Fact-checked by Grok 2 weeks ago

Electric field

The electric field is a fundamental concept in , defined as the electric force per unit positive test charge exerted on a small test charge placed at a given point in space, allowing the description of electrostatic interactions at a without direct contact between charges. It is a quantity, with its direction defined as that of the force on a positive test charge, pointing away from positive source charges and toward negative ones. The magnitude of the electric field \mathbf{E} due to a point charge q at a r is given by E = \frac{k q}{r^2}, where k = 8.99 \times 10^9 \, \mathrm{N \cdot m^2 / C^2} is Coulomb's constant, and the SI unit is newtons per coulomb (N/C), equivalently volts per meter (V/m). Introduced conceptually by Michael Faraday in the early 19th century as a way to model how charges influence one another through space, the electric field provides a physical framework for understanding phenomena like electrostatic attraction and repulsion, later formalized mathematically by James Clerk Maxwell in his equations of electromagnetism. Key properties include the superposition principle, where the total field from multiple charges is the vector sum of individual fields, and its conservative nature, meaning the work done by the field on a charge is path-independent and related to the electric potential V via \mathbf{E} = -\nabla V. Electric field lines, another visualization tool pioneered by Faraday, illustrate the field's direction (tangent to the lines) and relative strength (density of lines), never crossing and originating from positive charges while terminating on negative ones. This concept underpins diverse applications, from the operation of capacitors and electronic devices to the behavior of charged particles in accelerators and natural atmospheric electricity.

Fundamentals

Definition and Physical Description

The electric field is a quantity that describes the experienced by a due to the presence of other charges. It is formally defined as the electrostatic \vec{F} exerted on a small positive test charge q divided by that charge, expressed as \vec{E} = \frac{\vec{F}}{q}, where the test charge is assumed to be infinitesimally small to avoid perturbing the field significantly. This definition captures the field's role as a mediator of electrostatic interactions between charges, independent of the test charge itself. Physically, the electric field permeates the space surrounding electric charges, creating an invisible region where other charges would experience a force if placed there. This field arises from the charges producing it and extends outward in all directions, with its direction defined as that of the force on a positive test charge. To aid visualization, electric field lines are employed: these are continuous curves tangent to the field vector at every point, originating from positive charges and terminating on negative charges (or extending to for net positive or negative distributions). The of field lines in a region is proportional to the field's magnitude, providing an intuitive representation of varying strength. The strength of the electric field is measured in SI units of newtons per (N/C), equivalent to volts per meter (V/m), reflecting its dual interpretation as force per unit charge or negative gradient of . For instance, around an isolated point charge q, the field is radial—pointing radially outward if q is positive and inward if negative—with magnitude decreasing inversely proportional to the square of the distance from the charge, illustrating the field's rapid falloff with separation.

Historical Development

The earliest observations of electrostatic phenomena trace back to , where around 600 BC noted that , when rubbed with , could attract lightweight objects such as straw and feathers, marking the initial recognition of . In the late , William Gilbert expanded on these ideas through systematic experiments detailed in his 1600 work , coining the term "electric" from the Greek ēlektron for and distinguishing electrical attraction from while proposing an "electric fluid" released by . During the , investigations intensified with Charles François du Fay's 1733 experiments identifying two distinct types of —vitreous (from glass) and resinous (from amber)—and demonstrating that like charges repel while opposites attract, laying groundwork for understanding charge polarity. advanced this in the 1750s with his one-fluid theory, positing that consists of a single fluid whose excess or deficiency in bodies causes attraction or repulsion, as outlined in his 1747 letters and 1751 publication Experiments and Observations on , where he also named the charges positive and negative. In 1785, quantified the electrostatic force using a torsion balance, publishing his in the Mémoires de Mathématique et de Physique de la Société Royale des Sciences, which enabled the for combining forces from multiple charges. Michael Faraday revolutionized the conceptual framework in the 1830s and 1840s by rejecting action-at-a-distance in favor of a continuous field permeating space, introducing "lines of force" to represent the 's direction and intensity as a physical medium, as elaborated in his multi-volume (1839–1855). James Clerk Maxwell formalized this in the 1860s, unifying , magnetism, and through his theory, detailed in the 1865 paper "A Dynamical Theory of the ," where the emerges as a key component propagating as waves. In the 20th century, Albert Einstein's 1905 theory of further refined the electric field by showing it and the as intertwined aspects of a single , with their manifestations depending on the observer's , as demonstrated in his paper "On the Electrodynamics of Moving Bodies."

Mathematical Formulation in

and Field of a Point Charge

Coulomb's law describes the electrostatic force \mathbf{F} between two stationary point charges q_1 and q_2 separated by a distance r in as \mathbf{F} = \frac{1}{4\pi\epsilon_0} \frac{q_1 q_2}{r^2} \hat{\mathbf{r}}, where \hat{\mathbf{r}} is the unit vector pointing from q_1 to q_2, \epsilon_0 is the , and the force is along the line joining the charges. The magnitude of the force is inversely proportional to the square of the separation distance, and its direction is repulsive for like charges and attractive for opposite charges. The constant \frac{1}{4\pi\epsilon_0} arises from the constant k \approx 8.99 \times 10^9 \, \mathrm{N \cdot m^2 / C^2}, with \epsilon_0 \approx 8.85 \times 10^{-12} \, \mathrm{F/m}. The electric field \mathbf{E} produced by a single point charge q at a point in space is defined as the electrostatic force per unit positive test charge that would be experienced at that point, in the limit as the test charge q_0 approaches zero to avoid perturbing the source charge. Applying to this setup, the force on the test charge is \mathbf{F} = \frac{1}{4\pi\epsilon_0} \frac{q q_0}{r^2} \hat{\mathbf{r}}, so the electric field is \mathbf{E} = \frac{\mathbf{F}}{q_0} = \frac{1}{4\pi\epsilon_0} \frac{q}{r^2} \hat{\mathbf{r}}. This expression shows that the field is radial, with magnitude decreasing as $1/r^2 from the charge. For a positive charge q > 0, \mathbf{E} points away from the charge along \hat{\mathbf{r}}; for q < 0, it points toward the charge. At the location of the point charge itself (r = 0), the electric field strength formally diverges to due to the $1/r^2 dependence, representing a mathematical inherent to the idealization of a zero-size point charge. In physical reality, this is avoided because elementary charges possess finite spatial extent or are distributed, preventing infinite field values.

