Fact-checked by Grok 2 weeks ago

Light scattering by particles

Light scattering by particles refers to the elastic redirection of electromagnetic , such as visible , by small objects suspended in a medium, where the particles' dimensions are comparable to or smaller than the of the incident . This interaction arises from the induced oscillations in the particle's , which reradiates the energy in all directions, altering the light's path, intensity, and . The phenomenon is quantified by the scattering cross-section, which measures the effective area of the particle responsible for deflecting the light beam. The theoretical foundation of light scattering by particles is rooted in , with key developments including Rayleigh's approximation for particles much smaller than the (size parameter x = 2\pi a / \lambda \ll 1, where a is the particle radius and \lambda is the ). In this regime, the scattered intensity is proportional to the of the frequency, leading to stronger scattering of shorter s and explaining natural observations like the blue color of the sky. For larger particles where x \approx 1, Mie theory provides an exact solution for spherical particles by solving using , accounting for both scattering and potential absorption. Beyond basic models, scattering encompasses more complex scenarios involving nonspherical particles, ensembles, and multiple scattering events, often analyzed using methods like the T-matrix approach for arbitrary shapes or the discrete dipole approximation for irregular forms. These interactions also involve , where is converted to , and from thermally excited particles, influencing the overall of the beam. effects, described by the and scattering matrices, play a critical role in characterizing the process, with applications in of atmospheric aerosols and planetary atmospheres. In practical contexts, light scattering by particles is essential for techniques like (DLS) to measure distributions in colloids and suspensions, as well as in of pollutants. It underpins phenomena in , such as the formation of rainbows and glories, and informs models in for climate simulations. Advances in computational tools, including publicly available codes for Mie and T-matrix solutions, enable precise predictions for diverse particle systems.

Overview

Definition and physical basis

Light scattering by particles refers to the redirection of electromagnetic waves, commonly known as , upon interaction with particles whose dimensions are much smaller than or comparable to the of the . This phenomenon encompasses both , where the scattered maintains the same as the incident with no energy loss to the particle, and inelastic processes such as , though predominates in many applications. The physical basis of light scattering arises from the interaction between the oscillating electric field of the incident light and the electrons within the particle, which induces oscillating dipoles that serve as secondary sources of radiation. These induced dipoles re-emit electromagnetic waves in all directions, effectively redirecting the original light. This process can be understood through the Huygens-Fresnel principle, which describes the scattered field as the coherent superposition of secondary wavelets emanating from the induced dipoles across the particle's surface or volume. Particles involved in light scattering can be dielectric (non-absorbing), absorbing, or metallic, and they may be suspended in gases, liquids, or even vacuums. A fundamental characterizing the scattering is the size \alpha = \frac{2\pi a}{\lambda}, where a is the particle and \lambda is the of the incident ; this determines the relative scale of the particle to the light wave and influences the scattering regime. The scattered electric field \mathbf{E}_s in the far field is related to the incident field \mathbf{E}_i by the general expression \mathbf{E}_s = \frac{e^{ikr}}{r} f(\theta) \mathbf{E}_i, where r is the distance from the particle, k = 2\pi / \lambda is the wave number, and f(\theta) is the , a complex that depends on the scattering \theta and encapsulates the particle's properties. Light scattering by particles plays a crucial role in , where it explains phenomena such as the blue color of the sky due to preferential of shorter wavelengths by air molecules and aerosols. In biomedical imaging, it enables non-invasive techniques for probing cellular structures and tissue properties through elastic light interactions. Additionally, in , understanding helps design materials with tailored optical responses, such as in plasmonic nanoparticles for sensing and light manipulation.

Historical development

In the mid-19th century, early experimental investigations into light by particles focused on atmospheric phenomena. In 1869, conducted experiments using a beam of passed through a tube filled with air containing suspended dust particles, observing that the scattered light appeared and was polarized, which he attributed to by these small particles and linked to the color of the sky. Two years later, in 1871, Lord Rayleigh developed a theoretical framework explaining the of light by molecules much smaller than the , deriving the inverse fourth-power dependence on that accounts for the predominance of in clear skies. The early 20th century saw significant theoretical advances for scattering by larger particles. In 1908, Gustav Mie published a rigorous solution to for the scattering and absorption of electromagnetic waves by a homogeneous of arbitrary size, providing exact expressions for scattering cross-sections and angular distributions. Independently in the same year, derived an equivalent formulation for electromagnetic scattering by spheres, emphasizing and confirming Mie's results through a approach. Mid-century developments integrated these theories into broader applications, particularly in atmospheric and colloidal science. The Rayleigh-Gans approximation, developed in the early by and Gans for large, optically soft particles where phase shifts are small, enabling simplified calculations for non-spherical and inhomogeneous scatterers, was further advanced by Hendrik van de Hulst in his 1957 book, Light Scattering by Small Particles. Van de Hulst's seminal 1957 book, Light Scattering by Small Particles, synthesized these and other theories, providing a comprehensive reference that bridged exact solutions like Mie theory with approximations and surveyed applications to planetary atmospheres and aerosols. In 1969, Milton Kerker's book The Scattering of Light and Other compiled and expanded upon these foundations, emphasizing experimental validations and extensions to multiple scattering in complex media. From the 1980s onward, computational advancements facilitated practical implementations and extensions beyond spherical particles. In 1983, Craig Bohren and Donald Huffman published Absorption and Scattering of Light by Small Particles, which included widely adopted code for efficient numerical evaluation of Mie theory, enabling rapid simulations for atmospheric and material science applications. During the , significant progress occurred in methods for non-spherical particles, including refinements to the T-matrix approach for axisymmetric shapes and the discrete dipole approximation for arbitrary geometries, allowing accurate modeling of ice crystals, dust, and biological aerosols in climate and studies.

