Classical electromagnetism is the fundamental theory in physics that describes the behavior of electric and magnetic fields, their interactions with charged particles and currents, and the propagation of electromagnetic waves, all unified within a classical framework without quantum effects.[1] It encompasses the forces between stationary and moving charges, the generation of fields by charges and currents, and the synthesis of electricity, magnetism, and optics into a single coherent system.[2]The theory originated in the 19th century through the work of scientists such as Charles-Augustin de Coulomb, who formulated the inverse-square law for electric forces; Hans Christian Ørsted and André-Marie Ampère, who linked electric currents to magnetic effects; and Michael Faraday, who discovered electromagnetic induction.[1] James Clerk Maxwell synthesized these empirical laws into a set of four partial differential equations in the 1860s, predicting that changing electric fields produce magnetic fields and vice versa, leading to the concept of electromagnetic waves traveling at the speed of light.[2] This unification demonstrated that light itself is an electromagnetic phenomenon, resolving long-standing puzzles in optics and paving the way for modern technologies.[1]At its core, classical electromagnetism is governed by Maxwell's equations, which relate the electric field E and magnetic field B to charge density ρ and current density j:These equations, combined with the Lorentz force law F = q(E + v × B) for the force on a charge q moving with velocity v, fully describe electromagnetic interactions in classical settings.[2]The theory's predictions, including electromagnetic radiation from accelerating charges and the invariance of the speed of light c ≈ 3 × 10⁸ m/s in vacuum, were later confirmed by experiments and integrated with special relativity by Albert Einstein in 1905, ensuring consistency across inertial frames.[1] Classical electromagnetism underpins applications from electric power generation to radio communication and remains essential for understanding macroscopic phenomena, though it transitions to quantum electrodynamics at atomic scales.[2]
Historical Development
Early Observations and Experiments
The earliest recorded observations of electrical phenomena date back to ancient Greece around 600 BCE, when the philosopher Thales of Miletus noted that amber, after being rubbed with fur or cloth, could attract lightweight objects such as feathers or straw.[3] This effect, now understood as static electricity, was a qualitative discovery without explanation, marking the first documented instance of frictional electrification.[4]Magnetic attractions were similarly observed in antiquity, with lodestones—naturally magnetized pieces of magnetite—known to draw iron objects, as reported by early Chinese texts from the 4th century BCE and by Greek philosophers including Thales around 600 BCE.[5] These lodestones were used practically in China for early compass-like devices by the 2nd century BCE, while Greeks described their properties in philosophical treatises, distinguishing them from electrical effects but without quantitative analysis.[3]In the 17th century, English physician William Gilbert advanced these ideas in his 1600 treatise De Magnete, systematically differentiating electric forces—produced by rubbing insulating materials like amber—from magnetic forces inherent to lodestones and the Earth itself, which he demonstrated acts as a large magnet.[6] Gilbert's experiments involved versorium devices to detect attractions, establishing electricity as a distinct phenomenon and laying groundwork for controlled studies.[7]Around 1663, German engineer Otto von Guericke invented the first electrostatic generator, a rotating sulfur sphere rubbed by hand to produce stronger electrical effects, including sparks and attractions observable over distances, which facilitated more reliable demonstrations of frictional electricity.[8] This device marked a shift toward instrumental experimentation, though still qualitative in nature.By the early 18th century, French chemist Charles du Fay in 1733 identified two distinct types of electricity through experiments with rubbed glass and resin: "vitreous" electricity from glass-like materials and "resinous" from amber-like substances, which attracted each other but repelled like kinds, leading to the two-fluid theory of electricity with vitreous and resinous electric fluids, and laying the groundwork for the concept of opposing charges (later termed positive and negative).[9]American polymath Benjamin Franklin contributed in 1752 with his kite experiment during a thunderstorm, using a silk kite with a key to collect atmospheric electricity into a Leyden jar, confirming lightning as an electrical discharge and supporting his single-fluid theory where excess or deficiency of a subtle electrical fluid caused attractions and repulsions.[10] Franklin's work, reported in letters to European scientists, popularized safe electrical investigations and influenced charge nomenclature.[11]English experimenter Henry Cavendish in 1771 conducted precise measurements of electrical forces using a torsion balance-like apparatus with charged spheres, obtaining results that hinted at an inverse-square dependence of the force on distance—though his detailed manuscript remained unpublished until 1879.[12] These experiments provided early quantitative data on electrostatic repulsion and attraction, predating formal laws like Coulomb's but establishing empirical foundations for them.[13]In 1785, French physicist Charles-Augustin de Coulomb used a torsion balance to quantitatively measure the force between charged objects, establishing that the electrostatic force follows an inverse-square law proportional to the product of charges and inversely proportional to the square of distance. He extended similar findings to magnetic forces between poles.[14]
19th-Century Unification
In 1800, Italian physicist Alessandro Volta invented the voltaic pile, the first electrochemical battery consisting of stacked disks of zinc and copper separated by brine-soaked cardboard, producing a steady electric current. This breakthrough enabled reliable experiments with continuous currents, setting the stage for discoveries in electromagnetism.[15]In 1820, Danish physicist Hans Christian Ørsted discovered the link between electricity and magnetism during a lecturedemonstration when he observed that a compass needle deflected perpendicular to a wire carrying electric current, establishing that electric currents produce magnetic fields.[16] This serendipitous finding, published in July 1820, overturned the prevailing view that electricity and magnetism were unrelated phenomena and sparked intense research across Europe.[17]Building on Ørsted's result, French mathematician and physicist André-Marie Ampère rapidly developed a quantitative theory of electrodynamics between 1820 and 1827, formulating the force law governing interactions between current-carrying wires, which predicts attraction between parallel currents in the same direction and repulsion otherwise.[18]Ampère's work, detailed in his 1827 memoir Mémoire sur la théorie mathématique des phénomènes électro-dynamiques uniquement déduite de l'expérience, introduced the concept of the ampere as a unit of current and laid the foundation for understanding magnetic forces as arising from electric currents rather than hypothetical magnetic fluids.[19]In 1831, British chemist and physicist Michael Faraday achieved a reciprocal connection by discovering electromagnetic induction, demonstrating that a changing magnetic field induces an electric current in a nearby conductor, as shown in his experiments with coils and moving magnets reported in his paper "On Electromagnetic Induction" presented to the Royal Society.[20] Faraday also introduced the concept of lines of force to visualize magnetic and electric fields as continuous entities permeating space, and between 1832 and 1834, he established the laws of electrolysis, quantifying the relationship between electric current and chemical decomposition at electrodes.[21] Independently in 1832, American physicist Joseph Henry observed electromagnetic induction while experimenting with electromagnets at the Albany Academy, producing induced currents over longer distances than Faraday's initial setups and publishing his findings in the American Journal of Science.[22] In 1834, Russian physicist Heinrich Lenz clarified the direction of these induced currents, stating that the induced current flows in a direction opposing the change in magnetic flux that produced it, a principle essential for energy conservation in inductive processes.[23]Further advances in applying these principles to practical systems came in 1857 when German physicist Gustav Kirchhoff derived the telegraph equation, modeling the propagation of electric signals along transmission lines by incorporating inductance and capacitance, which predicted signal speeds comparable to that of light and enabled reliable long-distance communication.[24] Culminating these developments, Scottish physicist James Clerk Maxwell synthesized the disparate experimental results into a unified theoretical framework from 1861 to 1865, most notably in his 1865 paper "A Dynamical Theory of the Electromagnetic Field," where he introduced the displacement current term to account for changing electric fields in regions without conduction currents, resolving inconsistencies in Ampère's law and predicting electromagnetic waves.[25] This synthesis connected electricity, magnetism, and optics, demonstrating their interdependence through field equations that govern dynamic interactions.[26]