Superposition Principle

The superposition principle states that the total electric field \vec{E} at any point in space due to a collection of point charges is the vector sum of the electric fields \vec{E}_i produced by each individual charge: \vec{E} = \sum_i \vec{E}_i. This principle holds because the governing equations of electrostatics, derived from Maxwell's equations, are linear in the charge density and thus permit such additive solutions. The principle was implicitly evident in Charles-Augustin de Coulomb's 1785 experiments measuring forces between charged objects, where forces from multiple charges were observed to add without mutual among the sources. It was later formalized within the broader electromagnetic framework by James Clerk Maxwell in his 1861–1865 , where the linearity of the field equations explicitly supports superposition for electrostatic configurations. A key implication of the is that electric field sources act independently in , with no direct interaction between them mediated by the field itself; this allows complex systems, such as assemblies of many charges, to be analyzed by breaking them down into contributions from simpler, isolated elements. Consequently, calculations for intricate charge arrangements become tractable through stepwise addition rather than solving nonlinear coupled equations. The nature of the addition requires considering both magnitude and direction. For instance, with two point charges of equal magnitude q but opposite sign, separated by distance d to form a , the individual fields \vec{E}_1 and \vec{E}_2 at a distant point can be added graphically by placing the of one vector at the head of the other, or algebraically by summing their components: \vec{E} = \vec{E}_1 + \vec{E}_2 = (E_{1x} + E_{2x})\hat{i} + (E_{1y} + E_{2y})\hat{j}. Along the axis, the fields align and reinforce, yielding a stronger total field, whereas in the equatorial plane, they oppose each other, leading to partial cancellation and a weaker, reversed-direction field. As a concrete example, consider calculating the electric field at a point midway between two point charges: one of +2q and one of -q, separated by $2a. The field from +2q points away from it with magnitude proportional to $2q / a^2, while the field from -q points toward it (and thus in the same direction as the first) with magnitude q / a^2; the total field is their sum, (2q / a^2 + q / a^2) = 3q / a^2 in that direction, demonstrating reinforcement. If both charges were positive and equal, the fields would point oppositely at the midpoint, resulting in complete cancellation for identical magnitudes. The extends to continuous charge distributions by integrating the infinitesimal field contributions from each , yielding the total field as \vec{E}(\vec{r}) = \int \vec{E}(\vec{r}, \vec{r}') \rho(\vec{r}') dV'.

Fields from Charge Distributions

For discrete distributions consisting of a finite number of point charges, the electric field \mathbf{E} at a point in space is obtained by applying the , summing the contributions from each individual charge. The magnitude and direction of the field from each point charge follow , resulting in the vector expression \mathbf{E} = \sum_i \frac{1}{4\pi\epsilon_0} \frac{q_i}{r_i^2} \hat{\mathbf{r}}_i, where q_i is the magnitude of the i-th charge, r_i is the distance from that charge to the field point, \hat{\mathbf{r}}_i is the unit vector pointing from the charge to the field point, and \epsilon_0 is the vacuum permittivity. This summation accounts for both scalar magnitudes and vector directions, requiring careful resolution of components in Cartesian or other coordinates for computation. When dealing with continuous charge distributions, the discrete summation is replaced by an integral over the distribution, treating the charge as composed of infinitesimal elements dq. The general formula for the electric field becomes \mathbf{E} = \frac{1}{4\pi\epsilon_0} \int \frac{dq}{r^2} \hat{\mathbf{r}}, where r is the distance from each charge element to the field point and \hat{\mathbf{r}} is the corresponding unit vector. The infinitesimal charge dq depends on the geometry: for a line distribution, dq = \lambda \, dl with linear charge density \lambda; for a surface distribution, dq = \sigma \, dA with surface charge density \sigma; and for a volume distribution, dq = \rho \, dV with volume charge density \rho. Evaluating this integral typically involves choosing appropriate coordinates (e.g., cylindrical for lines, spherical for volumes) and resolving the vector components, often simplifying through symmetry in the charge arrangement./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/05%3A_Electric_Charges_and_Fields/5.06%3A_Calculating_Electric_Fields_of_Charge_Distributions) Symmetry in the charge plays a crucial role in simplifying these integrals by allowing cancellation of certain field components or reduction of the integration limits. For instance, in rotationally symmetric cases like a uniformly charged or , azimuthal components cancel, leaving only radial contributions. , as in infinite lines or planes, further constrains the field direction, making it perpendicular to the distribution and independent of position along the symmetric axis./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/05%3A_Electric_Charges_and_Fields/5.06%3A_Calculating_Electric_Fields_of_Charge_Distributions) Such exploitations reduce complex vector integrals to scalar ones, enhancing computational feasibility without altering the underlying physics. Representative examples illustrate these principles. For an infinite straight line with uniform linear \lambda, dictates a radial to the line; integrating the contributions yields E = \frac{\lambda}{2\pi\epsilon_0 r}, directed outward for positive \lambda, where r is the from the line. This result decays as $1/r, slower than the $1/r^2 for point charges, reflecting the extended . Similarly, for an infinite plane with uniform surface \sigma, the is uniform and to the plane, with magnitude E = \frac{\sigma}{2\epsilon_0}, constant on either side and pointing away from the plane for positive \sigma; the integration over the infinite area shows no distance dependence due to symmetric pairwise contributions. For arbitrary charge distributions lacking high symmetry, analytical integration is often intractable, necessitating numerical methods such as discretizing the distribution into point charges and approximating the sum, or using quadrature rules for the integral. Modern computational approaches, like the trapezoidal rule or more advanced techniques in software packages, enable accurate field calculations for complex geometries in simulations.

Relation to Electric Potential

The electric potential V at a point in an electrostatic field is defined as the negative of the electric field \mathbf{E} from a point to that location, given by V(\mathbf{r}) = -\int_{\text{ref}}^{\mathbf{r}} \mathbf{E} \cdot d\mathbf{l}, where the path is arbitrary due to the conservative nature of . For localized charge distributions, the point is conventionally chosen at , where V = 0, ensuring the potential vanishes far from the charges. This definition arises from the work required to assemble the charge configuration, divided by the test charge. In electrostatics, the electric field is irrotational, satisfying \nabla \times \mathbf{E} = 0, which implies that \mathbf{E} can be derived from a scalar potential function via \mathbf{E} = -\nabla V. This relation holds because any vector field with zero curl is the gradient of a scalar potential, allowing the vectorial electric field to be obtained by differentiating the scalar V in three dimensions. The negative sign ensures that the field points from higher to lower potential, consistent with the force on positive charges. For a single point charge q at the , the potential simplifies to V(r) = \frac{1}{4\pi\epsilon_0} \frac{q}{r}, where r is the distance from the charge and \epsilon_0 is the ; this follows directly from integrating the field along a radial path from infinity. In general, for multiple charges, the total potential is the algebraic sum of individual contributions, reflecting its scalar nature. Equipotential surfaces, where V is constant, are always perpendicular to the electric field lines, as the gradient \nabla V is normal to surfaces of constant V. Along such a surface, the line integral of \mathbf{E} is zero, meaning no net work is done when moving a test charge tangentially to the field. This geometric property aids in visualizing field configurations and solving boundary value problems. The scalar electric potential offers significant advantages over directly computing the electric field, particularly for complex geometries or distributed charges, as scalar addition avoids the need to resolve vector components and directions. This simplification is especially useful in calculating fields in symmetric systems, such as spherical or cylindrical arrangements, where direct integration can be cumbersome.