Theoretical foundations

Electromagnetic scattering principles

The theoretical foundation of light scattering by particles rests on , which govern the behavior of electromagnetic fields in the presence of matter. For time-harmonic fields with \omega, assuming non-magnetic media where the permeability \mu = \mu_0 is constant and uniform, the equations simplify to the vector Helmholtz forms: \nabla \times \mathbf{E} = i \omega \mu_0 \mathbf{H}, \quad \nabla \times \mathbf{H} = -i \omega \epsilon_0 \epsilon(\mathbf{r}) \mathbf{E}, where k_0 = \omega \sqrt{\mu_0 \epsilon_0} is the vacuum wavenumber, \epsilon(\mathbf{r}) is the position-dependent , \mathbf{E} is the , and \mathbf{H} is the . These equations must be solved subject to boundary conditions at the interfaces between the particle and the surrounding medium, ensuring continuity of the tangential components of \mathbf{E} and \mathbf{H}, as well as the normal components of \mathbf{D} = \epsilon_0 \epsilon \mathbf{E} and \mathbf{B} = \mu_0 \mathbf{H}. This framework captures the interaction of electromagnetic waves with or absorbing particles without approximations to particle geometry. In the scattering formulation, an incident plane wave \mathbf{E}_i = \mathbf{E}_0 e^{i k_0 z} (assuming propagation along the z-axis) illuminates the particle, and the total field is expressed as \mathbf{E} = \mathbf{E}_i + \mathbf{E}_s, where \mathbf{E}_s is the scattered field. The scattered field \mathbf{E}_s satisfies the at infinity, \lim_{r \to \infty} r \left( \frac{\partial \mathbf{E}_s}{\partial r} - i k_0 \mathbf{E}_s \right) = 0, ensuring outgoing spherical waves. For dielectric particles embedded in a non-absorbing host medium, the problem can be recast as a volume integral equation using the Lippmann-Schwinger formalism. The total electric field inside the particle volume V obeys \mathbf{E}(\mathbf{r}) = \mathbf{E}_i(\mathbf{r}) + k_0^2 \int_V \mathbf{G}(\mathbf{r}, \mathbf{r}') \chi(\mathbf{r}') \mathbf{E}(\mathbf{r}') \, dV', where \mathbf{G} is the dyadic Green's function for the background medium, representing the field due to a point dipole source, and the susceptibility \chi(\mathbf{r}) = \epsilon(\mathbf{r}) - 1 accounts for the particle's contrast. This integral equation, derived directly from Maxwell's equations via the Green's theorem, allows numerical solution for arbitrary shapes by discretizing the volume and solving for the induced field iteratively or via matrix inversion. Surface integral methods offer an alternative by reformulating the problem on the particle's boundary \partial V, avoiding volume discretization for efficiency with smooth surfaces. These approaches, such as the , express the scattered field in terms of equivalent electric \mathbf{J} and magnetic \mathbf{M} surface currents on \partial V, satisfying electric field integral equations (EFIE) or magnetic field integral equations (MFIE): \mathbf{E}_s(\mathbf{r}) = i \omega \mu_0 \int_{\partial V} \mathbf{G}(\mathbf{r}, \mathbf{r}') \cdot \mathbf{M}(\mathbf{r}') \, ds' - \frac{i}{k_0 \eta_0} \nabla \times \int_{\partial V} \mathbf{G}(\mathbf{r}, \mathbf{r}') \cdot \mathbf{J}(\mathbf{r}') \, ds', with \eta_0 = \sqrt{\mu_0 / \epsilon_0} the free-space impedance. The currents are determined by enforcing conditions, leading to a of equations solved via Galerkin or techniques. This method is particularly suited for impenetrable or high-contrast scatterers, as it reduces the dimensionality of the problem. effects are inherently ial in electromagnetic , distinguishing it from scalar approximations valid only for unpolarized or isotropic cases. The full treatment requires solving for the vector fields, yielding polarization-dependent scattered intensities described by the I, Q, U, V, which fully characterize the partial polarization state. For instance, the scattering matrix relates the incident and scattered Stokes vectors, capturing phenomena like linear or conversion, essential for analyzing non-spherical or oriented particles. Scalar treatments, while simpler for total intensity, neglect these effects and are limited to Rayleigh-like regimes or averaged .

