Fact-checked by Grok 2 weeks ago

Electromagnetic wave equation

The electromagnetic wave equation is a fundamental partial differential equation in classical electromagnetism that governs the propagation of electric and magnetic fields as coupled waves in free space, predicting that these fields travel at the speed of light c = \frac{1}{\sqrt{\mu_0 \epsilon_0}} \approx 3 \times 10^8 m/s, where \mu_0 and \epsilon_0 are the permeability and permittivity of vacuum, respectively. Derived from Maxwell's equations in source-free regions (where charge density \rho = 0 and current density \mathbf{j} = 0), the equation arises by taking the curl of Faraday's law and Ampère's law with Maxwell's correction, leading to the coupled wave equations for the electric field \mathbf{E} and magnetic field \mathbf{B}: \nabla^2 \mathbf{E} - \frac{1}{c^2} \frac{\partial^2 \mathbf{E}}{\partial t^2} = 0 and \nabla^2 \mathbf{B} - \frac{1}{c^2} \frac{\partial^2 \mathbf{B}}{\partial t^2} = 0. These equations reveal that electromagnetic disturbances propagate as transverse waves, with \mathbf{E} and \mathbf{B} oscillating perpendicular to each other and to the direction of propagation, and their magnitudes related by E = cB. Plane wave solutions take the form \mathbf{E}(\mathbf{r}, t) = \mathbf{E}_0 \cos(\mathbf{k} \cdot \mathbf{r} - \omega t + \phi) and \mathbf{B}(\mathbf{r}, t) = \frac{1}{c} \hat{n} \times \mathbf{E}(\mathbf{r}, t), where \mathbf{k} is the wave vector, \omega is the angular frequency, and the dispersion relation \omega = kc holds, confirming the wave's speed as that of light. This unification demonstrated by James Clerk Maxwell in 1865 that light itself is an electromagnetic wave, encompassing phenomena from radio waves to gamma rays across the electromagnetic spectrum. The equation's implications extend to energy transport via the Poynting vector \mathbf{S} = \frac{1}{\mu_0} \mathbf{E} \times \mathbf{B}, which points in the propagation direction and quantifies the wave's intensity, and it obeys the superposition principle, allowing complex waveforms to be built from simpler ones. In media, modifications account for refractive indices, but the vacuum form remains the cornerstone for understanding electromagnetic radiation in physics and engineering applications like optics and wireless communication.

Derivation from Maxwell's Equations

Classical Derivation in Vacuum

In vacuum, where there are no charges or currents, Maxwell's equations simplify to the following set of partial differential equations: \nabla \cdot \mathbf{E} = 0, \quad \nabla \cdot \mathbf{B} = 0, \nabla \times \mathbf{E} = -\frac{\partial \mathbf{B}}{\partial t}, \quad \nabla \times \mathbf{B} = \mu_0 \epsilon_0 \frac{\partial \mathbf{E}}{\partial t}. These equations describe the electric field \mathbf{E} and magnetic field \mathbf{B} in free space. To derive the wave equation for \mathbf{E}, take the curl of Faraday's law \nabla \times \mathbf{E} = -\partial \mathbf{B}/\partial t: \nabla \times (\nabla \times \mathbf{E}) = -\frac{\partial}{\partial t} (\nabla \times \mathbf{B}). Substitute the vacuum form of Ampère's law \nabla \times \mathbf{B} = \mu_0 \epsilon_0 \partial \mathbf{E}/\partial t: \nabla \times (\nabla \times \mathbf{E}) = -\mu_0 \epsilon_0 \frac{\partial^2 \mathbf{E}}{\partial t^2}. Apply the vector identity \nabla \times (\nabla \times \mathbf{E}) = \nabla (\nabla \cdot \mathbf{E}) - \nabla^2 \mathbf{E}. Since \nabla \cdot \mathbf{E} = 0 in vacuum, this simplifies to: -\nabla^2 \mathbf{E} = -\mu_0 \epsilon_0 \frac{\partial^2 \mathbf{E}}{\partial t^2}, or equivalently, \nabla^2 \mathbf{E} - \mu_0 \epsilon_0 \frac{\partial^2 \mathbf{E}}{\partial t^2} = 0. This is the homogeneous wave equation for the electric field. A similar derivation yields the wave equation for the magnetic field. Taking the curl of Ampère's law \nabla \times \mathbf{B} = \mu_0 \epsilon_0 \partial \mathbf{E}/\partial t: \nabla \times (\nabla \times \mathbf{B}) = \mu_0 \epsilon_0 \frac{\partial}{\partial t} (\nabla \times \mathbf{E}). Substitute Faraday's law \nabla \times \mathbf{E} = -\partial \mathbf{B}/\partial t: \nabla \times (\nabla \times \mathbf{B}) = -\mu_0 \epsilon_0 \frac{\partial^2 \mathbf{B}}{\partial t^2}. Using the identity \nabla \times (\nabla \times \mathbf{B}) = \nabla (\nabla \cdot \mathbf{B}) - \nabla^2 \mathbf{B} and \nabla \cdot \mathbf{B} = 0: \nabla^2 \mathbf{B} - \mu_0 \epsilon_0 \frac{\partial^2 \mathbf{B}}{\partial t^2} = 0. Thus, both \mathbf{E} and \mathbf{B} satisfy the same form of the wave equation. The propagation speed c of these waves is given by c = 1 / \sqrt{\mu_0 \epsilon_0}, where \mu_0 is the permeability of free space and \epsilon_0 is the permittivity of free space. Numerical evaluation yields c \approx 3 \times 10^8 m/s, which matches the measured speed of light in vacuum, indicating that light is an electromagnetic wave. Solutions to these wave equations must also satisfy the divergence-free conditions \nabla \cdot \mathbf{E} = 0 and \nabla \mathbf{B} = 0, implying that the electric and magnetic fields are transverse to the direction of propagation.