Properties of Electrostatic Fields

Uniform and Non-Uniform Fields

In , a electric field is characterized by a magnitude and direction throughout the , represented as a \mathbf{E}. Such fields arise in configurations where the charge distribution produces no spatial variation in the field strength, such as between two large, plates with opposite surface charge densities \sigma and -\sigma. In this setup, the electric field between the plates has magnitude E = \sigma / \epsilon_0, directed from the positive to the negative plate, assuming the plate separation is small compared to their dimensions. Non-uniform electric fields, in contrast, vary in magnitude and/or direction depending on position. For a single point charge q, the field is radial and decreases with distance r as E \propto 1/r^2, following , with lines emanating outward for positive charges and inward for negative ones. For an electric dipole consisting of charges +q and -q separated by distance d, the far-field strength falls off more rapidly as E \propto 1/r^3, where r is the distance from the center much larger than d. Electric fields are often visualized using field lines, which indicate direction (tangent to the field ) and relative strength (denser lines for stronger fields). In fields, these lines are straight and , reflecting the constant \mathbf{E}. In non-uniform fields, such as those from a point charge or , the lines diverge or converge radially, spreading out with increasing distance to show the field's decay. Uniform electric fields find application in particle accelerators, where they provide consistent to charged particles, as in the gaps between electrodes in cyclotrons or linear accelerators. The strength of an electric field, uniform or non-uniform, can be measured using a small positive test charge q_0, where \mathbf{E} = \mathbf{F}/q_0 and \mathbf{F} is the force on q_0, ensuring q_0 is negligible to avoid disturbing the field. Alternatively, for uniform fields, the magnitude is determined from the potential difference V over distance d via E = V/d, measured using a voltmeter across points in the field.

Parallels with Gravitational Fields

The electric field and the gravitational field exhibit striking mathematical and conceptual similarities, rooted in their descriptions as vector fields governed by inverse-square laws. The magnitude of the electric field \mathbf{E} produced by a point charge q at a distance r follows E \propto q / (4\pi\epsilon_0 r^2), mirroring the gravitational field \mathbf{g} due to a point mass M, where g = GM / r^2. This shared dependence on $1/r^2 arises from the underlying force laws—Coulomb's law for electricity and Newton's law of universal gravitation—both of which describe interactions that diminish with the square of the separation distance. Both fields are conservative vector fields, meaning the work done by the field on a along any closed path is zero, and they can be derived as the negative of a function. For the electric field, \mathbf{E} = -\nabla V, where V is the ; analogously, the gravitational field is \mathbf{g} = -\nabla \Phi, with \Phi as the . This property ensures path independence for the of the field, facilitating analyses in both systems. Additionally, the applies to both, allowing the total field from multiple sources to be the sum of individual contributions, due to the linearity of the governing equations. A further parallel is seen in their integral forms: for states that the flux of \mathbf{E} through a closed surface is Q_\text{enc}/\epsilon_0, while the analogous gives the flux of \mathbf{g} as -4\pi G M_\text{enc}. Despite these similarities, key differences distinguish the fields. Electric fields arise from charges that can be positive or negative, enabling both attractive and repulsive interactions, whereas s stem from masses that are always positive, resulting in purely attractive forces. Moreover, electric fields can be shielded—such as by placing a around a charge, which redistributes charges to cancel the internal field—while s permeate all matter without known shielding mechanisms, as no negative masses exist to produce cancellation. Historically, Michael Faraday's conceptualization of fields as continuous media influenced Albert Einstein's formulation of , where gravitational effects are treated as curvatures in analogous to field lines in . This analogy underscored the shift from action-at-a-distance to field-based descriptions in physics. Near the surface of , the approximates uniformity over small distances, much like the uniform electric field between parallel plates.

Continuous versus Discrete Charges

In electrostatics, charges are frequently modeled as discrete point charges, especially when representing fundamental particles such as electrons or protons in atoms, where the distances between charges are finite and on the order of atomic scales. This approach treats each charge as infinitesimally small compared to the separation from the observation point, allowing the electric field to be calculated precisely using for individual contributions and the for multiple charges. The discrete model is particularly suitable for systems with well-separated charges, providing exact results for field strengths at any finite distance away from the charges. In contrast, the continuous charge distribution model approximates a large collection of closely spaced charges as a smooth density spread over a , surface, or line, which is valid when the characteristic size of the or the relevant greatly exceeds the separation between individual charges. This is characterized by a volume charge density \rho(\mathbf{r}), where the total charge is obtained by integrating \rho over the distribution, and the electric field is found by integrating contributions from infinitesimal charge elements. Such approximations simplify calculations for macroscopic systems like charged conductors or plasmas, where treating each microscopic charge separately would be impractical. The transition between discrete and continuous models is facilitated by the , which approximates the potential due to a localized charge distribution in the far field—where the distance to the observation point is much larger than the distribution's size—by a series of terms starting with the (total charge), followed by the (net separation of positive and negative charges), and higher-order and multipole contributions that capture asymmetries. However, the point charge model exhibits a fundamental limitation: the electric field diverges to at r = 0, creating a that is unphysical and complicates calculations. At and subatomic scales, quantum mechanical effects further blur these classical distinctions, as the charge of an is described by a probability |\psi|^2 rather than a discrete point, rendering the continuous model more appropriate for wavefunction-based descriptions. A representative example is the modeling of an : at distances much larger than its radius (approximately $10^{-15} m), the nucleus can be treated as a continuous charge with a uniform or Fermi-type density \rho(r) to compute the external electric field accurately, reflecting the collective effect of its protons. In contrast, at shorter ranges probing the nuclear interior, the discrete nature of individual protons must be considered, as their finite separations dominate the field behavior.

Dynamic and Electromagnetic Aspects

Time-Varying Electric Fields

In , electric fields are assumed to be time-independent, satisfying ∇ × E = 0 and thus being conservative, with the potential difference between two points independent of the path taken. However, real-world scenarios often involve time-varying electric fields, where the E/∂t ≠ 0, extending the description beyond electrostatics to full electrodynamics as governed by . The key relation for such dynamics is Faraday's law, expressed as ∇ × E = -∂B/∂t, which indicates that a changing induces a electric field. Time-varying electric fields originate from two primary sources: the acceleration of electric charges, which produces radiation fields that propagate outward, and time-dependent magnetic fields, which directly induce electric fields via Faraday's law. For instance, an accelerating charge generates a non-static electric field component that decreases with distance and depends on the retardation time from the emission point. To ensure consistency in Ampère's law for time-dependent cases, Maxwell introduced the concept of displacement current, defined as J_d = ε₀ ∂E/∂t, which acts as an effective current density arising from the changing electric flux and enables the generation of magnetic fields even in charge-free regions. This term completes the Ampère-Maxwell law: ∇ × B = μ₀(J + ε₀ ∂E/∂t), resolving inconsistencies observed in circuits with capacitors. A classic example of a time-varying electric field occurs in a parallel-plate being charged by a I. The electric field between the plates increases linearly with time as E(t) = I t / (A ε₀), where A is the plate area, leading to a uniform ∂E/∂t = I / (A ε₀) that constitutes the density. This changing field produces a circling around the region between the plates, analogous to the field from conduction on the wires. The primary consequences of time-varying electric fields are that they become non-conservative, with non-zero implying path-dependent line integrals ∮ E · dl ≠ 0, and they drive induced electromotive forces (EMFs) in closed loops according to Faraday's law: ℰ = -dΦ_B/dt, where Φ_B is the . These effects underpin phenomena like in transformers and generators, where the induced E field opposes changes in per .

