Fact-checked by Grok 2 weeks ago

Vector spherical harmonics

Vector spherical harmonics are a set of vector-valued defined on the unit sphere, extending the scalar to facilitate the expansion of vector fields in spherical coordinates. They serve as solutions to the vector , \nabla^2 \mathbf{V} + k^2 \mathbf{V} = 0, where the components are constructed from scalar Y_{\ell m}(\theta, \phi) combined with differential operators or unit vectors. Typically denoted as \mathbf{Y}_{j \ell m} or similar, these functions are simultaneous eigenfunctions of the total \mathbf{J}^2, orbital angular momentum \mathbf{L}^2, and J_z, with eigenvalues j(j+1), \ell(\ell+1), and m, respectively, where j = \ell or j = \ell \pm 1, and |m| \leq j. The construction of vector spherical harmonics involves applying the \mathbf{L} = -i \mathbf{r} \times \nabla to scalar harmonics or using vector couplings. For the transverse electric mode (often labeled \mathbf{\Psi}_{j \ell m}), one form is \mathbf{\Psi}_{j \ell m} = \frac{1}{\sqrt{j(j+1)}} \mathbf{L} Y_{j m}, which is divergence-free and perpendicular to the radial direction. Other modes include radial-longitudinal and transverse-magnetic types. Key properties include orthonormality over the sphere, \int \mathbf{Y}_{j \ell}^{m*} \cdot \mathbf{Y}_{j' \ell'}^{m'} \, d\Omega = \delta_{j j'} \delta_{\ell \ell'} \delta_{m m'}, and completeness for expanding tangential vector fields, enabling separation of radial and angular dependencies in partial differential equations. They also exhibit transformations and properties inherited from scalar harmonics, with specific sets being solenoidal (\nabla \cdot \mathbf{V} = 0) or irrotational. Vector spherical harmonics find extensive applications in for expanding radiation fields from multipole sources, as detailed in classical treatments of wave propagation and by spherical objects. In magnetostatics, they decompose magnetic fields from current distributions, satisfying for efficient multipole expansions. Beyond physics, they are employed in for angular momentum states of particles with , in for solving the Navier-Stokes equations around spheres (e.g., ), and in astronomy for analyzing vector fields on the , such as proper motions or magnetic fields.

Mathematical Foundations

Scalar Spherical Harmonics

Scalar spherical harmonics, denoted Y_{l}^{m}(\theta, \phi), where l is the degree (a non-negative integer) and m is the order (an integer satisfying -l \leq m \leq l), arise as the solutions to the angular part of Laplace's equation in spherical coordinates. These functions provide a complete orthonormal basis for the space of square-integrable scalar functions on the unit sphere, enabling the expansion of any such function in terms of these harmonics. Pierre-Simon Laplace first introduced them in 1782 as part of his work on gravitational potentials, representing the coefficients in the expansion of the Newtonian potential. The explicit form of the scalar spherical harmonics is given by Y_{l}^{m}(\theta, \phi) = (-1)^{m} \sqrt{ \frac{(2l+1)(l-m)!}{4\pi (l+m)!} } \, P_{l}^{m}(\cos \theta) \, e^{i m \phi}, for m \geq 0, with Y_{l}^{-m}(\theta, \phi) = (-1)^{m} (Y_{l}^{m})^{*}(\theta, \phi) for negative orders, where P_{l}^{m} are the associated Legendre functions of the first kind. The factor (-1)^{m} for positive m incorporates the Condon-Shortley phase convention, which ensures consistency in applications such as calculations in . Key properties include their orthogonality over the sphere: \int_{0}^{2\pi} \int_{0}^{\pi} Y_{l}^{m*}(\theta, \phi) \, Y_{l'}^{m'}(\theta, \phi) \, \sin \theta \, d\theta \, d\phi = \delta_{l l'} \, \delta_{m m'}, which follows from the Sturm-Liouville theory applied to the associated Legendre equation, along with their completeness, allowing any L^{2} function on the sphere to be uniquely expanded as f(\theta, \phi) = \sum_{l=0}^{\infty} \sum_{m=-l}^{l} f_{l m} Y_{l}^{m}(\theta, \phi). Additionally, the addition theorem states that P_{l}(\cos \gamma) = \frac{4\pi}{2l+1} \sum_{m=-l}^{l} Y_{l}^{m*}(\theta', \phi') \, Y_{l}^{m}(\theta, \phi), where \gamma is the angle between the directions (\theta, \phi) and (\theta', \phi'), facilitating rotations and multipole expansions. In , these harmonics describe the angular dependence of the atom's wave functions, \psi_{n l m}(r, \theta, \phi) = R_{n l}(r) \, Y_{l}^{m}(\theta, \phi), where they correspond to the eigenfunctions of the operators.

Vector Fields on the Sphere

A \mathbf{V} defined on the unit sphere S^2 in three-dimensional is represented in spherical coordinates (\theta, \phi) as \mathbf{V} = V_r \hat{r} + V_\theta \hat{\theta} + V_\phi \hat{\phi}, where V_r, V_\theta, and V_\phi are the components along the local vectors \hat{r}, \hat{\theta}, and \hat{\phi}, respectively. The radial component V_r \hat{r} points normal to the sphere, while the tangential component \mathbf{V}_T = V_\theta \hat{\theta} + V_\phi \hat{\phi} lies entirely within the tangent plane at each point. The basis vectors \hat{\theta} and \hat{\phi} are defined as the directions of increasing polar angle \theta and azimuthal angle \phi, with explicit Cartesian expressions \hat{\theta} = \cos\theta \cos\phi \, \hat{i} + \cos\theta \sin\phi \, \hat{j} - \sin\theta \, \hat{k} and \hat{\phi} = -\sin\phi \, \hat{i} + \cos\phi \, \hat{j}. The tangential vector fields on S^2 admit an irreducible decomposition under the action of the rotation group SO(3) into two distinct parts: a polar (irrotational) component and an axial (solenoidal) component. This decomposition aligns with the Helmholtz-Hodge theorem adapted to the sphere, which uniquely expresses any tangential \mathbf{V}_T as the sum of a curl-free part \nabla_s f (the surface of a f) and a divergence-free part \hat{r} \times \nabla_s [g](/page/G) (the surface curl of another g), where \nabla_s denotes the surface projecting the Euclidean onto the tangent via the P_s = I - \hat{r} \hat{r}^T. The polar part transforms as a polar under rotations, while the axial part transforms as an axial , reflecting their distinct behaviors. For fields of fixed degree l, the space of such tangential vector fields decomposes into two irreducible SO(3)-representations, each of dimension $2l + 1, yielding a total dimension of $2(2l + 1). This structure necessitates two separate families of basis functions to span the space completely, with the components derived by applying the surface gradient and curl operators to scalar functions of degree l.