Derivation with Sources

The full set of Maxwell's equations in SI units, incorporating charge density \rho and current density \mathbf{J} as sources in vacuum, are given by \nabla \cdot \mathbf{E} = \frac{\rho}{\varepsilon_0}, \quad \nabla \cdot \mathbf{B} = 0, \quad \nabla \times \mathbf{E} = -\frac{\partial \mathbf{B}}{\partial t}, \quad \nabla \times \mathbf{B} = \mu_0 \mathbf{J} + \mu_0 \varepsilon_0 \frac{\partial \mathbf{E}}{\partial t}. These equations describe the fundamental behavior of electromagnetic fields driven by sources. To derive the inhomogeneous wave equation for the electric field \mathbf{E}, begin by taking the curl of Faraday's law: \nabla \times (\nabla \times \mathbf{E}) = -\frac{\partial}{\partial t} (\nabla \times \mathbf{B}). Substitute the Ampère-Maxwell law into the right-hand side: \nabla \times (\nabla \times \mathbf{E}) = -\mu_0 \frac{\partial \mathbf{J}}{\partial t} - \mu_0 \varepsilon_0 \frac{\partial^2 \mathbf{E}}{\partial t^2}. Apply the vector identity \nabla \times (\nabla \times \mathbf{E}) = \nabla (\nabla \cdot \mathbf{E}) - \nabla^2 \mathbf{E} and insert Gauss's law \nabla \cdot \mathbf{E} = \rho / \varepsilon_0: \nabla \left( \frac{\rho}{\varepsilon_0} \right) - \nabla^2 \mathbf{E} = -\mu_0 \frac{\partial \mathbf{J}}{\partial t} - \mu_0 \varepsilon_0 \frac{\partial^2 \mathbf{E}}{\partial t^2}. Rearranging yields the inhomogeneous wave equation for \mathbf{E}: \nabla^2 \mathbf{E} - \mu_0 \varepsilon_0 \frac{\partial^2 \mathbf{E}}{\partial t^2} = \mu_0 \frac{\partial \mathbf{J}}{\partial t} + \frac{1}{\varepsilon_0} \nabla \rho, or equivalently, using c^2 = 1/(\mu_0 \varepsilon_0), \nabla^2 \mathbf{E} - \frac{1}{c^2} \frac{\partial^2 \mathbf{E}}{\partial t^2} = \mu_0 \frac{\partial \mathbf{J}}{\partial t} + \frac{\mu_0 c^2}{\varepsilon_0} \nabla \rho = \mu_0 \left( \frac{\partial \mathbf{J}}{\partial t} + c^2 \nabla \rho \right). The right-hand side represents the driving terms from currents and charges, with the \partial \mathbf{J}/\partial t term contributing to transverse waves and the \nabla \rho term to longitudinal components. A similar procedure derives the equation for the magnetic field \mathbf{B}. Take the curl of the Ampère-Maxwell law: \nabla \times (\nabla \times \mathbf{B}) = \mu_0 \nabla \times \mathbf{J} + \mu_0 \varepsilon_0 \frac{\partial}{\partial t} (\nabla \times \mathbf{E}). Substitute Faraday's law: \nabla \times (\nabla \times \mathbf{B}) = \mu_0 \nabla \times \mathbf{J} - \mu_0 \varepsilon_0 \frac{\partial^2 \mathbf{B}}{\partial t^2}. Using the identity \nabla \times (\nabla \times \mathbf{B}) = \nabla (\nabla \cdot \mathbf{B}) - \nabla^2 \mathbf{B} and Gauss's law for magnetism \nabla \cdot \mathbf{B} = 0, -\nabla^2 \mathbf{B} = \mu_0 \nabla \times \mathbf{J} - \mu_0 \varepsilon_0 \frac{\partial^2 \mathbf{B}}{\partial t^2}, yielding \nabla^2 \mathbf{B} - \mu_0 \varepsilon_0 \frac{\partial^2 \mathbf{B}}{\partial t^2} = -\mu_0 \nabla \times \mathbf{J}, or \nabla^2 \mathbf{B} - \frac{1}{c^2} \frac{\partial^2 \mathbf{B}}{\partial t^2} = -\mu_0 \nabla \times \mathbf{J}. The source term here arises solely from the curl of the current density, ensuring transverse propagation consistent with \nabla \cdot \mathbf{B} = 0. The continuity equation \nabla \cdot \mathbf{J} + \partial \rho / \partial t = 0, which expresses local charge conservation, ensures consistency between the source terms in the wave equations. Taking the divergence of the Ampère-Maxwell law confirms this relation, as \partial / \partial t (\nabla \cdot \mathbf{E}) = (1/\varepsilon_0) \partial \rho / \partial t = -\nabla \cdot \mathbf{J} / \varepsilon_0. Without it, the equations would be overconstrained for arbitrary \rho and \mathbf{J}. Although the field equations are gauge-invariant, introducing scalar \phi and vector \mathbf{A} potentials via \mathbf{B} = \nabla \times \mathbf{A} and \mathbf{E} = -\nabla \phi - \partial \mathbf{A} / \partial t requires a gauge choice to simplify the resulting equations. The Lorenz gauge \nabla \cdot \mathbf{A} + (1/c^2) \partial \phi / \partial t = 0 decouples them into independent inhomogeneous wave equations: \nabla^2 \phi - (1/c^2) \partial^2 \phi / \partial t^2 = -\rho / \varepsilon_0 and \nabla^2 \mathbf{A} - (1/c^2) \partial^2 \mathbf{A} / \partial t^2 = -\mu_0 \mathbf{J}. This derivation highlights how sources \rho and \mathbf{J} drive electromagnetic wave propagation, a central result from Maxwell's unification of electricity, magnetism, and optics in his 1865 paper. When \rho = 0 and \mathbf{J} = 0, the equations reduce to the homogeneous vacuum case.

Historical and Physical Context

Origins in 19th-Century Electromagnetism

The foundations of electromagnetic theory in the 19th century were laid by André-Marie Ampère, who in 1820 formulated a law describing the magnetic force between electric currents, establishing a quantitative link between electricity and magnetism. This work built on Hans Christian Ørsted's 1820 discovery of electromagnetism but introduced mathematical precision without predicting wave propagation. Michael Faraday advanced the conceptual framework in the 1830s through his experiments on electromagnetic induction, discovered in 1831, and his introduction of field lines as a way to visualize magnetic and electric influences as continuous entities rather than discrete actions. Faraday's qualitative ideas, detailed in his 1832 paper, emphasized fields permeating space, shifting away from instantaneous action-at-a-distance models prevalent in Newtonian mechanics, though they lacked a full mathematical synthesis. This paradigm began challenging the dominant view of forces acting directly across distances without intermediaries, fostering skepticism among proponents of contact-action theories who saw fields as mere metaphors. James Clerk Maxwell synthesized these elements between 1861 and 1865, culminating in his 1864 paper "A Dynamical Theory of the Electromagnetic Field," where he predicted the existence of electromagnetic waves propagating at the speed of light by incorporating a displacement current term into Ampère's law. Maxwell's equations, as presented in this work and refined in 1865, unified electricity, magnetism, and optics, implying transverse waves in the electromagnetic field. Initial reception was mixed, with some physicists like William Thomson (Lord Kelvin) expressing doubt due to the theory's departure from mechanical analogies and action-at-a-distance intuitions, marking a gradual shift toward field-based electromagnetism. In the 1880s, Oliver Heaviside reformulated Maxwell's twenty equations into the compact vector form still used today, emphasizing the wave-like propagation of fields in his 1885 work. Concurrently, John Poynting developed the Poynting vector in 1884 to describe electromagnetic energy flow, further illuminating the wave nature of the theory. The definitive confirmation came in 1887 from Heinrich Hertz's experiments, which generated and detected electromagnetic waves traveling at the speed of light, validating Maxwell's predictions and solidifying the field theory paradigm over action-at-a-distance models. This timeline—from Faraday's 1831 induction, to Maxwell's 1865 equations, Heaviside and Poynting's 1880s clarifications, to Hertz's 1887 waves—underscored the revolutionary transition to a unified electromagnetic field theory.