Electric Fields in Electromagnetic Waves

In electromagnetic waves, the electric field \mathbf{E} oscillates perpendicularly to both the magnetic field \mathbf{B} and the direction of wave propagation, forming a transverse wave structure that arises as a solution to Maxwell's equations in free space. This perpendicular orientation ensures that the fields mutually sustain each other: the changing electric field induces the magnetic field, and vice versa, allowing the wave to propagate without a medium. For a monochromatic plane wave traveling in the z-direction, the electric field can be expressed as \mathbf{E}(z, t) = \mathbf{E}_0 \cos(kz - \omega t + \phi), where \mathbf{E}_0 is the amplitude vector perpendicular to the propagation direction, k = 2\pi / \lambda is the wave number, \omega = 2\pi f is the angular frequency, and \phi is the phase constant. The corresponding magnetic field follows as \mathbf{B}(z, t) = \frac{1}{c} \hat{\mathbf{k}} \times \mathbf{E}(z, t), with magnitude B_0 = E_0 / c. The propagation speed of these waves in is the , given by c = 1 / \sqrt{\epsilon_0 \mu_0}, where \epsilon_0 is the and \mu_0 is the permeability of free space; this value, approximately $3 \times 10^8 m/s, emerges directly from the wave equations derived from Maxwell's curl equations. Energy transport in the wave occurs via the , \mathbf{S} = \frac{1}{\mu_0} \mathbf{E} \times \mathbf{B}, which points in the direction of propagation and represents the instantaneous power flux density. For a sinusoidal plane wave, the time-averaged intensity (power per unit area) is I = \frac{1}{2} c \epsilon_0 E_0^2, highlighting the quadratic dependence on the electric field amplitude, which scales the wave's energy density and thus its brightness or signal strength. The polarization of an electromagnetic wave is defined by the orientation and behavior of the \mathbf{E} vector in the plane perpendicular to propagation. Linear polarization occurs when \mathbf{E} oscillates along a fixed direction, such as horizontal or vertical, common in antennas for radio transmission. Circular polarization arises when the \mathbf{E} vector tip traces a circle, resulting from two linear components of equal amplitude and 90° phase difference, rotating either clockwise (right-handed) or counterclockwise (left-handed) as viewed along the propagation direction; this is useful in satellite communications to mitigate signal fading. Elliptical polarization is a general case combining these, with unequal amplitudes or phases. Examples of such waves span the spectrum: radio waves (frequencies ~10 kHz to 300 GHz) carry information in broadcasting, while visible light (~400–700 THz) exhibits polarization in natural phenomena like sky scattering and is manipulated in optics for displays and lasers. In , electromagnetic waves experience , a gradual decrease in due to (energy conversion to heat via interactions with atoms or molecules) and (redirection by inhomogeneities), leading to of as I = I_0 e^{-\alpha z}, where \alpha is the dependent on and material properties. For instance, radio waves attenuate more in conductive like due to induced currents, while attenuates in via minor , enabling fiber optics with low over long distances. This behavior contrasts with vacuum propagation, where no such occurs.

Energy Stored in Electric Fields

The energy stored in an electrostatic field arises from the work required to assemble the charge distribution that produces the field, and this energy can be expressed either in terms of the charges and their potentials or directly in terms of the field itself. In vacuum, the energy density u, or energy per unit volume, associated with the electric field \mathbf{E} is given by u = \frac{1}{2} \epsilon_0 E^2, where \epsilon_0 is the vacuum permittivity. This expression quantifies how the field's strength determines the local energy concentration, with the quadratic dependence reflecting the pairwise interactions among charges. The total electrostatic energy U for a given charge distribution is obtained by integrating the over all space: U = \frac{\epsilon_0}{2} \int E^2 \, dV, where the integral extends over the entire volume. This formulation treats the as residing in the field, independent of the specific charges, and is equivalent to the alternative expression derived from the work to assemble the charges. To derive this, consider building the charge distribution incrementally by bringing in charge elements dq from , where the potential is zero. The work done to add dq at a point where the existing potential is V is V \, dq, so the total is U = \frac{1}{2} \int \rho V \, dV, with the factor of \frac{1}{2} accounting for double-counting pairwise interactions (each pair contributes once). Using the relation \nabla \cdot \mathbf{E} = \rho / \epsilon_0 and , this simplifies to the field integral form above, confirming their equivalence and highlighting that the is stored in . A practical example is a parallel-plate capacitor with plate area A, separation d, charge Q, capacitance C = \epsilon_0 A / d, and potential difference V = Q / C. The stored energy is U = \frac{1}{2} C V^2 = \frac{1}{2} Q V, and since the uniform field between the plates is E = V / d, substituting yields U = \frac{\epsilon_0}{2} E^2 (A d), matching the volume integral of the energy density over the capacitor's volume. This illustrates how field-based expressions simplify calculations for symmetric systems. These formulations underpin in electrostatic interactions, as the work done by or on the field during charge rearrangements equals changes in the stored energy, ensuring no net energy loss in isolated systems.

Electric Displacement Field

The , denoted as D, is a introduced to describe the electric field in the presence of , particularly dielectrics, by accounting for the effects of . It is defined as D = ε₀ E + P, where E is the electric field, P is the (representing the per unit volume induced in the material), and ε₀ is the . In , where there is no (P = 0), this simplifies to D = ε₀ E. A key property of the displacement field arises from in its differential form: ∇ · D = ρ_free, where ρ_free is the free . This form isolates the influence of free charges (those that can move, such as conduction electrons) from bound charges (those fixed in the material's molecular structure), unlike the standard for E (∇ · E = ρ_total / ε₀), which includes both. The integral form is ∮ D · dA = Q_free,enc, allowing the use of to compute D directly from enclosed free charge in Gaussian surfaces. At interfaces between media, boundary conditions for D ensure continuity of the normal component: the difference in the normal components of D across the boundary equals the free surface charge density, D_{above,⊥} - D_{below,⊥} = σ_free. Thus, if no free surface charge is present (σ_free = 0), the normal component of D is continuous. The tangential component involves polarization effects but is generally discontinuous unless the tangential polarizations match. The displacement field simplifies calculations in dielectrics by focusing solely on free charges, making it particularly useful for problems involving capacitors filled with insulating materials or waveguides. For example, in linear isotropic media—where polarization is directly proportional to the field and uniform in all directions—D = ε E, with ε = ε₀ ε_r, the material's permittivity and ε_r its relative permittivity (dielectric constant). This relation highlights how materials enhance the effective field response compared to vacuum. In materials, the energy density associated with the electric field generalizes to (1/2) D · E.