Fundamental Concepts
Electric Charge and Coulomb's Law
Electric charge is a fundamental scalar property of matter that underlies electromagnetic interactions, manifesting as either positive or negative types that attract or repel accordingly.[27] In classical electromagnetism, charge is treated as a continuous quantity, though modern quantum views reveal it to be quantized in discrete units, such as multiples of the elementary charge e = 1.602176634 \times 10^{-19} C (exact in SI since the 2019 redefinition).[28][29] Charge is conserved in isolated systems, meaning the total amount remains constant regardless of interactions or transformations.[28]The force between two stationary point charges q_1 and q_2 separated by a distance r is described by Coulomb's law, which states that the magnitude of the electrostatic force is proportional to the product of the charges and inversely proportional to the square of the distance between them:F = k_e \frac{|q_1 q_2|}{r^2},where k_e = \frac{1}{4\pi \epsilon_0} is Coulomb's constant and \epsilon_0 is the vacuum permittivity.[30] The force is directed along the line joining the charges, repulsive for like signs and attractive for opposite signs, expressed vectorially as\vec{F} = k_e \frac{q_1 q_2}{r^2} \hat{r},with \hat{r} the unit vector from one charge to the other.[30] This law was experimentally established in 1785 by Charles-Augustin de Coulomb using a torsion balance, a device consisting of a suspended rod with charged spheres at each end that twisted under electrostatic repulsion, allowing precise measurement of the force as a function of distance.[31] Earlier qualitative experiments, such as those by Henry Cavendish around 1771, had hinted at the inverse-square dependence but lacked the quantitative precision of Coulomb's apparatus.[32]For systems with multiple charges, the principle of superposition holds: the total force on any charge is the vector sum of the forces exerted by each individual charge, unaffected by the presence of others, as electrostatic interactions are linear.[33] This follows directly from the pairwise nature of Coulomb's law and enables the analysis of complex charge distributions by summing contributions.[34]Conceptually, the electric field \vec{E} arises from charges as the electrostatic force per unit positive test charge at a point in space:\vec{E} = \frac{\vec{F}}{q},where q is a small test charge that does not disturb the source configuration; this provides a framework for describing the influence of charges without direct reference to the force on specific objects.[35]In the International System of Units (SI), electric charge is measured in coulombs (C), defined as the amount of charge transported by a current of one ampere in one second, equivalent to exactly $6.241509074 \times 10^{18} elementary charges ($1/e).[36] The SI unit of electric current, the ampere (A), is defined by taking the elementary electrical charge e to be exactly $1.602176634 \times 10^{-19} coulombs (C), where $1 C = $1 A⋅s and the second is defined in terms of the caesium hyperfine transition frequency.[36] These definitions reflect the 2019 revision of the SI, which fixed fundamental constants like e and derived others like \epsilon_0 from measurements. The vacuum permittivity \epsilon_0 has the approximate value $8.8541878188 \times 10^{-12} F/m (relative uncertainty $1.6 \times 10^{-10}, per 2022 CODATA).[37]
Magnetic Poles, Currents, and Biot-Savart Law
In classical electromagnetism, magnetic phenomena are described in terms of poles that exhibit north-south attraction or repulsion, similar to electric charges, but isolated magnetic monopoles do not exist; magnetic fields always arise from dipoles or closed loops, as evidenced by the absence of sources or sinks in the magnetic field equations.[1] This contrasts with electric fields, where monopoles (charges) are fundamental. Permanent magnets, for instance, consist of aligned atomic dipoles, and cutting a bar magnet produces two new dipoles rather than isolated poles.[1]The source of magnetism was identified in 1820 when Hans Christian Ørsted discovered that steady electric currents produce magnetic fields, demonstrating that a current-carrying wire deflects a nearby compass needle, revealing the intimate connection between electricity and magnetism.[16] Electric currents, arising from the motion of electric charges, generate these fields, with the direction of the magnetic effect following a right-hand rule around the wire.[16]Building on Ørsted's finding, Jean-Baptiste Biot and Félix Savart formulated the Biot-Savart law in 1820, which quantifies the magnetic field \vec{B} produced by a steady current element. The contribution from an infinitesimal current element I d\vec{l} at a point is given byd\vec{B} = \frac{\mu_0}{4\pi} \frac{I d\vec{l} \times \hat{r}}{r^2},where \mu_0 is the permeability of free space, r is the distance from the element to the observation point, and \hat{r} is the unit vector in that direction.[38] To find the total field from a finite wire, this expression is integrated along the current path, yielding, for example, the circumferential field around a straight infinite wire: B = \frac{\mu_0 I}{2\pi r}.[38]The Biot-Savart law implies forces between currents, as elaborated by André-Marie Ampère shortly after 1820 through experiments showing attraction between parallel currents in the same direction and repulsion otherwise. The force on a length \ell of one wire due to a parallel infinite wire carrying current I_2 is\vec{F} = \frac{\mu_0 I_1 I_2}{2\pi d} \ell \hat{d},where d is the separation and \hat{d} the unit vector from the first to the second wire.[39] Consequently, two infinitely long parallel wires 1 m apart carrying 1 A each experience a force of approximately $2 \times 10^{-7} N/m.[36]The permeability of free space, \mu_0 \approx 4\pi \times 10^{-7} H/m (or $1.256637061 \times 10^{-6} N/A², measured with relative uncertainty $1.6 \times 10^{-10} per 2022 CODATA), scales the strength of these magnetic effects in vacuum and relates to the speed of light c via c = 1 / \sqrt{\mu_0 \epsilon_0}, where \epsilon_0 is the permittivity of free space, foreshadowing electromagnetic wave propagation.[40] In classical electromagnetism calculations, the value $4\pi \times 10^{-7} H/m is commonly used. The magnetic field \vec{B} is measured in teslas (T), where 1 T equals 1 Wb/m² or the field exerting 1 N on a 1 m wire carrying 1 A perpendicularly.[41]
Field Descriptions
Electric Field
