Fact-checked by Grok 2 weeks ago

Electric current

Electric current is the flow of electric charge through a conducting medium, defined as the rate at which charge passes through a given point or surface. In mathematical terms, the average current I is given by I = \Delta q / \Delta t, where \Delta q is the change in charge over time interval \Delta t, while the instantaneous current is I = dq/dt. The SI unit of electric current is the ampere (A), equivalent to one coulomb of charge per second (1 A = 1 C/s), named after the French physicist André-Marie Ampère. Electric current arises from the movement of charged particles, such as electrons in metallic conductors, ions in electrolytes, or holes in semiconductors, driven by an electric field or potential difference. In metals, these charge carriers acquire a drift velocity, typically on the order of millimeters to centimeters per second, despite their high thermal speeds, resulting in a net flow that constitutes the current. Conventionally, current direction is defined as the flow of positive charge from higher to lower potential, a historical convention established by Benjamin Franklin, though in most conductors like wires, actual electron movement is opposite to this direction. The magnitude of current depends on factors including the number density of charge carriers n, their charge q, the cross-sectional area A of the conductor, and drift velocity v, as expressed by I = n q A v. Currents are classified into direct current (DC), which flows steadily in one direction, and alternating current (AC), which periodically reverses direction, with AC being predominant in power distribution due to efficient transmission over long distances. There are two primary types: conduction current, involving physical movement of charges through a medium, and displacement current, an effective current due to changing electric fields in capacitors or dielectrics, as conceptualized by James Clerk Maxwell to complete electromagnetic theory. Electric current is fundamental to electrical engineering and physics, powering devices from household appliances to industrial machinery, while excessive currents can cause heating via Joule's law or hazards like short circuits. In specialized contexts, such as superconductors, currents can flow without resistance, enabling applications in magnets and quantum devices.

Basic Concepts

Definition

Electric current is the rate of flow of electric charge through a point or across a surface in an electrical conductor or medium. It is mathematically defined as the time derivative of the charge, expressed by the equation I = \frac{dQ}{dt}, where I represents the current, Q the electric charge, and t the time. This flow typically involves the movement of charged particles, such as electrons or ions, under the influence of an electric field. Electric current possesses both magnitude and direction, though it is fundamentally a scalar quantity that does not obey the parallelogram law of vector addition; the direction is conventionally specified along the path of the conductor. The magnitude of electric current is measured in amperes (A), the SI base unit, defined as one coulomb of charge passing through a point per second. In standard convention, electric current is taken as the directed flow of positive charge carriers from regions of higher to lower electric potential (positive to negative terminal). However, in metallic conductors, the actual charge carriers are negatively charged electrons, which move in the opposite direction—from negative to positive—resulting in the net current direction aligning with the conventional definition. The term "electric current" was coined in the early 19th century by French physicist André-Marie Ampère, who introduced it by analogy to the flow of water, formalizing the concept during his foundational work on electrodynamics following Hans Christian Ørsted's 1820 discovery of electromagnetism.

Symbol and Conventions

The conventional symbol for electric current, denoting its intensity or magnitude, is the uppercase italic letter I in steady-state, mean, or root-mean-square (RMS) contexts. For instantaneous or time-varying currents, such as those in alternating current systems, the lowercase italic i(t) is used to represent the value as a function of time. These notations follow international standards for electrical engineering symbols. The International System of Units (SI) designates the ampere, symbolized as A, as the base unit for electric current, equivalent to one coulomb of charge flowing per second ($1\ \mathrm{A} = 1\ \mathrm{C/s}). Since the 2019 redefinition of the SI, the ampere is precisely defined by fixing the elementary charge e to exactly $1.602176634 \times 10^{-19} coulombs, where the coulomb is expressed as \mathrm{A \cdot s}; thus, $1\ \mathrm{A} corresponds to the electric current produced by a flow of exactly $1 / (1.602176634 \times 10^{-19}) elementary charges per second. This definition ensures the ampere's constancy independent of experimental artifacts. In circuit conventions, positive current direction is defined as the flow from higher electric potential (positive terminal) to lower potential (negative terminal), aligning with the motion of hypothetical positive charges; actual electron flow is opposite this direction. This passive sign convention facilitates consistent analysis in electrical engineering. In more advanced electromagnetic contexts, current is often treated through current density, a vector quantity denoted by boldface \mathbf{J} or \vec{J} with an arrow, to specify both the magnitude of charge flow per unit area and its directional flow.

Reference Direction

In electrical circuits, the reference direction for current flow is an arbitrary convention that aligns with the notion of conventional current, defined as the movement of positive charges from the positive terminal to the negative terminal of a voltage source. This choice facilitates consistent analysis without regard to the actual motion of electrons, which flow in the opposite direction. The convention originated with Benjamin Franklin's 1747 experiments, where he labeled excess electrical fluid as "positive" and deficiency as "negative," inadvertently establishing the positive-to-negative flow as standard despite later discoveries about charge carriers. To apply this in practice, the passive sign convention governs the assignment of positive directions for components like resistors, capacitors, and voltage sources. Under this rule, current is deemed positive when it enters the component at the positive voltage terminal and exits at the negative terminal, ensuring that power calculations yield positive values for energy dissipation in passive elements. For instance, in a resistor, if the reference current flows into the positive-marked terminal, the voltage drop aligns with Ohm's law as a positive quantity. This convention simplifies power determination, where power P = VI is positive for absorbed energy and negative for supplied energy. The reference direction plays a critical role in Kirchhoff's laws, as inconsistent assignments can introduce sign errors that invalidate circuit solutions. In Kirchhoff's current law, currents are summed at a node with entering flows as positive; an erroneous direction for a branch current might flip its sign, leading to an unbalanced equation and incorrect node voltages. Similarly, in Kirchhoff's voltage law, traversing a loop requires adding voltage rises and subtracting drops based on the assumed current direction, which determines polarity across elements—if the reference current opposes the actual flow, voltage terms invert, potentially causing the loop sum to deviate from zero unless corrected. For example, in a simple series loop with a battery and resistor, assuming current opposite to the battery's polarity would assign a negative voltage to the source, requiring adjustment to avoid computational errors in finding branch currents.

Types of Current

Direct Current

Direct current (DC) is the unidirectional flow of electric charge carriers, such as electrons, through a conductor or medium. In ideal cases, this flow maintains a constant magnitude and direction over time, resulting in a steady current without periodic variations. Common sources of direct current include batteries, which generate DC through electrochemical reactions; solar cells, which produce DC via the photovoltaic effect; and rectifiers, which convert alternating current to DC. In these sources, the current can be represented mathematically as I = I_0, where I_0 is a constant value independent of time. Direct current finds widespread applications in electronics, where it powers devices requiring stable voltage, such as integrated circuits and light-emitting diodes (LEDs); in electrochemistry, for processes like electrolysis and battery charging that depend on consistent polarity; and in high-voltage direct current (HVDC) transmission lines, which enable efficient long-distance power transfer between asynchronous grids. Compared to alternating current, direct current offers advantages such as the absence of reactive power, which eliminates the need for power factor correction and reduces associated losses in transmission; additionally, it is particularly suitable for semiconductor-based devices like LEDs, which operate efficiently with unidirectional flow due to their diode nature.

Alternating Current

Alternating current (AC) is an electric current that periodically reverses direction, unlike direct current which flows steadily in one direction. This reversal occurs in a cyclical manner, most commonly following a sinusoidal waveform expressed as i(t) = I_m \sin(\omega t + \phi), where I_m is the peak current, \omega is the angular frequency, t is time, and \phi is the phase angle. The sinusoidal form arises naturally from the rotational motion in AC generators, making it the standard for most electrical power systems. Key parameters of AC include the peak value I_m, which represents the maximum amplitude of the current waveform; the root mean square (RMS) value I_{rms} = I_m / \sqrt{2}, used to quantify the effective current for power calculations as it equates to the DC value producing the same heating effect; the frequency f in hertz (Hz), indicating cycles per second; and the period T = 1/f, the time for one complete cycle. The angular frequency relates to the linear frequency by \omega = 2\pi f. AC is primarily generated by alternators or AC generators in power plants, which convert mechanical energy from sources like steam turbines, water, or wind into electrical energy through electromagnetic induction. These generators feed into interconnected power grids that distribute electricity over vast areas. Standard frequencies are 60 Hz in North America and 50 Hz in much of Europe and Asia, chosen for compatibility with equipment and to minimize losses in transmission lines. AC finds widespread applications in power distribution systems, electric motors, and transformers due to its ease of voltage transformation. In power grids, AC enables efficient long-distance transmission by stepping up voltages to hundreds of kilovolts, reducing resistive losses proportional to the square of the current, and then stepping down for safe consumer use. This advantage, pivotal in the historical adoption of AC over direct current, supports its dominance in modern electrical infrastructure for residential, industrial, and commercial needs.