Physical Interpretation and Significance

The solutions to the electromagnetic wave equation describe propagating disturbances in the electric and magnetic fields that travel through vacuum at the constant speed c = \frac{1}{\sqrt{\mu_0 \epsilon_0}} \approx 3 \times 10^8 m/s, representing self-sustaining oscillations where changing electric fields generate magnetic fields and vice versa. For plane wave solutions, the electric field \mathbf{E} and magnetic field \mathbf{B} are perpendicular to the direction of propagation and to each other, oscillating in phase with equal magnitudes related by B = E/c. The wave equation implies that electromagnetic waves in vacuum are purely transverse, with no longitudinal components, as the divergence conditions \nabla \cdot \mathbf{E} = 0 and \nabla \cdot \mathbf{B} = 0 prohibit field variations along the propagation direction for free-space solutions. These waves are dispersionless in vacuum, propagating at speed c independent of frequency, though they become dispersive in media where the refractive index varies with wavelength. The electromagnetic wave equation unifies the previously separate domains of electricity, magnetism, and optics by demonstrating that visible light and all optical phenomena are manifestations of electromagnetic waves, thereby resolving the longstanding conceptual gap in 19th-century physics. This unification is evident in the energy transport carried by these waves, quantified by the Poynting vector \mathbf{S} = \frac{1}{\mu_0} \mathbf{E} \times \mathbf{B}, which gives the directional energy flux (power per unit area) and aligns with the propagation direction, with its time-averaged magnitude yielding the wave intensity. The equation's prediction of a universal propagation speed c, invariant across inertial frames, provided the cornerstone for Albert Einstein's development of special relativity in 1905, as the constancy of light speed necessitated a reformulation of space and time to reconcile electromagnetism with the principle of relativity. In quantum electrodynamics (QED), the classical wave equation underpins the quantization of the electromagnetic field, where photons emerge as the discrete quanta of these waves, each carrying energy \hbar \omega and momentum \hbar \mathbf{k} in modes satisfying the dispersion relation \omega = c k. This framework enables precise descriptions of light-matter interactions at quantum scales. The electromagnetic wave equation remains foundational to modern technologies, including radio wave transmission, optical systems, and photonics devices that manipulate light for communication and sensing. It also drives computational electromagnetics, particularly through finite-difference time-domain (FDTD) methods, which numerically solve the wave equation on spatiotemporal grids to simulate complex wave propagation; these techniques originated with Kane Yee's 1966 algorithm and were advanced by Allen Taflove's work in the 1970s, with the term "FDTD" coined in his 1980 IEEE paper, enabling applications in antenna design and photonic structures.

Homogeneous Wave Equation

Form in Flat Spacetime

In flat spacetime, described by Minkowski space with metric signature (+, -, -, -), the homogeneous electromagnetic wave equation governs the propagation of electric and magnetic fields in vacuum, free from sources. The electric field \mathbf{E} and magnetic field \mathbf{B} each satisfy the scalar wave equation in three spatial dimensions plus time: \left( \nabla^2 - \frac{1}{c^2} \frac{\partial^2}{\partial t^2} \right) \mathbf{E} = 0, \quad \left( \nabla^2 - \frac{1}{c^2} \frac{\partial^2}{\partial t^2} \right) \mathbf{B} = 0, where c is the speed of light, \nabla^2 is the Laplacian operator, and the equations derive from Maxwell's equations in the absence of charges and currents. These are hyperbolic partial differential equations (PDEs), with characteristic surfaces forming light cones that define the causal structure of field propagation at speed c. To solve these equations, electromagnetic fields are expressed in terms of scalar and vector potentials, \phi and \mathbf{A}, respectively, such that \mathbf{B} = \nabla \times \mathbf{A} and \mathbf{E} = -\nabla \phi - \frac{\partial \mathbf{A}}{\partial t}. In the covariant formulation of special relativity, these potentials combine into a 4-vector A^\mu = \left( \frac{\phi}{c}, \mathbf{A} \right), and the electromagnetic field tensor is F_{\mu\nu} = \partial_\mu A_\nu - \partial_\nu A_\mu. The field tensor satisfies the homogeneous Maxwell equations \partial_\mu F^{\mu\nu} = 0 in vacuum, which, under an appropriate gauge, yield decoupled wave equations for the potentials. The Lorenz gauge condition, \partial_\mu A^\mu = 0, simplifies the equations by decoupling the components, leading to the Klein-Gordon-like form for each: \square A^\mu = 0, where the d'Alembertian operator is \square = \partial^\mu \partial_\mu = \frac{1}{c^2} \frac{\partial^2}{\partial t^2} - \nabla^2. This gauge is preferred in relativistic contexts for its Lorentz invariance, unlike the Coulomb gauge \nabla \cdot \mathbf{A} = 0, which is non-covariant but useful in non-relativistic approximations; the choice reflects the inherent gauge freedom in Maxwell's equations, where A^\mu \to A^\mu + \partial^\mu \Lambda for any scalar \Lambda leaves F_{\mu\nu} unchanged. In vacuum, solutions to these equations represent free electromagnetic waves that propagate harmonically in both time and space, with transverse polarizations orthogonal to the direction of propagation.

Covariant Formulation

The covariant formulation of the homogeneous electromagnetic wave equation expresses Maxwell's equations in terms of four-dimensional spacetime tensors, ensuring manifest invariance under Lorentz transformations. The homogeneous Maxwell equations in vacuum take the form \partial_\mu F^{\mu\nu} = 0, where F^{\mu\nu} is the antisymmetric electromagnetic field strength tensor, with components incorporating both electric and magnetic fields: F^{0i} = -E^i/c and F^{ij} = -\epsilon^{ijk} B_k in the mostly minus metric signature (+,-,-,-). This tensor is derived from the four-potential A^\mu = (\phi/c, \mathbf{A}) via F_{\mu\nu} = \partial_\mu A_\nu - \partial_\nu A_\mu, where the indices are raised and lowered with the Minkowski metric \eta_{\mu\nu}. Substituting the potential expression into the homogeneous equations yields the wave equation for A^\nu. In the Lorenz gauge, defined by the condition \partial_\mu A^\mu = 0, the equation simplifies to \square A^\nu = 0, where \square = \partial^\mu \partial_\mu = \frac{1}{c^2} \partial_t^2 - \nabla^2 is the d'Alembertian operator. Without the gauge condition, the general form is \square A^\nu - \partial^\nu (\partial_\mu A^\mu) = 0. This formulation highlights the relativistic structure of electromagnetism, as the wave equation propagates at speed c invariantly across inertial frames. The covariant approach offers key advantages, including explicit Lorentz invariance, which facilitates applications in particle physics, such as the photon propagator in quantum electrodynamics, where \square A^\nu = 0 underlies the Feynman rules for photon exchange. Albert Einstein's work from 1905 to 1908 demonstrated that this covariance reconciles electromagnetism with special relativity, resolving apparent conflicts between Maxwell's equations and Newtonian mechanics. Additionally, Noether's theorem connects the gauge invariance of the electromagnetic action to the conservation of electric charge, as the global U(1) symmetry generates a conserved current j^\mu satisfying \partial_\mu j^\mu = 0. For hypothetical magnetic monopoles, the dual field tensor *F^{\mu\nu} = \frac{1}{2} \epsilon^{\mu\nu\rho\sigma} F_{\rho\sigma} (with \epsilon^{0123} = +1) interchanges electric and magnetic roles, leading to symmetric equations \partial_\mu (*F)^{\mu\nu} = 0 in the absence of sources, preserving the wave equation's form under duality transformations.

Extension to Curved Spacetime

In curved spacetime, the homogeneous Maxwell equations are expressed covariantly as \nabla_\mu F^{\mu\nu} = 0, where \nabla_\mu denotes the covariant derivative associated with the Levi-Civita connection of the spacetime metric g_{\mu\nu}, and F^{\mu\nu} is the electromagnetic field strength tensor. This formulation replaces the flat-space partial derivatives with covariant ones to account for gravitational effects on electromagnetic propagation. For the four-potential A^\nu, the generalized Lorenz gauge condition \nabla_\mu A^\mu = 0 leads to a wave equation of the form \Box_g A^\nu - R^\nu{}_\lambda A^\lambda = 0, where \Box_g = g^{\mu\nu} \nabla_\mu \nabla_\nu is the covariant d'Alembertian operator, and R^\nu{}_\lambda is the Ricci curvature tensor. The explicit expansion of \Box_g A^\nu incorporates Christoffel symbols \Gamma^\lambda_{\mu\nu} through the second covariant derivative, \nabla_\mu \nabla_\nu A^\rho = \partial_\mu (\nabla_\nu A^\rho) - \Gamma^\sigma_{\mu\nu} \nabla_\sigma A^\rho. In weak-field approximations, where g_{\mu\nu} \approx \eta_{\mu\nu} + h_{\mu\nu} with |h_{\mu\nu}| \ll 1, the equation approximates the flat-spacetime wave equation plus first-order corrections from metric perturbations h_{\mu\nu} and their derivatives. This covariant extension of the electromagnetic wave equation was integrated into general relativity by Einstein in 1915, with electromagnetic fields entering the theory via their stress-energy tensor in the Einstein field equations. Key physical implications include gravitational lensing, where electromagnetic waves (such as light) follow null geodesics bent by spacetime curvature around massive bodies, altering observed paths and intensities. In cosmological settings, like the expanding universe described by the Friedmann-Lemaître-Robertson-Walker metric, propagating electromagnetic waves undergo gravitational redshift, with frequency shifts proportional to the scale factor. In black hole physics, solutions to this wave equation describe electromagnetic perturbations around horizons, underpinning classical aspects of phenomena like Hawking radiation, where quantum field modes in curved geometry lead to particle emission. Since the 1990s, numerical relativity simulations have advanced the study of these equations in strong-field regimes, such as magnetized binary black hole mergers, by evolving the coupled Einstein-Maxwell system on adaptive meshes to capture dynamical interactions. This framework generalizes the flat-spacetime covariant formulation as a zeroth-order limit when the Ricci tensor vanishes.