The electric field \vec{E} at a point in space is defined as the force \vec{F} per unit positive test charge q placed at that point, \vec{E} = \vec{F}/q, representing the influence of surrounding electric charges on a test charge without significant disturbance to the source configuration.[42] This concept emerged from Michael Faraday's experimental investigations into electrical phenomena, where he conceptualized forces acting through space via continuous distributions rather than direct action at a distance.[43]The electric field arises fundamentally from electric charge distributions, with the field due to a single point charge q at distance r given by Coulomb's law as \vec{E} = \frac{1}{4\pi\epsilon_0} \frac{q}{r^2} \hat{r}, where \epsilon_0 is the vacuum permittivity and \hat{r} is the unit vector from the charge to the point.[44] For a continuous charge distribution with density \rho(\vec{r}'), the field at position \vec{r} is obtained by integrating over the volume via the principle of superposition, yielding the formal expression\vec{E}(\vec{r}) = \frac{1}{4\pi\epsilon_0} \int \frac{\rho(\vec{r}') (\vec{r} - \vec{r}')}{|\vec{r} - \vec{r}'|^3} \, dV',which sums the contributions from all infinitesimal charge elements dq = \rho(\vec{r}') dV'.[45] This integral formulation, derived from Coulomb's empirical law and the linearity of electrostatic interactions, allows computation of the field from arbitrary charge arrangements.[42]A key property for calculating the electric field from symmetric charge distributions is Gauss's law, which states that the total electric flux through any closed surface is equal to the enclosed charge divided by \epsilon_0,\oint \vec{E} \cdot d\vec{A} = \frac{Q_{\text{encl}}}{\epsilon_0}.Formulated by Carl Friedrich Gauss in the context of attractive forces, this theorem applies to electrostatics and simplifies field determination when symmetry (spherical, cylindrical, or planar) makes \vec{E} constant in magnitude and normal to the surface.[42] For example, for an infinite line charge with linear density \lambda, a cylindrical Gaussian surface yields E = \frac{\lambda}{2\pi\epsilon_0 r}; for an infinite plane with surface density \sigma, a pillbox surface gives E = \frac{\sigma}{2\epsilon_0}, independent of distance.[45] These symmetry-based results highlight how Gauss's law efficiently captures the field's divergence from sources without full integration.[42]In electrostatics, the electric field is conservative, meaning the work done by the field on a charge moving between two points is path-independent, implying \nabla \times \vec{E} = 0.[45] Consequently, \vec{E} can be expressed as the negative gradient of a scalar electric potential V, \vec{E} = -\nabla V, where V is defined up to a constant and relates to the field's directional decrease in potential energy per unit charge.[42] This irrotational property stems from the absence of time-varying magnetic fields, ensuring closed loops yield zero line integral of \vec{E}.[45]Electric field lines, introduced by Faraday as "lines of force," provide a visual representation of the field's direction and relative strength, originating from positive charges and terminating on negative ones, with density proportional to field magnitude.[43] The principle of superposition ensures that fields from multiple sources add vectorially, allowing complex configurations to be decomposed into simpler components, such as combining uniform fields or point-charge contributions.[42]In electrostatic equilibrium, the electric field inside a conductor vanishes, \vec{E} = 0, as any internal field would cause charge redistribution until forces balance, confining excess charge to the surface.[45] This follows from Gauss's law applied to a Gaussian surface within the conductor, where zero flux implies zero enclosed charge, and free charges rearrange to cancel internal fields.[42]
Magnetic Field
The magnetic field, denoted by the vector field \vec{B}(\vec{r}), quantifies the magnetic influence produced by electric currents and magnetic materials in classical electromagnetism. Unlike the electric field, which arises from charges, the magnetic field originates solely from moving charges or currents in steady-state conditions. The Biot–Savart law provides the fundamental means to compute \vec{B} from a current distribution, treating each infinitesimal current element as a source. This law was empirically established through experiments by Jean-Baptiste Biot and Félix Savart in 1820, building on Ørsted's discovery of current-induced magnetism.[38][1]For a steady current I along a wire, the magnetic field at position \vec{r} is given by the integral\vec{B}(\vec{r}) = \frac{\mu_0}{4\pi} \int \frac{I \, d\vec{l}' \times (\vec{r} - \vec{r}')}{|\vec{r} - \vec{r}'|^3},where \mu_0 = 4\pi \times 10^{-7} N A^{-2} is the vacuum permeability, and the integral is over the current path.[1] In the SI system, the unit of \vec{B} is the tesla (T), defined as 1 T = 1 kg s^{-2} A^{-1}, reflecting its relation to force on currents.[46] Magnetic field lines, visualized as tangent to \vec{B}, form continuous closed loops that neither begin nor end, a consequence of the experimental absence of isolated magnetic monopoles. This property is mathematically captured by the divergence-free condition \nabla \cdot \vec{B} = 0, one of the cornerstone equations of magnetostatics.[1]Ampère's circuital law offers an alternative, often more efficient, method to determine \vec{B} for symmetric current configurations in static cases. Formulated by André-Marie Ampère in 1820 based on extensive experiments on current interactions, it states that the line integral of \vec{B} around any closed loop equals \mu_0 times the total enclosed current: \oint \vec{B} \cdot d\vec{l} = \mu_0 I_{\rm encl}.[19][1] For a long solenoid with n turns per unit length carrying current I, the law yields a uniform internal field \vec{B} = \mu_0 n I parallel to the axis and zero outside, idealizing infinite length. Similarly, for a toroidal solenoid with mean radius much larger than cross-sectional radius, \vec{B} = \mu_0 N I / (2\pi r) inside, where N is the total turns and r the radial distance, again zero outside. These examples highlight the law's utility for confined fields in devices like electromagnets.[1]In regions devoid of currents, \vec{B} is irrotational (\nabla \times \vec{B} = 0), allowing representation via a magnetic scalar potential \phi_m such that \vec{B} = -\mu_0 \nabla \phi_m, analogous to the electric scalar potential but scaled by \mu_0.[1] More generally, across all space, the magnetic field derives from a vector potential \vec{A} via \vec{B} = \nabla \times \vec{A}, ensuring \nabla \cdot \vec{B} = 0 automatically. The Coulomb gauge, \nabla \cdot \vec{A} = 0, simplifies calculations while preserving Lorentz invariance in static limits, with \vec{A} directly integrable from the current density as in the Biot–Savart form.[1]
Forces in Electromagnetic Fields
Electrostatic and Magnetostatic Forces
In electrostatics, the force experienced by a point charge q placed in an electric field \vec{E} is given by\vec{F} = q \vec{E},where the electric field is defined as the force per unit charge on a test charge.[44] This relation stems from experimental measurements of the repulsive or attractive forces between stationary charges, which follow an inverse-square law.[44] A representative example is the electrostatic force between two uniformly charged conducting spheres, where the field due to one sphere exerts a force on the charge of the other, leading to attraction or repulsion depending on the signs of the charges.[44]For an electric dipole consisting of two equal and opposite charges separated by a small distance, the net force in a uniform electric field is zero, but a torque arises that tends to align the dipole with the field. The torque \vec{\tau} on an electric dipole with moment \vec{p} = q \vec{d} (where \vec{d} is the displacement vector from the negative to the positive charge) is\vec{\tau} = \vec{p} \times \vec{E}.This torque is derived directly from the differential forces acting on the individual charges of the dipole.[44]The total electrostatic potential energy stored in the electric field throughout a volume is expressed asU = \frac{1}{2} \int \epsilon_0 E^2 \, dV,where \epsilon_0 is the vacuum permittivity; this formulation represents the work done to assemble the charge distribution producing the field.[47]In magnetostatics, the force on a current-carrying conductor in a magnetic field \vec{B} is given by\vec{F} = I \int d\vec{l} \times \vec{B},where I is the current and the integral is along the length of the wire.[48] This arises from the interaction between the moving charges in the current and the magnetic field. A key example is the attractive or repulsive force between two long, straight, parallel wires carrying currents I_1 and I_2 separated by distance r, where the force per unit length is F/l = \mu_0 I_1 I_2 / (2\pi r) (with \mu_0 the vacuum permeability), leading to attraction for like-directed currents and repulsion for opposite directions.[48]For a magnetic dipole with moment \vec{m}, such as a current loop, the torque in a uniformmagnetic field is\vec{\tau} = \vec{m} \times \vec{B},which acts to align the dipole with the field, analogous to the electric case.[48]The magnetic energy stored in the field isU = \frac{1}{2} \int \frac{B^2}{\mu_0} \, dV,representing the energy associated with the magnetic field configuration.[47]
Lorentz Force Law