Ohm's Law

Statement and Applications

Ohm's law states that the electric current I through a conductor between two points is directly proportional to the voltage V across the two points, with the proportionality constant being the inverse of the conductor's resistance R, provided the temperature remains constant. This relationship is expressed by the equation V = IR, where V is in volts, I is in amperes, and R is in ohms. The law was empirically derived by German physicist Georg Simon Ohm through experiments on metallic conductors, as detailed in his 1827 publication Die galvanische Kette, mathematisch bearbeitet. However, Ohm's law applies only to ohmic devices and materials, where the current-voltage relationship is linear; non-ohmic devices, such as diodes, exhibit nonlinear behavior and do not obey this proportionality. In practical applications, Ohm's law is fundamental for analyzing simple circuits, enabling the calculation of current in series configurations—where total resistance is the sum of individual resistances—and parallel configurations—where total resistance is the reciprocal of the sum of reciprocals of individual resistances. It also facilitates the determination of power dissipation in resistors via the formula P = I^2 R, which quantifies the rate of energy conversion to heat. The inverse of resistance is conductance G, defined as G = 1/R, measured in siemens, which describes the ease with which current flows through a component and allows the current to be expressed as I = VG.

Current Density Formulation

In materials where the electric current flows uniformly across a cross-sectional area A, the current density \vec{J} is defined as a vector quantity whose magnitude is the total current I divided by A, and whose direction aligns with the flow of positive charge carriers. For such uniform cases, \vec{J} = \frac{I}{A} \hat{n}, where \hat{n} is the unit normal vector perpendicular to the surface. At the microscopic level, current density can also be expressed as \vec{J} = \rho \vec{v_d}, where \rho is the charge carrier density and \vec{v_d} is the average drift velocity of the carriers. The relationship between current density and the local electric field \vec{E} is given by the microscopic form of Ohm's law: \vec{J} = \sigma \vec{E}, where \sigma is the material's electrical conductivity, a scalar measure of how easily charge carriers move under an applied field. Conductivity \sigma and resistivity \rho (the inverse of conductivity) are related by \rho = 1/\sigma, with resistivity quantifying the material's opposition to current flow. This formulation extends the macroscopic Ohm's law V = IR to local, vector-based descriptions within the material. In cases of non-uniform current distribution, the total current I through a surface is obtained by integrating the current density over the area: I = \int \vec{J} \cdot d\vec{A}. This integral accounts for variations in \vec{J} across the cross-section, ensuring the total charge flow is accurately captured. Electrical conductivity \sigma depends on temperature, with distinct behaviors in different materials. In metals, \sigma typically decreases as temperature increases due to enhanced lattice vibrations scattering charge carriers more effectively. For example, copper's conductivity drops by about 0.4% per Kelvin rise near room temperature, impacting applications like power transmission where cooling is used to maintain efficiency.

Measurement

Instruments and Methods

Ammeters are essential instruments for directly measuring electric current in a circuit, typically connected in series to ensure the full current passes through the device without significantly altering the circuit's behavior due to their low internal resistance. Analog ammeters, such as those based on the permanent magnet moving coil (PMMC) principle, operate by passing current through a coil suspended in a magnetic field, where the resulting torque causes a pointer to deflect proportionally to the current strength on a calibrated scale. This design provides a continuous analog readout and is particularly useful for observing trends in current over time. Digital ammeters, often integrated into multimeters, convert the current to a voltage drop across a shunt resistor and use an analog-to-digital converter to display the value numerically, offering higher precision and resolution for quantitative measurements. For non-invasive measurements, clamp meters encircle a conductor without requiring circuit interruption, detecting the magnetic field generated by the current flow to infer its magnitude. In AC applications, these devices employ a current transformer principle, where the conductor acts as a single-turn primary winding, inducing a proportional voltage in the meter's core via Faraday's law of induction. For DC currents, clamp meters utilize Hall effect sensors within the jaws to measure the steady magnetic field, enabling accurate readings up to several thousand amperes in industrial settings. Hall effect sensors provide a versatile method for current measurement, particularly in applications requiring vectorial information about the current direction and magnitude, by exploiting the Hall voltage generated perpendicular to both the current-induced magnetic field and the sensor's plane. These sensors, often integrated into integrated circuits, detect the Lorentz force on charge carriers in a semiconductor material, producing an output voltage linearly proportional to the magnetic flux density from the current, which can be calibrated to yield current values with bidirectional sensitivity. Precision calibration of current-measuring instruments relies on quantum standards derived from fundamental physical effects to ensure traceability to the international ampere definition. The Josephson effect, observed in superconductor junctions, generates stable voltage steps at fixed frequencies, serving as a primary voltage standard that, when combined with resistance measurements, enables accurate current calibration via Ohm's law. Complementarily, the quantum Hall effect in two-dimensional electron systems under strong magnetic fields and low temperatures produces quantized resistance plateaus, providing an invariant resistance standard essential for deriving precise current values independent of material properties. Recent developments include primary quantum current standards based on the Josephson and quantum Hall effects, capable of generating sizeable quantized currents with relative uncertainties below 10^{-10}, offering direct realization of the ampere. These quantum-based methods achieve uncertainties below parts per million, forming the basis for national metrology institutes' calibration hierarchies.

Units and Standards

The ampere (A) is the SI base unit of electric current, defined as the flow of charge per unit time. In the 2019 revision of the SI, the ampere was redefined by fixing the elementary charge e to the exact value $1.602176634 \times 10^{-19} coulombs (C), such that one ampere corresponds to the passage of exactly $1/e elementary charges per second. This quantum-based definition replaced the previous mechanical one, which relied on the force between two parallel current-carrying wires, ensuring greater precision and universality tied to fundamental constants. Submultiples and multiples of the ampere use standard SI prefixes to express currents across various scales. Common units include the milliampere (mA, $10^{-3} A) for low-power electronics, microampere (μA, $10^{-6} A) for sensors and biomedical applications, and picoampere (pA, $10^{-12} A) for leakage currents in semiconductors; larger units like the kiloampere (kA, $10^{3} A) apply to high-power transmission lines. Practical measurements span from picoamperes in nanoscale devices to kiloamperes in industrial power systems, accommodating the wide range of phenomena involving electric current. The International Bureau of Weights and Measures (BIPM) plays a central role in maintaining SI standards for electric current, coordinating global comparisons among national metrology institutes to ensure traceability and consistency. Traceability to the ampere is achieved through realizations involving the ohm (derived from the quantum Hall effect) and volt (from the Josephson effect), applying Ohm's law to calibrate current standards with uncertainties below parts per million. Historically, the ampere evolved from early electrochemical definitions at the 1893 International Electrical Congress, where the international ampere was based on the deposition of silver in electrolysis. By 1948, the 9th General Conference on Weights and Measures (CGPM) adopted an absolute definition using the magnetic force between parallel conductors, fixing the permeability of vacuum \mu_0 = 4\pi \times 10^{-7} H/m. The 2019 redefinition by the 26th CGPM shifted to a constant-based system, incorporating the elementary charge to align electrical metrology with quantum mechanics and eliminate reliance on physical artifacts.

Physical Effects

Joule Heating

Joule heating, also known as ohmic heating, refers to the process by which electrical energy is converted into thermal energy when an electric current passes through a conductor with resistance. This phenomenon was first quantitatively described by James Prescott Joule in his 1840 paper, where he experimentally demonstrated that the rate of heat production is proportional to the square of the current and the resistance of the conductor. The power dissipated as heat, denoted as P, is given by the equation P = I^2 R, where I is the current and R is the resistance. This energy dissipation leads to a temperature rise \Delta T in the conductor, which is proportional to the total energy released over time, as the heat capacity of the material determines how much the temperature increases for a given amount of thermal energy input. At the microscopic level, Joule heating arises from collisions between charge carriers, such as electrons in metals, and the ions of the conductor's lattice. As the electric field accelerates the charge carriers, they gain kinetic energy, but frequent inelastic collisions with lattice vibrations (phonons) transfer this energy to the lattice, manifesting as thermal agitation and increased temperature. This resistive dissipation is a fundamental consequence of Ohm's law, where the voltage drop across the resistor drives the current against material imperfections. Joule heating finds practical applications in devices designed to exploit thermal effects. Electric heaters, such as space and baseboard units, rely on high-resistance elements to generate warmth efficiently from electrical power. Similarly, incandescent light bulbs use a tungsten filament where the heating causes incandescence, producing visible light as a byproduct, though much of the energy is lost as heat. Fuses protect circuits by intentionally incorporating a low-melting-point material that overheats and melts when excessive current flows, interrupting the circuit to prevent damage. However, unintended Joule heating poses risks, such as wire overheating that can lead to insulation degradation or fires in overloaded electrical systems. In superconductors, where electrical resistance drops to zero below a critical temperature, Joule heating is effectively minimized, allowing current to flow without energy loss as heat. This property enables efficient power transmission and high-field magnets, though any disruption to the superconducting state can rapidly induce heating.