Key scattering parameters

In light scattering by particles, key parameters characterize the extent and nature of the interaction between electromagnetic waves and scatterers, enabling quantitative descriptions of phenomena such as attenuation, angular redistribution, and changes. These parameters are derived from fundamental solutions to and are essential for interpreting scattering data across various regimes, from small dielectric particles to complex aggregates. Standard references establish these quantities as the scattering cross-section for total redirected power, the differential cross-section for angular details, the extinction cross-section linking removal to forward scattering, the phase function for directional preferences, and the Mueller matrix for transformations. The scattering cross-section \sigma_s quantifies the total power scattered out of the incident beam per unit incident , effectively representing the particle's "scattering area." It is defined as the of the scattering cross-section over all angles: \sigma_s = \int_{4\pi} \frac{d\sigma}{d\Omega} \, d\Omega, with units of m², analogous to geometric cross-sections but accounting for effects. This parameter directly measures how much is redistributed by the particle without . The differential scattering cross-section \frac{d\sigma}{d\Omega} provides the angular distribution of scattered power, specifying the fraction scattered into a particular direction. It relates the far-field scattered intensity I_s to the incident intensity I_i at distance r via I_s = \frac{I_i}{r^2} \frac{d\sigma}{d\Omega}, allowing reconstruction of \sigma_s through integration; variations in \frac{d\sigma}{d\Omega} reveal interference patterns, such as forward lobes in larger particles. The extinction cross-section \sigma_\text{ext} encompasses both and , defined as \sigma_\text{ext} = \sigma_s + \sigma_a, where \sigma_a is the cross-section representing energy dissipated as . A fundamental relation, the optical theorem, connects \sigma_\text{ext} to the imaginary part of the forward f(0): \sigma_\text{ext} = \frac{4\pi}{k} \operatorname{Im}[f(0)], with k = 2\pi / \lambda the wave number; this theorem underscores that total beam attenuation is tied to forward , holding for any scatterer in . The phase function P(\theta) describes the normalized probability density of scattering into angle \theta relative to the incident direction, normalized such that \int_{4\pi} P(\theta) \, d\Omega = 4\pi. It captures the angular pattern, often peaking forward for particles comparable to the ; the parameter g = \langle \cos \theta \rangle = \frac{1}{4\pi} \int P(\theta) \cos \theta \, d\Omega quantifies this, with g > 0 indicating forward preference (e.g., g \approx 0.8 for cloud droplets) and g = 0 for symmetric cases. In the regime for small particles, P(\theta) is nearly isotropic. For partially polarized light, the Mueller matrix \mathbf{M} is a 4×4 real matrix that linearly transforms the incident Stokes vector \mathbf{S}_i = (I, Q, U, V)^T to the scattered Stokes vector \mathbf{S}_s = \mathbf{M} \mathbf{S}_i, fully encoding changes in intensity, , and due to . Elements of \mathbf{M} relate to the differential cross-section for (M_{11}) and depolarization effects (e.g., M_{12}, M_{34}), with symmetry properties like M_{ij} = M_{ji} for rotationally invariant scatterers; this formalism extends scalar theories to vectorial wave behavior.

Analytical approximation methods

Rayleigh scattering regime

The Rayleigh scattering regime describes the interaction of electromagnetic waves with particles whose characteristic dimensions are much smaller than the wavelength of the incident , specifically when the size parameter α = 2πa/λ ≪ 1, where a is the particle radius and λ is the , often valid for α < 0.3 or a ≪ λ/10. In this quasi-static approximation, higher-order multipole contributions are negligible, and the particle responds primarily as an induced electric dipole to the uniform incident field, neglecting retardation effects across the particle volume. This regime is foundational for understanding scattering in dilute gases, aerosols, and molecular atmospheres, as originally formulated by Lord Rayleigh in his analysis of skylight polarization and color. The derivation begins with the incident electric field E_i inducing a dipole moment p = α_p E_i in the particle, where α_p is the complex polarizability. For a homogeneous dielectric sphere with relative permittivity ε, the polarizability is given by α_p = 4π ε_0 a³ (ε - 1)/(ε + 2), derived from the electrostatic solution for a sphere in a uniform field, extended to the optical domain under the assumption of non-magnetic materials (μ = 1) and negligible absorption (real ε). The induced dipole radiates a scattered field akin to that of a , with the far-field scattered electric field proportional to the second time derivative of p. Integrating the radiated power over all directions yields the total scattering cross-section σ_s = (8π/3) k⁴ |α_p|² / (4π ε_0)², where k = 2π/λ is the wavenumber; this expression simplifies to σ_s ∝ a⁶ / λ⁴ | (ε - 1)/(ε + 2) |², highlighting the strong inverse fourth-power wavelength dependence that preferentially scatters shorter (blue) wavelengths, as observed in the clear daytime sky. The differential scattering cross-section exhibits angular dependence due to the dipole radiation pattern. For unpolarized incident light, dσ/dΩ ∝ |α_p|² k⁴ (1 + cos²θ)/2, where θ is the scattering angle relative to the incident direction, showing forward-backward symmetry and maximum intensity perpendicular to the incident propagation (θ = 90°). This polarization-sensitive pattern arises from the vector nature of the dipole oscillation, with linearly polarized incident light producing scattered light polarized parallel to the incident polarization in the forward and backward directions. For small non-absorbing spheres, the scattered light is partially polarized, with the degree of polarization increasing toward the sides. This approximation has limitations: it assumes electrically small particles where phase shifts are minimal and fails for absorbing materials (complex ε), where imaginary components alter the polarizability and introduce asymmetry, or when α > 0.3, necessitating higher multipole terms as in Mie theory for spheres. Extensions like the Rayleigh-Gans approximation address cases of larger but optically soft particles (|ε - 1| ≪ 1), maintaining the dipole-like response but incorporating variations across the particle.