Definitions

Standard Definition

Vector spherical harmonics are defined for degrees l \geq 1 and orders m with |m| \leq l; they vanish for l = 0. The two primary transverse families are the magnetic-type and electric-type harmonics, both tangential to the unit sphere and constructed from the scalar spherical harmonics Y_{lm}(\theta, \phi). The magnetic-type vector spherical harmonics are given by \mathbf{Y}_{lm}^{(M)}(\theta, \phi) = \frac{1}{\sqrt{l(l+1)}} \mathbf{L} Y_{lm}, where \mathbf{L} = -i \mathbf{r} \times \nabla is the . These functions are divergence-free on the sphere and form an for tangential fields with that property. The electric-type vector spherical harmonics are \mathbf{Y}_{lm}^{(E)}(\theta, \phi) = \frac{1}{\sqrt{l(l+1)}} \nabla_1 Y_{lm}, where \nabla_1 denotes the surface gradient operator on the unit sphere. These are curl-free on the sphere (up to a factor) and complete the for tangential fields when combined with the magnetic type. An alternative notation uses \Psi_{lm} for the magnetic type and \Phi_{lm} for the electric type. In the spherical basis \{\hat{r}, \hat{\theta}, \hat{\phi}\}, the components can be expressed explicitly; for example, the \theta-component of \Psi_{lm} is \Psi_{lm}^\theta = -\frac{m}{\sqrt{l(l+1)} \sin \theta} Y_{lm}. Both transverse families satisfy the normalization condition \int |\mathbf{Y}_{lm}^{(J)}|^2 \, d\Omega = 1, \quad J = M, E, where the integral is over the unit sphere. For completeness, a longitudinal type is defined as \mathbf{Y}_{lm}^{(L)}(\theta, \phi) = \hat{r} Y_{lm}(\theta, \phi), which is non-tangential and includes only a radial component.

Alternative Definitions

In electrodynamics, particularly for expanding solutions to in spherical coordinates, an alternative pair of vector spherical harmonics is employed to separate transverse electric () and transverse magnetic (TM) modes. The magnetic-type harmonic is given by \mathbf{M}_{lm}(\theta, \phi) = -\frac{i}{\sqrt{l(l+1)}} \mathbf{r} \times \nabla Y_{lm}(\theta, \phi), while the electric-type harmonic is \mathbf{N}_{lm}(\theta, \phi) = \frac{1}{k \sqrt{l(l+1)}} \nabla \times \mathbf{L} Y_{lm}(\theta, \phi), where k is the wave number and \mathbf{L} = -i \mathbf{r} \times \nabla. In quantum mechanics, vector spherical harmonics are often denoted as \mathbf{Y}_{J L M}(\theta, \phi), where L is the orbital angular momentum quantum number, J is the total angular momentum (J = L \pm 1 or J = L for vector coupling with spin-1), and M is the projection along the z-axis. These are constructed via angular momentum addition: \mathbf{Y}_{J L M}(\theta, \phi) = \sum_{m_L, m_S} \langle L m_L, 1 m_S | J M \rangle Y_{L m_L}(\theta, \phi) \boldsymbol{\chi}_{1 m_S}, with \boldsymbol{\chi}_{1 m_S} as the spherical basis vectors for spin-1. This notation emphasizes the coupling of orbital and intrinsic angular momentum, useful for describing particle states with definite total J. For J = L, the form is transverse like \mathbf{M}_{lm}; for J = L \pm 1, it includes longitudinal components analogous to \mathbf{N}_{lm}. A geometrical approach defines vector spherical harmonics using spin-weighted spherical harmonics {}_s Y_{lm}(\theta, \phi) with spin weight s = \pm 1, which naturally describe fields on the sphere via a local dyad basis (e.g., \mathbf{m}, \bar{\mathbf{m}} for ). A \mathbf{F} decomposes into components F_{+1} = \mathbf{F} \cdot \mathbf{m} (spin +1) and F_{-1} = \mathbf{F} \cdot \bar{\mathbf{m}} (spin -1), expanded as F_{\pm 1} = \sum_{l m} a_{\pm 1, lm} {}_{\pm 1} Y_{lm}. These relate to standard vector harmonics through basis transformations, with {}_{\pm 1} Y_{lm} \propto (\nabla_\theta \pm i \nabla_\phi / \sin\theta) Y_{lm}, facilitating analysis of and gravitational perturbations. Conventions vary across literature, notably in normalization and phase factors. Jackson includes the factor of -i in \mathbf{M}_{lm} for convenience in radiation problems, ensuring real-valued fields for certain modes, while Stratton omits the i and uses \sqrt{(l \pm 1)/(2l+1)} factors in some decompositions for consistency with scalar harmonics. These differences arise from choices in orthonormal bases but preserve integrals.

Properties

Orthogonality and Normalization

The inner product for vector fields on the unit is defined as \langle \mathbf{U}, \mathbf{V} \rangle = \int \mathbf{U}^* \cdot \mathbf{V} \, d\Omega, where the integral extends over the d\Omega = \sin\theta \, d\theta \, d\phi. This inner product induces an L^2 structure on vector fields, including radial and tangential components. The vector spherical harmonics \mathbf{Y}_{lm}^{(J)} satisfy relations with respect to this inner product. For the modes J, J' = M, E, \langle \mathbf{Y}_{lm}^{(J)}, \mathbf{Y}_{l'm'}^{(J')} \rangle = \delta_{ll'} \delta_{mm'} \delta_{JJ'}, while cross terms between different modes vanish. The radial mode \mathbf{Y}_{lm}^{(L)} is orthogonal to both M and E modes and satisfies \langle \mathbf{Y}_{lm}^{(L)}, \mathbf{Y}_{l'm'}^{(L)} \rangle = \delta_{ll'} \delta_{mm'}. These properties extend the of scalar to the vector case. The tangential vector spherical harmonics \mathbf{Y}_{lm}^{(M)} and \mathbf{Y}_{lm}^{(E)} form a complete orthonormal basis for the space of square-integrable tangential vector fields on the sphere. Any such field \mathbf{V}_T admits the expansion \mathbf{V}_T(\theta, \phi) = \sum_{l=1}^\infty \sum_{m=-l}^l \left[ a_{lm}^{(M)} \mathbf{Y}_{lm}^{(M)}(\theta, \phi) + a_{lm}^{(E)} \mathbf{Y}_{lm}^{(E)}(\theta, \phi) \right], where the coefficients are given by the projections a_{lm}^{(J)} = \langle \mathbf{Y}_{lm}^{(J)}, \mathbf{V}_T \rangle = \int \mathbf{Y}_{lm}^{(J)*} \cdot \mathbf{V}_T \, d\Omega for J = M, E. Together with the radial modes \mathbf{Y}_{lm}^{(L)}, they provide a complete basis for all square-integrable vector fields on the sphere. The radial harmonics \mathbf{Y}_{lm}^{(L)} provide a complete basis for the subspace of radial vector fields. The normalization of the vector spherical harmonics derives from integrals involving scalar spherical harmonics. In particular, the surface operator \nabla_1 on the sphere satisfies \int |\nabla_1 Y_{lm}|^2 \, d\Omega = l(l+1) \int |Y_{lm}|^2 \, d\Omega, which follows from the eigenvalue equation for the spherical Laplacian -\Delta_1 Y_{lm} = l(l+1) Y_{lm} and . This relation ensures that the tangential harmonics \mathbf{Y}_{lm}^{(M)} and \mathbf{Y}_{lm}^{(E)}, constructed via and of Y_{lm}, are normalized with factors such as $1/\sqrt{l(l+1)}. The radial mode \mathbf{Y}_{lm}^{(L)} is normalized directly from the scalar harmonic without the l(l+1) factor. Orthonormality implies the Parseval identity for tangential vector fields: \int |\mathbf{V}_T|^2 \, d\Omega = \sum_{l=1}^\infty \sum_{m=-l}^l \sum_{J=M,E} |a_{lm}^{(J)}|^2. An analogous identity holds for the radial subspace.