Magnetic Fields and Electromagnets

Steady electric currents generate magnetic fields, a phenomenon first observed experimentally and later formalized through fundamental laws of electromagnetism. The Biot-Savart law provides the mathematical description for the magnetic field produced by a small current element, stating that the infinitesimal magnetic field d\mathbf{B} at a point due to a current element I d\mathbf{l} is given by d\mathbf{B} = \frac{\mu_0}{4\pi} \frac{I d\mathbf{l} \times \hat{\mathbf{r}}}{r^2}, where \mu_0 is the permeability of free space, \mathbf{r} is the vector from the current element to the observation point, r its magnitude, and \hat{\mathbf{r}} the unit vector in that direction. This law, derived from experiments by Jean-Baptiste Biot and Félix Savart in 1820, applies to arbitrary steady current distributions and serves as the basis for calculating fields from wires, loops, and coils. For configurations with high symmetry, such as straight wires or solenoids, Ampère's circuital law offers a more efficient approach. This law relates the line integral of the magnetic field \mathbf{B} around a closed loop to the total current I_{\text{enc}} enclosed by that loop: \oint \mathbf{B} \cdot d\mathbf{l} = \mu_0 I_{\text{enc}}. Formulated by André-Marie Ampère in 1826 based on his electrodynamic experiments, it holds for steady currents where the fields are static and divergence-free. Applying Ampère's law to an infinite straight wire yields a circumferential magnetic field B = \frac{\mu_0 I}{2\pi r} at distance r from the wire, illustrating how current intensity and proximity determine field strength. Electromagnets exploit these principles to create controllable magnetic fields, typically using solenoids—long coils of wire wound in a helical fashion. Inside an ideal long solenoid with n turns per unit length carrying current I, Ampère's law gives a uniform magnetic field B = \mu_0 n I directed along the axis, independent of position within the core. This design allows the field to be turned on or off by controlling the current, enabling applications such as electromagnetic relays, where a solenoid actuates a switch to control higher-power circuits remotely. In medical imaging, superconducting solenoids in MRI machines generate fields up to 3 T or more, aligning atomic nuclei for precise tissue visualization without permanent magnets. The interaction between currents and magnetic fields also produces mechanical forces, essential for electromagnet operation. A straight wire of length \mathbf{L} carrying current I in a uniform magnetic field \mathbf{B} experiences a force \mathbf{F} = I \mathbf{L} \times \mathbf{B}, perpendicular to both the current direction and the field. This Lorentz force on the moving charges within the wire causes deflection or motion, as seen in galvanometers or the armature movement in solenoid-based devices, converting electrical energy into mechanical work efficiently.

Electromagnetic Induction

Electromagnetic induction refers to the process by which a changing magnetic field produces an electromotive force (EMF) in a conductor, leading to the generation of electric current when the conductor forms a closed circuit. This phenomenon was discovered by Michael Faraday in 1831 through experiments involving coils and magnets, where he observed that moving a magnet near a coil or changing the current in one coil induced a current in a nearby coil. Faraday's law quantifies this effect, stating that the magnitude of the induced EMF \epsilon in a closed loop equals the absolute value of the time rate of change of the magnetic flux \Phi_B through the loop: \epsilon = -\frac{d\Phi_B}{dt} where \Phi_B = \int \mathbf{B} \cdot d\mathbf{A} is the magnetic flux, \mathbf{B} is the magnetic field, and d\mathbf{A} is the differential area vector of the surface bounded by the loop. The negative sign indicates the direction of the induced EMF, as determined by Lenz's law, formulated by Heinrich Lenz in 1834, which asserts that the induced current creates a magnetic field opposing the change in flux that produced it. This opposition ensures conservation of energy, as the induced current requires work against the opposing force to maintain the flux change. A key application of electromagnetic induction is in electric generators, which convert mechanical energy into electrical energy by rotating conductors in magnetic fields to produce continuous EMF. Faraday's disk, or homopolar generator, constructed in 1831, exemplifies this: a copper disk rotating perpendicular to a uniform magnetic field induces a radial EMF between the center and edge, generating direct current when connected to brushes. Transformers, another vital application, rely on mutual induction between two coils wound around a common core to step up or step down alternating current voltages without significant energy loss, enabling efficient power transmission over long distances. Self-inductance arises when a changing current in a single coil induces an EMF within itself due to the varying magnetic flux it produces. Discovered by Joseph Henry in 1832, self-inductance L is defined by the relation V = -L \frac{di}{dt}, where V is the induced voltage and i is the current; L depends on the coil's geometry, such as the number of turns and core material, and opposes changes in current, as quantified in units of henries. Mutual inductance extends this to two coupled circuits, where a changing current in one induces EMF in the other; the mutual inductance M is given by \epsilon_2 = -M \frac{dI_1}{dt}, with M determined by the coils' relative positions and orientations, and it forms the basis for devices like transformers where energy transfer efficiency reaches over 99% in ideal cases.

Generation of Radio Waves

Time-varying electric currents, which involve accelerating charges, generate electromagnetic waves according to Maxwell's equations. These equations predict that changing electric fields produce magnetic fields and vice versa, leading to self-propagating waves when charges accelerate, such as in oscillating currents. In particular, an alternating current in a conductor creates oscillating electric and magnetic fields that detach from the source and propagate as transverse electromagnetic waves, commonly known as radio waves in the low-frequency regime. Antenna theory explains how these time-varying currents are harnessed to efficiently radiate radio waves. A basic example is the dipole antenna, consisting of two conductive elements driven by an oscillating current, which produces a characteristic radiation pattern. The pattern for a short dipole (length much less than the wavelength) is doughnut-shaped, with maximum intensity in the plane perpendicular to the antenna axis and nulls along the axis, varying as \sin^2 \theta where \theta is the angle from the axis. Antenna efficiency, defined as the ratio of radiated power to input power, is low for short dipoles due to ohmic losses dominating over radiation, but improves for resonant lengths like half-wavelength dipoles where the radiation resistance is about 73 ohms. The wavelength \lambda of the emitted radio waves is given by \lambda = c / f, where c is the speed of light and f is the frequency of the current oscillation, ensuring efficient radiation when the antenna dimensions match \lambda. These principles underpin key applications in radio transmission and wireless communication. Radio waves generated by antennas carry information over long distances, enabling broadcasting and data transfer. Amplitude modulation (AM) varies the amplitude of the carrier wave according to the signal, used in medium-wave radio (540–1600 kHz), while frequency modulation (FM) varies the frequency for higher fidelity in VHF bands (88–108 MHz). For a short dipole antenna of length l carrying peak current I_m at frequency f, the average radiated power is P = \frac{\pi \mu_0 I_m^2 f^2 l^2}{12 c}, where \mu_0 is the permeability of free space; this formula highlights the quadratic dependence on frequency and length, emphasizing the need for resonant designs to achieve practical power levels.

Conduction Mechanisms

In Metals

In metals, electric current is primarily conducted by the movement of free electrons, which form a delocalized "sea" within the metallic lattice, allowing for high electrical conductivity. The classical Drude model, proposed in 1900, describes this conduction as a gas of free electrons that acquire a drift velocity in response to an applied electric field, with their mobility determined by the mean time between collisions. In this model, electrons are treated as classical particles obeying Newton's laws between scattering events, leading to a conductivity given by the product of electron density, charge, and mobility. The high conductivity of metals arises from the abundance of these delocalized conduction electrons, typically on the order of 10^{22} to 10^{23} per cubic centimeter, which can respond readily to electric fields. However, electrical resistivity in metals originates from electron scattering mechanisms, including collisions with lattice impurities and phonons (quantized lattice vibrations). Impurity scattering contributes a temperature-independent residual resistivity, while phonon scattering dominates at higher temperatures, increasing the overall resistivity as thermal vibrations intensify. Resistivity in metals exhibits a nearly linear increase with temperature, approximated by the empirical relation \rho(T) = \rho_0 [1 + \alpha (T - T_0)], where \rho_0 is the resistivity at reference temperature T_0, and \alpha is the positive temperature coefficient of resistivity, typically around $0.003 to $0.006 per kelvin for common metals. This behavior reflects the enhanced phonon scattering at elevated temperatures, which reduces electron mean free path and mobility. Among metals, silver exhibits the highest room-temperature electrical conductivity at approximately $6.30 \times 10^7 S/m, followed closely by copper at $5.96 \times 10^7 S/m, making them ideal for applications like wiring and interconnects due to their low resistivity. In alternating current (AC) scenarios, the skin effect further influences conduction, where high-frequency currents concentrate near the conductor's surface because of induced eddy currents opposing penetration deeper into the material. This effect, prominent above a few kilohertz, increases effective resistance and is quantified by the skin depth \delta = \sqrt{\frac{2}{\omega \mu \sigma}}, where \omega is angular frequency, \mu is permeability, and \sigma is conductivity.