Constitutive Relations in Materials

In linear dielectrics, the electric displacement field \mathbf{D} is related to the electric field \mathbf{E} by the constitutive relation \mathbf{D} = \epsilon \mathbf{E}, where \epsilon is the of the material. This permittivity is expressed as \epsilon = \epsilon_0 (1 + \chi_e), with \epsilon_0 denoting the and \chi_e the , which quantifies the material's response to the applied field through induced polarization from bound charges. Such relations hold for isotropic materials under weak fields, where the response is proportional and independent of field direction. For anisotropic materials, like crystals without cubic symmetry, the relation becomes tensorial: D_i = \epsilon_{ij} E_j, where \epsilon_{ij} is the permittivity tensor, reflecting direction-dependent polarization. This form accounts for varying dielectric responses along principal axes, as seen in materials such as calcite or quartz. In nonlinear dielectrics, the relation includes higher-order terms, such as \mathbf{D} = \epsilon \mathbf{E} + \boldsymbol{\chi}^{(2)} : \mathbf{E}\mathbf{E} + \cdots, leading to effects like second-harmonic generation or electro-optic modulation where refractive index changes with field strength. Ferroelectric materials exhibit pronounced nonlinearity, characterized by hysteresis in the polarization-electric field loop, enabling spontaneous polarization and switchable domains, as in barium titanate. Conductors represent an extreme limit where the permittivity approaches (\epsilon \to \infty), resulting in zero electric field inside the under electrostatic conditions, as charges rearrange to cancel any internal . In fields, the is frequency-dependent and complex, \epsilon(\omega) = \epsilon'(\omega) - i\epsilon''(\omega), capturing where the real part \epsilon' describes and the imaginary part \epsilon'' accounts for losses, arising from resonant responses of bound charges. This frequency variation explains phenomena like dielectric relaxation in polymers or in .

Relativistic Effects

Fields from Moving Point Charges

The electric field produced by a point charge q moving with constant velocity \vec{v} is modified by relativistic effects compared to the static Coulomb field. In the lab frame, where the charge's position at the observation time is taken as the origin, the electric field \vec{E} at a point with position vector \vec{r} (of magnitude r) is given by \vec{E} = \frac{1}{4\pi \epsilon_0} \frac{q (1 - \beta^2) \vec{r} }{ r^3 (1 - \beta^2 \sin^2 \theta)^{3/2} }, where \beta = v/c, c is the , and \theta is the angle between \vec{v} and \vec{r}. This expression arises from transforming the Coulomb field from the charge's instantaneous using the for electromagnetic fields. For low speeds (\beta \ll 1), the factor (1 - \beta^2) \approx 1 and (1 - \beta^2 \sin^2 \theta)^{3/2} \approx 1, reducing the field to the familiar form \vec{E} = \frac{1}{4\pi \epsilon_0} \frac{q \vec{r}}{r^3}. The velocity dependence introduces anisotropy: the field strength is enhanced in the direction perpendicular to \vec{v} by a factor approaching \gamma = (1 - \beta^2)^{-1/2} as \beta \to 1, while it diminishes along the direction of motion. This results in field line contraction, or "bunching," where the lines of \vec{E} are compressed into a disk-like pattern transverse to the velocity, becoming increasingly flattened as the speed increases. For instance, a charge moving at v = 0.99c (\beta = 0.99, \gamma \approx 7) exhibits field lines that are significantly denser in the equatorial plane relative to the polar direction, illustrating the relativistic compression. For arbitrary motion, the fields are more generally described by the Liénard–Wiechert potentials, which incorporate retardation from the charge's past position and acceleration terms; however, for uniform motion, these simplify to the steady-state form above without radiation contributions. Additionally, the moving charge generates a magnetic field \vec{B} = \frac{1}{c^2} \vec{v} \times \vec{E}, which in the non-relativistic limit approximates \vec{B} = \frac{\mu_0}{4\pi} \frac{q \vec{v} \times \hat{r}}{r^2}, linking the electric field modification to the emergence of magnetism.

Lorentz Transformation of Fields

In special relativity, the electric and magnetic fields transform between inertial reference frames under Lorentz boosts to ensure the invariance of Maxwell's equations. For a boost along the positive x-direction with velocity v, the components parallel to the boost direction remain unchanged, while the perpendicular components mix electric and magnetic contributions. The transformation equations in SI units are: \begin{align} E_x' &= E_x, \\ E_y' &= \gamma (E_y - v B_z), \\ E_z' &= \gamma (E_z + v B_y), \\ B_x' &= B_x, \\ B_y' &= \gamma \left( B_y + \frac{v}{c^2} E_z \right), \\ B_z' &= \gamma \left( B_z - \frac{v}{c^2} E_y \right), \end{align} where \gamma = \frac{1}{\sqrt{1 - \beta^2}} is the , \beta = v/c, and c is the . These transformations imply that a purely electric field in one frame generally appears as a combination of electric and in another frame moving relative to the first. This mixing demonstrates that magnetic phenomena, such as the around a , can be understood as relativistic effects arising from the transformation of produced by moving charges. Two key Lorentz invariants remain unchanged under such boosts: the scalar \mathbf{E} \cdot \mathbf{B} and the difference E^2 - c^2 B^2, which characterize the intrinsic nature of the regardless of the observer's frame. As an example, consider a point charge at rest in one frame, producing a purely electric Coulomb field with no magnetic component. Applying the to obtain the fields in a frame where the charge moves uniformly along the x-direction yields both electric and magnetic fields, with the magnetic field perpendicular to the and electric field directions.

Propagation of Field Disturbances

In classical electrodynamics, changes in do not propagate instantaneously but at the finite c in , a consequence of that ensures in electromagnetic interactions. This finite propagation speed implies that the electric field at a given point \mathbf{r} and time t depends on the positions and motions of charges at earlier times, specifically the t_r = t - |\mathbf{r} - \mathbf{r}'|/c, where \mathbf{r}' is the position of the source charge. The accounts for the travel time of the electromagnetic disturbance from the source to the observation point, preventing any influence from future events and upholding that information cannot travel faster than c. For a single point charge in arbitrary motion, the electric field is described by the Liénard-Wiechert expression, which incorporates both and of the charge evaluated at the . The general form of the electric field \mathbf{E} is \mathbf{E}(\mathbf{r}, t) = \frac{q (1 - \beta^2) (\mathbf{n} - \boldsymbol{\beta}) }{4\pi \epsilon_0 (1 - \mathbf{n} \cdot \boldsymbol{\beta})^3 R^2} \bigg|_{\rm ret} + \frac{q}{4\pi \epsilon_0 c} \frac{ \mathbf{n} \times [(\mathbf{n} - \boldsymbol{\beta}) \times \dot{\boldsymbol{\beta}} ] }{(1 - \mathbf{n} \cdot \boldsymbol{\beta})^3 R} \bigg|_{\rm ret}, where q is the charge, \boldsymbol{\beta} = \mathbf{v}/c is the in units of c, \dot{\boldsymbol{\beta}} is the term, \mathbf{n} is the unit vector from the retarded position to the observation point, R is the distance from the retarded position, and all quantities are evaluated at t_r. The first term represents the velocity-dependent "generalized Coulomb field," which falls off as $1/R^2, while the second term involves and decreases as $1/R, contributing to radiation fields at large distances. This formulation arises from the retarded potentials and satisfies for moving sources. In the static limit, where the charge velocity v \ll c and acceleration a = 0, the Liénard-Wiechert field reduces to the familiar Coulomb field \mathbf{E} = q \mathbf{n} / (4\pi \epsilon_0 R^2), confirming consistency with for slow or stationary charges. The propagation of disturbances at speed c ensures no signaling, as any change in the source—such as a sudden motion—creates a that expands spherically at c, leaving the field unchanged beyond this front until the disturbance arrives. For instance, consider a charge moving at constant velocity $0.8c along the x-axis that suddenly stops at the origin at t = 0; inside a sphere of radius ct, the field reflects the stopped charge as if stationary, while outside, it appears as if the charge continues moving, with the transition shell propagating outward at c and carrying a transverse radiation component.