Symmetry and Parity

Vector spherical harmonics transform under s as irreducible s of the SO(3) group, carrying l. Specifically, for a R, the unitary representation operator U(R) acts on the basis functions as U(R) \mathbf{Y}_{l m}^{(J)} = \sum_{m'} D_{m' m}^l (R) \mathbf{Y}_{l m'}^{(J)}, where D^l_{m'm}(R) are the Wigner D-matrices and J denotes the type (electric, magnetic, or longitudinal). Under parity transformation \mathbf{r} \to -\mathbf{r}, which on the unit corresponds to (\theta, \phi) \to (\pi - \theta, \phi + \pi), the magnetic vector \mathbf{Y}_{l m}^{(M)} are even, acquiring a P = +1 relative to the scalar ' parity (-1)^l, while the electric \mathbf{Y}_{l m}^{(E)} and longitudinal \mathbf{Y}_{l m}^{(L)} are with P = -1. This behavior links directly to multipole expansions, where even- modes correspond to certain patterns and odd- to others. The distinction between types arises from their vector nature: magnetic harmonics \mathbf{Y}^{(M)} behave as axial vectors (pseudovectors), transforming without an extra sign under inversion compared to polar vectors, whereas electric \mathbf{Y}^{(E)} and radial longitudinal \mathbf{Y}^{(L)} are true polar vectors. Under full spatial inversion, these properties ensure consistent classification in multipole parity, with magnetic modes preserving orientation relative to the and electric/radial modes flipping. Tangential vector spherical harmonics, both electric and magnetic, have zero radial component, ensuring they lie in the tangential plane. This orthogonality to \mathbf{\hat{r}} facilitates their use in divergence-free or curl-free decompositions.

Differential Operator Relations

Vector spherical harmonics interact with differential operators defined on the sphere through relations that reflect their construction from scalar spherical harmonics and their role in decomposing vector fields into irreducible representations under rotations. These relations are particularly useful for solving partial differential equations on spherical geometries, such as those in electromagnetism and fluid dynamics. The surface gradient operator \nabla_1, acting on a scalar spherical harmonic Y_{lm}, yields a tangential proportional to the electric-type vector spherical harmonic \mathbf{Y}_{lm}^{(E)}: \nabla_1 Y_{lm} = \sqrt{l(l+1)} \, \mathbf{Y}_{lm}^{(E)}. This relation follows from the definition of \mathbf{Y}_{lm}^{(E)} as the normalized tangential of the scalar harmonic, ensuring and unit norm on the unit . The surface curl operator \mathrm{curl}_1, applied to a radial vector field f \hat{r}, produces a tangential vector field aligned with the magnetic-type vector spherical harmonic \mathbf{Y}_{lm}^{(M)}. Specifically, for f = Y_{lm}, \mathrm{curl}_1 (Y_{lm} \hat{r}) = \hat{r} \times \nabla_1 Y_{lm} = i \sqrt{l(l+1)} \, \mathbf{Y}_{lm}^{(M)}, up to normalization conventions that may absorb the for phase consistency in applications. This arises because \mathbf{Y}_{lm}^{(M)} is defined via the normalized \hat{r} \times \nabla_1 Y_{lm}, making it divergence-free and solenoidal. The surface divergence operator \mathrm{div}_1 highlights the properties of the tangential vector spherical harmonics. For the magnetic component, \mathrm{div}_1 \mathbf{Y}_{lm}^{(M)} = 0, reflecting its incompressibility on the sphere, whereas the electric harmonic satisfies \mathrm{div}_1 \mathbf{Y}_{lm}^{(E)} = -\sqrt{l(l+1)} \, Y_{lm}. These properties stem from the on the sphere, where \mathbf{Y}_{lm}^{(M)} forms the solenoidal basis, and \mathbf{Y}_{lm}^{(E)} the irrotational part. The radial mode \mathbf{Y}_{lm}^{(L)} has no tangential components, so \mathrm{div}_1 does not apply. The spherical Laplacian \Delta_1 acts as an eigenvalue operator on the tangential vector spherical harmonics: \Delta_1 \mathbf{Y}_{lm}^{(J)} = -l(l+1) \, \mathbf{Y}_{lm}^{(J)}, for J = E, M, analogous to the scalar case \Delta_1 Y_{lm} = -l(l+1) Y_{lm}. This eigenvalue equation underscores their role as eigenfunctions of the on the sphere. Extending to three dimensions, the full and operators relate radial and tangential components. For a radial , \nabla \times (\mathbf{r} f) with f = Y_{lm} yields a vector field proportional to \mathbf{Y}_{lm}^{(M)}, specifically \nabla \times (\mathbf{r} Y_{lm}) = i r \sqrt{l(l+1)} \, \mathbf{Y}_{lm}^{(M)}, mirroring the surface up to radial scaling. Similarly, the of a general \mathbf{A} involves \nabla (\mathbf{r} \cdot \mathbf{A}), which couples to the longitudinal mode through the . These 3D relations facilitate expansions of solutions to vector wave equations.