In Electrolytes

In electrolytes, electric current is conducted through the migration of ions in a solution or molten state under an applied electric field. Unlike metallic conduction, which involves free electrons, electrolytic conduction relies on the movement of positively charged cations toward the cathode and negatively charged anions toward the anode, resulting in a net charge transfer. This process occurs in both aqueous solutions and molten salts, where the electrolyte dissociates into free ions capable of mobility. The quantitative relationship governing the amount of substance altered during electrolysis is described by Faraday's laws. The first law states that the mass m of a substance deposited or liberated at an electrode is directly proportional to the quantity of electric charge Q passed through the electrolyte, expressed as m = \frac{Q}{F} \cdot \frac{M}{n}, where F is Faraday's constant (approximately 96,485 C/mol), M is the molar mass, and n is the number of electrons transferred per ion. The second law asserts that the masses of different substances deposited by the same quantity of charge are proportional to their equivalent weights, meaning equal charges produce masses inversely proportional to the charge of the ions involved. These laws, experimentally established in the 1830s, underpin the stoichiometry of electrochemical reactions in electrolytes. The electrical conductivity \sigma of an electrolyte arises from the collective drift of ions and is given by \sigma = n q \mu, where n is the ion density (number of charge carriers per unit volume), q is the charge per ion, and \mu is the ionic mobility (drift velocity per unit electric field strength). In practice, for multi-ion systems like salts, this sums over all species: \sigma = \sum_i n_i |q_i| \mu_i. Mobility \mu depends on factors such as ion size, solvent viscosity, and temperature, leading to conductivities typically ranging from 0.01 to 10 S/m in common solutions, far lower than in metals due to slower ionic diffusion. The current density J in the electrolyte is then J = \sigma E, where E is the electric field. Electrolytes can be aqueous solutions, such as sodium chloride (NaCl) dissolved in water, where Na⁺ and Cl⁻ ions conduct current; at the cathode, water reduction to hydrogen and hydroxide often competes with Na⁺ deposition due to overpotential effects, while Cl⁻ oxidizes to chlorine gas at the anode. In molten salts, like pure NaCl heated above 800°C, conduction occurs via fully dissociated Na⁺ and Cl⁻ ions without solvent interference, enabling direct production of sodium metal at the cathode and chlorine at the anode. These examples illustrate how electrolyte composition influences ion availability and reaction selectivity. Practical applications of electrolytic conduction include batteries, where ions shuttle between electrodes through the electrolyte to store and release energy, as in lithium-ion cells using liquid or solid electrolytes for Li⁺ migration. Electroplating employs controlled ion deposition to coat surfaces, such as applying a thin copper layer from a CuSO₄ solution onto objects for corrosion protection or aesthetics. Electrode reactions in these systems often involve overpotentials—the extra voltage beyond the thermodynamic minimum needed to drive the reaction due to kinetic barriers at the electrode-electrolyte interface—which can increase energy losses but are minimized through catalyst selection.

In Gases and Plasmas

Electric conduction in gases occurs primarily through ionization processes, where neutral gas molecules are transformed into charged particles capable of carrying current. Unlike in solids or liquids, gases are excellent insulators at low voltages, requiring a sufficiently high electric field to initiate breakdown and sustain conduction. This breakdown is governed by Paschen's law, which states that the minimum voltage required for electrical discharge depends solely on the product of gas pressure (p) and electrode gap distance (d), exhibiting a characteristic curve with a minimum value for each gas species. The Paschen minimum breakdown voltage for air is approximately 327 V at a pressure-distance product of about 0.75 torr·cm. At atmospheric pressure, for gaps of ~1 mm, the breakdown voltage is around 3 kV, enabling conduction only above this threshold. The initiation of conduction in gases involves the Townsend avalanche mechanism, where free electrons—often produced by cosmic rays, photoemission, or field emission—accelerate in the electric field, colliding with neutral atoms to ionize them and create additional electrons. This multiplicative process, characterized by the first Townsend ionization coefficient (α), leads to an exponential increase in charge carriers, culminating in a self-sustaining discharge if the secondary emission coefficient (γ) from the cathode provides sufficient feedback. When the avalanche gain exceeds about 10^8, the resulting space charge distorts the field, transitioning to a spark discharge with rapid current rise to amperes or more, as observed in phenomena like lightning where stepped leaders propagate through air plasma channels formed by such avalanches. In controlled settings, this mechanism underpins spark gaps used in high-voltage protection devices. Plasmas, defined as quasi-neutral gases consisting of ionized particles with equal numbers of positive ions and electrons, facilitate conduction through collective motion of these charges. The quasi-neutrality arises because any charge imbalance is screened over the Debye length (λ_D), a characteristic distance beyond which electric fields from individual charges are shielded by surrounding oppositely charged particles, typically on the order of micrometers to millimeters in laboratory plasmas. Plasma conductivity is dominated by electron motion, limited by collisions with ions (electron-ion collisions), which scatter electrons and determine the resistivity via the Spitzer formula, yielding conductivities up to 10^4-10^7 S/m in hot, low-collisionality plasmas—far higher than in neutral gases. In magnetized plasmas, such as those in fusion devices, the Hall parameter (β_H = ω_ce τ_e, where ω_ce is the electron cyclotron frequency and τ_e the collision time) quantifies the degree of magnetization; values β_H >> 1 indicate that electrons gyrate many times between collisions, decoupling their motion from ions and enabling anisotropic transport perpendicular to the field. This is critical in tokamak fusion reactors, where high-temperature plasmas (10-100 keV) achieve near-ideal conductivity to confine currents for magnetic confinement, though neoclassical effects from trapped particles can enhance losses. Glow discharges represent a stable conduction regime in partially ionized gases at low pressures (0.1-10 Torr), commonly used in applications like neon signs, where low-current (mA) operation produces luminous plasma columns. The discharge structure includes the cathode fall region near the cathode, a high-voltage drop (100-300 V) over a short distance (mm) where intense ionization occurs via ion bombardment and electron avalanches, accelerating electrons to energies sufficient for excitation and further ionization. This transitions to the positive column, a longer, quasi-neutral region with uniform low-field (1-10 V/cm) conduction sustained by ambipolar diffusion and volume ionization, emitting characteristic glow from excited species like neon's red-orange light at 585-640 nm. In lightning, similar plasma dynamics create transient high-conductivity channels (10^4 S/m) over kilometers, dissipating gigajoules of energy.

In Vacuum

In vacuum, electric current flows primarily through the emission of charged particles, such as electrons or ions, from a cathode across an evacuated space to an anode, without a conducting medium like gas or solid. This process, known as vacuum conduction, relies on mechanisms that overcome the material's work function—the energy barrier preventing particle escape. Unlike conduction in other media, vacuum currents are non-ohmic and limited by emission rates and space charge effects. Thermionic emission is one key mechanism, where thermal energy from heating the cathode enables electrons to gain sufficient kinetic energy to surmount the work function. The current density J for thermionic emission is described by the Richardson-Dushman equation: J = A T^2 e^{-W / kT} where A is the Richardson constant (approximately 120 A/cm²K²), T is the cathode temperature in Kelvin, W is the work function in electron volts, k is Boltzmann's constant, and e is the base of the natural logarithm. This equation, derived from statistical mechanics applied to the Fermi-Dirac distribution of electrons in metals, predicts emission rates that increase exponentially with temperature, making it suitable for moderate vacuum environments. Field emission, or cold emission, occurs at room temperature under high electric fields that tunnel electrons through the potential barrier via quantum mechanical effects. The Fowler-Nordheim equation governs this process, giving the current density as: J = \frac{A E^2}{\phi} \exp\left( -\frac{B \phi^{3/2}}{E} \right) where E is the applied electric field strength, \phi is the work function, and A and B are constants depending on electron mass and Planck's constant (with B \approx 6.83 \times 10^9 V/m·eV^{3/2}). This tunneling mechanism allows significant currents at fields around 10^9 V/m, without requiring heat, and is prominent in sharp-tip cathodes. These emission processes enable key applications in vacuum devices. Vacuum tubes, such as triodes and tetrodes, utilize thermionic emission to control electron flow for amplification and oscillation in early electronics, with cathodes heated to 800–2500 K to achieve currents up to several amperes. Cathode ray tubes (CRTs) employ thermionic or field emission to generate focused electron beams that strike phosphors for display purposes, as seen in oscilloscopes and televisions. Electron microscopes, including scanning and transmission types, rely on field emission sources for high-brightness, low-energy-spread beams, achieving resolutions below 1 nm by accelerating electrons to 1–300 keV across vacuum chambers. However, the emitted current is often limited by space charge—the electrostatic repulsion among charged particles that reduces the effective accelerating field near the cathode. The Child-Langmuir law quantifies this maximum space-charge-limited current density in a planar vacuum diode: J = \frac{4 \epsilon_0}{9} \sqrt{\frac{2e}{m}} \frac{V^{3/2}}{d^2} where \epsilon_0 is the vacuum permittivity, e and m are the electron charge and mass, V is the anode voltage, and d is the electrode spacing. Originally derived for ion flows but extended to electrons, this law shows current scaling with V^{3/2}, setting fundamental limits in high-current devices like microwave tubes. Ion currents in vacuum also play a crucial role, particularly in mass spectrometry, where positive or negative ions are generated via electron impact or electrospray and accelerated through vacuum paths for separation by mass-to-charge ratio. These beams, typically in the nanoampere to microampere range, travel collision-free in high vacuum (10^{-5} to 10^{-10} Torr), enabling precise isotopic analysis and molecular identification in fields like proteomics and environmental monitoring.