Inhomogeneous Wave Equation

General Form and Sources

The inhomogeneous electromagnetic wave equation arises directly from Maxwell's equations in the presence of sources, specifically the charge density \rho and current density \mathbf{J}. In covariant form, the source equation is \partial_\mu F^{\mu\nu} = \mu_0 J^\nu, where F^{\mu\nu} is the electromagnetic field strength tensor and J^\nu = (c\rho, \mathbf{J}) is the four-current density in SI units. This equation describes how sources generate the fields, with the left-hand side encoding the wave-like propagation. To solve Maxwell's equations, it is often convenient to introduce the four-potential A^\nu = (\phi/c, \mathbf{A}), where \phi is the scalar potential and \mathbf{A} is the vector potential. In the Lorenz gauge, defined by \partial_\mu A^\mu = 0, the four-potential satisfies the inhomogeneous wave equation \square A^\nu = -\mu_0 J^\nu, where \square = \partial_\mu \partial^\mu = \nabla^2 - \frac{1}{c^2} \frac{\partial^2}{\partial t^2} is the d'Alembertian operator. Decomposing into components, the scalar potential obeys \square \phi = -\frac{\rho}{\epsilon_0}, and the vector potential satisfies \square \mathbf{A} = -\mu_0 \mathbf{J}. These decoupled equations highlight how charges and currents directly drive the potentials, from which the fields are derived via \mathbf{E} = -\nabla \phi - \frac{\partial \mathbf{A}}{\partial t} and \mathbf{B} = \nabla \times \mathbf{A}. The wave equations can also be expressed directly in terms of the fields. For the electric field, taking the curl of Faraday's law and substituting Ampère's law yields \nabla^2 \mathbf{E} - \frac{1}{c^2} \frac{\partial^2 \mathbf{E}}{\partial t^2} = \frac{1}{\epsilon_0} \nabla \rho + \mu_0 \frac{\partial \mathbf{J}}{\partial t}. Similarly, for the magnetic field, \nabla^2 \mathbf{B} - \frac{1}{c^2} \frac{\partial^2 \mathbf{B}}{\partial t^2} = -\mu_0 \nabla \times \mathbf{J}. Here, c = 1/\sqrt{\mu_0 \epsilon_0} is the speed of light in vacuum. The right-hand sides act as source terms: the \nabla \rho term arises from spatial variations in charge density, while the \partial \mathbf{J}/\partial t and \nabla \times \mathbf{J} terms capture time-varying and curl components of the current, respectively. These sources generate electromagnetic waves, with radiation particularly prominent from accelerating charges, as the fields propagate disturbances away from the source at speed c. The general solutions to these equations are the retarded potentials, which integrate the sources over the past light cone to enforce causality: \phi(\mathbf{r}, t) = \frac{1}{4\pi \epsilon_0} \int \frac{\rho(\mathbf{r}', t - |\mathbf{r} - \mathbf{r}'|/c)}{|\mathbf{r} - \mathbf{r}'|} d^3\mathbf{r}', and analogously for \mathbf{A}. This form ensures that the fields at a point depend only on sources at retarded times, reflecting the finite propagation speed of electromagnetic disturbances. In SI units, the explicit appearance of \mu_0 and \epsilon_0 ties the equations to measurable constants, with the source strengths scaled by these factors. In contrast, Gaussian units eliminate \mu_0 and \epsilon_0 (setting them to unity in vacuum) but introduce factors of $4\pi in the source terms and retain c explicitly throughout, simplifying some relativistic expressions at the cost of non-standard force laws. When sources vanish (J^\nu = 0), these reduce to the homogeneous wave equations describing free propagation. Beyond classical electrodynamics, quantum electrodynamics (QED) introduces vacuum polarization as an effective source term, arising from virtual electron-positron pairs that modify the propagation of electromagnetic waves. This quantum correction, first quantified in the effective Heisenberg-Euler Lagrangian in the 1930s and refined in the 1940s with the development of renormalization, alters Maxwell's equations by adding nonlinear terms proportional to powers of the field strengths, effectively treating the vacuum as a polarizable medium. These effects become significant in strong fields, such as near black holes or in laser experiments, providing a bridge between classical wave equations and quantum field theory.

Green's Function Approach

The Green's function approach provides an integral solution to the inhomogeneous electromagnetic wave equation, which arises from time-varying sources such as charges and currents. In the Lorentz gauge, the scalar potential \phi and vector potential \mathbf{A} satisfy the inhomogeneous wave equations \square \phi = -\rho / \epsilon_0 and \square \mathbf{A} = -\mu_0 \mathbf{J}, where \square = \nabla^2 - \frac{1}{c^2} \frac{\partial^2}{\partial t^2} is the d'Alembertian operator. To solve these, one introduces the Green's function G(\mathbf{x}, t; \mathbf{x}', t') that satisfies \square G = -\delta^3(\mathbf{x} - \mathbf{x}') \delta(t - t'). The retarded Green's function, which enforces causality, is given by G(\mathbf{x}, t; \mathbf{x}', t') = \frac{\delta\left(t - t' - \frac{|\mathbf{x} - \mathbf{x}'|}{c}\right)}{4\pi |\mathbf{x} - \mathbf{x}'|}. This form ensures that the potential at a point depends only on sources at earlier times, consistent with the finite speed of light. The general solutions for the potentials are then expressed as retarded integrals over the sources: \phi(\mathbf{x}, t) = \frac{1}{4\pi \epsilon_0} \int \frac{[\rho(\mathbf{x}', t - \frac{|\mathbf{x} - \mathbf{x}'|}{c})]}{|\mathbf{x} - \mathbf{x}'|} d^3\mathbf{x}', \mathbf{A}(\mathbf{x}, t) = \frac{\mu_0}{4\pi} \int \frac{[\mathbf{J}(\mathbf{x}', t - \frac{|\mathbf{x} - \mathbf{x}'|}{c})]}{|\mathbf{x} - \mathbf{x}'|} d^3\mathbf{x}', where the square brackets denote evaluation at the retarded time t_r = t - |\mathbf{x} - \mathbf{x}'|/c. In four-vector notation, this becomes A^\mu(x) = \mu_0 \int G(x, x') J^\mu(x') d^4 x', with G(x, x') = \frac{\delta\left( (t - t') - \frac{|\mathbf{x} - \mathbf{x}'|}{c} \right)}{4\pi |\mathbf{x} - \mathbf{x}'|}. An advanced Green's function exists with \delta(t - t' + |\mathbf{x} - \mathbf{x}'|/c), but it violates causality by implying effects precede causes; physical solutions require the retarded form to respect the principle that influences propagate forward in time. For a single point charge q moving with velocity \mathbf{v}(t'), the retarded potentials take the explicit Liénard-Wiechert form: \phi(\mathbf{x}, t) = \frac{1}{4\pi \epsilon_0} \left[ \frac{q}{ (1 - \mathbf{n} \cdot \boldsymbol{\beta}) R} \right]_{t_r}, \quad \mathbf{A}(\mathbf{x}, t) = \frac{\mu_0}{4\pi} \left[ \frac{q \mathbf{v}}{ (1 - \mathbf{n} \cdot \boldsymbol{\beta}) R} \right]_{t_r}, where \boldsymbol{\beta} = \mathbf{v}/c, \mathbf{n} is the unit vector from the retarded position to the observation point, R = |\mathbf{x} - \mathbf{x}'(t_r)|, and all quantities are evaluated at the retarded time solving t - t_r = R(t_r)/c. In the non-relativistic limit v \ll c, \boldsymbol{\beta} \to 0, reducing to the familiar Coulomb and Biot-Savart forms with retardation. These potentials were pivotal in Hendrik Lorentz's 1900 electron theory, where they described the electromagnetic fields of accelerating charges and enabled calculations of radiation from electric dipoles. For complex, extended sources beyond simple point charges, the integral expressions are often evaluated numerically. The method of moments emerged as a key technique in antenna design, discretizing the retarded potential integrals into matrix equations solved for current distributions on wire or surface structures, facilitating accurate predictions of radiation patterns for practical devices.