The Lorentz force law describes the force experienced by a charged particle moving through electromagnetic fields, combining the effects of both electric and magnetic fields on the particle's motion. In classical electromagnetism, this force \vec{F} acting on a particle of charge q with velocity \vec{v} in an electric field \vec{E} and magnetic field \vec{B} is given by\vec{F} = q \left( \vec{E} + \vec{v} \times \vec{B} \right).This expression, known as the Lorentz force, was fully derived by Hendrik Lorentz in his 1895 treatise on the electromagnetic theory of moving bodies, building on earlier insights from Maxwell's work on the forces between moving charges. The law captures the total electromagnetic force in the non-relativistic regime, where the electric term q\vec{E} arises from the field's direct interaction with the charge, while the magnetic term q(\vec{v} \times \vec{B}) depends on the particle's velocity and is always perpendicular to both \vec{v} and \vec{B}, thus doing no work on the particle.The Lorentz force reduces to simpler cases under specific conditions. When the magnetic field \vec{B} = 0, it simplifies to the electrostatic force \vec{F} = q\vec{E}, as derived from Coulomb's law for stationary charges. Similarly, if the electric field \vec{E} = 0 or the velocity \vec{v} = 0, it becomes the magnetostatic force \vec{F} = q(\vec{v} \times \vec{B}), consistent with the force on moving charges in steady currents observed in early experiments. These reductions highlight how the full law unifies previously separate electrostatic and magnetostatic descriptions into a velocity-dependent framework applicable to dynamic situations.A key application of the Lorentz force is in the motion of charged particles in uniform magnetic fields, leading to cyclotron motion. For a particle with perpendicular velocity component v_\perp in a uniform \vec{B} field (with \vec{E} = 0), the magnetic force provides the centripetal acceleration for circular motion, yielding a radius r = \frac{m v_\perp}{q B}, where m is the particle's mass. This helical path, with cyclotron frequency \omega_c = \frac{q B}{m}, forms the basis for particle accelerators like the cyclotron, invented by Ernest Lawrence in 1930, and is derived by balancing the Lorentz force against the centripetal force \frac{m v_\perp^2}{r} = q v_\perp B./21%3A_Magnetism/21.4%3A_Motion_of_a_Charged_Particle_in_a_Magnetic_Field)Another important example is the Hall effect, discovered by Edwin Hall in 1879, where the Lorentz force deflects charge carriers in a current-carrying conductor placed in a perpendicular magnetic field. In a conductor with current density \vec{J} along the x-direction and \vec{B} along z, the magnetic force q(\vec{v} \times \vec{B}) (with \vec{v} as drift velocity) pushes positive carriers toward one side, creating a transverse electric field E_y that balances the force in steady state, resulting in the Hall voltage V_H = \frac{I B}{n e t}, where I is current, n is carrier density, e is charge, and t is thickness. This effect, explained through the Lorentz force, enables measurement of carrier type and density in materials.For macroscopic currents, the Lorentz force on a current-carrying wire is obtained by integrating the microscopic forces over the charge distribution. For a wire segment of length \vec{l} carrying current I in a uniform \vec{B}, the total force is \vec{F} = I \vec{l} \times \vec{B}, derived by summing d\vec{F} = dq (\vec{v} \times \vec{B}) along the wire, where dq = n e A v_d dl and I = n e A v_d, with A as cross-sectional area and v_d as drift speed. This integral form, equivalent to \int \vec{J} \times \vec{B} \, dV for volume currents, explains the deflection of wires in magnetic fields observed in Ampère's experiments.[49]In the non-relativistic limit, the Lorentz force law exhibits invariance under Galilean transformations, ensuring that the measured force remains consistent across inertial frames moving at constant velocity relative to each other. For two frames S and S' with relative velocity \vec{u} along x, the transformed fields and velocities adjust such that \vec{F}' = q (\vec{E}' + \vec{v}' \times \vec{B}') yields the same \vec{F} components parallel to \vec{u}, while perpendicular components transform appropriately to preserve Newton's second law \vec{F} = m \vec{a}. This Galilean covariance holds for velocities much less than the speed of light, aligning with classical mechanics before the advent of special relativity.[50]
Electrodynamic Laws
Faraday's Law of Induction
Faraday's law of induction, discovered through a series of experiments in 1831, describes the fundamental relationship between changing magnetic fields and the induction of electric fields, serving as a cornerstone of classical electromagnetism. Michael Faraday observed that moving a magnet near a wire loop or varying the current in a nearby coil produced a transient electric current in the loop, even without direct contact, leading to the recognition that a time-varying magnetic flux through a circuit generates an electromotive force (EMF). This phenomenon, later formalized mathematically, underpins the generation of electricity in generators and the operation of inductive devices.The integral form of Faraday's law states that the line integral of the electric field \vec{E} around a closed loop equals the negative rate of change of the magnetic flux \Phi_B through the surface bounded by that loop:\oint_C \vec{E} \cdot d\vec{l} = -\frac{d\Phi_B}{dt},where \Phi_B = \int_S \vec{B} \cdot d\vec{A} is the magnetic flux, with \vec{B} the magnetic field and d\vec{A} the differential area element. This equation implies that the induced EMF \mathcal{E} in the loop is \mathcal{E} = -\frac{d\Phi_B}{dt}, quantifying the voltage generated by the changing flux. The law applies to both motional EMF, arising from the motion of a conductor in a static magnetic field (where the flux change results from relative motion), and transformer EMF, due to a time-varying magnetic field in a stationary circuit. For instance, in a simple generator, a conducting rod moving perpendicular to a uniform magnetic field B with velocity v over length l induces an EMF of \mathcal{E} = B l v, as the flux through the swept area changes at rate B l v.Lenz's law, formulated by Heinrich Lenz in 1834, provides the physical interpretation of the negative sign in Faraday's law, stating that the induced current creates a magnetic field that opposes the change in flux responsible for it. This conservation principle ensures that energy is not created from nothing; for example, if the flux through a loop is increasing, the induced current flows to produce a field that tries to decrease the flux, requiring work against the opposing force. In a moving conductor scenario, the induced current in the rod experiences a Lorentz force \vec{F} = I \vec{l} \times \vec{B} that opposes the motion, converting mechanical energy into electrical energy. Lenz's law thus dictates the direction of induced currents and is essential for understanding the efficiency limits in inductive systems.The differential form of Faraday's law, derived from the integral version via Stokes' theorem, expresses the local relationship between the electric and magnetic fields:\nabla \times \vec{E} = -\frac{\partial \vec{B}}{\partial t}.This curl equation reveals that a time-varying magnetic field acts as a source for the rotational component of the electric field, enabling non-conservative electric fields even in regions without charges. It highlights the interdependence of electric and magnetic fields in dynamic situations, distinguishing electrodynamics from statics.Faraday's law also introduces the concept of inductance, which quantifies how a circuit's own or another circuit's current produces flux linked to it. Self-inductance L for a coil is defined as L = \frac{\Phi_B}{I}, where \Phi_B is the flux linkage due to current I, leading to an induced EMF of \mathcal{E} = -L \frac{dI}{dt} when the current changes. Mutual inductance M between two circuits measures the flux in one due to current in the other, given by M = \frac{\Phi_{21}}{I_1}, enabling energy transfer in transformers where an alternating current in the primary coil induces voltage in the secondary via \mathcal{E}_2 = -M \frac{dI_1}{dt}. In a basic transformer, a primary coil with N_1 turns around an iron core links nearly all flux to a secondary with N_2 turns, achieving voltage transformation ratios of approximately V_2 / V_1 = N_2 / N_1 for ideal coupling. These inductance concepts are crucial for designing inductors, motors, and power transmission systems, where the opposition to current changes (back-EMF) regulates performance.