In Superconductors

Superconductivity is a quantum mechanical state in which certain materials allow electric current to flow with zero electrical resistance below a critical temperature T_c, enabling persistent currents that circulate indefinitely without energy dissipation. This phenomenon was first observed by Dutch physicist Heike Kamerlingh Onnes in 1911 during experiments on mercury cooled to near absolute zero using liquid helium, where the resistivity dropped abruptly to zero at approximately 4.2 K. A defining characteristic is the Meissner effect, discovered in 1933 by Walther Meissner and Robert Ochsenfeld, in which superconductors expel external magnetic fields from their interior, behaving as perfect diamagnets and preventing flux penetration up to a critical magnetic field strength. This expulsion arises from induced supercurrents that generate an opposing field, distinguishing superconductivity from mere zero resistance in normal conductors at low temperatures. The microscopic mechanism of superconductivity in conventional materials is described by the BCS theory, developed in 1957 by John Bardeen, Leon Cooper, and John Robert Schrieffer, which posits that electrons form bound pairs known as Cooper pairs through attractive interactions mediated by lattice vibrations (phonons). These pairs condense into a single quantum state, allowing coherent motion without scattering and thus zero resistance. Superconductors are categorized into Type I and Type II based on their response to magnetic fields. Type I superconductors, typically pure metals like lead and mercury, maintain the Meissner state up to a single critical magnetic field B_c (around 0.1 T at low temperatures), beyond which superconductivity is abruptly quenched; they support lower critical current densities J_c. Type II superconductors, such as alloys like NbTi, exhibit two critical fields: a lower field H_{c1} where magnetic flux begins penetrating as quantized vortices, and an upper field H_{c2} (often exceeding 10 T) where the normal state is restored; they tolerate higher J_c values, up to $10^{10} A/m² in optimized materials, making them suitable for practical high-field applications. Key applications leverage these properties for high-efficiency current transport. Superconducting magnets, employing Type II materials cooled by liquid helium, generate intense fields in particle accelerators like the Large Hadron Collider at CERN, where currents exceed 11 kA to bend proton beams at energies up to 7 TeV. Superconducting quantum interference devices (SQUIDs), which exploit interference of supercurrents in loops containing Josephson junctions, provide magnetic sensitivity down to 10^{-15} T, enabling applications in magnetocardiography for non-invasive heart diagnostics and geophysical surveys for mineral exploration. A major advance came with high-temperature cuprate superconductors, discovered in 1986 by J. Georg Bednorz and K. Alex Müller in the Ba-La-Cu-O system with T_c \approx 35 K—far above the previous limit of ~23 K for conventional superconductors—allowing cooling with liquid nitrogen (77 K) and opening paths to room-temperature prospects, though the pairing mechanism remains unconventional and not fully explained by BCS theory. In alternating current (AC) scenarios, superconductors incur AC losses primarily from hysteresis as magnetic flux vortices move under time-varying fields or currents, and from eddy currents in stabilizing matrices like copper; these losses, which scale with frequency and field amplitude, limit efficiency in power transmission cables and motors but can be minimized through filamentary designs in multifilamentary wires. The Josephson effect, theoretically predicted by Brian Josephson in 1962 and experimentally verified shortly thereafter, occurs in junctions formed by two superconductors separated by a thin insulating barrier (typically 1-2 nm), permitting a supercurrent to tunnel without voltage (DC effect) or generating microwave oscillations under applied voltage (AC effect) at frequencies f = 2eV / h \approx 483.6 MHz/μV. This effect underpins SQUIDs, precision voltage standards, and superconducting qubits in quantum computing, where phase coherence enables gate operations with minimal dissipation.

In Semiconductors

In semiconductors, electric current arises primarily from the movement of charge carriers—electrons and holes—across an energy band gap, as described by band theory. The valence band consists of filled electron states at absolute zero, while the conduction band lies above it, separated by a forbidden energy gap E_g typically ranging from 0.1 to 3 eV. Thermal energy or doping can excite electrons from the valence band to the conduction band, leaving behind positively charged holes in the valence band; both electrons in the conduction band and holes in the valence band contribute to conduction under an applied electric field. Intrinsic semiconductors, such as pure silicon or germanium, rely solely on thermal generation of carriers. The intrinsic carrier concentration n_i, which equals both electron density n and hole density p, is given by n_i = \sqrt{N_c N_v} \, e^{-E_g / 2kT}, where N_c and N_v are the effective densities of states in the conduction and valence bands, respectively, E_g is the band gap energy, k is Boltzmann's constant, and T is the absolute temperature. This exponential dependence on temperature means that carrier density, and thus conductivity, increases dramatically with rising temperature in intrinsic materials, often doubling every 10–15 K near room temperature. Extrinsic semiconductors are doped with impurities to enhance conductivity at lower temperatures. In n-type doping, pentavalent atoms like phosphorus introduce donor levels just below the conduction band, releasing extra electrons (n \gg p); in p-type doping, trivalent atoms like boron create acceptor levels above the valence band, generating holes (p \gg n). The total current density J in extrinsic semiconductors is J = n e \mu_e E + p e \mu_h E, where e is the elementary charge, \mu_e and \mu_h are the electron and hole mobilities (typically \mu_e > \mu_h, e.g., 1400 cm²/V·s and 450 cm²/V·s in silicon at 300 K), and E is the electric field. Mobility reflects scattering by phonons, impurities, and defects, decreasing with increasing impurity concentration. A key application of carrier dynamics in semiconductors is the p-n junction, formed by adjoining p-type and n-type regions, which creates a depletion layer with a built-in potential barrier. Under forward bias, the barrier reduces, allowing minority carriers to inject and recombine, yielding exponential current flow described by the Shockley diode equation: I = I_s \left( e^{eV / kT} - 1 \right), where I_s is the reverse saturation current, V is the applied voltage, and the ideality factor is assumed unity for ideal junctions. This behavior enables diodes for rectification and forms the basis for bipolar junction transistors (BJTs), invented in 1947, which amplify or switch currents by controlling base-emitter junctions. Semiconductors also drive photovoltaic applications, such as solar cells, where absorbed photons generate electron-hole pairs, producing a photocurrent proportional to light intensity; crystalline silicon solar cells have achieved lab efficiencies up to 27.8% as of 2025 under standard test conditions, with commercial modules typically around 22-25%. Overall, the temperature-activated nature of semiconductor conduction—contrasting with metals—enables precise control via doping and bias, underpinning modern electronics from integrated circuits to renewable energy systems.

Microscopic Aspects

Drift Velocity

Drift velocity is the average velocity attained by charged particles, such as conduction electrons in a metal, due to the influence of an applied electric field, superimposed on their random thermal motions. In the classical Drude model, this directed motion arises from the balance between acceleration by the electric field and frequent collisions with lattice ions, resulting in a steady-state average velocity. The magnitude of the drift velocity v_d is given by v_d = \frac{e \tau}{m} E, where e is the elementary charge, \tau is the mean time between scattering events (relaxation time), m is the effective mass of the electron, and E is the electric field. Equivalently, v_d = \mu E, where \mu = \frac{e \tau}{m} represents the mobility of the charge carriers, quantifying their responsiveness to the field. The relaxation time \tau, typically on the order of $10^{-14} seconds at room temperature for metals, determines the scale of v_d through scattering processes. This microscopic motion connects to the macroscopic current via the relation I = n e A v_d, where I is the electric current, n is the number density of conduction electrons, A is the cross-sectional area of the conductor, and e is the charge magnitude. For instance, in copper with n \approx 8.5 \times 10^{28} electrons per cubic meter, a current density of 1 A/mm² yields v_d \approx 10^{-4} m/s. In contrast, the random thermal velocities of these electrons, characterized by the root-mean-square speed in the Drude model, reach approximately $10^5 m/s at room temperature, highlighting that the directed drift is a small perturbation on chaotic thermal motion.

Relation to Conductivity

The macroscopic electrical conductivity \sigma of a material arises from the collective motion of charge carriers under an applied electric field, as described by the Drude model, where \sigma = n e \mu and \mu = e \tau / m, leading to \sigma = n e^2 \tau / m with n as the carrier density, e the elementary charge, \tau the relaxation time, and m the effective mass. This expression connects the average drift velocity of carriers to the current density J = \sigma E, highlighting how frequent collisions limit conductivity in classical treatments. The reciprocal of conductivity, resistivity \rho = 1/\sigma = m / (n e^2 \tau), quantifies opposition to current flow and inversely scales with carrier density and relaxation time, explaining why metals exhibit low resistivity due to high n and moderate \tau. In this framework, resistivity increases with decreasing \tau, which governs how quickly carriers lose momentum to scattering events. The Hall effect provides an experimental link between microscopic carrier properties and macroscopic transport, where a magnetic field B perpendicular to current I induces a transverse voltage V_H = I B / (n e d) across a sample of thickness d, allowing measurement of carrier density n and type (electrons or holes) from the sign of V_H. This voltage arises from the Lorentz force deflecting carriers, establishing a steady-state electric field that balances magnetic deflection, and is particularly useful in semiconductors where n varies. While the classical Drude model captures essential conductivity features, quantum mechanics refines it by emphasizing the role of the Fermi level in metals, where only electrons near the Fermi energy E_F contribute to transport due to Pauli exclusion, as deeper electrons lack available states for acceleration. Anomalies like the temperature-independent resistivity at low temperatures or the Wiedemann-Franz law's validity reveal the model's incompleteness, necessitating quantum treatments like the Sommerfeld model for full accuracy. The relaxation time \tau is influenced by temperature and impurities: at high temperatures, phonon scattering dominates, making \tau \propto 1/T and increasing resistivity linearly with temperature in metals; at low temperatures, impurity scattering prevails, rendering \tau nearly constant and resistivity temperature-independent. These effects underscore how material purity and thermal environment tune conductivity through scattering rates.