References

  1. [1]
    Electric field - HyperPhysics
    Electric field is defined as the electric force per unit charge. The direction of the field is taken to be the direction of the force it would exert on a ...
  2. [2]
    18.4 Electric Field: Concept of a Field Revisited - UCF Pressbooks
    A field is a way of conceptualizing and mapping the force that surrounds any object and acts on another object at a distance without apparent physical ...
  3. [3]
    [PDF] Chapter 18 - Physics
    The Electric Field. • The concept of electric fields was invented by Michael. Faraday to describe his model of how charges interact. • Charges interact by ...
  4. [4]
    [PDF] Chapter 2 Electric Fields
    It follows from Coulomb's law that the electric field at point r due to a charge q located at the origin is given by. E = k q r2 r. (2.3) where r is the unit ...
  5. [5]
    Electric field lines
    Electric field lines provide a means to visualize the electric field. Since the electric field is a vector, electric field lines have arrows showing the ...
  6. [6]
    5.6 Electric Field Lines – University Physics Volume 2
    Summary. Electric field diagrams assist in visualizing the field of a source charge. The magnitude of the field is proportional to the field line density.
  7. [7]
    [PDF] Chapter 15.1-2 The Electric Field - General Physics II
    we place in its vicinity. • We define the electric field in terms of the electric force that acts on a “test charge”: =.
  8. [8]
    [PDF] A Brief History of The Development of Classical Electrodynamics
    1867: Ludwig Lorenz develops an electromagnetic theory of light in which the scalar and vector potentials, in retarded form, are the starting point. He shows ...
  9. [9]
    V. Experimental researches in electricity - Journals
    (2015) The birth of the electric machines: a commentary on Faraday (1832) 'Experimental researches in electricity', Philosophical Transactions of the Royal ...
  10. [10]
    VIII. A dynamical theory of the electromagnetic field - Journals
    ... publication date: 14-Nov-2025. Liu H and Du B Effects of Magnetic ... Bork A (2024) Maxwell and the development of electromagnetic theory, American Journal ...
  11. [11]
    [PDF] ON THE ELECTRODYNAMICS OF MOVING BODIES - Fourmilab
    This edition of Einstein's On the Electrodynamics of Moving Bodies is based on the English translation of his original 1905 German-language paper. (published as ...Missing: magnetism | Show results with:magnetism
  12. [12]
    [PDF] Chapter 2 Coulomb's Law - MIT
    Coulomb's law applies to any pair of point charges. When more than two ... the equation above implies that the electric field at the center of a non ...
  13. [13]
    18.3 Coulomb's Law – College Physics - UCF Pressbooks
    To compare the two forces, we first compute the electrostatic force using Coulomb's law, F = k | q 1 q 2 | r 2 . We then calculate the gravitational force ...
  14. [14]
    vacuum electric permittivity - CODATA Value
    vacuum electric permittivity $\varepsilon_0$ ; Numerical value, 8.854 187 8188 x 10-12 F m ; Standard uncertainty, 0.000 000 0014 x 10-12 F m ...
  15. [15]
    [PDF] Electric field due to point charge - UMD Physics
    Lecture 18. • Electric field due to point charge. • Chapter 27 (The Electric Field) due to configuration of (source) charges (today).
  16. [16]
    [PDF] Electric Field due to a point charge - Physics Courses
    OVERVIEW: Gauss' Law: relates electric fields and the charges from which they emanate. Technique for calculating electric field for a given distribution of ...
  17. [17]
    [PDF] Spherical Coordinates
    We have a point charge at the origin. ρ is zero everywhere except there, where it is infinite. That's the singularity. We know how to deal with it, because the ...
  18. [18]
    [PDF] Mass Renormalization and Radiation Damping for a Charged ...
    Now a point charge involves fields falling off as r-2 where r is ... The energy in the electro- magnetic field has a singularity at the point charge. ... point ...
  19. [19]
    The Feynman Lectures on Physics Vol. II Ch. 4: Electrostatics
    This fact is called “the principle of superposition.” That's all there is to electrostatics. If we combine the Coulomb law and the principle of superposition, ...Missing: development | Show results with:development
  20. [20]
    [PDF] principle of superposition in electrodynamics - ResearchGate
    Notice that this generalized form of the superposition principle for time-dependent e/m fields rests on the linearity of Maxwell's differential equations.
  21. [21]
    Generalization of Coulomb's law to Maxwell's equations using ...
    Jul 1, 1986 · Maxwell's equations are obtained by generalizing the laws of electrostatics, which follow from Coulomb's law and the principle of superposition,<|control11|><|separator|>
  22. [22]
    [PDF] All Electrostatics follows from Coulomb's law and the poinciple of
    the electrostatic Maxwell equations follow from. Coulomb's Law + superposition pple. An alternative form to this is to integrate over all volume. SEE d³r.
  23. [23]
    [PDF] Electric Fields; Dipoles; Continuous Charge
    calculating the electric field generated by a continuous distribution of charge. ... Electric field from a discrete charge distribution: E = ∑ ri. 2. 3. 4πε0 i.Missing: formula | Show results with:formula
  24. [24]
    5.5 Calculating Electric Fields of Charge Distributions
    Electric fields of continuous charge distributions are calculated by dividing charge into infinitesimal pieces, using integrals for line, surface, and volume  ...
  25. [25]
    [PDF] Problem Solving 2: Continuous Charge Distributions
    In order to calculate the electric field created by a continuous charge distribution we must break the charge into a number of small pieces dq, each of ...
  26. [26]
    Electrostatic fields from distributed charges - Ximera
    The electric field from distributed charges is found by dividing the charge into small pieces, treating each as a point charge, and summing the fields.
  27. [27]
    Electric Potential Energy - Richard Fitzpatrick
    When a charged particle is taken from point $A$ to point $B$ in an electric field its electric potential energy increases by the amount specified in Eq. (79).
  28. [28]
    Electric Potential and Electric Field - Richard Fitzpatrick
    The electric field is the negative gradient of electric potential. Electric potential is a scalar field, while the electric field is a vector field.
  29. [29]
    [PDF] Chapter 3 Electric Potential
    the direction of the electric field E leads to a lower electric potential. ... gradient of the electric potential V . Physically, the negative sign implies ...
  30. [30]
    Concept: The curl of the electric field is zero in electrostatics
    Concept: The curl of the electric field is zero in electrostatics (editing) · The curl of the magnetic vector potential is the magnetic field · The negative ...
  31. [31]