Solution Techniques

Plane Wave Solutions

Plane wave solutions represent the fundamental propagating modes of the homogeneous electromagnetic wave equation in vacuum, characterized by infinite, flat wavefronts that advance without changing shape or spreading. These solutions are monochromatic, assuming a single frequency, and serve as building blocks for more complex wave phenomena through linear superposition. The electric field of a plane wave is expressed in complex form as \mathbf{E}(\mathbf{r}, t) = \mathbf{E}_0 \, \operatorname{Re} \left[ \exp \left( i (\mathbf{k} \cdot \mathbf{r} - \omega t) \right) \right], where \mathbf{E}_0 is the constant complex amplitude vector, \mathbf{k} is the real wave vector pointing in the propagation direction with magnitude k = |\mathbf{k}|, \mathbf{r} is the position vector, and \omega > 0 is the angular frequency. Substituting this ansatz into the homogeneous wave equation \nabla^2 \mathbf{E} - \frac{1}{c^2} \frac{\partial^2 \mathbf{E}}{\partial t^2} = 0 (with c = 1/\sqrt{\mu_0 \epsilon_0}) yields the dispersion relation k^2 = \omega^2 / c^2, or equivalently \omega = c k, which dictates that the phase velocity equals the speed of light in vacuum. From Maxwell's equations in vacuum, specifically \nabla \cdot \mathbf{E} = 0 and \nabla \cdot \mathbf{B} = 0, the plane wave solutions exhibit transversality: \mathbf{k} \cdot \mathbf{E}_0 = 0 and \mathbf{k} \cdot \mathbf{B}_0 = 0, meaning both fields are perpendicular to the propagation direction. The associated magnetic field is \mathbf{B}(\mathbf{r}, t) = \frac{1}{\omega} \mathbf{k} \times \mathbf{E}(\mathbf{r}, t), or in amplitude form \mathbf{B}_0 = \frac{1}{c} \hat{k} \times \mathbf{E}_0, ensuring \mathbf{E}, \mathbf{B}, and \mathbf{k} form a mutually orthogonal triad with |\mathbf{B}_0| = |\mathbf{E}_0| / c. Polarization describes the orientation of the electric field oscillation in the plane perpendicular to \mathbf{k}. For linear polarization, \mathbf{E}_0 is fixed in direction, resulting in \mathbf{E} oscillating along a straight line. Circular polarization occurs when \mathbf{E}_0 has equal components along two orthogonal directions with a 90-degree phase difference, such as \mathbf{E}_0 = E_0 (\hat{e}_1 \pm i \hat{e}_2)/\sqrt{2}, where \hat{e}_1 and \hat{e}_2 are unit vectors perpendicular to \mathbf{k}; the + sign corresponds to left-handed (counterclockwise when looking toward the source) and the - to right-handed polarization, defined by the helicity relative to the propagation direction. Elliptical polarization generalizes this with unequal amplitudes and arbitrary phase. The energy transport is quantified by the time-averaged Poynting vector \langle \mathbf{S} \rangle = \frac{\epsilon_0 c}{2} |\mathbf{E}_0|^2 \hat{k}, which points in the \mathbf{k} direction and gives the intensity (energy flux) of the wave; the associated momentum density is \mathbf{g} = \langle \mathbf{S} \rangle / c^2, reflecting the electromagnetic field's momentum per unit volume. In vacuum, all solutions to the homogeneous wave equation can be decomposed as a Fourier superposition of such plane waves, leveraging the linearity of Maxwell's equations under the monochromatic approximation for simplicity. At dielectric interfaces, plane waves incident beyond the critical angle undergo total internal reflection, producing an evanescent wave in the lower-index medium where the perpendicular wave vector component becomes imaginary, leading to exponential decay without net energy propagation.

Monochromatic and Steady-State Solutions

In the analysis of electromagnetic waves, monochromatic solutions assume a sinusoidal time dependence, allowing the separation of spatial and temporal variables. For the electric field, this is expressed as \mathbf{E}(\mathbf{r}, t) = \operatorname{Re} \left[ \mathbf{E}(\mathbf{r}) e^{-i \omega t} \right], where \mathbf{E}(\mathbf{r}) is a complex vector amplitude and \omega is the angular frequency. Substituting this form into the homogeneous wave equation \nabla^2 \mathbf{E} - \frac{1}{c^2} \frac{\partial^2 \mathbf{E}}{\partial t^2} = 0 yields the Helmholtz equation \nabla^2 \mathbf{E} + k^2 \mathbf{E} = 0, with k = \omega / c the wavenumber and c the speed of light in vacuum. This frequency-domain formulation simplifies the study of wave propagation for single-frequency fields. Steady-state solutions arise when sources vary sinusoidally at frequency \omega, leading to time-independent complex amplitudes that satisfy the Helmholtz equation. The phasor method, analogous to steady-state analysis in AC circuits, replaces time derivatives with multiplication by -i \omega, transforming Maxwell's equations into algebraic relations for the complex fields. This approach is particularly useful for driven systems, where the response reaches a periodic steady state after transients decay. For unbounded propagation, monochromatic solutions include spherical waves, which represent outgoing radiation from a point source. The scalar form for such a wave is \psi(\mathbf{r}) \propto \frac{e^{i k r}}{r}, satisfying the Helmholtz equation in spherical coordinates and the Sommerfeld radiation condition \lim_{r \to \infty} r \left( \frac{\partial}{\partial r} - i k \right) \psi = 0 to ensure energy flows outward without incoming waves from infinity. Plane waves emerge as a special limiting case of these spherical waves at large distances in a specific direction. In optics, monochromatic solutions are essential for modeling coherent sources like lasers, where the output is nearly single-frequency, enabling precise interference and phase control. The impedance of free space, Z_0 = \sqrt{\mu_0 / \epsilon_0} \approx 376.73 \, \Omega, relates the magnitudes of the electric and magnetic fields in a plane monochromatic wave via |\mathbf{E}| = Z_0 |\mathbf{H}|, characterizing the wave's intrinsic properties in vacuum. Boundary value problems for monochromatic waves often involve interfaces between dielectrics, where continuity of tangential \mathbf{E} and \mathbf{H}, and normal \mathbf{D} and \mathbf{B}, determines reflection and transmission. The Fresnel coefficients quantify these amplitudes for plane waves incident on a planar boundary, depending on the angle of incidence, polarization (s- or p-wave), and refractive indices of the media. In modern ultrafast optics since the 1990s, the monochromatic approximation extends to femtosecond pulses when their spectral bandwidth \Delta \omega satisfies \Delta \omega \ll \omega, allowing treatment as quasi-monochromatic waves with a slowly varying envelope modulating the carrier frequency. This facilitates analysis of pulse propagation and nonlinear effects in media.