Geometric optics regime

The geometric optics regime applies to light scattering by particles where the size parameter \alpha \gg 1, defined as \alpha = 2\pi a / \lambda with a the particle and \lambda the , such that a \gg \lambda and the wave nature of becomes negligible. In this limit, scattering is modeled using ray-tracing techniques that treat as rays propagating in straight lines, governed by for refraction at interfaces and for the amplitude and of reflected and transmitted rays. This approximation is particularly suitable for large, non-absorbing particles like cloud droplets or ice crystals, where interference effects are minimal. Ray paths in the geometric optics regime include external from the particle surface, internal and multiple reflections within transparent particles, and edge around the particle boundary. External reflections contribute to backscattering, while refracted rays undergo internal bounces that can lead to specific angular features such as (near-backward scattering from rays grazing the surface) and rainbows (from rays with one or more internal reflections). Edge diffraction is approximated using to determine ray deviations at boundaries, supplementing the primary and contributions. These paths are traced numerically to compute the scattered intensity, accounting for the particle's shape and . In this regime, the extinction cross-section approaches approximately $2\pi a^2, reflecting the where the total removed power equals twice the geometric cross-section due to shadowing and effects. For opaque spheres, the cross-section simplifies to σ_s ≈ 2π a², consisting of contributions from surface reflections and (each approximately π a²), without . These values highlight the scale of interaction, where alone accounts for the forward lobe roughly equal to the geometric area. Angular distributions feature a prominent forward diffraction peak from rays bent around the particle, broad backscattering from external and internal reflections, and sharp peaks at rainbow angles, such as approximately 138° for primary rainbows in water droplets with refractive index near 1.33. The glory appears as a backward enhancement from surface-reflected rays. Limitations of the geometric optics regime include its neglect of , which underestimates oscillations in the scattering patterns for finite sizes, making it approximate even for \alpha \sim 100 and best for \alpha > 100 in non-absorbing cases. It can be integrated briefly with Mie theory for hybrid models in intermediate regimes to capture partial wave effects.

Exact analytical methods for simple geometries

Mie theory for spherical particles

Mie theory provides the exact analytical solution to Maxwell's equations for the scattering of a plane electromagnetic wave by a homogeneous, isotropic sphere of radius a and relative refractive index m = n + ik, where n is the real part and k the imaginary part accounting for absorption. The formulation involves solving the vector wave equation in spherical coordinates, expanding both the incident and scattered fields in terms of vector spherical harmonics. The incident plane wave is expressed as a sum over these harmonics, and the scattered field is similarly expanded but with coefficients determined by boundary conditions at the sphere's surface. The scattering coefficients a_n and b_n, which govern the electric and magnetic multipole contributions, are obtained by matching the tangential components of the electric and at the boundary r = a. These are given by a_n = \frac{m \psi_n(mx) \psi_n'(x) - \psi_n(x) \psi_n'(mx)}{m \psi_n(mx) \xi_n'(x) - \xi_n(x) \psi_n'(mx)}, b_n = \frac{\psi_n(mx) \psi_n'(x) - m \psi_n(x) \psi_n'(mx)}{\psi_n(mx) \xi_n'(x) - m \xi_n(x) \psi_n'(mx)}, where x = ka = 2\pi a / \lambda is the size parameter with k the in the surrounding medium, \psi_n(z) = z j_n(z) the Riccati-Bessel function with spherical Bessel function j_n, and \xi_n(z) = z h_n^{(1)}(z) the Riccati-Hankel function of the first kind with h_n^{(1)}; primes denote derivatives with respect to the argument. For non-absorbing spheres (k=0), m reduces to the real ratio. Key parameters include the scattering efficiency Q_s = \sigma_s / (\pi a^2), extinction efficiency Q_{\rm ext} = \sigma_{\rm ext} / (\pi a^2), and absorption efficiency Q_a = \sigma_a / (\pi a^2), where \sigma_s, \sigma_{\rm ext}, and \sigma_a are the respective cross sections. These are computed as Q_s = \frac{2}{x^2} \sum_{n=1}^\infty (2n+1) (|a_n|^2 + |b_n|^2), \quad Q_{\rm ext} = \frac{2}{x^2} \sum_{n=1}^\infty (2n+1) \operatorname{Re}(a_n + b_n), \quad Q_a = Q_{\rm ext} - Q_s. Plots of Q_s and Q_{\rm ext} versus x exhibit characteristic ripple structures due to interference between diffracted and reflected/refracted waves, with oscillations becoming more pronounced for larger x > 10. In the large-particle limit (x \to \infty), Q_{\rm ext} \to 2, known as the extinction paradox, where the extinction cross section approaches twice the geometric area because diffraction contributes equally to shadowing. The theory also describes polarization effects through the amplitude functions for perpendicular and parallel polarizations: S_1(\theta) = \sum_{n=1}^\infty \frac{2n+1}{n(n+1)} \left( a_n \pi_n(\cos\theta) + b_n \tau_n(\cos\theta) \right), S_2(\theta) = \sum_{n=1}^\infty \frac{2n+1}{n(n+1)} \left( a_n \tau_n(\cos\theta) + b_n \pi_n(\cos\theta) \right), where \pi_n and \tau_n are angular functions derived from . The scattered intensity is proportional to |S_1|^2 and |S_2|^2, enabling computation of the full function and degree of . For small x \ll 1, the n=1 terms dominate, recovering the approximation. Computational implementation of Mie theory requires careful handling of the series summation to avoid numerical overflow or cancellation errors, particularly for large x. Stable algorithms, such as the one developed by Bohren and Huffman, use logarithmic derivatives and recursive relations for the Bessel functions to ensure accuracy across a wide range of parameters, with the number of terms needed approximately x + 4x^{1/3} + 2.