    19.3 Electrical Potential Due to a Point Charge - UCF Pressbooks
    Electric potential of a point charge is V = k Q / r . Electric potential is a scalar, and electric field is a vector. Addition of voltages as numbers gives the ...
  32. [32]
    [PDF] Chapter 4 The Electric Potential
    ... formula for the electric field due a nonconducting sheet of charge. From Eq. 3.5, we had: Ez = σ/(2 0), where σ is the charge density of the sheet, which ...
  33. [33]
    7.5 Equipotential Surfaces and Conductors - UCF Pressbooks
    Equipotential lines are always perpendicular to electric field lines. No work is required to move a charge along an equipotential, since Δ V = 0.
  34. [34]
    19.4 Equipotential Lines – College Physics
    Equipotential lines are always perpendicular to electric field lines. No work is required to move a charge along an equipotential.
  35. [35]
    [PDF] 24-1 electric potential - UF Physics Department
    of potential over electric field: It is a lot easier to sum several scalar quantities than to sum several vector quantities whose directions and components must.
  36. [36]
    [PDF] 08. Electric potential and potential energy - DigitalCommons@URI
    Sep 25, 2020 · Here is an application of electric potential to point charges that demonstrates the advantage of working with scalars rather than vectors. There ...
  37. [37]
    [PDF] Chapter 2 The Electric Field
    practical definition of the E field as. E = F. Q where Q is a small charge ... We're done with the physics. First off, we can cancel the constant k in Eq ...
  38. [38]
    6 The Electric Field in Various Circumstances - Feynman Lectures
    The potential, and hence the field, which is its derivative, is proportional to qd, the product of the charge and the separation. This product is defined as the ...
  39. [39]
    Electric field
    ### Visualization of Electric Fields
  40. [40]
    Particle Accelerators and Detectors – University Physics Volume 3
    A particle accelerator is a machine designed to accelerate charged particles. This acceleration is usually achieved with strong electric fields, magnetic ...
  41. [41]
    [PDF] 2 Electric Field Mapping
    Electric fields are mapped by using lines of force, calculating resultant vectors, and plotting equipotential lines, which are perpendicular to lines of force.
  42. [42]
    16.2. A Different Way Out: Einstein's Relativity Paper
    A Different Way Out: Einstein's Relativity Paper#. Einstein had been committed to Maxwell's (and Faraday's) notion of a field from his earliest days and ...
  43. [43]
    29.6 Continuous Charge Distributions - Physics Bootcamp
    Therefore, rather than treat such large collection of charges individually, we model them as distributed continuously with a charge density or volume charge ...
  44. [44]
    The Multipole Expansion - Physics LibreTexts
    Jan 3, 2021 · A multipole expansion is a mathematical series representing a function that depends on angles—usually the two angles on a sphere.Setting up the System · The Expansion · The Monopole Scalar · The Dipole Vector
  45. [45]
    Effect of realistic nuclear charge distributions on isotope shifts and ...
    Nuclear charge distribution models. The nuclear charge distribution can be approximated by an analytical expression such as the Fermi distribution ... nucleus ...
  46. [46]
    Nuclear Charge Distribution Measurements May Solve Outstanding ...
    Jul 7, 2023 · Scientists can measure this distribution by scattering electrons off nuclear targets or by studying energy levels through atomic spectroscopy.The Science · The Impact · Summary<|control11|><|separator|>
  47. [47]
    Maxwell's Equations and Electromagnetic Waves
    Maxwell's four equations describe the electric and magnetic fields arising from distributions of electric charges and currents, and how those fields change in ...
  48. [48]
    17 The Laws of Induction - Feynman Lectures - Caltech
    The general law for the electric field associated with a changing magnetic field is ∇×E=−∂B∂t. We will call this Faraday's law.<|control11|><|separator|>
  49. [49]
    The Production of EM waves
    Accelerating charges produce changing electric and magnetic fields. Changing electric fields produce magnetic fields and changing magnetic fields produce ...
  50. [50]
    [PDF] Lecture 6 Notes, Electromagnetic Theory II
    - Any acceleration of an electric charge creates a non-linearly changing electric field. Therefore, accelerating charges radiate electromagnetic waves. - The ...
  51. [51]
    The displacement current - Richard Fitzpatrick
    Of course, the displacement current is not a current at all. It is, in fact, associated with the generation of magnetic fields by time-varying electric fields.
  52. [52]
    [PDF] Displacement current. Maxwell's equations. - MIT
    Apr 26, 2005 · If we imagine we have a time varying electric field, this shows that it will “source” some kind of magnetic field. That magnetic field will be ...
  53. [53]
    [PDF] Chapter 13 Maxwell's Equations and Electromagnetic Waves
    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, ...
  54. [54]
    [PDF] Magnetic Field in a Time-Dependent Capacitor 1 Problem 2 Solution
    Since the charge on the capacitor plates is time dependent, there must be currents flowing on those plates. What contribution, if any, do these “radial” ...
  55. [55]
    Maxwell's Third Equation: the Faraday's Law - USM Digital Commons
    Jan 2, 2025 · Faraday's Law states that the voltage induced in a closed loop is equal to the rate of change of magnetic flux through that loop.
  56. [56]
    Electromagnetic waves - Richard Fitzpatrick
    Thus, the electric field is perpendicular to the direction of propagation of the wave. ... electromagnetic wave is smaller in magnitude than the electric field by ...
  57. [57]
    [PDF] Section 3: Electromagnetic Waves in Vacuum and Simple Matter
    E. B. = = . (3.25). That is, electromagnetic waves are transverse: the electric and magnetic fields are perpendicular to the direction of propagation. Moreover ...
  58. [58]
    [PDF] Chapter 13 Maxwell's Equations and Electromagnetic Waves - MIT
    Direction of wave propagation: The argument of the sine form in the electric field ... Magnetic field B : The magnetic field B is perpendicular to both E which ...Missing: structure | Show results with:structure
  59. [59]
    [PDF] Electricity and Magnetism Summary of topics/formulae for final exam
    Coulomb's law: F = QE; Electric field, E due to point charge = q. 4π 0. ˆ. ˆr‡. ˆr ... direction n: E(r,t) = E0 cos(k.r−ωt+δ)n and B(r,t) = 1 c. E0 cos(k.r−ωt+δ).
  60. [60]
    [PDF] Electromagnetic Waves - Physics Courses
    Electric field perpendicular to the direction of propagation. Magnetic field perpendicular to direction of propagation and to the Electric field. c. B. E. = The ...
  61. [61]
    Maxwell's Equations: Electromagnetic Waves Predicted and Observed
    Maxwell calculated that electromagnetic waves would propagate at a speed given by the equation. c =1μ0ε0.c =1μ0ε0. size 12{“c “= { {1} over { sqrt {μ rSub ...