Multipole Moments

Vector multipole moments serve as coefficients in the expansion of the generated by a localized vector source, such as a distribution confined within a finite . For points outside the sources, the vector potential \mathbf{A}(\mathbf{r}) admits the \mathbf{A}(\mathbf{r}) = \sum_{l=1}^{\infty} \sum_{m=-l}^{l} \left[ \mathbf{M}_{l m}(\hat{\mathbf{r}}) \frac{q_{l m}^{(M)}}{r^{l+1}} + \mathbf{N}_{l m}(\hat{\mathbf{r}}) \frac{q_{l m}^{(E)}}{r^{l+1}} \right], where \mathbf{M}_{l m} and \mathbf{N}_{l m} denote the toroidal and poloidal vector spherical harmonics, respectively, and the multipole moments q_{l m}^{(J)} (with J = M, E) are defined by the volume integral q_{l m}^{(J)} = \int \mathbf{j}(\mathbf{r}') \cdot \mathbf{Y}_{l m}^{(J)*} (\hat{\mathbf{r}}') \, r'^l \, dV', with \mathbf{j}(\mathbf{r}') the current density and \mathbf{Y}_{l m}^{(J)} the appropriate vector spherical harmonics conjugate to \mathbf{M}_{l m} or \mathbf{N}_{l m}. The electric (polar) multipole moments q_{l m}^{(E)} originate from charge distributions and the longitudinal components of the current, contributing to the poloidal part of the field, whereas the magnetic (toroidal) moments q_{l m}^{(M)} arise from transverse currents or equivalent magnetic sources, driving the toroidal field structure. In the context of far-field radiation for time-harmonic sources, the static radial dependence $1/r^{l+1} generalizes to outgoing spherical waves via spherical Hankel functions of the first kind, h_l^{(1)}(kr), which incorporate phase factors e^{ikr} and scaling with kr for large kr \gg 1, enabling the description of radiating multipoles. For the dipole case l=1, these vector multipole moments reduce to the familiar scalar electric and magnetic dipole moments, bridging the vector expansion to the standard scalar multipole theory. This expansion is unique, owing to the completeness and orthogonality of the vector spherical harmonics, and it converges for all r larger than the radial extent of the sources.

Explicit Forms and Examples

Construction Methods

Vector spherical harmonics are typically constructed from scalar spherical harmonics by applying tangential components of the gradient and curl operators, yielding the electric (poloidal) and magnetic (toroidal) modes, respectively. The electric-type vector spherical harmonic is expressed in spherical coordinates as \mathbf{Y}_{l m}^{(E)}(\theta, \phi) = \frac{1}{\sqrt{l(l+1)}} \left( \hat{\theta} \frac{\partial Y_{l m}}{\partial \theta} + \hat{\phi} \frac{i m Y_{l m}}{\sin \theta} \right), where Y_{l m} denotes the scalar spherical harmonic. This form arises from the tangential gradient of the scalar harmonic, normalized to ensure unit norm over the sphere. The magnetic-type counterpart is obtained via azimuthal derivatives, given by \mathbf{Y}_{l m}^{(M)}(\theta, \phi) = \frac{1}{\sqrt{l(l+1)}} \left( \hat{\theta} \frac{i m Y_{l m}}{\sin \theta} - \hat{\phi} \frac{\partial Y_{l m}}{\partial \theta} \right), corresponding to the tangential curl and ensuring orthogonality to the electric mode. Recursive relations facilitate computation across magnetic quantum numbers m. Raising and lowering operators J_{\pm} act on the scalar Y_{l m} as J_{\pm} Y_{l m} = \hbar \sqrt{(l \mp m)(l \pm m + 1)} Y_{l, m \pm 1}, after which the vector construction is applied to generate \mathbf{Y}_{l, m \pm 1}^{(E/M)}. These ladder operators preserve the vector harmonic properties under rotation. Integral representations employ addition theorems, expressing vector spherical harmonics in terms of zonal harmonics for special cases like m=0. For instance, the addition theorem for vector modes links them to scalar addition formulas via tensor products, enabling efficient evaluation in multipole expansions. Numerical implementation relies on fast recursion for associated Legendre functions underlying Y_{l m}, using three-term relations like (l - m + 1) P_l^m(x) = (2l - 1) x P_{l-1}^m(x) - (l + m - 1) P_{l-2}^m(x) to compute values stably from the toward the poles, avoiding singularities at \theta = 0, \pi. Derivatives for the vector components are then evaluated analytically from these. As of 2025, libraries such as Mathematica provide built-in construction via SphericalHarmonicY combined with operators, while Python's integrates with packages like Windspharm for vector spherical harmonic transforms and computations.

Low-Degree Examples

The lowest-degree vector spherical harmonics illustrate the distinct structures of electric and magnetic modes, with the electric modes corresponding to poloidal fields and the magnetic modes to toroidal fields. For the case of degree l=1 and order m=0, using Y_{10} = \sqrt{\frac{3}{4\pi}} \cos \theta, the electric mode is \mathbf{Y}_{10}^{(E)} = \sqrt{\frac{3}{8\pi}} \sin \theta \, \hat{\theta}, while the magnetic mode is \mathbf{Y}_{10}^{(M)} = \sqrt{\frac{3}{8\pi}} \sin \theta \, \hat{\phi}, (up to sign convention). Both exhibit the characteristic dipole pattern with a single nodal line at the equator and maximum intensity in the xy-plane. For l=1 and m=1, using Y_{11} = -\sqrt{\frac{3}{8\pi}} \sin \theta \, e^{i\phi}, the electric mode has components \mathbf{Y}_{11}^{(E)} = -\sqrt{\frac{3}{8\pi}} \frac{1}{\sqrt{2}} \left( \cos \theta \, e^{i\phi} \, \hat{\theta} + i \, e^{i\phi} \, \hat{\phi} \right), with the m=-1 counterpart obtained by complex conjugation; real linear combinations of these, such as those proportional to \sin \theta \cos \phi \, \hat{\theta} and \sin \theta \sin \phi \, \hat{\phi}, correspond to Cartesian-oriented dipoles like the p_x pattern, featuring tilted lobes and a nodal plane. The associated magnetic mode has components \mathbf{Y}_{11}^{(M)} = -i\sqrt{\frac{3}{8\pi}} \frac{1}{\sqrt{2}} \left( e^{i\phi} \, \hat{\theta} + \cos \theta \, e^{i\phi} \, \hat{\phi} \right), showing a toroidal structure with azimuthal variation. At degree l=2 and order m=0, the quadrupole examples demonstrate higher complexity; the magnetic mode has components \mathbf{Y}_{20}^{(M)} \propto \sin \theta \cos \theta \, \hat{\phi}, derived from the associated Legendre function P_2(\cos \theta) = \frac{1}{2}(3 \cos^2 \theta - 1), while the electric mode involves \mathbf{Y}_{20}^{(E)} \propto \frac{d P_2}{d \theta} \, \hat{\theta} = -3 \sin \theta \cos \theta \, \hat{\theta}. These exhibit four lobes symmetric about the z-axis, with nodal lines at \theta = \pi/2 and \theta = \cos^{-1}(1/\sqrt{3}) for the electric case, contrasting the purely azimuthal flow of the magnetic mode. Qualitatively, plots of these functions reveal the electric modes as divergence-free with poloidal streamlines concentrated along symmetry axes, featuring doughnut-shaped nodal surfaces, whereas magnetic modes display curl-dominated toroidal loops with zero divergence and nodal circles perpendicular to the axis; for instance, the l=1 electric shows equatorial nulls, while the l=2 has additional meridional nodes. A special case arises in representing a uniform , which can be decomposed as a of the l=1 modes, specifically involving the radial scalar coupled with the m=0 electric to yield constant magnitude across the sphere.