T-matrix method for axisymmetric particles

The T-matrix method provides an exact semi-analytical approach for computing electromagnetic scattering by particles possessing axial symmetry, such as spheroids, cylinders, and Chebyshev particles, by generalizing the vector spherical wave function expansions used in Mie theory through the application of addition theorems that account for translations and rotations of coordinate systems. This enables the method to handle deviations from spherical geometry while maintaining analytical tractability for the far-field scattering properties. In the spherical case, the T-matrix reduces to the diagonal form characteristic of Mie theory, confirming its consistency as a limiting case. The core of the method lies in the , or T-matrix, which linearly relates the coefficients of the scattered field expansion to those of the incident field in a basis of vector spherical wave functions. Specifically, if \mathbf{a} and \mathbf{b} denote the column vectors of scattered and incident field coefficients, respectively, then \mathbf{a} = \mathbf{T} \mathbf{b}. The T-matrix is derived from the boundary conditions at the particle surface using the extended boundary condition method, expressed as \mathbf{T} = \mathbf{A}^{-1} \mathbf{B}, where \mathbf{A} and \mathbf{B} are infinite matrices obtained from surface integrals enforcing continuity of the tangential field components. For axisymmetric particles, the azimuthal (invariance under rotations about the ) renders the T-matrix block-diagonal, with blocks corresponding to each azimuthal mode index m, significantly reducing computational complexity by decoupling the equations for different m. This finds broad applications in modeling by non-spherical particles, including prolate and spheroids, finite-length cylinders, and aggregates of such particles, where multiple inclusions can be treated via superposition using the T-matrix addition theorems for their relative translations and orientations. It is particularly valuable for atmospheric and oceanic , such as simulations of light by deformed raindrops or crystals, and extends to ensembles of randomly oriented particles by averaging the T-matrix over orientations. Key advantages include its efficiency for particles with size parameters up to approximately 50 (in units of ka, where k is the and a the characteristic dimension) and moderate aspect ratios, allowing rapid computation of the full matrix, including the Mueller matrix elements that describe effects. Unlike fully numerical methods, it leverages analytical expansions to achieve high accuracy with fewer resources for symmetric shapes. The method was originally formulated by Waterman in 1971, establishing the symmetry properties and unitarity of the T-matrix for electromagnetic . Mishchenko extended this in the , particularly in 1991, by deriving efficient algorithms for randomly oriented axisymmetric particles and providing publicly available implementations that handle coated and composite structures, facilitating widespread adoption in and research. Limitations arise for highly aspherical particles with extreme aspect ratios (beyond about 3:1), where ill-conditioning of the matrices \mathbf{A} and \mathbf{B} can lead to numerical instability, necessitating regularization techniques.