  62. [62]
    [PDF] Chapter 33 Electromagnetic waves Masatsugu Sei Suzuki ... - bingweb
    Jul 29, 2019 · This equation is usually expressed as S. = (1/μ0)ExB where S is the Poynting vector, is the permeability of the medium, E is the electric field ...
  63. [63]
    24.4 Energy in Electromagnetic Waves – College Physics
    ... electromagnetic wave, the average intensity I ave is given by. I ave = c ϵ 0 E 0 2 2 ,. where c is the speed of light, ϵ 0 is the permittivity of free space, ...
  64. [64]
    [PDF] Chapter 12 - Polarization - MIT OpenCourseWare
    The general elec- tromagnetic plane wave has two polarization states, corresponding to the two directions that the electric field can point transverse to the ...
  65. [65]
    [PDF] Lecture 14: Polarization
    The magnetic field will be polarized in direction. 90◦ behind the electric field. Remember, we always implicitly want to have the real part of these fields. So, ...
  66. [66]
  67. [67]
    Wave Propagation in Conducting Media - Richard Fitzpatrick
    An electromagnetic wave propagating through a good conductor has markedly different properties to a wave propagating through a conventional dielectric.
  68. [68]
    [PDF] Electromagnetic Waves in Conductors and Dispersive Matter
    The attenuated plane waves (4.26) and (4.27) ... As we saw earlier, the propagation of the plane electromagnetic wave in a dispersive medium is described.<|control11|><|separator|>
  69. [69]
    The Feynman Lectures on Physics Vol. II Ch. 8: Electrostatic Energy
    So the energy density at the distance r from the charge is ϵ0E22=q232π2ϵ0r4. We can take for an element of volume a spherical shell of thickness dr ...
  70. [70]
    Electrostatic energy - Richard Fitzpatrick
    This is the potential energy (ie, the difference between the total energy and the kinetic energy) of a collection of charges.
  71. [71]
    6. Electrostatics I: Fields, Potentials, Energy - Galileo and Einstein
    The mathematical structures of the electric field from a “point” charge and that mapping fluid flow from a “point” source are identical! But we know that for ...
  72. [72]
    [PDF] 3.1 Energy in the field - MIT
    Feb 8, 2005 · Now, using E = 4πσ, we can write this formula in terms of the electric field: dW = E2 8π dV . The quantity E2/8π is the energy density of the ...
  73. [73]
    Energy Density - AstroBaki - CASPER
    Sep 17, 2018 · Deriving Energy Density in an Electric Field Using a Capacitor. Recall that for a parallel-plate capacitor, two plates close together create a ...
  74. [74]
    None
    ### Summary of Electric Displacement Field D from http://physicspages.com/pdf/Electrodynamics/Electric%20displacement.pdf
  75. [75]
    None
    ### Summary of Electric Displacement D from Phys 332 Ch 4 D3.pdf
  76. [76]
    [PDF] Chapter 9: Electromagnetic Waves - MIT OpenCourseWare
    May 9, 2011 · D = εE = εo E + P = εo E + nqd. (9.5.21) where q = -e is the electron charge, d is the mean field-induced displacement of the electrons from ...Missing: ε₀ | Show results with:ε₀
  77. [77]
    [PDF] Chapter 3 Constitutive Relations
    CONSTITUTIVE RELATIONS. 3.1 Linear Materials. The most general linear relationship between D and E can be written as. D(r,t) = ε0. ∞. Z. −∞. 0. Z−∞. ˜ε(r−r′, t ...
  78. [78]
    6.4 Constitutive Laws of Polarization - MIT
    All isotropic media behave as linear media and obey (2) if the applied E field is sufficiently small. As long as E is small enough, any continuous function P(E ) ...
  79. [79]
    [PDF] Constitutive Relations | EMPossible
    The dielectric response of a material arises due to the electric field displacing charges. Due to structural and bonding effects at the atomic scale, ...
  80. [80]
    [PDF] Nonlinear & Anisotropic Materials | EMPossible
    Jan 28, 2020 · Tensor Relations for Anisotropic Materials. Slide 12. This is most often treated through a dielectric tensor. ( ). ( ) r e ε ω χ ω. = +.
  81. [81]
    [PDF] Nonlinear Dielectric Materials - ET JAYNES,† SENIOR MEMBER, IRE
    By far the most important nonlinear dielectrics, however, are the ferroelectric crystals or ceramics. FERROELECTRICS. From a phenomenological point of view, ...
  82. [82]
    Quantum critical electro-optic and piezo-electric nonlinearities
    Oct 23, 2025 · Unfortunately, ferroelectric nonlinearities, plasma dispersion and thermo-optic tuning mechanisms usually decrease at cryogenic temperatures (6 ...
  83. [83]
    The conductor limit of dielectrics - IOPscience
    In the limit that the relative dielectric constant of a body in an electrostatic field approaches infinity, the influence of that body on the field becomes ...Missing: permittivity | Show results with:permittivity<|separator|>
  84. [84]
    [PDF] Lorentz Oscillator Model of the Dielectric Function
    And more than that, it also predicts that the dielectric function (and hence index of refraction) will have frequency dependence: there is dispersion. In ...
  85. [85]
    [PDF] The electromagnetic fields of a uniformly moving charge
    The electromagnetic fields of a moving charge are given by ~E = γq r′ 3 xk − vt + x⊥ and ~B = γq cr′ 3 v × x⊥.Missing: formula | Show results with:formula
  86. [86]
    Fields due to a moving charge - Richard Fitzpatrick
    The fields generated by a uniformly moving charge can be calculated from the expressions (1525) and (1526) for the potentials, it is simpler to calculate them ...
  87. [87]
    26: Lorentz Transformations of the Fields - Feynman Lectures
    For a slow-moving charge (v≪c), we can take for E the Coulomb field; then B=q4πϵ0c2v×rr3. This formula corresponds exactly to equations for the magnetic field ...
  88. [88]
    [PDF] ON THE ELECTRODYNAMICS OF MOVING BODIES
    For if the magnet is in motion and the conductor at rest, there arises in the neighbour- hood of the magnet an electric field with a certain definite energy, ...Missing: magnetism | Show results with:magnetism
  89. [89]
    [PDF] 5. Electromagnetism and Relativity - DAMTP
    Under Lorentz transformations, electric and magnetic fields will transform into each other. ... The Maxwell equations are not invariant under Lorentz ...
  90. [90]
    21 Solutions of Maxwell's Equations with Currents and Charges
    The second term says that there is a “correction” to the retarded Coulomb field which is the rate of change of the retarded Coulomb field multiplied by r′/c, ...Missing: disturbances | Show results with:disturbances
  91. [91]
  92. [92]
    [PDF] Classical Electrodynamics - Duke Physics
    With relativity in hand, relativistic electrodynamics is developed, including the properties of radiation emitted from a point charge as it is accelerated.
  93. [93]
    [PDF] Accelerating charges, radiation, and electromagnetic waves in solids
    Dec 2, 2011 · In this case, our charge is moving with velocity v = 0.8c until reaching the origin at t = 0, and which point it suddenly stops. Outside a ...