Spectral Decomposition Methods

Spectral decomposition methods represent general solutions to the electromagnetic wave equation by expanding arbitrary waveforms into superpositions of basis functions, typically plane waves, which serve as fundamental modes satisfying the dispersion relation. The electric field \mathbf{E}(\mathbf{r}, t) can be expressed via the Fourier transform as \mathbf{E}(\mathbf{r}, t) = \int \tilde{\mathbf{E}}(\mathbf{k}, \omega) \exp[i(\mathbf{k} \cdot \mathbf{r} - \omega t)] \, d^3k \, d\omega, where the spectral support is constrained to the light cone \omega = c |\mathbf{k}| for solutions in vacuum, ensuring causality and the speed-of-light propagation. This integral representation decomposes broadband or transient fields into their frequency-wavenumber components, facilitating analysis of wave propagation and interaction. The inverse transform recovers the spatial-temporal field from the spectrum \tilde{\mathbf{E}}(\mathbf{k}, \omega), which encodes amplitude and phase information across all frequencies. In one dimension, the homogeneous wave equation admits the d'Alembert solution for initial conditions E(x,0) = \phi(x) and E_t(x,0) = \psi(x), given by E(x,t) = \frac{1}{2} [\phi(x - ct) + \phi(x + ct)] + \frac{1}{2c} \int_{x-ct}^{x+ct} \psi(s) \, ds, which separates the field into arbitrary right- and left-propagating profiles f(x - ct) and g(x + ct). This form highlights the non-dispersive nature of 1D electromagnetic waves, where pulse shapes remain undistorted during propagation. For electromagnetic contexts, such as transmission lines or plane-stratified media, this solution directly applies to the scalar components of \mathbf{E} and \mathbf{B}. In three dimensions, the initial value problem for the wave equation is solved using Kirchhoff's formula, which expresses the field at (\mathbf{x}, t) as a surface integral over the sphere of radius ct centered at \mathbf{x}: E(\mathbf{x}, t) = \frac{1}{4\pi c^2 t} \int_{|\mathbf{y} - \mathbf{x}| = ct} \left[ t \frac{\partial E}{\partial n}(\mathbf{y}, 0) + E(\mathbf{y}, 0) + (\mathbf{y} - \mathbf{x}) \cdot \nabla E(\mathbf{y}, 0) \right] dS(\mathbf{y}), where the integrals involve initial data E(\mathbf{y}, 0) and \partial_t E(\mathbf{y}, 0), with the normal derivative \partial_n on the light cone surface. This formula embodies Huygens' principle, localizing the influence of initial conditions to the backward light cone, and is essential for modeling transient electromagnetic disturbances like pulses from antennas. Spectral methods also enable analysis of bandwidth and pulse shaping in electromagnetic signals, where Gaussian pulses, characterized by \exp(-t^2 / 2\tau^2), exhibit minimal time-bandwidth product \Delta t \Delta \omega \approx 1, optimizing for short-duration, broad-spectrum waves in applications like ultrafast optics. Chirped pulses, with linearly varying instantaneous frequency, allow dispersion management by compensating material effects, extending pulse propagation distances. Parseval's theorem ensures energy conservation across domains, equating time-domain integral \int |E(t)|^2 dt to frequency-domain \int |\tilde{E}(\omega)|^2 d\omega / 2\pi, crucial for quantifying pulse energy in radar and laser systems. The foundations of these spectral techniques trace to Joseph Fourier's 1822 solution of the heat equation using series expansions, which inspired their adaptation to hyperbolic wave equations for separating variables and integral representations in electromagnetism. This approach became essential for signal processing in electromagnetic contexts, enabling efficient computation of wave spectra. For non-stationary waves, where Fourier methods assume stationarity, wavelet decompositions provide time-frequency localization, as developed in 1990s radar signal analysis for handling transient reflections and scattering. Electromagnetic wavelets, such as those based on localized monopoles or dipoles, decompose fields into scalable, translatable basis functions suited to irregular pulses in subsurface radar.

Multipole Expansion

The multipole expansion provides a powerful method for analyzing the far-field radiation from localized electromagnetic sources, such as oscillating charges or currents confined to a small region compared to the wavelength. This expansion decomposes the solutions to the electromagnetic wave equation into angular and radial components using spherical harmonics, separating the interior (near-field) and exterior (radiation) behaviors. For radiating systems, the scalar potential \phi can be expressed in the exterior region as a series involving outgoing spherical waves: \phi(\mathbf{r}, t) = \sum_{l=0}^\infty \sum_{m=-l}^l B_l^m h_l^{(1)}(kr) Y_l^m(\theta, \phi) e^{-i\omega t}, where h_l^{(1)}(kr) is the spherical Hankel function of the first kind representing outgoing waves, k = \omega / c is the wavenumber, and the coefficients B_l^m are determined by the source multipole moments. In the interior region, regular spherical Bessel functions j_l(kr) replace the Hankel functions to ensure finiteness at the origin. This form arises from solving the Helmholtz equation \nabla^2 \phi + k^2 \phi = 0 in spherical coordinates, with the radiation term dominating in the far field (kr \gg 1) where h_l^{(1)}(kr) \approx (-i)^{l+1} e^{i kr}/(kr). In the near field (kr \ll 1), h_l^{(1)}(kr) \approx -i \frac{(2l-1)!!}{(kr)^{l+1}}, recovering the static multipole form \propto 1/r^{l+1}. The lowest-order term in the expansion, the electric dipole (l=1), is typically dominant for non-symmetric sources unless forbidden by symmetry. For an electric dipole moment \mathbf{p} = p \hat{z} e^{-i\omega t}, the far-field radiation electric field is \mathbf{E}_\mathrm{rad} \approx \frac{\mu_0 \omega^2 p}{4\pi r c} \sin\theta \, \hat{\theta} \, \exp[i(kr - \omega t)], with the corresponding magnetic field \mathbf{B}_\mathrm{rad} = \mathbf{E}_\mathrm{rad} / c in the transverse direction. Higher-order terms include the magnetic dipole (l=1, transverse electric mode) and electric quadrupole (l=2), which contribute when the dipole vanishes, such as in parity-symmetric systems. Selection rules from parity conservation dictate that electric $2^l-pole transitions require a parity change (odd l), while magnetic transitions preserve parity (even l); these rules explain forbidden transitions in atomic spectra. The power radiated by an electric dipole provides a key measure of the expansion's lowest term. For a non-relativistic harmonic dipole, the time-averaged power is given by the Larmor formula adapted to multipoles: P = \frac{\mu_0 \omega^4 p^2}{12 \pi c}, where p is the dipole amplitude; higher multipoles radiate less efficiently, scaling as $1/(2l+1)! in the coefficients. This formalism was systematically developed in Jackson's Classical Electrodynamics (first edition, 1962), which formalized the vector spherical harmonic expansion for EM fields and remains foundational for radiation problems. In atomic physics, the multipole expansion underpins the analysis of radiative transitions, where dipole terms dominate allowed decays (e.g., \Delta J = 0, \pm 1) and higher multipoles govern forbidden lines observable in spectra. Modern applications extend this to engineered sources, such as phased array antennas developed in the late 1950s, which allow control of multipole content to shape radiation patterns for radar and communications.