References

  1. [1]
    Electric Current - The Physics Hypertextbook
    Electric current is defined as the rate at which charge flows through a surface (the cross section of a wire, for example).Missing: authoritative | Show results with:authoritative
  2. [2]
  3. [3]
    Electric current | McGraw Hill's AccessScience
    An electric current is the movement of electrically charged particles through a conducting medium or empty space, resulting in a net transfer of electric charge ...Missing: authoritative | Show results with:authoritative
  4. [4]
    Physics Tutorial: Electric Current
    Current is the rate at which charge flows. Charge will not flow in a circuit unless there is an energy source capable of creating an electric potential ...Missing: authoritative | Show results with:authoritative
  5. [5]
    André-Marie Ampère and the two hundred years of electrodynamics
    Jul 2, 2020 · At the time there was still no discussion of 'electric current' and it was André-Marie Ampère who first clearly distinguished between the ...
  6. [6]
    [PDF] Units & Symbols for Electrical & Electronic Engineers - IET
    In the expression I = 16 mA, I is the quantity symbol for the physical phenomenon of electric current, and 16 is its numerical value in terms of the decimal ...
  7. [7]
    - ampere - BIPM
    The ampere, symbol A, is the SI unit of electric current. It is defined by taking the fixed numerical value of the elementary charge e to be 1.602 176 634 x 10 ...
  8. [8]
    1.9: Sign conventions - Engineering LibreTexts
    Jan 22, 2024 · Current Sign Convention: The assumed direction of current flow is typically from the positive terminal to the negative terminal. This convention ...
  9. [9]
    Electric Current - Summary - The Physics Hypertextbook
    The symbol for current density is J (bold). As a vector, current density has magnitude and direction. By definition, current density is the product of charge ...
  10. [10]
    None
    - **Origin of Positive and Negative Notation**: Benjamin Franklin introduced the terms "positive" and "negative" in electrical theory in 1747, based on his single-fluid theory.
  11. [11]
    3.2: Conventional Current Flow and Electron Flow
    Sep 12, 2021 · Franklin surmised that the “electrical flow” moved from positive to negative. This idea was accepted and became the conventional view. Today we ...Missing: 1747 | Show results with:1747
  12. [12]
    Sign convention for passive components and sources (article) | Khan Academy
    ### Summary of Sign Convention for Passive Components and Sources
  13. [13]
    Kirchhoff's laws (article) | Khan Academy
    ### Summary: Reference Directions and Kirchhoff's Laws in Circuit Analysis
  14. [14]
    Kirchhoff's Voltage Law (KVL) | Electronics Textbook
    Kirchhoff’s Voltage Law (KVL) states: “The algebraic sum of all voltages in a loop must equal zero”.
  15. [15]
    20.5: Alternating Current versus Direct Current - Physics LibreTexts
    Aug 19, 2025 · Direct current (DC) is the flow of electric charge in only one direction. It is the steady state of a constant-voltage circuit.
  16. [16]
  17. [17]
    How is DC Power Generated? - ANS Advanced Network Services
    Nov 14, 2023 · Solar cells, fuel cells, rectifiers, and batteries are some of the most common ways of generating and storing DC power.
  18. [18]
    Alternating Current (AC) vs. Direct Current (DC) - SparkFun Learn
    Both AC and DC describe types of current flow in a circuit. In direct current (DC), the electric charge (current) only flows in one direction.
  19. [19]
    Electrolytic Cell - Electrochemistry - MCAT Content - Jack Westin
    – A direct current (DC) supply: provides the energy necessary to create or discharge the ions in the electrolyte. Electric current is carried by electrons in ...
  20. [20]
    [PDF] Applications of HVDC Technologies - Department of Energy
    Research on converters, switches, and configurations is currently ongoing, and such projects have made steady progress in reducing losses and lowering prices.Missing: electrochemistry | Show results with:electrochemistry
  21. [21]
    Difference DC and AC power| Tech - Matsusada Precision
    Aug 12, 2021 · Therefore, no reactive power is generated in the steady state, allowing for efficient power delivery. A key advantage of DC is its suitability ...
  22. [22]
    Knowledge AC LED vs. DC LED - Global Lighting Forum
    Dec 16, 2023 · LEDs are intrinsically direct current (DC) devices that only pass current in one polarity and are typically driven by DC voltage sources ...
  23. [23]
  24. [24]
    15.2 Simple AC Circuits – University Physics Volume 2
    In this section, we study simple models of ac voltage sources connected to three circuit components: (1) a resistor, (2) a capacitor, and (3) an inductor.Missing: waveform parameters
  25. [25]
    Electricity 101 | Department of Energy
    AC had the advantage of being converted via transformers to higher voltages which, at the time, allowed electricity to be transmitted over long distances at ...
  26. [26]
    Counting Down to the New Ampere | NIST
    Aug 25, 2016 · Dating from the 1820s, it states that the amount of current (I) in a conductor is equal to the voltage (V) divided by the resistance (R): I=V/R.<|control11|><|separator|>
  27. [27]
    Die galvanische Kette - Smithsonian Libraries
    63. Ohm's Law This is the first publication of Ohm's law, that is, the discovery that the current in an electric circuit is directly proportional to the voltage ...
  28. [28]
    9.4 Ohm's Law – University Physics Volume 2 - UCF Pressbooks
    Nonohmic devices do not exhibit a linear relationship between the voltage and the current. One such device is the semiconducting circuit element known as a ...Missing: limitations non-
  29. [29]
    21.1 Resistors in Series and Parallel – College Physics
    Summary. Draw a circuit with resistors in parallel and in series. Calculate the voltage drop of a current across a resistor using Ohm's law.
  30. [30]
    9.5 Electrical Energy and Power – University Physics Volume 2
    Power dissipated by a resistor depends on the square of the current through the resistor and is equal to P = I 2 R = V 2 R . The SI unit for electric power is ...
  31. [31]
    [PDF] Current, continuity equation, resistance, Ohm's law. - MIT
    Feb 24, 2005 · ~J = σ~E . In other words, the current density is directly proportional to the electric field. The constant of proportionality σ is called the ...
  32. [32]
    [PDF] Current and Resistance
    the electron current is. I = dq dt. = edN dt. dV. dV. = neA dx dt where n = dN/dV is the number density. The term vd = dx/dt is referred as the drift speed: it ...<|control11|><|separator|>
  33. [33]
    [PDF] Electric Current
    • The current density J is defined as the current I flowing per unit area ... • This velocity is called the electron drift velocity. Current density.
  34. [34]
    Factors affecting electrical conduction - DoITPoMS
    For pure metals at around room temperature, the resistivity depends linearly on temperature.
  35. [35]
    Electrical Conductivity and Resistivity - NDE-Ed.org
    The conductivity of most materials decreases as temperature increases. Alternately, the resistivity of most material increases with increasing temperature.
  36. [36]
    10.4 Electrical Measuring Instruments – University Physics Volume 2
    An ammeter is placed in series to get the full current flowing through a branch and must have a small resistance to limit its effect on the circuit. Standard ...Missing: moving | Show results with:moving
  37. [37]
    Moving Coil Meters
    An ammeter is an instrument for measuring the electric current in amperes in a branch of an electric circuit. It must be placed in series with the measured ...
  38. [38]
    [PDF] Low Level Measurements Handbook - 7th Edition - Tektronix
    1.1 Introduction. DC voltage, DC current, and resistance are measured most often with digital multimeters (DMMs). Generally, these instruments are adequate ...
  39. [39]
  40. [40]
    Clamp meter measurement principles | HIOKI
    Clamp meters measure current by detecting the magnetic field of flowing current, then converting it to current. CT, Hall element, and Rogowski methods are used.
  41. [41]
    22.6 The Hall Effect – College Physics - UCF Pressbooks
    The creation of a voltage across a current-carrying conductor by a magnetic field is known as the Hall effect, after Edwin Hall, the American physicist who ...<|control11|><|separator|>
  42. [42]
    All In One: NIST Develops Single Device to Realize Electrical ...
    Aug 12, 2025 · The precision voltage standard, known as a programmable Josephson voltage standard, consists of an integrated circuit chip containing pairs of ...
  43. [43]
    Hall Effect Measurements Comments | NIST
    Since the Hall voltage levels that are generated are in the low uV range it suggests that the measurement system is incapable of accurately measuring them.
  44. [44]
    The Ampere and Electrical Standards - PMC - PubMed Central - NIH
    NIST research has helped to develop and refine electrical standards using the quantum Hall effect and the Josephson effect, which are both based on quantum ...