Ampère's Circuital Law with Displacement Current
Ampère's circuital law, in its original formulation, states that the line integral of the magnetic field \vec{B} around a closed loop is equal to \mu_0 times the total conduction current I_{\text{encl}} passing through the surface bounded by the loop:\oint \vec{B} \cdot d\vec{l} = \mu_0 I_{\text{encl}}.This relation, derived from experimental observations of magnetic fields produced by steady currents, holds for magnetostatic situations where fields do not vary with time.[51]However, this form leads to inconsistencies when applied to time-varying fields, such as during the charging of a capacitor. In a parallel-plate capacitor connected to a current source, conduction current flows in the wires but appears to abruptly stop between the plates, where no conduction current exists. Without modification, Ampère's law would predict zero circulation of \vec{B} between the plates, contradicting the continuity of magnetic fields and the observed magnetic field generation in such setups.[52]To resolve this, James Clerk Maxwell introduced the concept of displacement current in 1865, adding a term that accounts for the time rate of change of the electric flux through the surface. The displacement current density is given by I_d = \epsilon_0 \frac{d\Phi_E}{dt}, where \Phi_E is the electric flux. The corrected law becomes\oint \vec{B} \cdot d\vec{l} = \mu_0 \left( I_{\text{encl}} + \epsilon_0 \frac{d\Phi_E}{dt} \right).This amendment ensures continuity, as the displacement current between the capacitor plates equals the conduction current in the wires, maintaining a consistent magnetic field circulation.[53]In differential form, the Maxwell-Ampère law is expressed as\nabla \times \vec{B} = \mu_0 \vec{J} + \mu_0 \epsilon_0 \frac{\partial \vec{E}}{\partial t},where \vec{J} is the conduction current density and the second term represents the displacement current density. This form highlights the symmetry with Faraday's law of induction, where changing magnetic fields induce electric fields.[53]The inclusion of displacement current is physically necessary to uphold charge conservation and prevent violations of the continuity equation \nabla \cdot \vec{J} + \frac{\partial \rho}{\partial t} = 0 in time-dependent scenarios. It also provides the key insight for electromagnetic wave propagation, as the coupled time derivatives suggest oscillatory solutions for \vec{E} and \vec{B}.[54]
Maxwell's Equations
Integral Formulation
The integral formulation of Maxwell's equations provides a global description of electromagnetic phenomena, relating fields to their sources through surface and line integrals over closed paths. These equations are derived from experimental laws and are particularly advantageous for problems with high symmetry, such as spherical or cylindrical geometries, where the integrals simplify calculations of flux and circulation. They also facilitate the application of boundary conditions at interfaces between different media, such as in capacitors or waveguides, by considering enclosed charges and currents. Furthermore, the integral forms inherently encode conservation principles, linking field behaviors to the continuity of charge and the absence of magnetic monopoles.[55]The first of these equations is Gauss's law for electricity, which asserts that the total electric flux through any closed surface S is proportional to the total electric charge Q_{\text{encl}} enclosed within the volume bounded by S:\oint_S \vec{E} \cdot d\vec{A} = \frac{Q_{\text{encl}}}{\epsilon_0},where \vec{E} is the electric field, d\vec{A} is the outward-pointing area element, and \epsilon_0 is the vacuum permittivity. This law, originally developed by Carl Friedrich Gauss to describe the flux due to attracting ellipsoids, quantifies how electric charges act as sources or sinks for the electric field lines. Applying the divergence theorem to this equation reveals the local conservation of charge, as the surface integral transforms into a volume integral of the divergence of \vec{E}, equating to the enclosed charge density integrated over the volume, consistent with the continuity equation \nabla \cdot \vec{J} + \frac{\partial \rho}{\partial t} = 0.[55]The second equation, Gauss's law for magnetism, states that the total magnetic flux through any closed surface is zero:\oint_S \vec{B} \cdot d\vec{A} = 0,where \vec{B} is the magnetic field. This reflects the experimental observation that magnetic field lines form continuous loops with no beginning or end, implying the nonexistence of magnetic monopoles or isolated "north" and "south" magnetic charges. Formulated by James Clerk Maxwell as part of his unification of electromagnetic laws, this equation ensures that magnetic fields arise solely from currents and changing electric fields, without independent magnetic sources.[55]Faraday's law of induction constitutes the third equation, expressing that the electromotive force around a closed loop is equal to the negative rate of change of the magnetic flux through the surface bounded by that loop:\oint_C \vec{E} \cdot d\vec{l} = -\frac{d\Phi_B}{dt},where C is the closed contour, d\vec{l} is the line element along C, and \Phi_B = \int_S \vec{B} \cdot d\vec{A} is the magnetic flux. Discovered experimentally by Michael Faraday through demonstrations involving moving magnets and coils, this law captures the induction of electric fields by time-varying magnetic fields, foundational to devices like transformers and generators. In boundary applications, it determines the tangential component of \vec{E} across interfaces, aiding analysis in regions with abrupt field changes.The fourth equation, the Ampère-Maxwell law, relates the circulation of the magnetic field around a closed loop to the enclosed conduction current and the rate of change of electric flux:\oint_C \vec{B} \cdot d\vec{l} = \mu_0 \left( I_{\text{encl}} + \epsilon_0 \frac{d\Phi_E}{dt} \right),where \mu_0 is the vacuum permeability, I_{\text{encl}} is the total current passing through the surface, and \Phi_E = \int_S \vec{E} \cdot d\vec{A} is the electric flux. André-Marie Ampère originally established the relation to conduction currents in 1826, but Maxwell's 1865 addition of the displacement current term \epsilon_0 \frac{d\Phi_E}{dt} resolved inconsistencies with charge conservation in time-varying situations, enabling the prediction of electromagnetic waves.[55] This form is essential for boundary conditions on the tangential magnetic field, particularly in scenarios involving capacitors where no conduction current flows but displacement current does.