References

  1. [1]
    [PDF] Chapter 13 Maxwell's Equations and Electromagnetic Waves - MIT
    To see how magnetic fields can be created by a time-varying electric field, consider a capacitor which is being charged. During the charging process, ...
  2. [2]
    [PDF] 18 The Electromagnetic Wave Equation - DigitalCommons@USU
    The Electromagnetic Wave Equation. Let us now see how the Maxwell equations (17.2)–(17.5) predict the existence of elec- tromagnetic waves. For simplicity ...
  3. [3]
    Maxwell's equations and light
    The speed c of an electromagnetic wave is determined by the constants of electricity and magnetism that you know so well: c = 1/(e0m0)1/2 = 2.998 X 108m/s.
  4. [4]
    [PDF] Maxwell's Equations
    Mar 15, 2013 · Recall, for a traveling EM wave in vacuum the electric and magnetic field amplitudes are related by cEB. = . Equation (8) thus shows that equal ...
  5. [5]
    [PDF] Classical Electromagnetism - Richard Fitzpatrick
    ... CLASSICAL ELECTROMAGNETISM. Page 7. Maxwell's Equations ... This configuration of electric and magnetic fields is characteristic of an electromagnetic wave.
  6. [6]
    [PDF] Classical Electrodynamics - Duke Physics
    (the 4–vector inhomogeneous electromagnetic wave equation constructed from the 4–scalar D'Lambertian wave operator – the set of four wave equations for φ.
  7. [7]
    VIII. A dynamical theory of the electromagnetic field - Journals
    The theory assumes bodies have states that act at a distance, and the force depends on their relative position and electric/magnetic condition.
  8. [8]
    André-Marie Ampère - Magnet Academy - National MagLab
    Ampère became one of the first people to measure, rather than simply detect, electric currents by using a device of his own invention, quantitatively described ...
  9. [9]
    1830 - 1839 - Magnet Academy - National MagLab
    1830 - 1839 · General theory of electricity · Michael Faraday develops a general theory of electricity by appropriating his model of magnetic lines of induction.
  10. [10]
    The birth of the electric machines: a commentary on Faraday (1832 ...
    Michael Faraday's 1832 paper on electromagnetic induction sits proudly among these works and in a sense can be regarded as having an almost immediate effect ...
  11. [11]
    [PDF] The Origins of the Field Concept in Physics
    Much depends on whether a theory consistent with the notion of action at a distance ought qualify as a ''field'' theory. Roger Boscovich and Immanuel Kant ...
  12. [12]
    James Clerk Maxwell - Magnet Academy - National MagLab
    Maxwell's equations first appeared in 1864 in a paper entitled “A Dynamical Theory of the Electromagnetic Field,” but were more completely addressed in his ...
  13. [13]
    Oliver Heaviside's electromagnetic theory - PMC - NIH
    He reformulated James Clerk Maxwell's field equations in terms of the electric and magnetic forces ('murdering' Maxwell's potentials along the way, due to ...Missing: 1885 | Show results with:1885
  14. [14]
    [PDF] A Brief History of The Development of Classical Electrodynamics
    1884: John Poynting shows that Maxwell's equations predict that energy flows through empty space with the energy flux given by . He also investigates energy ...
  15. [15]
    Heinrich Hertz - Magnet Academy - National MagLab
    Amazingly, the waves were moving at the speed of light. Thus, it appeared to Hertz that he had discovered a previously unknown form of electromagnetic radiation ...
  16. [16]
    Discovering the Electromagnetic Spectrum - Imagine the Universe!
    In 1887 Heinrich Hertz demonstrated the existence of the waves predicted by Maxwell by producing radio waves in his laboratory. It took a bit longer for ...Missing: confirmation speed
  17. [17]
    [PDF] LECTURE NOTES 5 - High Energy Physics
    ELECTROMAGNETIC WAVES IN VACUUM. THE WAVE EQUATION(S) FOR E.. AND B.. In ... Macroscopic EM plane waves propagating in free space are purely transverse waves, ...
  18. [18]
    16.3 Energy Carried by Electromagnetic Waves - UCF Pressbooks
    What is the physical significance of the Poynting vector? Show Solution. It has the magnitude of the energy flux and points in the direction of wave propagation ...
  19. [19]
    Einstein's Pathway to Special Relativity - University of Pittsburgh
    Light waves just are electromagnetic waves. This identification of light with electromagnetic waves was one of the crowning achievements of Maxwell's theory.
  20. [20]
    [PDF] Quantization of the Free Electromagnetic Field: Photons and Operators
    The main ideas and equations for quantized free electromagnetic fields are developed and summarized here, based on the quantization procedure for coordinates ( ...
  21. [21]
    McCormick's Allen Taflove: 40 Years Solving Maxwell's Equations
    In a 1980 paper in IEEE Transactions on Electromagnetic Compatibility resulting from this work, Taflove coined “finite difference time domain” and its FDTD ...
  22. [22]
    [PDF] Notes on gauge theory (S. Naculich, July 2024) 1 Electromagnetism
    Jun 2, 2025 · One may use gauge invariance to enforce the Lorenz gauge condition. ∂µAµ = 0. (1.16). Show that Maxwell equations in free space (jµ = 0) become.
  23. [23]
    [PDF] Space and Time - UCSD Math
    This volume includes Hermann Minkowski's three papers on relativity: The Relativity. Principle, The Fundamental Equations for Electromagnetic Processes in ...
  24. [24]
    The covariant formulation of Maxwell's equations expressed in a ...
    Jan 1, 2009 · The covariant formulation of Maxwell's equations can be expressed in a form independent of the usual systems of units by introducing the constants alpha, beta ...
  25. [25]
    Covariant Formulation of Electrodynamics
    We have developed a beautiful, geometric system for describing the coordinates in terms of which electrodynamics must be formulated for the speed of light to be ...
  26. [26]
    [PDF] 6. Quantum Electrodynamics - DAMTP
    The D00 piece of the propagator doesn't look a whole lot different from the transverse photon propagator. But wouldn't it be nice if they were both part of ...
  27. [27]
    [PDF] Einstein's Investigations of Galilean Covariant Electrodynamics Prior ...
    Jul 30, 2004 · Einstein learned from the magnet and conductor thought experiment how to use field transformation laws to extend the covariance of Maxwell's ...
  28. [28]
    [PDF] Noether's Theorems and Gauge Symmetries - arXiv
    Consideration of the Noether variational problem for any theory whose action is invariant under global and/or local gauge transformations leads to three ...
  29. [29]
    [1608.00659] An axiomatic approach to Maxwell's equations - arXiv
    Aug 2, 2016 · The theorem is also formulated in the Minkowski space-time and applied to obtain the covariant form of Maxwell's equations with magnetic ...
  30. [30]
    Maxwell equations in curved spacetime | The European Physical ...
    Oct 27, 2023 · In curved spacetime, Maxwell's equations can be expressed in forms valid in Minkowski background, with the effect of the metric (gravity) ...
  31. [31]
    [PDF] GRAVITATION F10 Lecture 12 1. Maxwell's Equations in Curved ...
    This leads to a wave equation with source for the electromagnetic potential. ∂µ∂µAν − ∂ν [∂µAµ] = jν. It is common to impose the condition ∂µAµ = 0,(the ...
  32. [32]
    [PDF] Electromagnetic potentials in curved spacetimes - arXiv