Applications

Electromagnetism

In the static case, vector spherical harmonics provide a natural basis for expanding magnetostatic fields generated by localized current distributions. The magnetic induction field \mathbf{B} outside the source region can be expressed as \mathbf{B}(\mathbf{r}) = \sum_{l,m} \nabla \times \left( \frac{\mathbf{M}_{lm}}{r^{l+1}} \right), where the multipole moments \mathbf{M}_{lm} are determined by integrals over the current density \mathbf{J}(\mathbf{r}'): \mathbf{M}_{lm} = \int \mathbf{J}(\mathbf{r}') r'^l Y_{lm}(\theta',\phi') \, d^3\mathbf{r}'. This expansion parallels the scalar multipole series for electrostatics but accounts for the vector nature of \mathbf{B} through the curl operator and the tangential vector spherical harmonics \mathbf{M}_{lm}. For dynamic electromagnetic radiation from time-harmonic sources, the far-field electric and magnetic fields are decomposed into transverse electric (TE) and transverse magnetic (TM) modes using vector spherical harmonics. The outgoing wave solutions involve the first-kind spherical Hankel functions h_l^{(1)}(kr) to capture the spherical wave propagation. Specifically, the TM (electric multipole) mode contributes to the electric field as \mathbf{E}^{\text{TM}}(\mathbf{r}) \propto \nabla \times \left[ \mathbf{r} \mathbf{N}_{lm}(\hat{\mathbf{r}}) h_l^{(1)}(kr) \right], while the TE (magnetic multipole) mode is \mathbf{B}^{\text{TE}}(\mathbf{r}) \propto \nabla \times \left[ \mathbf{r} \mathbf{M}_{lm}(\hat{\mathbf{r}}) h_l^{(1)}(kr) \right], with the corresponding orthogonal field components derived from Maxwell's equations. In the far zone (kr \gg 1), these simplify to transverse plane-wave-like forms, enabling the computation of radiation patterns and power via mode coefficients related to source moments. In Mie scattering theory for electromagnetic plane waves incident on a homogeneous spherical particle, the incident field is expanded in regular vector spherical wave functions centered at the particle: \mathbf{E}^{\text{inc}}(\mathbf{r}) = \sum_{l=1}^\infty \sum_{m=-l}^l \left[ a_{lm} \mathbf{M}_{lm}(k r) + b_{lm} \mathbf{N}_{lm}(k r) \right], where \mathbf{M}_{lm} and \mathbf{N}_{lm} are the TE and TM modes with radial dependence given by spherical j_l(kr), and coefficients a_{lm}, b_{lm} depend on the wave's and direction. The scattered field employs outgoing Hankel functions h_l^{(1)}(kr), while the internal field uses adjusted for the particle's . Boundary conditions at the sphere's surface yield scattering coefficients, facilitating exact solutions for cross-sections and phase functions. Electromagnetic transitions between quantum states obey selection rules derived from the and properties of vector spherical harmonics. Angular momentum conservation requires that the photon's l satisfies |\mathbf{J}_i - \mathbf{J}_f| \leq l \leq \mathbf{J}_i + \mathbf{J}_f, where \mathbf{J}_{i,f} are the initial and final state angular momenta, excluding l=0 for J_i = J_f = 0. For , electric $2^l-pole transitions (TM modes) change parity by (-1)^l, while magnetic $2^l-pole transitions (TE modes) change it by (-1)^{l+1}, ensuring only allowed multipoles contribute to transition rates. These rules, rooted in the irreducible representations of the rotation group and parity operator, dictate forbidden and relative strengths in atomic spectra. Recent advances in the have extended vector spherical harmonics to multipole expansions in anisotropic media, relevant for designing exhibiting . By linking Cartesian and spherical formulations, these expansions enable precise modeling of field interactions in materials with spatially varying and permeability, such as those achieving negative refractive indices through resonant structures. This approach supports inverse design for control, addressing limitations in isotropic Mie-like theories for complex geometries.

Fluid Dynamics

In , vector spherical harmonics provide a natural basis for decomposing divergence-free velocity fields on spherical domains, particularly in geophysical applications such as atmospheric and oceanic circulations or flows in planetary interiors. This decomposition separates the flow into poloidal and toroidal components, which are orthogonal and satisfy the incompressibility condition \nabla \cdot \mathbf{u} = 0 by construction, leveraging the transverse properties of the vector harmonics. The velocity field \mathbf{u} is expanded as \mathbf{u} = \sum_{l,m} \left[ \nabla \times \left( \mathbf{Y}_{l m}^{(M)} \chi_{l m} \right) + \nabla \times \nabla \times \left( \mathbf{Y}_{l m}^{(E)} \psi_{l m} \right) \right], where \mathbf{Y}_{l m}^{(M)} and \mathbf{Y}_{l m}^{(E)} are the magnetic () and electric (poloidal) vector spherical harmonics, respectively, and \chi_{l m} and \psi_{l m} are scalar coefficients representing the and poloidal potentials. The term \nabla \times (\mathbf{Y}_{l m}^{(M)} \chi_{l m}) describes swirl-like motions without radial flow, while the poloidal term \nabla \times \nabla \times (\mathbf{Y}_{l m}^{(E)} \psi_{l m}) captures compressive and expansive flows aligned with meridional planes. This form builds on the differential relations of vector spherical harmonics to ensure the basis vectors are solenoidal. The vorticity \boldsymbol{\omega} = \nabla \times \mathbf{u} inherits this structure and can be expressed directly in terms of the magnetic vector spherical harmonics, facilitating the analysis of rotational flows in models of geophysical . In applications to Earth's core dynamics, such decompositions are central to magnetohydrodynamic (MHD) dynamo models, where Chandrasekhar functions—derived from vector spherical harmonics—describe the generation of the geomagnetic field through modes, particularly the l=1 dipole-dominant configurations that align with observed . These modes capture the antisymmetric flows essential for sustaining the geodynamo against ohmic . Recent advances have extended this framework to ocean circulation modeling, incorporating vector spherical harmonics to resolve flows and their interactions with mean currents. For instance, in 2023 studies of thin-shell tidal dynamics on ocean worlds, the harmonics enable efficient spectral representation of velocity perturbations from tidal forcing, improving simulations of dissipation and circulation patterns in models. This approach addresses limitations in resolving non-hydrostatic effects, enhancing predictions for contributions to mixing.