Differential Formulation
The differential formulation of classical electromagnetism provides a local description of electromagnetic fields through partial differential equations that apply at every point in space-time, enabling precise analysis of field behavior and interactions with charges and currents. This approach contrasts with the integral formulation by emphasizing point-wise relations rather than global constraints over volumes or surfaces. The equations relate the electric field \vec{E}, magnetic field \vec{B}, charge density \rho, and current density \vec{J}, assuming vacuum or free space conditions. Oliver Heaviside first presented these equations in their modern vectorial differential form in his 1884–1885 series of articles, streamlining Maxwell's original integral expressions into a compact set of four equations.[56]The two divergence equations express the sources of the fields. Gauss's law for the electric field states that the divergence of \vec{E} is proportional to the local charge density:\nabla \cdot \vec{E} = \frac{\rho}{\epsilon_0},where \epsilon_0 is the vacuum permittivity. Gauss's law for magnetism asserts the absence of magnetic monopoles, with the divergence of \vec{B} vanishing everywhere:\nabla \cdot \vec{B} = 0.These relations imply that electric field lines originate from positive charges and terminate on negative ones, while magnetic field lines form closed loops.The curl equations describe the rotational aspects of the fields and their time evolution. Faraday's law of induction links the curl of \vec{E} to the time rate of change of \vec{B}:\nabla \times \vec{E} = -\frac{\partial \vec{B}}{\partial t}.This captures how a changing magnetic field induces an electric field, fundamental to electromagnetic induction. Ampère's circuital law, augmented by Maxwell's displacement current term, relates the curl of \vec{B} to currents and the time-varying electric field:\nabla \times \vec{B} = \mu_0 \vec{J} + \mu_0 \epsilon_0 \frac{\partial \vec{E}}{\partial t},where \mu_0 is the vacuum permeability; the displacement current ensures consistency with charge conservation and enables wave propagation.These differential equations derive directly from their integral counterparts using fundamental theorems of vector calculus. The divergence theorem, proved by Carl Friedrich Gauss in 1813, equates the volume integral of a vector field's divergence to the flux through the enclosing surface: \int_V (\nabla \cdot \vec{F}) \, dV = \oint_S \vec{F} \cdot d\vec{a}. Applying this to the integral forms of Gauss's laws yields the differential versions for arbitrary volumes. Similarly, Stokes' theorem, established by George Gabriel Stokes in 1850, relates the surface integral of a curl to the line integral around the boundary: \int_S (\nabla \times \vec{F}) \cdot d\vec{a} = \oint_C \vec{F} \cdot d\vec{l}. This transforms the line and surface integrals in Faraday's and Ampère's laws into local curl expressions valid at every point. These derivations assume smooth fields and suitable boundary conditions, confirming the equivalence between integral and differential formulations.The differential equations are often solved using electromagnetic potentials to exploit gaugefreedom and simplify computations. The magnetic field derives from a vector potential \vec{A} via \vec{B} = \nabla \times \vec{A}, satisfying \nabla \cdot \vec{B} = 0 automatically due to the curl's divergence-free property. The electric field expresses as \vec{E} = -\nabla \phi - \frac{\partial \vec{A}}{\partial t}, where \phi is the scalar potential; this form satisfies Faraday's law identically. Substituting into the remaining equations yields coupled partial differential equations for \phi and \vec{A}, such as the inhomogeneous wave equations in Lorentz gauge. However, the potentials are not unique: a gauge transformation \vec{A}' = \vec{A} + \nabla \chi and \phi' = \phi - \frac{\partial \chi}{\partial t}, with arbitrary scalar \chi, leaves \vec{E} and \vec{B} unchanged, reflecting the redundancy in the potential description. This gauge freedom allows choices like the Coulomb gauge (\nabla \cdot \vec{A} = 0) or Lorentz gauge (\nabla \cdot \vec{A} + \frac{1}{c^2} \frac{\partial \phi}{\partial t} = 0, with c = 1/\sqrt{\mu_0 \epsilon_0}) to simplify solutions while preserving physical observables.
Electromagnetic Waves
Derivation and Wave Equation
The derivation of the electromagnetic wave equation begins with Maxwell's equations in their differential form, which unify electricity, magnetism, and optics. To obtain the wave equation for the electric field \vec{E}, start by taking the curl of Faraday's law of induction:\nabla \times (\nabla \times \vec{E}) = -\frac{\partial}{\partial t} (\nabla \times \vec{B}).This follows directly from the curl form of Faraday's law, \nabla \times \vec{E} = -\frac{\partial \vec{B}}{\partial t}.[53] Next, substitute the Ampère-Maxwell law, \nabla \times \vec{B} = \mu_0 \vec{J} + \mu_0 \epsilon_0 \frac{\partial \vec{E}}{\partial t}, into the right-hand side, yielding:\nabla \times (\nabla \times \vec{E}) = -\mu_0 \frac{\partial \vec{J}}{\partial t} - \mu_0 \epsilon_0 \frac{\partial^2 \vec{E}}{\partial t^2}.Applying the vector identity \nabla \times (\nabla \times \vec{E}) = \nabla (\nabla \cdot \vec{E}) - \nabla^2 \vec{E} and using Gauss's law \nabla \cdot \vec{E} = \rho / \epsilon_0 results in the inhomogeneous wave equation:\nabla^2 \vec{E} - \frac{1}{c^2} \frac{\partial^2 \vec{E}}{\partial t^2} = \frac{1}{\epsilon_0} \nabla \rho + \mu_0 \frac{\partial \vec{J}}{\partial t},where the speed c = 1 / \sqrt{\mu_0 \epsilon_0} emerges as the propagation velocity of electromagnetic disturbances in vacuum.[53] This equation describes how electric field variations propagate as waves, influenced by charge density \rho and current density \vec{J}.A similar derivation applies to the magnetic field \vec{B}. Taking the curl of the Ampère-Maxwell law and substituting Faraday's law leads to:\nabla^2 \vec{B} - \frac{1}{c^2} \frac{\partial^2 \vec{B}}{\partial t^2} = -\mu_0 \nabla \times \vec{J}.In regions free of sources (\rho = 0, \vec{J} = 0), the equations simplify to the homogeneous forms \nabla^2 \vec{E} - \frac{1}{c^2} \frac{\partial^2 \vec{E}}{\partial t^2} = 0 and \nabla^2 \vec{B} - \frac{1}{c^2} \frac{\partial^2 \vec{B}}{\partial t^2} = 0, indicating wave propagation without dissipation.[53]Solutions to these homogeneous equations include plane waves, such as \vec{E} = \vec{E_0} \cos(\vec{k} \cdot \vec{r} - \omega t), where \vec{k} is the wave vector, \omega is the angular frequency, and \omega = c k with k = |\vec{k}|. From Gauss's law in vacuum (\nabla \cdot \vec{E} = 0) and the relations \vec{B} = \frac{1}{c} \hat{k} \times \vec{E}, the waves exhibit transverse polarization: both \vec{E} and \vec{B} are perpendicular to \vec{k}, and \vec{E} \perp \vec{B}.[53]James Clerk Maxwell first derived these results in 1865, calculating c \approx 3.107 \times 10^8 m/s using contemporary values of \mu_0 and \epsilon_0, which closely matched the known speed of light, implying that light is an electromagnetic wave.[53] This prediction was experimentally confirmed by Heinrich Hertz in 1887–1888 through generation and detection of radio waves at speeds matching c.[57]
Properties and Propagation