    Sep 23, 2021 · Starting from Maxwell's equations and adopting the Lorenz gauge, we arrive at a pair of wave- like formulae for the vector and the scalar ...
  33. [33]
    Electromagnetic potentials in curved spacetimes - IOPscience
    Starting from Maxwell's equations and adopting the Lorenz gauge, we arrive at a pair of wave-like formulae for the vector and the scalar potentials of the ...<|control11|><|separator|>
  34. [34]
    [PDF] Die Feldgleichungen der Gravitation.
    Von A. EINSTEIN. In zwei vor kurzem erschienenen Mitteilungen' habe ich gezeigt, wie man zu Feldgleichungen der Gravitation gelangen kann, die dem ...
  35. [35]
  36. [36]
    Fundamentals of numerical relativity for gravitational wave sources
    Jul 27, 2018 · To model such systems, we need to perform “multi-physics” simulations. Modeling electromagnetic fields in GRHD can be accomplished by coupling ...
  37. [37]
    [PDF] 6. Electromagnetic Radiation - DAMTP
    6.1.2 Green's Function for the Wave Equation. The expression for the retarded potential (6.7) is important. In this section, we provide a slightly different ...
  38. [38]
    Retarded potentials - Richard Fitzpatrick
    ... equations are easily found using the Green's function for Poisson's equation (480): ... After all, electromagnetic waves travel at the speed of light, and no wave ...
  39. [39]
    39 Solving Maxwell's Equations: Green's Functions, Jefimenko
    ... (R,τ)=1Rδ(τ∓Rc). The choice G(+)(R,τ)=1Rδ(τ−Rc) is called the retarded Green's function: the scalar potential equation ∇2φ−1c2∂2φ∂t2= ...
  40. [40]
    [PDF] Direct derivation of Li nard–Wiechert potentials, Maxwell's ... - arXiv
    Dec 18, 2020 · The first hint that Maxwell's equations could be derived from Coulomb's law and. Lorentz transformation can be found in Einstein's original 1905 ...
  41. [41]
    [PDF] Maxwell's Objection to Lorenz' Retarded Potentials - Kirk T. McDonald
    An early application of the retarded potentials in electrodynamics was made in 1883 by. FitzGerald [25, 26], who apparently reinvented them in a derivation of ...
  42. [42]
    20 Solutions of Maxwell's Equations in Free Space
    ... solutions of the wave equation which correspond to plane waves, and that any electromagnetic wave can be described as a superposition of many plane waves.<|control11|><|separator|>
  43. [43]
    [PDF] Chapter 2 The Wave Equation
    Fig. 2.1 illustrates the characteristic features of plane waves. The corresponding magnetic field is readily found by using Maxwell's equation ∇ × E = iωµ0 H. ...
  44. [44]
    Plane Wave Solutions - Ocean Optics Web Book
    Oct 13, 2021 · P is proportional to · E and is commonly written P = 𝜖 o χ E , where χ is the electric susceptibility (i.e., χ measures how susceptible the ...
  45. [45]
    [PDF] SPHERICAL WAVES - UT Physics
    Waves like these are called divergent spherical waves because their wave-fronts are spheres spreading out from the center as r = vphaset + const.
  46. [46]
    [PDF] Spherical Wave Reflection and Transmission - Open Research Online
    represents an outgoing spherical wave when exp(-iwt) is used for the time ... This condition is called the "radiation condition". It specified that no.
  47. [47]
    How laser light is different from ordinary light - Gentec-EO
    Sep 10, 2021 · Unlike the light emitted by common sources, such as light bulbs, laser light is mostly monochromatic, ie only one specific wavelength (color).
  48. [48]
    characteristic impedance of vacuum - CODATA Value
    Click symbol for equation. characteristic impedance of vacuum $Z_0$. Numerical value, 376.730 313 412 OmegaRed. Standard uncertainty, 0.000 000 059 Omega.
  49. [49]
    [PDF] 1 Electromagnetic Waves at Interface –Fresnel Equations
    1.5 Use Boundary Conditions to: Solve Maxwell's equations for incident plane wave. – result depends on incident angle 0 and on vector direction of electric.Missing: coefficients value
  50. [50]
    [PDF] Femtosecond Pulse Shaping for Synthesis, Processing, and Time-to ...
    monic generation, ultrafast optics. I. INTRODUCTION. REVOLUTIONARY breakthroughs have occurred in the field of ultrafast laser technology during the 1990's.
  51. [51]
    [PDF] Ultrafast Optics - MIT OpenCourseWare
    TR: pulse repetition time. W : pulse energy. Pave = W/TR : average power. τFWHM is the Full Width at Half Maximum of the intensity envelope of the.
  52. [52]
    [PDF] The method of Fourier transforms applied to electromagnetic knots
    Sep 13, 2018 · Abstract. The Fourier transform method can be applied to obtain electromagnetic knots, which are solutions of Maxwell equations in vacuum ...
  53. [53]
    [PDF] arXiv:physics/0309079v3 [physics.optics] 23 Aug 2020
    Aug 23, 2020 · ... Fourier transform of the solution of the wave equation. (8). It can be shown that the function of the general type ψ (k,ω)=Ξ(χ, β)δ kz − β ...<|separator|>
  54. [54]
    (PDF) D'Alembert and the Wave Equation - ResearchGate
    The solution of the vibrating string problem is one of d'Alembert's fundamental contributions to mathematics and physics.<|separator|>
  55. [55]
    [PDF] 7 Wave Equation in Higher Dimensions
    [φ(y) + ∇φ(y) · (y − x) + tψ(y)]dS(y). This is known as Kirchoff's formula for the solution of the initial value problem for the wave equation in R3. ut(x,0) = ...
  56. [56]
    [PDF] Fundamentals - Diels Research Group
    Femtosecond light pulses are electromagnetic wave packets and as such are fully described by the time and space dependent electric field.
  57. [57]
    [PDF] Fundamentals - Elsevier
    For Gaussian pulses, τp ωp = t2 2 = 1/2 is true, while for the products of the FWHM of the intensity and spectral intensity. cB = τp vp = 0.441. In fact, the ...
  58. [58]
    A tutorial on THz pulse analysis: accurate retrieval of pulse arrival ...
    May 7, 2025 · Parseval's theorem can be used to verify numerically whether the amplitude spectrum of the discrete Fourier transform is scaled correctly ...
  59. [59]
    [PDF] Fourier's Heat Equation and the Birth of Fourier Series
    Feb 7, 2022 · The goal of this project is to give the reader an insight into the techniques Fourier employed therein, as they have become the basis of modern ...Missing: wave spectral
  60. [60]
    The Use of Fourier Transform in Heat Equation and Wave Equation
    The passage mainly discussed about using Fourier transform to solve standing waves equation heat equation in steady state and in the time dependent state.
  61. [61]
    A friendly guide to wavelets. Reprint of the 1994 original
    In this chapter we propose an application of electromagnetic wavelets to radar signal analysis and electromagnetic scattering. The goal in radar, as well as ...
  62. [62]
    [PDF] wavelet analysis of randomly reflected electromagnetic waves of ...
    using FDTD method in 1990's, and we found that the time response of subsurface radar has some reflected waves from individual unwanted scatterers in the ...Missing: stationary | Show results with:stationary
  63. [63]
    Multipole Expansion of Vector Electromagnetic Fields - KSU Physics
    Electromagnetic Radiation and Multipoles (Jackson, Ch. 9) Multipole Expansion of Vector Electromagnetic Fields -- CGS.
  64. [64]
    [PDF] The development of phased-array radar technology
    Lincoln Laboratory has been involved in the development of phased-array radar technology since the late 1950s. Radar research activities have included.