Numerical methods for complex geometries

Discrete dipole approximation

The discrete dipole approximation (DDA) is a method for computing electromagnetic scattering and absorption by particles of arbitrary shape, size, and material properties, including inhomogeneous, anisotropic, and chiral media. It discretizes the particle's volume into an array of N polarizable points, treated as point s, which respond to the local by developing induced dipole moments. This approach is particularly suited for non-spherical or irregular geometries where analytical methods fail, as it solves numerically in the without relying on surface discretizations. In the DDA formulation, the particle is represented by N dipoles located at positions \mathbf{r}_j (j = 1, \dots, N), each with polarizability \alpha_j determined by the local material properties. The induced dipole moment at site j is given by \mathbf{p}_j = \alpha_j \mathbf{E}_j, where \mathbf{E}_j is the total electric field at that location. The total field \mathbf{E}_j comprises the incident field \mathbf{E}_{\rm inc,j} and the scattered field from all other dipoles, mediated by the dyadic Green's function \mathbf{G}(\mathbf{r}_j, \mathbf{r}_k, \omega) for free space: \mathbf{E}_j = \mathbf{E}_{\rm inc,j} + \sum_{k \neq j} \mathbf{G}(\mathbf{r}_j, \mathbf{r}_k, \omega) \cdot \mathbf{p}_k. This leads to a system of $3N linear equations for the dipole moments \mathbf{P} = (\mathbf{p}_1, \dots, \mathbf{p}_N): \left( \alpha_j^{-1} \mathbf{I} - \mathbf{A} \right) \mathbf{P} = \mathbf{E}_{\rm inc}, where \mathbf{A} is the $3N \times 3N interaction matrix with elements A_{jk} = -\mathbf{G}(\mathbf{r}_j, \mathbf{r}_k, \omega) for j \neq k and zero diagonal (or a small self-term), and \mathbf{I} is the identity. For large N, direct solvers are impractical due to O(N^3) cost; instead, iterative methods like the conjugate gradient (CG) algorithm are employed, achieving O(N \log N) or better scaling with fast Fourier transform (FFT) acceleration for near-field interactions. The accuracy of DDA depends on the dipole spacing d, which must satisfy d \ll \lambda (where \lambda is the in the medium) to resolve variations. Convergence is typically achieved for d \approx \lambda / 20 to \lambda / 10, with relative errors in quantities scaling as O((k d)^2) (where k = 2\pi / \lambda) plus a linear term for larger d, bounded rigorously under conditions like |m| k d < 1 (m is the refractive index). For small particles in the Rayleigh limit, DDA reduces to the exact polarizability sum, validating its use for clusters. DDA finds wide application in modeling light scattering by irregular particles, such as atmospheric dust, biological cells, or nanoparticles, as well as in chiral media where handedness affects polarization, and near-field optics for subwavelength structures. Open-source implementations like , developed in the 1990s by Draine and Flatau, facilitate these calculations for particles up to sizes \sim 100\lambda with N \sim 10^5–$10^6. Despite its versatility, DDA is computationally intensive for very large N > 10^6, requiring significant memory and time even with iterative solvers, and becomes ill-conditioned for metallic particles due to near-resonant polarizabilities that amplify numerical errors.

Finite-difference time-domain method

The finite-difference time-domain (FDTD) method is a full-wave numerical technique for simulating electromagnetic wave propagation and scattering by discretizing Maxwell's equations on a spatiotemporal grid, making it particularly suitable for modeling light interactions with complex particle geometries. The approach employs the Yee grid, a staggered lattice where electric field components are positioned at the centers of cell edges and faces, while magnetic field components are offset to cell vertices, enabling a leapfrog time-stepping scheme that alternates updates between electric (E) and magnetic (H) fields. This discretization approximates spatial derivatives using central finite differences and advances the fields in time, capturing broadband responses from a single pulsed simulation. The core update equations derive from the curl forms of in isotropic media: \frac{\partial \mathbf{E}}{\partial t} = \frac{1}{\epsilon} \nabla \times \mathbf{H} - \frac{\sigma}{\epsilon} \mathbf{E}, \frac{\partial \mathbf{H}}{\partial t} = -\frac{1}{\mu} \nabla \times \mathbf{E} - \frac{\sigma^*}{\mu} \mathbf{H}, where \epsilon and \mu are the and permeability, and \sigma and \sigma^* account for electric and magnetic conductivities, respectively. These are discretized on the Yee grid to yield explicit finite-difference expressions, such as for the x-component of E at time step n+1/2: E_x^{n+1/2}(i+1/2, j, k) = E_x^{n-1/2}(i+1/2, j, k) + \frac{\Delta t}{\epsilon} \left[ \frac{H_z^{n}(i+1/2, j+1/2, k) - H_z^{n}(i+1/2, j-1/2, k)}{\Delta y} - \frac{H_y^{n}(i+1/2, j, k+1/2) - H_y^{n}(i+1/2, j, k-1/2)}{\Delta z} \right] - \frac{\sigma \Delta t}{\epsilon} E_x^{n+1/2}(i+1/2, j, k), with analogous updates for other components; H fields are updated similarly using the prior E values. Numerical stability requires adherence to the Courant-Friedrichs-Lewy condition, \Delta t \leq \frac{1}{c \sqrt{1/(\Delta x)^2 + 1/(\Delta y)^2 + 1/(\Delta z)^2}}, which for a uniform cubic grid simplifies to \Delta t < \Delta x / (c \sqrt{3}), where c is the speed of light. For scattering simulations, the total-field/scattered-field (TFSF) formulation introduces an incident by partitioning the grid into total-field and scattered-field regions separated by a Huygens surface, where the incident field is added to the scattered field across the boundary to enforce continuity. Absorbing boundaries are implemented using perfectly matched layers (PML), which attenuate outgoing with minimal by introducing anisotropic, lossy materials that match the impedance of the surrounding medium; originally formulated by Bérenger in , PML has evolved into uniaxial variants for enhanced stability. This setup allows efficient computation of scattering cross-sections and near-field patterns for particles embedded in absorbing or dispersive media. The FDTD method excels in applications involving light by nanoparticles, where it resolves subwavelength features like plasmonic resonances in spheres or rods, yielding efficiencies that match experimental spectra. It is also widely used for photonic structures, such as metamaterials or metasurfaces, to predict and time-resolved dynamics like pulse propagation through scattering media. Unlike frequency-domain methods, FDTD naturally handles material , nonlinearity, and gain, enabling simulations of active particles or time-varying responses in biological or atmospheric scattering scenarios. Pioneered by Allen Taflove in the 1970s building on Kane Yee's 1966 discretization, FDTD originated from early work on propagation and has since become a cornerstone for computational electrodynamics. Modern implementations include open-source tools like MEEP, developed at , which supports and scripting for of particle scattering setups. However, the method demands significant memory for fine grids—typically requiring resolutions of \lambda/50 or finer near scatterers to capture evanescent fields accurately—limiting its feasibility for very large or highly detailed structures without adaptive meshing. Validation studies often compare FDTD results for spherical particles against Mie theory, confirming agreement in scattering efficiencies to within 1% for sizes up to 10\lambda.