Electromagnetic waves are transverse in nature, with the electric field \vec{E}, magnetic field \vec{B}, and propagation direction given by the wave vector \vec{k} all mutually perpendicular, satisfying \vec{E} \perp \vec{B} \perp \vec{k}.[58] This orthogonality follows from the structure of plane wave solutions to Maxwell's equations in vacuum.[59] The intrinsic relationship between the fields is quantified by the impedance of free space, Z_0 = \sqrt{\mu_0 / \epsilon_0} \approx 377 \, \Omega, which sets the ratio |\vec{E}| / |\vec{B}| = c for plane waves, where c is the speed of light.[60]Polarization describes the orientation of the electric field oscillation in the plane perpendicular to \vec{k}. Linear polarization occurs when \vec{E} vibrates along a fixed direction, while circular polarization arises when \vec{E} rotates at constant magnitude, either clockwise (right-handed) or counterclockwise (left-handed), often as a superposition of two linear components phase-shifted by \pi/2.[61] The interaction of polarized waves with polarizing materials, such as ideal polarizers that transmit only the component parallel to their axis, follows Malus's law: the transmitted intensity is I = I_0 \cos^2 \theta, where \theta is the angle between the incident polarization and the polarizer axis.[62]Upon oblique incidence at an interface between two dielectric media, electromagnetic waves partially reflect and refract according to boundary conditions on the tangential field components. The angle of reflection equals the angle of incidence, while refraction obeys Snell's law: n_1 \sin \theta_1 = n_2 \sin \theta_2, with n the refractive index of each medium and \theta the respective angles from the normal.[63] For the field amplitudes, the Fresnel equations provide the reflection coefficient r and transmission coefficient t, which differ for s-polarization (electric field perpendicular to the plane of incidence) and p-polarization (parallel to it); for example, at normal incidence, r = (n_1 - n_2)/(n_1 + n_2).[64] These relations hold for all electromagnetic frequencies, from radio waves to visible light, assuming non-magnetic media.In material media, the propagation speed v = c / n varies with frequency due to dispersion, where the refractive index n(\omega) depends on the angular frequency \omega. Normal dispersion, common in transparent regions like glass for visible wavelengths, features dn/d\omega > 0 (or dn/d\lambda < 0), causing shorter wavelengths to travel slower. Anomalous dispersion occurs near strong absorption resonances, where dn/d\omega < 0, leading to counterintuitive speed increases with frequency, though absorption dominates in such bands.[65] This frequency dependence arises from the medium's dielectric response to the oscillating fields.[66]Representative examples of electromagnetic waves include radio waves, with wavelengths from meters to kilometers used in broadcasting and wireless communication, and visible light, spanning 400–700 nm wavelengths that enable human vision and optical phenomena like rainbows.[67]
Energy and Momentum
Poynting Theorem and Energy Flow
The Poynting theorem expresses the conservation of energy in electromagnetic fields, relating the rate of change of field energy density to the divergence of the energy flux and the work done on charges.[68] This theorem was originally derived by John Henry Poynting in 1884, building on Maxwell's equations to describe the flow of electromagnetic energy.[68]To derive the theorem, start from Maxwell's equations in differential form and consider the scalar product of the electric field \vec{E} with Ampère's law and the magnetic field \vec{B} with Faraday's law:\vec{E} \cdot (\nabla \times \vec{B}) = \mu_0 \vec{E} \cdot \vec{J} + \mu_0 \epsilon_0 \frac{\partial \vec{E}}{\partial t} \cdot \vec{E},\vec{B} \cdot (\nabla \times \vec{E}) = -\frac{\partial \vec{B}}{\partial t} \cdot \vec{B}.Subtracting the second equation from the first gives\vec{E} \cdot (\nabla \times \vec{B}) - \vec{B} \cdot (\nabla \times \vec{E}) = \mu_0 \vec{E} \cdot \vec{J} + \mu_0 \epsilon_0 \frac{\partial \vec{E}}{\partial t} \cdot \vec{E} + \vec{B} \cdot \frac{\partial \vec{B}}{\partial t}.Applying the vector identity \vec{E} \cdot (\nabla \times \vec{B}) - \vec{B} \cdot (\nabla \times \vec{E}) = -\nabla \cdot (\vec{E} \times \vec{B}) and recognizing that the right-hand side equals \mu_0 (\vec{J} \cdot \vec{E} + \frac{\partial u}{\partial t}), where u = \frac{1}{2} \left( \epsilon_0 E^2 + \frac{B^2}{\mu_0} \right) is the electromagnetic energy density, yields, after rearrangement,-\frac{\partial u}{\partial t} = \nabla \cdot \vec{S} + \vec{J} \cdot \vec{E},with the Poynting vector \vec{S} = \frac{1}{\mu_0} \vec{E} \times \vec{B}, representing the energy flux.[69] Integrating over a volume V with surface \partial V gives the integral form:-\frac{d}{dt} \int_V u \, dV = \oint_{\partial V} \vec{S} \cdot d\vec{A} + \int_V \vec{J} \cdot \vec{E} \, dV,indicating that the decrease in stored energy equals the outward energy flow plus the power delivered to charges.[70]The term \vec{J} \cdot \vec{E} represents the rate of work done by the fields on charges per unit volume, such as Joule heating in conductors. The Poynting vector \vec{S} points in the direction of energy propagation and its magnitude gives the power per unit area; for plane waves in vacuum, the time-averaged \langle |\vec{S}| \rangle = \frac{1}{2} c \epsilon_0 E_0^2, where c is the speed of light and E_0 the peak electric field.[70]In waveguides, the Poynting vector describes the directional energy flow along the guide, even though fields are transverse; for a rectangular waveguide in the TE_{10} mode, the time-averaged \vec{S} is primarily in the propagation direction, with its integral over the cross-section yielding the guided power.[71] Similarly, for radiation from a dipole antenna, the far-field Poynting vector radiates energy spherically, with the time-averaged magnitude \langle S_r \rangle = \frac{\mu_0 p_0^2 \omega^4 \sin^2 \theta}{32 \pi^2 c r^2} (where p_0 is the dipole moment, \omega the frequency, and r, \theta spherical coordinates), determining the radiation pattern and total radiated power.[72]In static limits, the Poynting theorem reduces to energy storage in capacitors and inductors. For a charging capacitor, the displacement current produces a time-varying \vec{B}, leading to a Poynting vector directing energy inward between plates, where the stored energy \frac{1}{2} C V^2 matches the integral of \vec{S} over the surface. For an inductor, the changing current induces \vec{E}, with \vec{S} flowing into the coil to account for the magnetic energy \frac{1}{2} L I^2.[73] In the quasistatic approximation, these flows confirm that field energy resides in space rather than the devices themselves.[74]
Electromagnetic Momentum and Radiation Pressure
Electromagnetic fields carry linear momentum, which can be transferred to matter upon interaction, leading to mechanical forces such as radiation pressure. The momentum density of the electromagnetic field in vacuum is given by \vec{g} = \epsilon_0 \vec{E} \times \vec{B}, where \epsilon_0 is the vacuum permittivity, \vec{E} is the electric field, and \vec{B} is the magnetic field.[75] This expression arises from the field's interaction with charges and currents as described by the Lorentz force law. The total momentum \vec{P} contained within a volume V is then the volume integral \vec{P} = \int_V \vec{g} \, dV.[76]Conservation of total momentum in classical electrodynamics follows from the Maxwell equations combined with the Lorentz force density \vec{f} = \rho \vec{E} + \vec{J} \times \vec{B}, where \rho is the charge density and \vec{J} is the current density. This yields a continuity equation analogous to that for energy: \frac{\partial}{\partial t} (\vec{p}_\text{mech} + \vec{g}) + \nabla \cdot \overleftrightarrow{T} = 0, where \vec{p}_\text{mech} is the mechanical momentum density and \overleftrightarrow{T} is the Maxwell stress tensor representing the momentum flux.[77] Integrating over a volume shows that changes in the total momentum (mechanical plus field) equal the net flux through the surface, ensuring conservation in closed systems.[78]When electromagnetic waves interact with matter, they exert radiation pressure due to momentum transfer. For a plane wave with intensity I incident normally on a perfect absorber, the pressure is P = \frac{I}{c}, where c is the speed of light; for a perfect reflector, the pressure doubles to P = \frac{2I}{c} because the reflected wave reverses the momentum direction.[79] This effect is evident in astronomical phenomena, such as the formation of comet dust tails, where solar radiation pressure pushes small dust particles away from the Sun, creating elongated structures trailing the comet nucleus.[80] A historical example is the Crookes radiometer, initially misinterpreted as demonstrating radiation pressure; in reality, its rotation arises from thermal transpiration effects in the rarefied gas, not direct momentum transfer from light, as the vanes rotate with the brighter side forward, opposite to the expected pressure direction.[81]Electromagnetic fields also carry angular momentum, which can be decomposed into orbital and spin components. The orbital angular momentum density is \vec{l} = \vec{r} \times \vec{g}. The spin angular momentum, particularly in circularly polarized waves, arises from the field's helicity—a measure of the screw-like handedness. For a circularly polarized plane wave, the time-averaged spin angular momentum density points along the propagation direction, with magnitude \langle s \rangle = \pm \frac{\langle u \rangle}{\omega}, where u is the energy density, \omega is the angular frequency, and the sign indicates left- or right-handed polarization. This intrinsic angular momentum can be transferred to matter, such as in optical tweezers, but remains a classical field property conserved under Maxwell's equations in free space.[82]