Quantum Mechanics

In quantum mechanics, vector spherical harmonics provide a natural basis for describing states with total angular momentum \vec{J} = \vec{L} + \vec{S}, where \vec{L} is the and \vec{S} is the with s=1. These harmonics, denoted \mathbf{Y}_{l,1,J,m_J}(\theta,\phi), are constructed by coupling scalar Y_{l m_l} with vector spin states \chi_{1 m_s} using Clebsch-Gordan coefficients: \mathbf{Y}_{l,1,J,m_J} = \sum_{m_l,m_s} \langle l m_l ; 1 m_s | J m_J \rangle Y_{l m_l}(\theta,\phi) \chi_{1 m_s}, where the sum satisfies m_J = m_l + m_s and |l-1| \leq J \leq l+1. This basis spans the possible total angular momenta J = l \pm 1, l, enabling the representation of vector fields or tensor operators in and systems. For particles obeying the in a central potential, such as the relativistic , the positive-energy wavefunctions incorporate vector spherical harmonics to account for spin-orbit coupling. The four-component solution takes the form \psi(\mathbf{r}) = \begin{pmatrix} g(r) \mathbf{Y}_{J L M}^{(E)}(\hat{r}) \\ i f(r) \mathbf{Y}_{J L M}^{(M)}(\hat{r}) \end{pmatrix}, where g(r) and f(r) are large and small radial components, respectively, L denotes the effective orbital index (L = J \pm 1), and the superscripts (E) and (M) distinguish even- and odd-parity vector harmonics related to electric and magnetic multipoles. These harmonics ensure the wavefunction transforms correctly under rotations, with parity determined by (-1)^{L+1}. In atomic photoionization, vector spherical harmonics facilitate the computation of for interactions between bound and states. The operator couples initial atomic states to outgoing waves via the expanded as \vec{A} \propto \sum \mathbf{Y}_{1 m}^{(E)}(\hat{r}) for electric (E1) transitions, yielding matrix of the form \langle \psi_f | \vec{\epsilon} \cdot \vec{r} | \psi_i \rangle, where \vec{\epsilon} is the . This expansion isolates selection rules (\Delta J = 0, \pm 1) and enables calculation of differential cross sections, with the of \mathbf{Y}_{J L M} simplifying integrals over the . Relativistic corrections to atomic energy levels, including , rely on vector spherical harmonics to incorporate spin-orbit and Darwin terms within the Dirac-Coulomb framework. The arises from the expectation value of the spin-orbit operator \vec{S} \cdot \vec{L}, expressed in the coupled basis where vector harmonics diagonalize total J, yielding shifts proportional to \alpha^2 / n^3 (j + 1/2) for hydrogen-like atoms, with \alpha the . Higher-order Breit interactions further refine these levels using tensor products of vector harmonics. Recent advances in employ vector spherical harmonics to describe vector vortex beams carrying orbital and singularities. These beams are expanded in multipole bases as \vec{E} = \sum a_{J L M} \mathbf{Y}_{J L M}(\hat{r}) e^{i k r}, enabling entanglement between spin and orbital in pairs.

Integral Relations

Expansion Theorems

Vector fields satisfying the vector in source-free regions admit expansions in terms of vector spherical wave functions, which incorporate radial dependencies via spherical and angular structure through vector spherical harmonics. For the interior of a of a (where r < a), the regular solution requires the spherical Bessel function of the first kind j_l(kr) to ensure finiteness at the origin, yielding the form \sum_{l=0}^\infty \sum_{m=-l}^l j_l(kr) \mathbf{Y}_{lm}^{(J)}(\hat{\mathbf{r}}), with \mathbf{Y}_{lm}^{(J)} denoting the vector spherical harmonics of type J (typically electric, magnetic, or longitudinal). For the exterior region (r > a), the expansion uses the spherical Hankel function of the first kind h_l^{(1)}(kr) to represent outgoing , as \sum_{l=0}^\infty \sum_{m=-l}^l h_l^{(1)}(kr) \mathbf{Y}_{lm}^{(J)}(\hat{\mathbf{r}}). These forms satisfy the radiation condition at infinity and the appropriate boundary behavior. The uniqueness of these expansions in source-free domains enclosing all singularities follows from the completeness and orthogonality of the vector spherical harmonics on the unit , mirroring the scalar spherical harmonic case. The Atkinson-Wilcox expansion theorem establishes that any time-harmonic in a homogeneous isotropic medium has a unique representation of the above form exterior to a containing all sources, with coefficients determined by surface integrals over an intermediate . This uniqueness holds provided the fields satisfy the and decay appropriately. In numerical implementations, the infinite series is truncated at a finite maximum l_{\max}, imposing a limitation that captures the essential features of smooth fields. For fields with k and enclosing a, truncation at l_{\max} \approx ka ensures that higher modes contribute minimally to the field's energy. of the expansion depends on the field's regularity; analytic fields converge exponentially fast with increasing l_{\max}, while C^\infty fields exhibit rapid polynomial decay. In the full 3D case, radial convergence follows from the decay of coefficients, yielding overall errors proportional to the angular truncation error for fixed k. Generalized expansions accommodate non-unit spheres by rescaling the radial coordinate r \to r/a and adjusting the argument of the accordingly, preserving the form while adapting to arbitrary radii. For weighted integrals, such as over spheres with density \mu(\hat{\mathbf{r}}), the theorems extend via weighted relations, enabling expansions for fields on manifolds with non-uniform measures; coefficients are then computed using weighted inner products, maintaining in L^2_\mu spaces. These generalizations rely on the closure properties of the harmonics and apply to tensor fields as well.