References

  1. [1]
    [PDF] Scattering, Absorption, and Emission of Light by Small Particles
    First, the general formalism of scattering, absorption, and emission of light and other electromagnetic radiation by arbitrarily shaped and arbitrarily oriented.
  2. [2]
    None
    ### Summary of Light Scattering by Particles (Lecture 34)
  3. [3]
    None
    Below is a merged summary of "Light Scattering by Small Particles" (Chapters 1-3, primarily based on Bohren & Huffman, 1998) that consolidates all information from the provided segments into a single, comprehensive response. To retain maximum detail and manage the complexity of the data, I will use a structured table format in CSV style for key recurring elements (e.g., definitions, physical basis, particle types, etc.), followed by a narrative summary for additional context and details that don’t fit neatly into a table. This approach ensures all information is preserved while maintaining clarity and density.
  4. [4]
    Mie Scattering References - OMLC
    ... alpha ( alpha =2 pi a/ lambda , where a=radius of the scattering sphere). It is shown that recurrent relationships can be found directly between the ...
  5. [5]
    Microscopic Imaging and Spectroscopy with Scattered Light - PMC
    Optical contrast based on elastic scattering interactions between light and matter can be used to probe cellular structure and dynamics, and image tissue ...Missing: atmospheric | Show results with:atmospheric
  6. [6]
    Application of Light Scattering Techniques to Nanoparticle ...
    Jun 25, 2018 · Dynamic light scattering is used to measure nanoparticles size, but also to evaluate their stability over time in suspension, at different pH ...
  7. [7]
    IV. On the blue colour of the sky, the polarization of skylight, and on ...
    Tyndall John. 1869IV. On the blue colour of the sky, the polarization of skylight, and on the polarization of light by cloudy matter generallyProc. R. Soc ...Missing: dust | Show results with:dust
  8. [8]
    [PDF] Peter Debye- Nobel Lecture
    We come to a second method of determining the dipole moment if the question is considered of the dispersion of the dielectric constant. If we now first confine ...Missing: 1918 | Show results with:1918
  9. [9]
    Light scattering by small particles : Hulst, H. C. van de (Hendrik ...
    Oct 23, 2021 · Light scattering by small particles. 470 p. : 22 cm Reprint. Originally published: New York : Wiley, 1957. (Structure of matter series) Includes ...
  10. [10]
    Light Scattering by Small Particles - Google Books
    It is an exhaustive study of light-scattering properties of small, individual particles, and includes a survey of all the relevant literature.
  11. [11]
    The Scattering of Light and Other Electromagnetic Radiation
    This book can be useful for physicists, chemists, biochemists, and engineers whose work includes research utilizing light scattering in the study of certain ...
  12. [12]
    [PDF] Light Scattering by Nonspherical Particles
    ... Hendrik C. van de Hulst. Part I. Introduction. Chapter 1. Concepts, Terms ... Approximation for Light Scattering by Irregular Targets. Bruce T. Draine. I ...
  13. [13]
    1.7: Time-Harmonic Maxwell Equations in Matter - Physics LibreTexts
    Sep 16, 2022 · We now return to the Maxwell equations and consider time-harmonic electromagnetic fields, because these are by far the most important fields in optics.
  14. [14]
    Absorption and Scattering of Light by Small Particles
    This self-contained, interdisciplinary study examines and illustrates how small particles absorb and scatter light.
  15. [15]
    [PDF] Part I: Electromagnetic Scattering, Absorption, Emission
    By directly solving the Maxwell equations, one can find the field scattered by an object in the form of a single body or a fixed cluster consisting of a limited ...Missing: seminal | Show results with:seminal
  16. [16]
    [PDF] Surface Integral Equations and the Boundary Element Method ...
    Mar 23, 2012 · Surface integral equation: Write an integral equation for effective electric and magnetic surface currents K(x), N(x) on the surface of the ...
  17. [17]
    Volume and surface integral equations for electromagnetic ...
    We derive and analyze two equivalent integral formulations for the time-harmonic electromagnetic scattering by a dielectric object.
  18. [18]
    Absorption and Scattering of Light by Small Particles - Google Books
    This self-contained, interdisciplinary study examines and illustrates how small particles absorb and scatter light.Missing: code | Show results with:code
  19. [19]
    [PDF] FORWARD SCATTERING MODEL FOR SPHERICAL ... - HAL
    Geometric Optics Approximation (GOA) valid in the near- and far-field ... large size parameter, i.e. 𝛼 = ka ≫ 1 where a is the particle radius, k = 2𝜋/λ ...
  20. [20]
  21. [21]
