mathcal{O}}_{1,-1}) = (\hat{S}_{x}, \hat{S}_{z}, \hat{S}_{y}).\]
The case $k=2$ gives the quadrupole operators,
\[(\hat{\mathcal{O}}_{2,2}, \dots, \hat{\mathcal{O}}_{2,-2}) = \left(\hat{S}_x^2 - \hat{S}_y^2, \frac{\hat{S}_x \hat{S}_z + \hat{S}_z \hat{S}_x}{2}, 2\hat{S}_z^2-\hat{S}_x^2-\hat{S}_y^2, \frac{\hat{S}_y \hat{S}_z + \hat{S}_z \hat{S}_y}{2}, \hat{S}_x \hat{S}_y + \hat{S}_y \hat{S}_x\right).\]
For each $k$ value, the set of operators $\hat{\mathcal{O}}_{k,q'}$ for $q' = -k, \dots, k$ form an irreducible representation of the group of rotations O(3). That is, rotation will transform $\hat{\mathcal{O}}_{k,q}$ into a linear combination of $\hat{\mathcal{O}}_{k,q'}$ where $q'$ varies but $k$ remains fixed.
In taking the large-$s$ limit, each dipole operator is replaced by its expectation value $\boldsymbol{\Omega} = \langle \hat{\mathbf{S}} \rangle$, and only leading-order terms are retained. The operator $\hat{\mathcal{O}}_{k,q}$ becomes a homogeneous polynomial $O_{k,q}(\boldsymbol{\Omega})$ of order $k$ in the spin components $\Omega^\alpha$. One can see these polynomials by constructing stevens_matrices
with the argument s = Inf
. Due to the normalization constraint, each dipole can be expressed in polar angles, $(\theta, \phi)$. Then the Stevens functions $O_{k,q}(\boldsymbol{\Omega})$ correspond to the spherical harmonic functions $Y_l^m(\theta, \phi)$ where $l=k$ and $m=q$; this correspondence is valid up to $k$ and $q$-dependent rescaling factors.