Transform Properties

Vector spherical harmonics form the basis for integral transforms that decompose vector fields into radial and angular components, particularly useful in frequency-domain analyses. The vector spherical transform projects a vector field \mathbf{V}(\mathbf{r}) onto the harmonics with a radial weighting by spherical Bessel functions, yielding coefficients \tilde{\mathbf{V}}_{l m}^{(J)}(k) = \int \mathbf{V}(\mathbf{r}) \cdot \mathbf{Y}_{l m}^{(J)*}(\hat{r}) \, j_l(kr) \, r^2 \, dr \, d\Omega, where j_l is the spherical Bessel function of the first kind, l is the degree, m the order, and J denotes the type (electric or magnetic). The inverse transform reconstructs the field as \mathbf{V}(\mathbf{r}) = \sum_{l=0}^\infty \sum_{m=-l}^l \sum_{J} \tilde{\mathbf{V}}_{l m}^{(J)}(k) \, \mathbf{Y}_{l m}^{(J)}(\hat{r}) \, j_l(kr), assuming suitable convergence conditions. This transform is orthogonal and complete for square-integrable transverse vector fields satisfying the Helmholtz equation, enabling efficient spectral representations in spherical geometries. An important application of this transform arises in the expansion of s, generalizing the scalar formula to vectors. For a plane wave propagating in direction \hat{k}, the expansion is \exp(i \mathbf{k} \cdot \mathbf{r}) \hat{\mathbf{e}} = 4\pi \sum_{l=0}^\infty \sum_{m=-l}^l i^l j_l(kr) \, \mathbf{Y}_{l m}^{(J)*}(\hat{k}, \hat{\mathbf{e}}) \, \mathbf{Y}_{l m}^{(J)}(\hat{r}), where \hat{\mathbf{e}} is the vector and the sum adapts the scalar form to vector harmonics by incorporating the appropriate J type (e.g., transverse electric or magnetic). This decomposition facilitates the analysis of wave propagation in spherical coordinates, particularly for far-field approximations. The Funk-Hecke theorem extends to vector spherical harmonics, providing a formula for integrals of the form \int_{S^2} \mathbf{Y}_{l' m'}^{(J')}(\hat{r}') \, P_n(\hat{r} \cdot \hat{r}') \, \mathbf{Y}_{l m}^{(J)*}(\hat{r}') \, d\Omega' = \delta_{l l'} \delta_{m m'} \delta_{J J'} \, c_{l n}^{(J)} \, \mathbf{Y}_{l m}^{(J)}(\hat{r}), where P_n is a Legendre polynomial and c_{l n}^{(J)} is a coefficient depending on the vector type J. This result simplifies convolutions on the sphere and is derived from the rotational invariance of the harmonics, analogous to the scalar case but accounting for vector parity. It is particularly valuable for evaluating kernel integrals in transform inversions. In scattering theory, these transform properties underpin the solution of the Lippmann-Schwinger equation in the vector spherical harmonic basis. The equation, \mathbf{\psi}(\mathbf{r}) = \mathbf{\psi}_0(\mathbf{r}) + \int G(\mathbf{r}, \mathbf{r}') V(\mathbf{r}') \mathbf{\psi}(\mathbf{r}') d\mathbf{r}', where G is the Green's function and V the potential, is expanded using vector harmonics: the incident wave \mathbf{\psi}_0 via the formula, and the scattered field in outgoing Hankel functions. This yields a equation for the coefficients, decoupling angular and reducing to radial integral equations per mode, efficient for numerical solutions in electromagnetic or acoustic . Recent advances include discrete vector spherical transforms with fast algorithms. The FaVeST algorithm, for instance, achieves O(L^3 \log L) complexity for degree-L expansions on nonequidistant grids, enabling reconstruction of vector fields by approximating continuous integrals via . These methods address in discrete settings.