  45. [45]
    Ampere: Introduction | NIST
    May 15, 2018 · The ampere is a measure of the amount of electric charge in motion per unit time ― that is, electric current.
  46. [46]
    SI prefixes - BIPM
    Decimal multiples and submultiples of SI units ; tera. T · 10 ; giga. G · 10 ; mega. M · 10 ; kilo. k. 10 ...
  47. [47]
    Metric (SI) Prefixes | NIST
    Jan 13, 2010 · The simplified table below shows common metric prefixes and the relationship with their place values.
  48. [48]
    Electrical metrology - BIPM
    Electrical metrology uses Josephson voltage and quantum Hall resistance standards. BIPM provides comparison and calibration services for voltage, resistance, ...
  49. [49]
  50. [50]
    Historical perspective: Unit of electric current, ampere - BIPM
    The definition of the ampere, chosen by the CIPM, was referenced to the force between parallel wires carrying an electric current.
  51. [51]
    (PDF) Joule's 1840 manuscript on the production of heat by voltaic ...
    In 1840, James Prescott Joule submitted to the Royal Society a paper describing experimental research on the heat produced by electric currents in metallic ...<|control11|><|separator|>
  52. [52]
  53. [53]
    [PDF] Kinetic Energy of Conduction Electrons in the Drude Model
    Oct 10, 2009 · due to inelastic collisions at frequency f = 1/τ of the conduction electrons with the lattice ... (8) is, of course, the power dissipated by Joule ...Missing: mechanism | Show results with:mechanism
  54. [54]
    [PDF] iv ELECTRONIC AND THERMAL BEHAVIOR OF GERMANIUM ...
    Jul 13, 2023 · When an external field is applied, the charge carriers accelerate in the direction of the current density j until they collide with a lattice ...
  55. [55]
    [PDF] Explain joule's law of heating mathematically - Jimars
    The amount of heat that is produced within an electric wire due to the flow of current is expressed in the unit of Joules. When the current flows through ...
  56. [56]
    [PDF] Tuesday May 8 Problem 1: Inefficiencies in early DC electrical syste
    May 8, 2018 · This loss is known as “Joule heating” or sometimes just “I2R heating”. It is Joule heating that lights the filament of a lightbulb, but also ...
  57. [57]
    [PDF] Non-Uniform Transition Conductivity of Superconducting Ceramic
    a current there is joule heating. For superconducting metals, joule heating effects are minimized by their high thermal conductivity. (1.0 W cm-lK-l) and by ...
  58. [58]
    [PDF] The experiments of Biot and Savart concerning the force exerted by ...
    The best source for the details of both the first and the second Biot–Savart experiments is the third edition (1824) of Biot's textbook, Précis E´lémentaire de ...
  59. [59]
    Solenoids as Magnetic Field Sources - HyperPhysics
    The magnetic field B is proportional to the current I in the coil. The expression is an idealization to an infinite length solenoid, but provides a good ...
  60. [60]
    Magnetic Resonance Imaging - Mathematics and Physics of ... - NCBI
    The heart of the apparatus is the magnet system, typically a superconducting solenoid. The second system is the transmit/receive assembly, typically consisting ...
  61. [61]
    The birth of the electric machines: a commentary on Faraday (1832 ...
    Apr 13, 2015 · The first experiment Faraday discusses in the paper demonstrates simple induction and is worth describing here. First, 26 feet of copper wire ...
  62. [62]
    [PDF] Chapter 10 Faraday's Law of Induction - MIT
    In. 1831, Michael Faraday discovered that, by varying magnetic field with time, an electric field could be generated. The phenomenon is known as electromagnetic ...
  63. [63]
    [PDF] Superconductor and Lenz's Law - Edinburgh Research Explorer
    Feb 27, 2020 · Lenz's law, named after the Russian physicist Heinrich Friedrich Emil Lenz who deduced it in 1834 [1,2], a fundamental law of physics ...
  64. [64]
    The birth of the electric machines: a commentary on Faraday (1832 ...
    Faraday's spinning disc—generating a continuous electric current in a conducting disc as it spins between two poles of a powerful permanent magnet. This ...
  65. [65]
    Mutual Inductance - HyperPhysics
    The mutual inductance M can be defined as the proportionalitiy between the emf generated in coil 2 to the change in current in coil 1 which produced it.
  66. [66]
    Self Inductance – Joseph Henry Project - McGraw Commons
    Using this method of measurement Joseph Henry discovered that uniformly coiled copper wire exhibits a stronger self inductance than straight wire of the same ...
  67. [67]
    Maxwell's equations
    Accelerating charges produce changing electric and magnetic fields. Changing ... Maxwell's equations imply that EM waves in free space are transverse waves.
  68. [68]
    Electromagnetic waves - Physics
    Jul 26, 1999 · An electromagnetic wave can be created by accelerating charges; moving charges back and forth will produce oscillating electric and magnetic fields.
  69. [69]
    16.2 Plane Electromagnetic Waves – University Physics Volume 2
    Accelerating charges create electromagnetic waves (for example, an oscillating current in a wire produces electromagnetic waves with the same frequency as the ...
  70. [70]
    [PDF] Electric dipole radiation and simple antennas - Galileo
    The radiation pattern of the half-wave antenna is similar to the dipole pattern (4.18), and the total power is larger than the simple formula (4.17) by a factor ...
  71. [71]
    Basic Antenna Theory - Richard Fitzpatrick
    Typically, in a Herztian dipole, the radiated power is swamped by ohmic losses that appear as heat. Thus, a ``short'' antenna is a very inefficient radiator.
  72. [72]
    [PDF] Module 6: Antennas - MSU College of Engineering
    Radiation efficiency - The radiation efficiency of an antenna is the ratio of the power radiated by the antenna to the total power supplied to the antenna.
  73. [73]
    Dipole Antennas - AstroBaki - CASPER
    Oct 15, 2012 · Dipole antennas are a type of radio antenna that is very common at lower frequencies, where wavelengths are long enough that such elements are reasonable to ...
  74. [74]
    [PDF] Radio Waves 05/16/2006 Lecture 13
    May 16, 2006 · Radio waves are electromagnetic radiation produced by accelerating charges, including AM/FM, cell phone, and microwave signals, traveling at c.Missing: applications | Show results with:applications
  75. [75]
    24.3 The Electromagnetic Spectrum - UCF Pressbooks
    AM radio waves are used to carry commercial radio signals in the frequency range from 540 to 1600 kHz. The abbreviation AM stands for amplitude modulation, ...Missing: generation | Show results with:generation
  76. [76]
    [PDF] Handout 1 Drude Model for Metals - Cornell University
    1) Metals have a large density of “free electrons” that can move about freely from atom to atom (“sea of electrons”). 2) The electrons move according to.
  77. [77]
    [PDF] 2 Chapter 1 The Drude Theory of Metals - SIUC Physics WWW2
    In this chapter we shall examine the theory of metallic conduction put forth by. P. Drude' at the turn of the century. The successes of the Drude model were con ...
  78. [78]
    [PDF] Current and Resistance
    ρ(T) = ρ0[1+α (T−T0)] where α is the temperature coefficient of resistivity ... For metals α is positive and is nearly constant over a large temperature range ...
  79. [79]
    [PDF] Phys 446 Solid State Physics Lecture 7 - NJIT
    The scattering by impurities is essentially independent of temperature. The number of phonons increases with temperature. ⇒ the scattering by phonons is ...
  80. [80]
    The Behavoir of the Resistivity of Metals at Low and High ... - Stanford
    Jun 16, 2007 · Electron-Phonon Scattering. In many materials, Electron-Phonon scattering is the dominatant source of the temperature-dependant resistivity.
  81. [81]
    Resistivity and Temperature Coefficient at 20 C - HyperPhysics
    Resistivity and Temperature Coefficient at 20 C ; Material. Resistivity ρ (ohm m) ; Silver. 1.59. x10 ; Copper. 1.68. x10 ; Copper. 1.724. x10 ; Copper, annealed ...Missing: dependence ρ_0 T)
  82. [82]
    Skin Effect in AC Conduction - HyperPhysics
    Skin effect in AC conduction causes current to concentrate on the outside of a wire, with 63% flowing within one skin depth and 98% within 4 skin depths.
  83. [83]
    [PDF] EDDY CURRENTS, DIFFUSION, AND SKIN EFFECT
    This decrease of the current amplitude with depth is called the skin effect: a high-frequency AC current flows only near the surface of the conductor, and δ — ...
  84. [84]
    [PDF] Physics, Chapter 28: Electrical Conduction in Liquids and Solids
    compared to a metallic conductor such as copper whose conductivity is about 108 mhos/m. The electrical conductivity of different liquids is shown in Table ...
  85. [85]
    Faraday's Laws - Student Academic Success - Monash University
    Faraday's first law states the mass of a substance deposited at an electrodeA solid conductor that allows the transfer of electrons and serves as the site of ...