Classical Models
Microscopic Electron Models
In the early 20th century, J.J. Thomson proposed a model of the atom known as the plum pudding model, envisioning it as a sphere of uniform positive charge with embedded electrons distributed like plums in a pudding.[83] This structure aimed to maintain overall neutrality while accounting for the recently discovered electron as a fundamental constituent.[83] The model suggested that electrostatic forces balanced the electrons within the positive sphere, providing stability for atomic structure.[83]However, the plum pudding model proved unstable under scattering experiments. In 1911, Ernest Rutherford's gold foil experiment demonstrated that alpha particles could penetrate thin foils with minimal deflection for most, but occasional large-angle scatters indicated a concentrated positive charge rather than a diffuse distribution, contradicting Thomson's uniform sphere.[84] This instability highlighted the inadequacy of the model for explaining atomic interactions at the microscopic level.[84]Building on such ideas, Hendrik Lorentz developed a classical model of the electron in 1904 as a rigid, uniformly charged sphere to address electromagnetic behavior in moving systems.[85] This rigid electron was intended to embody the particle's electromagnetic mass arising solely from its self-field, with the Lorentz force governing its motion under external fields.[85] The model treated the electron as an extended object to avoid singularities, assuming incompressibility for stability.[85]A key challenge in the Lorentz model was the self-force exerted by the electron's own electromagnetic field, leading to the Abraham-Lorentz formula for radiation reaction. Derived independently by Max Abraham in 1903 and Lorentz in 1904, this force accounts for energy loss due to radiation from acceleration:\vec{F}_{\mathrm{rad}} = \frac{\mu_0 q^2}{6\pi c} \frac{d^2 \vec{v}}{dt^2}.[86][85] The formula modifies the equation of motion to include this back-reaction, where q is the charge, \vec{v} the velocity, and c the speed of light.[86]These models encountered fundamental problems, including infinite self-energy for point-like limits and unphysical runaway solutions in the Abraham-Lorentz equation, where particles accelerate indefinitely without external forces.[86] The self-energy divergence arises from integrating the electromagnetic field energy over the electron's volume, yielding infinity as the radius approaches zero, rendering the classical electron's mass ill-defined.[87] Runaway solutions emerge from the third-order differential equation, implying pre-acceleration or exponential velocity growth, both physically untenable.[86] These issues underscored the limitations of classical theory, foreshadowing the need for quantum mechanics to resolve microscopic electromagnetic paradoxes.[87]Despite these flaws, the models enabled key applications, such as calculating radiation from accelerating charges via the Larmor formula, derived by Joseph Larmor in 1897. This gives the non-relativistic power radiated by a single charge:P = \frac{\mu_0 q^2 a^2}{6\pi c},where a is the acceleration magnitude.[88] The formula quantifies energy loss in atomic processes, like electron oscillations, providing a cornerstone for classical radiation theory.[88]
Macroscopic Dielectric and Magnetic Materials
In macroscopic electromagnetism, materials are described using average fields that incorporate the collective response of bound charges and currents, without resolving individual atomic or molecular interactions. This approach introduces auxiliary fields, the electric displacement \vec{D} and the magnetic field strength \vec{H}, to simplify the treatment of dielectrics and magnetic materials in Maxwell's equations. These fields account for the material's polarization and magnetization, respectively, allowing the equations to focus on free charges and currents as sources.In dielectric materials, an applied electric field \vec{E} induces a polarization \vec{P}, defined as the electric dipole moment per unit volume, arising from the displacement of bound charges within the material. This polarization generates bound charge densities that oppose the applied field, effectively screening it. The electric displacement field is then given by\vec{D} = \epsilon_0 \vec{E} + \vec{P},where \epsilon_0 is the vacuum permittivity, relating the total electric field to both free and bound contributions.For linear isotropic dielectrics, where the response is proportional to the applied field, the polarization is \vec{P} = \epsilon_0 \chi_e \vec{E}, with \chi_e the electric susceptibility (a dimensionless measure of the material's polarizability). The permittivity \epsilon then follows as \epsilon = \epsilon_0 (1 + \chi_e), and the constitutive relation simplifies to \vec{D} = \epsilon \vec{E}. Typical values of \chi_e > 0 for most dielectrics, such as water with \epsilon_r \approx 80 (where \epsilon_r = 1 + \chi_e), illustrate how materials can store significant electric energy.In magnetic materials, an applied magnetic field induces a magnetization \vec{M}, defined as the magnetic dipole moment per unit volume, which can originate from the alignment of atomic magnetic moments or, equivalently, from microscopic Amperian current loops per unit volume. These macroscopic quantities aggregate the underlying microscopic electron motions without detailing them. The magnetic flux density \vec{B} relates to the magnetization via\vec{B} = \mu_0 (\vec{H} + \vec{M}),where \mu_0 is the vacuum permeability and \vec{H} is the auxiliary field that sources free currents.[89][90]For linear isotropic magnetic materials, the magnetization is \vec{M} = \chi_m \vec{H}, with \chi_m the magnetic susceptibility. The permeability is then \mu = \mu_0 (1 + \chi_m), yielding the constitutive relation \vec{B} = \mu \vec{H}. Paramagnetic materials have small positive \chi_m (e.g., aluminum with \chi_m \approx 2 \times 10^{-5}), enhancing the field slightly, while diamagnetic materials exhibit weak negative \chi_m (e.g., copper with \chi_m \approx -10^{-5}), opposing it. Ferromagnetic materials, like iron, show nonlinear behavior with large \chi_m, but linear approximations apply in weak fields.[89]The constitutive relations \vec{D} = \epsilon \vec{E} and \vec{B} = \mu \vec{H} encapsulate the linear response of isotropic media, where \epsilon and \mu may be tensors in anisotropic cases. These relations bridge the microscopic origins—such as atomic dipoles—to bulk properties, enabling practical computations in devices like capacitors and inductors. In general, \epsilon_r > 1 and \mu_r \approx 1 for most non-magnetic dielectrics, highlighting their primary role in electric rather than magnetic effects.At interfaces between materials, boundary conditions ensure continuity of fields derived from Maxwell's equations. The normal component of \vec{D} is continuous across the boundary in the absence of free surface charge density \sigma_f, i.e., D_{2\perp} - D_{1\perp} = \sigma_f = 0. Similarly, the tangential component of \vec{H} is continuous without free surface currents, H_{2\parallel} - H_{1\parallel} = 0. These conditions, applied via pillbox and loop integrals, maintain consistency in energy and momentum flow.[91]Maxwell's equations in matter incorporate these auxiliary fields to describe fields sourced by free charges \rho_f and currents \vec{J}_f:\nabla \cdot \vec{D} = \rho_f, \quad \nabla \cdot \vec{B} = 0,\nabla \times \vec{E} = -\frac{\partial \vec{B}}{\partial t}, \quad \nabla \times \vec{H} = \vec{J}_f + \frac{\partial \vec{D}}{\partial t}.Bound effects are absorbed into \vec{D} and \vec{H}, simplifying analysis while preserving the differential form of the vacuum equations.[92]