References

  1. [1]
    Vector Spherical Harmonic -- from Wolfram MathWorld
    ### Definition and Key Formulas for Vector Spherical Harmonics
  2. [2]
    Vector Spherical Harmonics and Multipoles
    The vector spherical harmonics to be constructed out of sums of products of spherical harmonics and the eigenvectors of the operator $S_z$ defined above.
  3. [3]
    [PDF] The tensor spherical harmonics Howard E. Haber
    Feb 25, 2024 · The first two vector spherical harmonics, n × ∇ Yjm(θ, φ) and n × (n × ~∇)Yjm(θ, φ), are transverse (i.e., perpendicular to n), whereas the ...
  4. [4]
    [PDF] Vector spherical harmonics
    1. Vector spherical harmonics. In mathematics, vector spherical harmonics (VSH) are an extension of the scalar spherical harmonics for the use with vector ...
  5. [5]
    Vector spherical harmonics and their application to magnetostatics
    The new set of vector spherical harmonics satisfies the properties of orthogonality and completeness, and is compared with other existing definitions.
  6. [6]
    Spherical functions - Book chapter - IOPscience
    conserve their sign. Remark: Pierre Simon de Laplace first introduced the set of spherical harmonics ... , known as the Laplace spherical harmonics, in 1782.
  7. [7]
    [PDF] The Spherical Harmonics
    The angle between the two vectors is denoted by γ. Note that if one sets ℓ = 1 in the addition theorem and employs P1(cosγ) = cosγ, then the result ...
  8. [8]
    [PDF] Notes on Spherical Harmonics and Linear Representations of Lie ...
    Nov 13, 2013 · The Legendre polynomials can be defined in various ways. One definition is in terms of Rodrigues' formula: Pn(t) = 1. 2nn! dn dtn. (t2 − 1)n ...
  9. [9]
    [PDF] Spherical Harmonics and Legendre Polynomials
    The different spherical harmonics are orthogonal to each other. 23. Page 24. Spherical Harmonics. We saw that by considering just rotations about the y-axis ...
  10. [10]
    Spherical Harmonic Addition Theorem -- from Wolfram MathWorld
    "The Addition Theorem for Spherical Harmonics." §12.8 in Mathematical Methods for Physicists, 3rd ed. Orlando, FL: Academic Press, pp. 693-695, 1985 ...
  11. [11]
    [1612.08062] Modeling Tangential Vector Fields on a Sphere - arXiv
    Dec 23, 2016 · These models are constructed by applying the surface gradient or the surface curl operator to scalar random potential fields defined on the unit ...<|separator|>
  12. [12]
    [PDF] Isotropic random tangential vector fields on spheres
    Jun 3, 2024 · The tangential vector field can be decomposed into a curl-free part and a divergence-free part by the Helmholtz– Hodge decomposition.
  13. [13]
    [PDF] Vector Spherical Harmonics - Rutgers Physics
    ∗. ` m Y`m = δ`` δmm . So the vector spherical harmonics are an orthonormal set: Z. dΩ~X∗. ` m · ~X`m = δ`` δmm . We also have. Z. dΩ~X∗. ` m · r × ~X`m. = 1 p.
  14. [14]
    [PDF] arXiv:1711.08806v2 [hep-th] 10 May 2018
    May 10, 2018 · B.2 Vector spherical harmonics. The pure-spin vector spherical harmonics of magnetic (B−), electric (E−) and radial (R−) type are defined as [9].<|control11|><|separator|>
  15. [15]
    [PDF] Spin-weighted spherical harmonics and their applications
    Spin-weighted spherical harmonics, introduced by Newman and Penrose, are used in solving partial differential equations, especially for nonzero spin fields.
  16. [16]
    [PDF] Revisiting the concentration problem of vector fields within a ...
    ... vector spherical harmonics. They are defined as ... complete basis of the Hilbert space of square-integrable ... tangential vector fields localized to C.
  17. [17]
    [PDF] The Concentration Problem for Vector Fields.
    Vector spherical harmonics form a complete basis set on a sphere and can be used to express an arbitrary vector function, f(θ, φ), tangential to the surface ...
  18. [18]
    Angular momentum and spherical harmonics
    The angular part of the Laplacian is related to the angular momentum of a wave in quantum theory. In units where $\hbar = 1$, the angular momentum operator is:
  19. [19]
    [PDF] arXiv:2302.00529v1 [gr-qc] 1 Feb 2023
    Feb 1, 2023 · ... vector spherical harmonics, denoted here by Ea,`m, ... with Q1 and Ql being the electric and magnetic charges, respectively, while Q is defined.
  20. [20]
    Vector spherical harmonics and their application to classical ...
    A set of vector spherical harmonics which was earlier introduced and shown to be applicable to magnetostatics problems, is employed to decompose and solve ...
  21. [21]
    [PDF] arXiv:2010.09433v1 [physics.class-ph] 13 Oct 2020
    Oct 13, 2020 · We first develop a new method for calculating the scalar spherical harmonics Yjm in terms of components of the direction unit vector n ...
  22. [22]
  23. [23]
    Argument‐recursive computation of Legendre polynomials and its ...
    Jul 22, 2009 · An efficient x-recursive numerical scheme is presented to compute Legendre polynomials P n (x) and their derivatives P′ n (x) on the interval (0, 1) for a ...
  24. [24]
    SphericalHarmonicY - Wolfram Language Documentation
    Mathematical function, suitable for both symbolic and numerical manipulation. The spherical harmonics are orthonormal with respect to integration over the ...
  25. [25]
    Windspharm: A High-Level Library for Global Wind Field ...
    Aug 2, 2016 · It provides a high-level interface for computing derivatives and integrals of vector wind fields over a sphere using spherical harmonics. The ...Missing: SciPy | Show results with:SciPy
  26. [26]
    [PDF] Vector spherical harmonics and their application to magnetostatics
    The new set of vector spherical harmonics satisfies the properties of orthogonality and complete- ness. and is compared with other existing definitions of.
  27. [27]
    A survey of the application of the spherical vector wave mode ...
    Jun 9, 2014 · This paper has provided a tutorial on the application of the spherical vector wave mode expansion of the electromagnetic field and the antenna ...
  28. [28]
    [PDF] Part II Calculation and Measurement of Scattering and Absorption ...
    vector spherical wave functions and the expansion coefficients, which are determined through the requirement of the standard boundary conditions, are ...
  29. [29]
    Photon angular momentum: Selection rules and multipolar transition ...
    The representation of electromagnetic radiation by means of vector spherical harmonics and Bessel functions is described and used to derive selection rules ...
  30. [30]
    Cartesian and spherical multipole expansions in anisotropic media
    Aug 22, 2024 · Title:Cartesian and spherical multipole expansions in anisotropic media. Authors:Elias Le Boudec, Toma Oregel-Chaumont, Farhad Rachidi ...
  31. [31]
  32. [32]
  33. [33]
    Vector representation on a sphere | Jingtao Min Homepage
    Here I compile a self-contained description of vector field representation and decomposition, and differential operator on a sphere.
  34. [34]
    Chandrasekhar-Kendall functions in astrophysical dynamos
    Aug 9, 2025 · Particular emphasis is placed on the Chandrasekhar-Kendall functions that allow a decomposition of a vector field into right- and left-handed ...
  35. [35]
    [PDF] Clebsch-Gordon coefficients and the tensor spherical harmonics
    which implies that if m 6= mℓ + ms then the corresponding Clebsch-Gordon coefficient must vanish. This is simply a consequence of Jz = Lz + Sz. Likewise, ...
  36. [36]
    [PDF] Solutions of the Dirac Equation and Their Properties
    To do this we freeze the radial coordinate r and work with the Hilbert space of spinors over a sphere, in which the spin spherical harmonics Yℓ jmj form a ...
  37. [37]
    [PDF] Atomic Structure Theory
    1.5 Spinor and Vector Spherical Harmonics. 22. 1.5.1 Spherical Spinors. 22. 1.5.2 Vector Spherical Harmonics. 24. Problems. 26. Central-Field Schrodinger ...
  38. [38]
    Vector-spherical-harmonics representation of vector complex source beams carrying vortices
    ### Summary of Vector Spherical Harmonics in Vector Vortex Beams for Quantum Optics and Entanglement
  39. [39]
    [PDF] High- dimensional spin- orbital single- photon sources
    Nov 6, 2024 · OAM beams to vector vortex beams and gives rise to spin- orbit vectorial states (12, 13), especially with the help of planar subwave- length ...
  40. [40]
    Convergence of vector spherical wave expansion method applied to ...
    For the interior of the sphere, we use p = 1 (regular vector spherical waves). For the exterior of the sphere, we use p = 3 (outgoing vector spherical waves).
  41. [41]
    [PDF] Accurate calculation of spherical and vector spherical harmonic ...
    This paper presents a spectrally accurate numerical method for computing spherical/vector spherical harmonic expansions using spectral element grids, using ...<|separator|>
  42. [42]
    Spherical Wave Expansion of Vector Plane Wave - Richard Fitzpatrick
    It is useful to be able to express a plane electromagnetic wave as a superposition of spherical waves.Missing: Rayleigh | Show results with:Rayleigh
  43. [43]
    [PDF] Diagonalizations of Vector and Tensor Addition Theorems
    In this section, we shall extend the. Funk-Hecke formula and Rayleigh plane-wave expansion formula for tensor fields. ... 3 vector spherical harmonics can be ...
  44. [44]
    [PDF] Polynomial bases for subspaces of potential and solenoidal vector ...
    Mar 20, 2006 · Applying the Funk-Hecke formula for vector spherical harmonics (Theorem 3.2) and identity (d) from Lemma 2.5 yields. A(n) n−1−2k,l(x) = S2.
  45. [45]
    Three-dimensional integral equation approach to light scattering ...
    We present a numerical formalism for solving the Lippmann–Schwinger equation for the electric field in three dimensions. The formalism may be applied to ...
  46. [46]
    Fourier Transform of the Lippmann-Schwinger Equation - MDPI
    Nov 18, 2023 · In the article, we rigorously computed the Fourier transform of the vectorial Lippmann–Schwinger equation in the space of tempered distributions.
  47. [47]
    [1908.00041] FaVeST: Fast Vector Spherical Harmonic Transforms
    Jul 31, 2019 · In the representation of a tangent vector field, one needs to evaluate the expansion and the Fourier coefficients of vector spherical harmonics.
  48. [48]
    Algorithm 1018: FaVeST—Fast Vector Spherical Harmonic Transforms
    Sep 28, 2021 · In this article, we develop fast algorithms (FaVeST) for vector spherical harmonic transforms on these evaluations. The forward FaVeST evaluates ...<|control11|><|separator|>