    Light scattering from water droplets in the geometrical optics ...
    In particular, geometrical optics methods are usually used for scatterers like water droplets and others that can be regarded as small spherical lenses. In this ...
  22. [22]
    [PDF] Identify the limits of geometric optics ray tracing by numerically ...
    Mar 30, 2020 · In the geometric optics method, the scattering field is decomposed into the diffracted field, and the fields resulting from reflection and ...
  23. [23]
    Single Scattering Properties of Atmospheric Ice Crystals
    Oct 1, 1996 · The present model describes the scattering at non- spherical ice crystals in the limit of classical geometric optics, that is, for particle ...
  24. [24]
    Mie Theory Examples - Ocean Optics Web Book
    Jan 6, 2021 · An equivalent statement is that the asymptotic value of the extinction cross section σ c is twice the particle's geometric cross section π ρ 2 .
  25. [25]
    Solar Radiative Transfer in Cirrus Clouds. Part I - AMS Journals
    The delta forward peak does not contribute to the normalization of the scattering phase function. If the diffracted light rays and geometric optics rays are ...
  26. [26]
    The mathematical physics of rainbows and glories - ScienceDirect.com
    The primary rainbow arises from scattering at about 138° from the forward direction, whereas the (backward) glory is associated with scattering very close to ...
  27. [27]
    Identify the limits of geometric optics ray tracing by numerically ...
    Mar 27, 2020 · It has been well known that geometric optics methods only give results with acceptable accuracy when the particle size parameter (defined as the ...
  28. [28]
    Geometrical Optics – light rays, ray tracing, paraxial ... - RP Photonics
    Geometrical optics uses rays to describe light propagation, useful for analyzing optical systems without tight confinement of light.
  29. [29]
  30. [30]
    Mie Theory Overview - Ocean Optics Web Book
    Oct 13, 2021 · Mie approached this problem by working out the solution to scattering of light by small spheres, starting with Maxwell's equations. His approach ...
  31. [31]
    A new explanation of the extinction paradox - ScienceDirect.com
    This paradox refers to the large size limit of a particle's extinction cross section. It is called a paradox because the geometrical optics approximation, which ...
  32. [32]
    [PDF] A new explanation of the extinction paradox - UNL Digital Commons
    Sep 9, 2010 · This paradox refers to the large size limit of a particle's extinction cross section. It is called a paradox because the geometrical optics.Missing: σ_ext π a²
  33. [33]
    Mie Scattering - OMLC
    Mie scattering calculations can be done using an interactive web page, C, Python, or MATLAB code. Fortran is considered the best.
  34. [34]
  35. [35]
    [PDF] T-matrix theory of electromagnetic scattering by particles and its ...
    Abstract. The T-matrix method is one of the most powerful and widely used theoretical techniques for the computation of electromagnetic scattering by single ...
  36. [36]
    Mishchenko 1991: Light scattering by randomly oriented axially ...
    Jul 10, 2023 · Waterman's T-matrix approach is used to develop a rigorous analytical method to compute the corresponding expansion coefficients. The main ...Missing: axisymmetric | Show results with:axisymmetric
  37. [37]
  38. [38]
    User Guide for the Discrete Dipole Approximation Code DDSCAT 7.3
    May 26, 2013 · This User Guide explains how to use DDSCAT 7.3 to carry out electromagnetic scattering calculations, including use of DDPOSTPROCESS, a Fortran-90 code.
  39. [39]
    Numerical solution of initial boundary value problems involving ...
    Numerical solution of initial boundary value problems involving maxwell's equations in isotropic media. Publisher: IEEE. Cite This. PDF. Kane Yee. All Authors.
  40. [40]
    Introduction - MEEP Documentation - Read the Docs
    Meep implements the finite-difference time-domain (FDTD) method for computational electromagnetics. This is a widely used technique in which space is divided ...<|control11|><|separator|>
  41. [41]
  42. [42]
    FDTD simulation of the optical properties for gold nanoparticles
    In this paper, the optical properties of gold nanoparticles were investigated by using the Finite-Difference Time-Domain (FDTD) method.
  43. [43]
    Finite-difference Time-domain (FDTD) Optical Simulations: A Primer ...
    This primer is intended to help researchers understand FDTD, implement the method to model optical effects, and learn about the benefits and limitations of ...
  44. [44]
    Allen Taflove - Engineering and Technology History Wiki
    Jan 13, 2016 · Since 1972, his FDTD research has pioneered fundamental theory, algorithms, and applications of the numerical technique for Maxwell's equations ...