  86. [86]
    [PDF] Electrical Conductivity of Rocks, Crust, Mantle, and Core
    Generally, conductivity in a material is given by σ = nqµ where q is the charge of the mobile species, n is the concentration of the charged species, and µ is ...
  87. [87]
    Electrolytic Cells
    Electrolysis of aqueous NaCl solutions gives a mixture of hydrogen and chlorine gas and an aqueous sodium hydroxide solution.
  88. [88]
    [PDF] Principles and implementations of electrolysis systems for water ...
    Feb 12, 2016 · The total kinetic overpotential of an electrolyzer is thus the sum of the overpotentials at each of the electrodes of the electrolysis cell.
  89. [89]
    [PDF] SPACECRAFT HIGH-VOLTAGE PASCHEN AND CORONA DESIGN ...
    Dec 29, 2020 · Paschen's Law dictates that for all gases, the function V = f (pd) has a typical form with a distinct minimum. Gas breakdown voltage has ...
  90. [90]
    [PDF] Electrical breakdown from macro to micro/nano scales
    Feb 7, 2020 · Paschen's law indicates that under a uniform electric field, for various discharge gaps with different values of gas pressure and gap distance, ...
  91. [91]
    [PDF] Fundamentals of Undervoltage Breakdown Through the Townsend ...
    The voltage required to sustain a gas discharge is always lower than the voltage required to form it, so a stable conducting state can be reached at a moderate ...
  92. [92]
    [PDF] The Plasma Nature of Lightning Channels and the Resulting ...
    Aug 19, 2019 · Lightning channels are made of plasma, which are filaments of ionized air. The current changes the channel's resistance in a nonlinear fashion.
  93. [93]
    [PDF] Chapter 3 Collisions in Plasmas
    Ion momentum loss to electrons is much lower collision frequency than e → i because ions possess so much more momentum for the same velocity. 3.4.3 i → i. Ion- ...
  94. [94]
    [PDF] Chapter 20: Particle Kinetics of Plasma [version 1220.1.K]
    ... ions shield out the electric field of a charge in a plasma (Debye shielding), and oscillations of a plasma's electrons relative to its ions (plasma oscillations) ...
  95. [95]
    [PDF] The Influence of the Hall Term on the Development of Magnetized ...
    Apr 6, 2018 · This latter parameter, the resistive Hall or magnetization parameter Ωce/νei = Ωceτei, can reach ∼ 100 in many strongly magnetized HED plasmas, ...
  96. [96]
    Creating loops of liquid lithium for fusion temperature control
    Jul 23, 2024 · Flowing liquid lithium over a series of slats could offer an ideal solution for drawing excess heat off of a fusing plasma.
  97. [97]
    [PDF] Hollow Cathode Discharges - Analytical Applications
    This proves that the positive column is not essential for maintaining the low pressure glow dis- charge, while the part of the discharge at the cathode,.
  98. [98]
    Secondary electron emission and glow discharge properties of ... - NIH
    The voltage drop of region b occurs in the body of the discharge plasma called the positive column. The sustaining voltage is the sum of the cathode fall ...
  99. [99]
    [PDF] Thermionic Phenomena and the Laws which Govern Them
    In its broadest aspect this subject may be summarized as the branch of Phys- ics which deals with the effect of heat on the interaction between electricity.
  100. [100]
    Electron emission in intense electric fields - Journals
    Electron emission in intense fields is extracted from cold metals, with current independent of temperature at ordinary temperatures, and blends into thermionic ...
  101. [101]
    Electron-emission materials: Advances, applications, and models
    Jul 10, 2017 · Thermionic emitters are used in a variety of applications, including high-frequency vacuum transistors for electronics, electron guns for ...
  102. [102]
    [PDF] Chapter 6 Field Emission - OSTI.GOV
    Deviations from the. Fowler-Nordheim model for total current have been observed in specific samples because of adsorbates and possibly quantized energy levels ...Missing: primary | Show results with:primary
  103. [103]
    Mass Spectrometry Basics - Jeol USA
    The ion source creates electrically charged particles (“ions”) from the atoms and molecules in the sample. There are many different kinds of ion sources – some ...
  104. [104]
    Heike Kamerlingh Onnes – Facts - NobelPrize.org
    In 1911 Kamerlingh Onnes discovered that the electrical resistance of mercury completely disappeared at temperatures a few degrees above absolute zero. The ...
  105. [105]
    Walther Meissner - Magnet Academy - National MagLab
    Walther Meissner discovered while working with Robert Ochsenfeld that superconductors expel relatively weak magnetic fields from their interior and are strongly ...
  106. [106]
    Theory of Superconductivity | Phys. Rev.
    A theory of superconductivity is presented, based on the fact that the interaction between electrons resulting from virtual exchange of phonons is attractive.
  107. [107]
    Type I vs Type II - DoITPoMS
    “Type I” superconductors have a sharp transition from the superconducting state where all magnetic flux is expelled to the normal state. Type II superconductors ...
  108. [108]
    [PDF] Theoretical derivation of critical current density and critical magnetic ...
    Aug 4, 2022 · However, a typical superconductor has three critical points: a critical temperature, a critical magnetic field, and a critical current density.
  109. [109]
    [PDF] Applications of Superconductivity - Princeton University
    Superconductors are used in magnets that bend and focus the particle beam, as well as in detectors that separate the collision fragments in the target area. ( ...
  110. [110]
    Superconducting quantum interference device instruments and ...
    Oct 11, 2006 · Superconducting quantum interference devices. SQUID uses Josephson effect phenomena to measure extremely small variations in magnetic flux. ...
  111. [111]
    ac Losses in Superconductors - AIP Publishing
    In this review, theories and models as well as experiments are discussed, and brief mention is made of actual and proposed ac applications.<|separator|>
  112. [112]
    Brian D. Josephson – Facts - NobelPrize.org
    In 1962 Brian Josephson predicted unexpected ... The 14 laureates' work and discoveries range from quantum tunnelling to promoting democratic rights.
  113. [113]
    What are Josephson junctions? How do they work?
    Nov 24, 1997 · A Josephson junction is made by sandwiching a thin layer of a nonsuperconducting material between two layers of superconducting material.
  114. [114]
    [PDF] Electrons and Holes in Semiconductors
    For simplicity, we can assume that all shallow donors and acceptors are ionized. EXAMPLE 1–6 Complete Ionization of the Dopant Atoms. In a silicon sample doped ...
  115. [115]
    [PDF] Intrinsic carrier concentration in semiconductors - Galileo
    Intrinsic carrier concentration (ni) is the number of electrons in the conduction band and holes in the valence band per unit volume in a semiconductor free of ...
  116. [116]
    [PDF] Motion and Recombination of Electrons and Holes
    When there is more than one scattering mechanism, the total scattering rate. (1/τ) and therefore the total mobility are determined by the sum of the inverses.
  117. [117]
    [PDF] Lecture 3 Introduction to Semiconductors and Energy Bandgaps
    Extrinsic Semiconductor: A doped semiconductor. Many electrical properties controlled by the dopants, not the intrinsic semiconductor. Donor: An impurity ...
  118. [118]
    Semiconductor Applications in Society
    Semiconductor applications range from transistors and microchips to solar cells and LEDs ... Photovoltaic (solar) cells are used to provide electrical power ...
  119. [119]
    [PDF] Temperature Dependence of Semiconductor Conductivity
    The approximate temperature dependence of mobility due to lattice scattering is T-3/2 , while the temperature dependence of mobility due to impurity scattering ...
  120. [120]
    [PDF] Lecture 4: London's Equations Drude Model of Conductivity
    Sep 20, 2005 · The conduction electrons are modeled as a gas of particles with no coulomb repulsion (screening). 2. Independent Electron Approximation.
  121. [121]
    [PDF] The Hall Effect
    Thus, the mobility can be found from a Hall effect experiment: show that µ = σRH. For metals, the exact value of the charge carrier density and the sign of the ...
  122. [122]
    [PDF] Classic Hall Effect primer for Quantum Hall Effect Experiment
    The relevant Hall Effect equation is: is current, B magnetic field, n current density of carrier electrons, e is electron charge, t is sample thickness.Missing: formula | Show results with:formula
  123. [123]
    [PDF] Free Electron Fermi Gas (Kittel Ch. 6) - SMU Physics
    • For typical metals the Fermi energy temperature is much greater than ordinary temperatures – transition from f(E)=1 to f(E)=0 is sharp at room temperature ...
  124. [124]
    [PDF] MSE 351-1: Introductory Physics of Materials
    Apr 6, 2025 · The main limitation is that the Drude model can only explain metals. ... small in metals and a lot smaller than the Fermi level. Therefore ...